ArticlePDF AvailableLiterature Review

Abstract

Commensal microorganisms (the microbiota) live on all the surface barriers of our body and are particularly abundant and diverse in the distal gut. The microbiota and its larger host represent a metaorganism in which the cross talk between microbes and host cells is necessary for health, survival, and regulation of physiological functions locally, at the barrier level, and systemically. The ancestral molecular and cellular mechanisms stemming from the earliest interactions between prokaryotes and eukaryotes have evolved to mediate microbe-dependent host physiology and tissue homeostasis, including innate and adaptive resistance to infections and tissue damage. Mostly because of its effects on metabolism, cellular proliferation, inflammation, and immunity, the microbiota regulates cancer at the level of predisposing conditions, initiation, genetic instability, susceptibility to host immune response, progression, comorbidity, and response to therapy. Here, we review the mechanisms underlying the interaction of the microbiota with cancer and the evidence suggesting that the microbiota could be targeted to improve therapy while attenuating adverse reactions. Expected final online publication date for the Annual Review of Immunology Volume 35 is April 26, 2017. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
R
E
V
I
E
W
S
I
N
A
D
V
A
N
C
E
Microbes and Cancer
Amiran Dzutsev,1,2 Jonathan H. Badger,1
Ernesto Perez-Chanona,1Soumen Roy,1
Rosalba Salcedo,1Carolyne K. Smith,1
and Giorgio Trinchieri1
1Cancer and Inflammation Program, Center for Cancer Research, National Cancer Institute,
National Institutes of Health, Bethesda, Maryland 20892, email: trinchig@mail.nih.gov
2Leidos Biomedical Research, Frederick, Maryland 21702
Annu. Rev. Immunol. 2017. 35:199–228
The Annual Review of Immunology is online at
immunol.annualreviews.org
This article’s doi:
10.1146/annurev-immunol-051116-052133
This is a work of the U.S. Government and is not
subject to copyright protection in the United
States.
Keywords
microbiota, cancer, carcinogenesis, metabolism and cancer, cachexia,
chemotherapy, immunotherapy
Abstract
Commensal microorganisms (the microbiota) live on all the surface barriers
of our body and are particularly abundant and diverse in the distal gut. The
microbiota and its larger host represent a metaorganism in which the cross
talk between microbes and host cells is necessary for health, survival, and
regulation of physiological functions locally, at the barrier level, and sys-
temically. The ancestral molecular and cellular mechanisms stemming from
the earliest interactions between prokaryotes and eukaryotes have evolved to
mediate microbe-dependent host physiology and tissue homeostasis, includ-
ing innate and adaptive resistance to infections and tissue damage. Mostly
because of its effects on metabolism, cellular proliferation, inflammation,
and immunity, the microbiota regulates cancer at the level of predisposing
conditions, initiation, genetic instability, susceptibility to host immune re-
sponse, progression, comorbidity, and response to therapy. Here, we review
the mechanisms underlying the interaction of the microbiota with cancer and
the evidence suggesting that the microbiota could be targeted to improve
therapy while attenuating adverse reactions.
199
Review in Advance first posted online
on January 30, 2017. (Changes may
still occur before final publication
online and in print.)
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
INTRODUCTION
Microbial communities have been living on the surface of the Earth for three-fourths of its his-
tory. They have developed a diversity of specialized lineages to adapt to different habitats and have
been instrumental in shaping the evolution of modern life (1). When eukaryotes and multicellular
organisms appeared over a billion years ago, their close interaction with microbes necessitated
their coevolution, establishing persistent relationships including mutualism, commensalism, par-
asitism and a dependency on each other for survival and control of homeostasis; this history is
clearly written in the genomes of eukaryotic hosts and associated microbiota. The phagocytic ca-
pability of eukaryotic cells was preserved from the most primitive organisms to the highest forms
of animal life and was required for both nutrition and defense against pathogens. Phagocytosis
was probably the key trait that eventually gave rise to eukaryotes, by enabling the acquisition
of bacterial endosymbionts that evolved into mitochondria (2, 3). Eventually, mobile phagocytic
cells in higher organisms developed as specialized cell types of the myeloid lineage (granulocytes,
monocytes, macrophages, dendritic cells) that play a central role in wound repair, innate resistance
to infections, and promotion of adaptive immunity (4, 5). Commensal microorganisms, including
eubacteria, archaea, protists, fungi, and viruses, inhabit all the epithelial barrier surfaces of our
body, where bacteria, in particular, are as numerous as human cells (6–8). The unique microbial
genes in our body outnumber human genes by a factor of at least 100, although many microbial
genes are functionally redundant (6). Thus, we are symbionts or metaorganisms composed of host
and microbial cells (9–11). The microbial genome is an integral part of the metaorganism genetic
framework (metagenome), and the wealth of metabolic processes and products (metabolome) it en-
codes profoundly influence all the physiological functions of the body and alter its pathology. Both
microbial and human cells act as sensors for chemical, physical, and biological environmental clues.
Through reciprocal communications they detect homeostatic changes and modify their compo-
sition and/or function accordingly (12). The signaling pathways by which human and microbial
cells in the metaorganism communicate involve small molecule metabolites, nucleic acids, and
physical protein-protein interactions. This cross talk includes molecular and cellular mechanisms
that are also key to innate resistance mediated by myeloid/phagocytic cell (5). Both commensal
microbiota and many pathogens produce “danger molecules” recognized by the immune system
to elicit a reaction. However, tolerance to microbiota is maintained, despite a substantial translo-
cation of microbial products and cells in a steady state, until pathogenic microorganisms breach
the epithelial barriers and then an immune response is mounted (13, 14).
METHODS OF MICROBIAL COMPOSITION IDENTIFICATION
AND FUNCTIONAL ANALYSIS
Elucidation of the interaction between the host and commensal microbiota has been hampered,
until recently, by the fact that only a small fraction of microbial cells could be isolated, cultured in
vitro, and analyzed (15). Although the leading laboratories have made great progress in culturing
previously uncultivable bacterial species (16), the real revolution in microbiota studies over the last
few years has been due to the widespread availability of next-generation sequencing technology.
This has allowed many investigators to analyze the composition of the commensal microbiota in
an unbiased way by sequencing all or part of their genome. The standard method of identifying
bacterial and archaeal taxa in the microbiota is to sequence one or more variable regions of
the gene encoding 16S rRNA (17), although the optimal region may vary by body site. Fungal
composition analysis relies on sequencing of the internal transcribed spacer (ITS) region between
the 18S and 28S rRNAs (18). Analysis of the protist fraction of the microbiota is in its early
stages, but 18S rRNA (which also has variable regions, as in 16S) appears to be the preferred
200 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
method (19). Alternatively, the microbial metagenome can be sequenced using standard shotgun
sequencing protocols. A further degree of information is available with metatranscriptomic data,
where not only the functional potential of the metagenome is revealed, but also the genes that are
actively expressed. Both metatranscriptomic and metagenomic information could be used to infer
relationships with host transcriptomes or proteomes by building trans-kingdom networks (20).
MICROBIAL CONTRIBUTION TO HOST HOMEOSTASIS
In the metaorganism, cross talk between the commensal microbes and the host is essential for main-
tenance of physiological homeostasis, response to environmental changes and survival. Evidence
accumulating in the last few years suggests that the composition of the microbiota at the epithelial
barrier affects systemic functions, including metabolism, energy balance, central nervous system
physiology including cognitive functions, cardiovascular functions, nutrition, circadian rhythm,
inflammation, innate resistance, and adaptive immunity (21–23). Microbes reside along the gas-
trointestinal tract, with the largest microbial population in the colon, ranging between 1013 to
1014 bacteria (24). Approximately 1012 bacteria live on the skin in various communities depending
on factors such as sebum or dryness of the environment (25). The composition of the microbiota
at various anatomical sites is controlled by host genetics, particularly by the polymorphisms in
immune-related genes, as well as by environmental factors, such as lifestyle and nutrition (26).
The initial composition of a newborn’s microbiota largely resembles either the vaginal microbiota
or, if the method of delivery was cesarean section, the skin and environmental microbiota (27).
The gut microbiota matures during the first three years of life. Concurrently, large changes in the
expression of genes related to vitamin biosynthesis and metabolism are also observed. Through-
out adulthood, the microbiota is relatively stable, although disease onset, use of antibiotics, and
changes in alimentation affect its composition (28). In individuals older than 60–70 years, the com-
position of the microbiota gradually changes; a striking negative association is observed between
frailty in elderly individuals and microbial diversity (29, 30).
Many of the effects of the microbiota occur at local barrier sites. For example, in the gut,
the microbiota modulates nutrient absorption, synthesis of vitamins, metabolism of bile and hor-
mones, and fermentation of carbohydrates. In addition, it contributes to barrier fortification and
the establishment of mucosal immunity (22, 31, 32). However, the microbiota also exerts systemic
effects, including modulation of metabolism, inflammation, and immunity (32). The microbiota
trains myeloid and lymphoid cells by providing instructive signals to regulate epigenetic mod-
ifications that calibrate the immune response to inflammatory stimuli, infections, vaccines, and
autoimmunity (33–36). In addition to the gut mucosa, commensals at other epithelial barriers
such as the skin also control local immune homeostasis, protective responses, and tissue pathology
(32). A compartmentalized control of immunity by the skin microbiota has been clearly defined
(37). Although many of the cellular and molecular mechanisms mediated by vicinal microbiota
on barrier homeostasis have been identified, the exact method of systemic influence is yet to be
determined. Traditionally the epithelial barriers and their immune system have been considered
to prevent in healthy organisms the translocation of bacteria into internal tissues. However, it is
now becoming evident that bacteria can penetrate the gut mucosa and skin, diffuse to the draining
lymph nodes, and disseminate systemically with protective rather than pathogenic effects (13, 38,
39). This translocation is amplified in conditions of immune deficiency or during diseases that
compromise the integrity of the epithelium such as colitis or atopic dermatitis (38, 39). Other
mechanisms also contribute to the systemic effects conferred by the microbiota: diffusion of bac-
terial products, host growth factors, cytokines, chemokine-induced migration of barrier immune
cells, and release of vesicles or exosomes from bacteria or immune cells.
www.annualreviews.org Microbes and Cancer 201
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
The effect of the microbiota on innate resistance and immunity is, in part, dependent on its effect
on hematopoiesis. Germfree mice are deficient in bone marrow–derived peripheral myeloid pop-
ulations (40). Recolonization of these mice with microbes provides cues from microbe-associated
molecular patterns (MAMPs) and short-chain fatty acids (SCFAs) to promote hematopoietic ho-
meostasis by activating innate immune receptors (40–42). The maturation of the yolk sac–derived
CNS microglia and their ability to respond to infections was also defective in germ-free mice. This
deficiency was similarly reproduced by treating conventionally raised mice with antibiotics, show-
ing that a complex microbiota is required to maintain the functions of tissue-resident myeloid cells
(43). The functional maturation of circulating neutrophils and their ability to form extracellular
traps is also regulated by the microbiota (44). Diurnal fluctuations of circulating inflammatory
monocytes are regulated by synchronous oscillations of gut microbial composition and through
the level of expression of Toll-like receptors (TLRs) on gut epithelial cells, which control the
expression of the circadian gene Bmal1 (45, 46). Alteration of the circadian clock (e.g., due to jet
lag) induces pathological alteration to the microbiota, termed dysbiosis, promoting a metabolic
syndrome that is transferable by fecal transplantation to germfree mice (47).
Bacterial products can induce expression of type II interferon (IFN-γ), which is known to
affect neutrophil survival and activation (48, 49). This suggests that activated barrier myeloid
cells, among others, can reenter circulation and contribute to systemic microbiota effects (50).
Interleukin-12 (IL-12) produced during intestinal Toxoplasma gondii infection affects monocyte
precursor differentiation in the bone marrow by inducing IFN-γproduction by natural killer
(NK) cells (51). Low concentrations of MAMPs, such as lipopolysaccharide, can also change
myeloid cell activation thresholds for an enhanced inflammatory response, also known as priming
(52). These mechanisms likely contribute to the microbiota’s influence on the responsiveness
of peripheral and tumor-infiltrating myeloid cells (34, 53) and to the generalized Shwartzman
reaction, which also depends on the priming of myeloid cells by low doses of lipopolysaccharide
and IL-12 (54).
MICROBIOTA ASSOCIATION WITH CANCER
Although tumors grow locally before invading other tissues or metastasizing to other parts of the
organism, cancer should be considered a systemic disease affected by and altering the homeostatic
mechanisms that control the physiology of the metaorganism. The growth characteristics of the
tumor cells are dependent on the genetic alterations resulting in the activation of oncogenes, the
silencing of tumor suppressor genes, and other classical hallmarks of cancer (55). However, it is
becoming evident that transformed cells cannot effectively grow without a suitable microenviron-
ment (56). The microbiota, with its ability to modulate the metaorganism’s physiology, contributes
to the establishment of a pro- or antitumor inflammatory milieu (Figure 1). Two findings exem-
plify the requirement of a suitable microenvironment for tumor growth. Rous sarcoma virus, the
first-discovered oncovirus, induces tumors in adult birds at the site of injection or injury, but not
in sterile embryos. This is true even if the infected cells in the embryo express the v-Src oncogene
and show a transformed phenotype in vitro (57). The other example is the finding that more than
one-quarter of the skin cells in aged, sun-exposed eyelid carry clonally expressed cancer-causing
driver mutations similar to those found in squamous cell carcinoma, yet they maintain the physi-
ological differentiation and functions of normal skin without evolving into cancer (58). Although
evidence for the impact of the skin microbiota on the malignant progression of keratinocyte car-
cinomas is still lacking, normal tissue homeostasis and architecture are known to restrain cancer
so that changes in the microenvironment are required for tumor progression (59). Indeed, the
microbiota affects tissue metabolism as well as the differentiation and function of immune cells
202 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
Comorbidity
Cachexia Toxic side eects
of therapy
Antibiotics
Xenobiotics
Probiotics,
prebiotics,
and symbiotics
Lifestyle
Host genetics
Fecal transplantation
Microbiota
Dysbiosis
Diet
Cancer
initiation and progression
Immunosuppression
Inammation
Genotoxicity
Anticancer therapy
Chemotherapy
Immunotherapy
Cell-based
therapy
Radiation
therapy
Lifetime
exposure to
commensals
Metabolic
disorder
Dened microbiota
transfer
Myeloid cells,
innate lymphocytes,
T cells, and B cells
Myeloid cells,
innate lymphocytes,
T cells, and B cells
Myeloid cells,
innate lymphocytes,
T cells, and B cells
Figure 1
Microbiota-dependent regulation of cancer development, progression and treatment. The microbiota plays a
dynamic role, from maintenance of healthy host physiology to disease development. Physiological factors
(blue) such as lifetime exposure to commensals, diet, lifestyle, and host genetics shape the composition of the
microbiota in each individual, which affects response to disease and therapy. Under pathogenic conditions,
changes in microbiota composition (dysbiosis; yellow) may contribute to disease. The microbiota regulates
cancer initiation and progression, comorbidity, and anticancer therapy in part by priming myeloid cells and
(directly or indirectly) lymphoid cells that mediate innate resistance and adaptive immunity. Treatments
targeting microbiota composition ( green), such as probiotics, prebiotics, symbiotics, antibiotics, xenobiotics,
transplantation of fecal or defined microbiota, have potential to modulate cancer progression, cancer-
associated comorbidity, response to therapy, and adverse reactions.
www.annualreviews.org Microbes and Cancer 203
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
in the tumor microenvironment, which may foster skin carcinogenesis. In turn, host genetics and
environmental factors may also affect the tumor microenvironment by modifying the composition
and diversity of the microbiota (12, 26). The incidence of keratinocyte cancer and cancer at other
barrier surfaces exposed to the microbiota is elevated in immunosuppressed recipients of organ
transplants (60), likely owing to changes in the microbiota composition at these sites and defective
tumor immunosurveillance.
Epidemiological cancer studies based on the analysis of oral, fecal, and tissue samples to evaluate
the role of the microbiota and dysbiosis have mostly been limited to gastrointestinal and lung
carcinoma (61). These studies have confirmed the role in stomach cancer of Helicobacter pylori,
the only bacterial species recognized as a group 1 human carcinogen by the International Agency
for Research on Cancer (IARC) (62, 63). These studies also identified other emerging candidate
oncogenic bacteria, such as Fusobacterium, enterotoxigenic Bacteroides fragilis (ETBF), and pks+
strains of Escherichia coli (64–67). Traditional epidemiological data showing an association of the
presence of certain species with cancer incidence provide important clues but are in many cases
difficult to interpret because evidence of causality is often lacking. The presence of dysbiosis or
shifts in the abundance of specific microbes may, in fact, be a consequence rather than a cause of
cancer. The microorganisms responsible for cancer initiation may no longer be present when the
patients are analyzed, and dysbiosis may cause homeostatic alterations or epigenetic effects that
contribute to cell transformation and tumor promotion (68, 69). Definitive evidence for the role of
particular species in cancer pathogenesis will come from more molecular-based epidemiological
studies as well as identification of the mechanisms involved in both clinical studies and experimental
animal studies.
ROLE OF THE MICROBIOTA IN DIET- AND OBESITY-ASSOCIATED
CANCER AND CANCER COMORBIDITY
High body mass index (BMI) is a risk factor at the population level for diabetes, cardiovascular
disease, and many common cancers, including colon, liver, gallbladder, postmenopausal breast,
uterus, and kidney cancers (70). High BMI is associated with the pathological inflammation respon-
sible for insulin resistance and altered energy metabolism (71). Colon cancer and gut dysbiosis,
too, are linked to obesity. High-fat diet (HFD) increases tumorigenicity of intestinal cell pre-
cursors by activating peroxisome proliferator-activated receptor delta (72). In hepatocarcinoma,
carcinogenesis is associated with Clostridium cluster XI– and cluster XIVa–induced production of
the secondary bile acid product deoxycholic acid, which induces DNA damage and senescence
in hepatic stellate cells (73, 74). Although it has been estimated for over 30 years that one-third
of all cancers and the majority of gastrointestinal cancers may be due to dietary factors (75),
the epidemiological evidence associating HFD with cancer is modest, unlike the experimental
data demonstrating that HFD causes cardiovascular disease and diabetes (74, 76). Conversely,
high-fiber diets increase the microbe-driven metabolism of SCFAs, such as butyrate, which have
anti-inflammatory properties. In addition, they may protect against gastrointestinal cancers and
lymphoma through the modulation of epigenetic changes by inhibiting histone deacetylases (74,
77). Exposure of female mice to HFD during pregnancy induced obesogenic and diabetogenic
traits in two generations of offspring and was associated with the development of lung and liver
cancers (69, 78). This phenotype was linked to epigenetic changes in genes encoding adiponectin
and leptin, and administration of a normal diet abolished it in three generations. Cancer suscep-
tibility was also induced in recipients of the microbiota from HFD-fed females but was reversed
by treatment with the beneficial probiotic species Lactobacillus reuteri (69, 78). The use of antibi-
otics can persistently reduce bacterial diversity, deplete protective niches that otherwise prevent
204 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
overgrowth of pathogens and pathobionts, and induce gut dysbiosis that may increase the risk of
disease (79, 80). The repeated use of antibiotics in young children resulted in an increased risk of
childhood obesity (81) and up to sevenfold increase in the risk of developing inflammatory bowel
disease (82); thus, it may also increase the risk of colitis-associated cancer. In breast cancer, a
meta-analysis of five case-control studies showed that antibiotic use was associated with a slightly
elevated risk of breast cancer, though this association is controversial (83). Mammary tissue is not
sterile and as the bacterial composition changes in the presence of cancer, the local microbiota
modulates antitumor immunity in the tumor microenvironment (84). The long-term effects of
antibiotics and diets such as HFD and high fiber diet, however, alter the microbial composition
and confound the identification of specific cancer-causing microbes.
Exposure to microbes during natural birth may help the newborn establish a proinflammatory
environment that favors tumor immunosurveillance (85, 86). Data from murine studies have
provided insight into the balance of microbial exposure and immune cell activity. One study
showed that acute intestinal infection induced intolerance to commensals because of dysbiosis-
driven alterations in mucosal integrity and immunity (14, 87). However, this resulted in the
accumulation of Th1-biased, microbe-specific memory T cells that also facilitated protective
responses to subsequent infections and tumor growth (14).
Whereas obesity may initiate certain cancers, anorexia-cachexia syndrome is a cancer-associated
comorbidity observed in a large majority of patients with late-stage, advanced cancer. This wast-
ing of muscle and adipose tissue is also triggered by infection, kidney and intestinal damage,
and chemotherapy (88). It dramatically affects quality of life, decreases the efficacy and tolerabil-
ity of anticancer treatments, and is the predominant cause of cancer-associated comorbidity and
mortality (88). Cancer-associated cachexia is linked to systemic inflammation and characterized
by elevated levels of chemokines and cytokines such as IL-1βand IL-6 (89). Cancer cells in-
duce cachexia through the secretion of proinflammatory cytokines and factors such as parathyroid
hormone–related protein (90). Different microbial species dominate in patients with metabolic
dysfunctions: the archaeon Methanobrevibacter smithii in patients with anorexia nervosa and Lacto-
bacillus spp. in the obese patients (91). Indeed, mice treated with the probiotic L. reuteri, prebiotics,
or both combined (symbiotics) were protected against cancer-associated cachexia, likely because
of enhanced anti-inflammatory, mucosa-protective effects (92, 93). In experimental models of in-
testinal damage or of infection-induced cachexia, E. coli O21:H+, present in the gut, protected
against cachexia (94). E. coli colonized white adipose tissue, where it activated the NLRC4 inflam-
masome and induced production of insulin-like growth factor 1, a hormone that prevents muscle
degradation (94). Changes in the microbial composition of the gut may be a cause or consequence
of cancer cachexia. Regardless, multimodal strategies, including the use of beneficial microbes,
aimed at fortifying barrier function may help reduce the incidence of cachexia (95). Therefore,
further investigation using experimental models may shed light on the role of the gut microbiota
in cancer anorexia/cachexia.
HUMAN CARCINOGENIC MICROORGANISMS
Infection-induced cancer accounts for approximately 16% of the global burden of all human can-
cers, corresponding to approximately 2 million new cases per year (96). The frequency varies by
region, with lower percentages, on average, in more developed countries (7.4%) compared to less
developed countries (22.9%) (96). The IARC classifies 10 microbial agents (7 viruses, 2 parasites,
and 1 bacterium) as group 1 human carcinogens; i.e., their status as carcinogens is based on either
strong evidence in humans, or limited evidence in humans supported by strong evidence in ani-
mals (97). H. pylori, hepatitis B and C viruses (HBV, HCV), and human papillomaviruses (HPV)
www.annualreviews.org Microbes and Cancer 205
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
together are responsible for over 90% of all infection-attributed cancers (96). The herpesvirus
Epstein-Barr virus (EBV) is the causative agent of Burkitt lymphoma, nasopharyngeal carcinoma,
and a subset of gastric carcinomas, whereas the Kaposi sarcoma–associated herpesvirus (HSV8)
causes Kaposi sarcoma and other pathologies in immunosuppressed or elderly individuals (98).
HBV and HCV are associated with hepatocellular carcinoma (HCC) (99). High-risk oncogenic
strains of HPV (HPV16, HPV18, and 11 others) are associated with anogenital cancers, a subset
of head and neck cancers, and skin cancers (100, 101). Human T cell lymphotropic virus type 1
is associated with the T cell lymphomas prevalent in certain geographical regions (102). Merkel
cell polyomavirus is the first human virus discovered by metagenomic sequencing and is associ-
ated with the majority of cases of Merkel cell carcinoma, an aggressive skin cancer observed in
immunosuppressed individuals (103).
With the exception of HCV, all human oncogenic viruses encode at least one oncogene and
may directly induce cell transformation (104). However, factors such as infection-associated geni-
tal tract inflammation and vaginal dysbiosis likely play a role in cancer progression even for viruses
such as HPV that have a potent transforming ability (105). HBV and HCV establish chronic liver
infection and account for over 80% of HCCs (106). In some patients, the immune response to
chronic viral infection progresses to fibrosis and cirrhosis and ultimately HCC. Whereas the initial
innate response to HBV infection is modest, HCV actively evades innate responses by inhibiting
both the production and the signaling of type I and type III interferon (106, 107). Unlike HCV,
HBV may directly transform hepatocytes. However, for both viruses, the pathogenesis of HCC
is dependent on immune-related inflammation (106). A higher degree of chronic HBV and liver
pathology correlates with greater gastrointestinal richness of Candida spp. and Saccharomyces cere-
visiae and with a less abundance and diversity of Bifidobacterium spp. (108, 109). The role of the
gut microbiota in regulating liver pathology and progression to HCC in mice has been clearly
documented (110). Interestingly, young mice, similarly to neonates or young children, fail to
clear HBV infection in a hydrodynamic transfection model until an adult-like gut microbiota is
established (111). Oncogene-carrying viruses (HPVs and HBV) require inflammation to promote
tumorigenesis, but this may indeed apply to all oncogenic viruses (112), including the previously
mentioned Rous sarcoma virus (57). As the limited data hitherto published suggest that the com-
mensal microbiota may be involved in regulating the response to infection and progression to
cancer, these represent possible targets for preventive and therapeutic interventions for cancer.
Gastric infection with H. pylori is strongly associated with noncardiac gastric carcinoma and
lymphoma (62). Yet, approximately half of the world population is infected with H. pylori,and
in most cases this infection only develops into well-tolerated gastritis. Only rarely does H. pylori
infection progress to serious lesions such as atrophy, metaplasia, and cancer (62). The virulence and
carcinogenicity of various strains of H. pylori have been associated with the variable expression of
two cytotoxin-encoding genes: cytotoxin-associated gene A (cagA) and vacuolating cytotoxin gene
A(vacA) (113). H. pylori infection, before any damage to the gastric mucosa occurs, cooperates with
the gut microbiota to control energy homeostasis by affecting circulating metabolic gut hormones
(114). H. pylori infection has profound effects on the host immune response. It activates the TLR4
and TLR2 receptors as well as the NLRP3 inflammasome, thus inducing the secretion of both
IL-1βand IL-18. This promotes the activation of both Th1 cell and regulatory T cell (Treg)
responses to protect against asthma, chronic inflammatory diseases, and tuberculosis (115, 116).
Whereas H. pylori directly affects gastric epithelial cells and can compromise genetic integrity by
inducing DNA damage, gastric carcinogenesis requires exposure to the bacterium over multiple
decades, with an initial inflammatory response, epithelium injury and atrophy, reduction in acid
secretory functions, and intestinal metaplasia (117, 118). Often, H. pylori is no longer detectable in
the stomach of seropositive individuals with atrophic body gastritis, suggesting that deterioration
206 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
of the gastric niche and the lowering of acidity due to long-term H. pylori infection cause gastric
dysbiosis dominated by the presence of cancer-provoking species of oropharyngeal or intestinal
origin (119). In developed countries, the incidence of H. pylori infection is decreasing owing to
frequent use of antibiotics, improved hygiene, and eradication protocols using broad-spectrum
antibiotics and proton pump inhibitors (120). Although this decrease correlates with a lower
incidence of gastric inflammation and carcinogenesis, the eradication protocols cause perturbation
of the gut microbiota with possible side effects (120). Also, the absence of H. pylori may remove some
of the beneficial effects of the infection, thereby increasing susceptibility to asthma and obesity
as well as gastroesophageal reflux with increased risk of esophageal and gastric cardia carcinoma
(63). This hypothesis, however, has been challenged by the observation that the incidence of
these pathologies is not increased in certain ethnic Malaysian populations that have a low natural
incidence of H. pylori infection and generally poor sanitation (85). Thus, these findings may suggest
that the relationship between the absence of H. pylori infection and the increased incidence of these
pathologies is not universal and that, in some populations, H. pylori infection may be a marker of
poor hygiene that has a protective effect on asthma, obesity, and esophageal carcinoma (85).
MICROBIOTA AND CANCER AT THE EPITHELIAL BARRIERS
At the epithelial barrier surfaces, the primary contact point between host commensals, the com-
position of the microbiota or the abundance of particular species affects both inflammation and
immunity, as well as the homeostasis of epithelial and stromal cells (22, 121). Although this cross
talk occurs at all barriers (32), it is particularly evident at the level of the lower gastrointesti-
nal tract, where the most bacteria are present (24) (Figure 2). Direct interactions of bacterial
structural components and their metabolites, for example, hydrogen sulfide and p-cresol, with
epithelial, stromal, and hematopoietic cells may have direct genotoxic effects and promote cancer
progression (122–126).
Mice deficient in genes controlling host-microbe cross talk have been extensively used for
studying the mechanisms by which the microbiota affects cancer. In particular, mice deficient in
immunologically relevant genes, such as Tlr5,Il10,Tbx1,andRag2, are not only susceptible to
colitis and colon carcinogenesis, but are also characterized by gut dysbiosis. This susceptibility to
cancer can be transferred to healthy mice by cohousing, fostering, or fecal transplant (65, 127,
128). These findings are relevant to humans because polymorphisms in immunologically relevant
genes affect human microbiota composition and cancer predisposition (26).
Clinical and epidemiological investigations as well as experimental studies in animal models
have already identified bacterial species putatively involved in carcinogenesis on the basis of either
their physical association with the neoplastic lesions or a positive correlation of their abundance
with cancer risk. Mechanisms of bacteria-mediated carcinogenesis have been well characterized
in mice using several microbial species, including Enterococcus faecalis,Streptococcus gallolyticus,en-
teropathogenic E. coli,ETBF,Helicobacter hepaticus,Salmonella enterica,andFusobacterium nucleatum
(64, 65, 129). The bacterial genera Odoribacter and Akkermansia are reportedly enriched in colon
cancer–bearing mice (130), and archaea, such as the order Methanobacteriales, have been found
in the fecal microbiota of patients with colorectal cancer (131).
Fusobacterium spp. are anaerobic, gram-negative bacteria that usually reside in the oropharynx,
where they are involved in oral pathology and participate in the formation of dental biofilms
(132). However, they are also found in the inflamed colon and are particularly enriched in human
colonic adenomas and adenocarcinomas (64, 129). The latter observation may be explained by
the binding of the F. nucleatum protein Fap2 to the carbohydrate moiety Gal-GalNac that is
overexpressed on colonic tumor cells (133). F. nucleatum is immunosuppressive, owing to its
www.annualreviews.org Microbes and Cancer 207
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
ability to recruit tumor-promoting myeloid cells to the intestine of APCMin/+mice and to inhibit
human NK and T cell activity via binding of its Fap2 protein to the TIGIT inhibitory receptor
(64, 134). It also activates β-catenin/Wnt signaling in epithelial cells by the association of its FadA
adhesin to E-cadherin (129). The level of expression of the gene encoding FadA is much higher in
colon tissues of patients with colon neoplasia compared to those of healthy individuals, suggesting
FadA as a potential diagnostic and therapeutic target (129). Fusobacteria have also been identified
IL-22
K
+
K
+
FadA
Fap2
GPR43
GPR109a
Il6, Tnf
pro-IL-18
IL-18
Macrophage
IL-17
Th17 ILC3
CTL
TIGIT
NK
IL-12
TNF IL-23
IL-22BP
SCFAs
VacA
CagA
Colibactin
ROS
IL-6
IL-11
RTK
PI3K/ AKT
MAPK/ ERK
E2F TCF
FadA
FadA
Ca
++
Il18
Treg
Chemokines
NF-ΚB
pSTAT 3
NF-ΚB
pSTAT 3
B-Cat
DCs
Neutrophils
H. pylori
S. enterica
E. coli
H. hepaticus
B. fragilis
Fusobacterium
NLRP6
NLRP3
Recruitment
of myeloid
suppressor cells
E-cadherin
E-cadherin
Xenobiotics
p-cresol
Deoxycholic acid
Clostridium cluster XIVa
Clostridium cluster XIVa
MAMPs
PRRs
MAMPs
PRRs
SCFAs
p53
MSH2
MLH1
ZRANB3
β-Catenin
β-Catenin
CDTH2S
Cag A
AvrA
BFT
208 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
in the biofilms observed in carcinomas of the ascending colon and the tumor-free mucosa of
the same patients (135). The biofilms show upregulation of the polyamine metabolite N1,N12-
diacetylspermine, which is likely responsible for the observed enhanced epithelial proliferation,
diminished E-cadherin, and activation of STAT3 and IL-6 (135, 136).
Patients with colitis and colon cancer also have an increased abundance of E. coli (67). A direct
role in mouse carcinogenesis was demonstrated for strains of E. coli expressing the pks pathogenic-
ity island, which encodes the genotoxin colibactin (67). Both induction of inflammation and a
direct effect of the pks+E. coli are needed for carcinogenesis (67). Colibactin induces alterations
in p53 SUMOylation, inducing cellular senescence associated with the production of growth fac-
tors leading to tumor-promoting effects (137). E. coli along with several other microbial species,
including Campylobacter jejuni,Aggregatibacter actinomycetemcomitans,Haemophilus ducreyi,Shigella
dysenteriae,Helicobacter hepaticus,andS. enterica, also has another type of genotoxic compound
that belongs to the family of cytolethal distending toxins (CDTs). CDT consists of three sub-
units: CdtA, CdtB, and CdtC. Its CdtB subunit has similarities with DNAse-I-like nucleases and
demonstrates potent DNAse activity that causes extensive DNA lesions and apoptosis in target
cells (138). Attaching and effacing E. coli has also been shown, similarly to attaching and effacing
H. pylori, to downregulate the key DNA mismatch proteins MSH2 and MLH1, which are also
mutated in hereditary nonpolyposis colorectal cancer (139, 140). In addition, enteropathogenic
E. coli utilizes another pathway to inhibit DNA repair via the secretory cysteine methyltransferase
NleE, which blocks the function of DNA annealing helicase and endonuclease ZRANB3 (141).
ETBF is a subclass of the human commensal B. fragilis. It is present in the intestine at relatively
low abundance; however, it can act as a pathobiont by causing diarrhea and has been associated
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
Figure 2
The role of prokaryotic microbes in cancer initiation and progression. (Far left) Microbial metabolites with direct and indirect
genotoxic activity include products of protein (e.g., H2S, p-cresol) and bile (e.g., deoxycholic acid) degradation as well as products of the
breakdown of liver-detoxified xenobiotics (e.g., sulfation and glucuronidation conjugates). A number of bacterial species can also
produce toxins that are injected into host cells via the type III secretion system. These induce cell toxicity via apoptosis and repression
of proton pump expression. Among the toxins that induce DNA damage are colibactin, produced by some strains of E. coli;and
cytolethal distending toxins (CDTs), produced by Escherichia coli,Campylobacter jejuni,Helicobacter hepaticus,andSalmonella enterica.In
addition to directly damaging DNA, some bacteria, including H. pylori and E. coli, downregulate genome stability–related and repair-
related genes (e.g., TP53, MSH2, MLH1, and ZRANB3). Bacterial species that translocate through the epithelial barrier induce
recruitment of myeloid cells, including reactive oxygen species (ROS)–producing neutrophils that contribute to DNA damage in the
host cell. (Center left) FadA protein from Fusobacterium nucleatum, CagA toxin from H. pylori,Bacteroides fragilis toxin (BFT), and
avirulence protein A (AvrA) from S. enterica Typhi can activate the β-catenin pathway by promoting detachment of β-catenin from
E-cadherin. Some of the same bacterial species, such as H. pylori and S. enterica, also activate the PI3K/AKT and MAPK/ERK pathways
via receptor tyrosine kinases. (Center and far right) Cytokines, such as IL-22, IL-11, and IL-6, promote development of colon cancer via
activation of STAT3. The intestinal microbiota and transmucosally translocated bacterial species, following mucosal damage or tumor
growth, regulate the production of many cytokines—such as IL-6, IL-11, IL-12, IL-18, IL-23, and TNF production by macrophages,
dendritic cells (DCs), and epithelial and mesenchymal cells and IL-22 and IL-17 production by T cells and innate lymphocytes. The
IL-18/IL-22 axis is particularly important in maintaining mucosal homeostasis, and it is tightly regulated. Microbial products induce
epithelial cell production of pro-IL-18, which is cleaved into active IL-18 by inflammasomes. These inflammasomes (NLRP3 and/or
NLRP6) are activated by bacteria-derived small-chain fatty acids (SCFAs) via GPR43 and GPR109a receptors and by commensal
protists. IL-18 then blocks macrophage production of the soluble IL-22 antagonist IL-22BP, and it is required for IL-22 production by
T cells and innate lymphoid cells, thus increasing production and bioavailability of IL-22. IL-22 also induces STAT3 phosphorylation
in epithelial cells, promoting proliferation and secretion of antibacterial peptides, and, in a positive feedback loop, enhances production
of IL-18. Bacteria also activate immunosuppressive mechanisms: F. nucleatum promotes accumulation of immunosuppressive myeloid
cells in colonic tumors and produces the Fap2 protein, which activates the inhibitory receptor TIGIT on natural killer (NK) and T
cells; SCFAs induce regulatory T cells (Tregs), which inhibit local immune response. Abbreviations: CTL, cytotoxic T lymphocyte;
MAMP, microbe-associated molecular pattern.
www.annualreviews.org Microbes and Cancer 209
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
with inflammatory bowel disease and colorectal cancer (142). B. fragilis toxin stimulates intestinal
epithelial cell proliferation, which depends in part on E-cadherin degradation and β-catenin
activation. ETBF induces mucosa permeabilization and STAT3 activation in both inflammatory
and epithelial cells, leading to colon carcinogenesis in APCMin/+mice, which is dependent on
simultaneous expansion of both Tregs and Th17 cells (143, 144).
In mouse colon carcinogenesis, the polyp-containing epithelium is more permeable than the
contiguous healthy tissue and allows transmucosal bacterial translocation (145). The translocated
bacteria induce the production of proinflammatory IL-6, IL-11, IL-23, IL-17, and IL-22, which
are required for cancer progression (145, 146).
Another pro-inflammatory cytokine, IL-18, is a potent inducer of IFN-γproduction and, when
combined with IL-12, results in type 1 polarization of both innate and adaptive lymphocytes. How-
ever, paradoxically, results in IL-18-deficient mice suggest that in the large intestine this cytokine
has anti-inflammatory properties and mediates mucosal protective mechanisms (147). Mice defi-
cient in IL-18, IL-18R, and MyD88 as well as those deficient for inflammasome-related genes,
which are unable to process pro-IL-18, are more susceptible to dextran sulfate sodium (DSS)-
induced colitis; azoxymethane (AOM)/DSS-induced, colitis-associated cancer; and diet-induced,
nonalcoholic steatohepatitis (148). These mice also have gut dysbiosis characterized by increased
abundance of the phyla Bacteroidetes (Prevotellaceae) and TM7 and greater susceptibility to colon
carcinogenesis that can be transferred to wild-type mice by cohousing or by fecal transplant (128).
Commensal bacteria and protists induce production of IL-18 in intestinal epithelial cells by tran-
scriptionally inducing the production of pro-IL-18, which is then cleaved into biologically active
IL-18 by inflammasomes (149, 150). Bacteria-induced SCFAs activate the epithelial cell inflam-
masome via the metabolite-sensing receptors GPR43 and GPR109A (149). SCFAs also act on
macrophages, dendritic cells, and T cells, in particular, causing the expansion of IL-10-producing
Tregs that limit colonic inflammation and carcinogenesis (151, 152). Whether NLRP3 and/or
NLRP6 are involved in IL-18 induction remains controversial and diametric results have been
published by different laboratories (153). The discordant data could be explained by variations in
the composition of the microbiota in different animal facilities and by the differential distribution
of the two inflammasomes in epithelial and hematopoietic cells.
The ability of IL-18 to protect the mucosa depends on its regulation of IL-22 production and
availability (154–156). IL-22 is produced by T cells and innate lymphoid cells in the intestinal
lamina propria and, through activation of STAT3, induces epithelial cell proliferation and produc-
tion of antibacterial peptides (155). Three regulatory mechanisms have been proposed: (a)IL-18
inhibits IL-22 binding protein (IL-22BP) production by macrophages; (b) IL-18 is required for
IL-22 production by T lymphocytes and innate lymphoid cells; and (c) in a positive feedback
loop, IL-22 induces pro-IL-18 but not its processing by inflammasomes in intestinal epithelial
cells (154–156). In different experimental models of colon carcinogenesis, IL-22, which promotes
mucosa repair, has been observed to be pro- or anticarcinogenic depending on the extent of mu-
cosal damage in the carcinogenic model (154). The paradoxical ability of the proinflammatory
cytokine IL-18 to protect from colitis and colon carcinogenesis (154) may be explained by a re-
cent report showing that IL-18 protects against mucosal inflammation by maintaining a protective
microbiota (157). This article reports that IL-18 signaling in intestinal epithelial cells prevented
mucus-forming goblet cells from maturing, and mice that were deficient for IL-18 or IL-18R in
intestinal epithelial cells and that were cohoused with wild-type mice to equalize the microbiota
composition were resistant to DSS-induced colitis (157).
In several experimental models, SCFAs acting through the GPR109A receptors are anti-
inflammatory and decrease incidence of colon and mammary cancer (152, 158). Some bacterial
species contribute more than others to SCFAs levels, such as the butyrate-producing members of
210 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
Clostridium cluster XIVa (159). SCFAs, particularly butyrate and propionate, are also inhibitors
of histone deacetylases and downregulate expression of proinflammatory genes such as IL-6
and tumor necrosis factor (TNF) (160). However, reducing the production of SCFAs using
either antibiotics or a carbohydrate-poor diet decreases incidence of colonic polyps in ApcMin/+/
Msh2 /mice (161). This procarcinogenic effect of SCFA-producing bacteria was not dependent
on inflammation or DNA damage but on induction of hyperproliferation and aberrant β-catenin
signaling in Msh2 /cells (161).
Although the microbiota is present on all barrier surfaces, scarce and mostly correlative data
are available for the role of the local microbiota in cancer development outside the gastrointestinal
tract. More studies deciphering the contribution of the microbiota to cancers of the skin, oropha-
ryngeal cavity, lung, and urogenital tract are needed. Also, relatively little attention has been
devoted to microorganisms other than bacteria and viruses, although production of carcinogenic
acetaldehyde by fungi has been proposed to play a role in oral cancer in patients with a mutation
in the AIRE gene and with autoimmune polyendocrinopathy-candidiasis-ectodermal dystrophy
(162). Additionally, other members of our microbiota could also contribute to cancer, including
protists and helminths, which have been shown to orchestrate gut immunity (150, 163).
The studies described here have allowed investigators to identify many bacterial species with
carcinogenic activity; however, with the exception of H. pylori, none of them have been formally
proven to be a human carcinogen, for example, by disease prevention upon their elimination from
the host (65).
EFFECTS OF THE GUT MICROBIOTA ON TUMORS
IN DISTANT ORGANS
The microbiota regulates cancer not only at the epithelial barriers that it inhabits but also systemi-
cally at distant sterile sites. One example is the regulation by the gut microbiota of metabolism and
energy balance, which contributes to cancer-predisposing conditions such as obesity and metabolic
disease. Bacteria may also modulate distant malignancies such as endometrial and breast cancer
through a noninflammatory pathway by expressing β-glucuronidases and β-glucuronides that
participate in estrogen metabolism. Thus, the use of antibiotics and dysbiosis may alter estrogen
metabolism and affect the progression of these tumors (119).
The microbiota also metabolizes xenobiotics by transforming heterocyclic aromatic amines
from burned meat or environmental pollutants into genotoxic and procarcinogenic metabo-
lites (164). The colonic procarcinogen AOM is processed by P450 enzymes in the liver and
in the small intestine into carcinogenic methylazoxymethanol (MAM) that also gets transformed
in the liver by UDP-glucuronosyltransferase into inactive MAM-glucuronide. Accumulation of
carcinogenic AOM products in the colon and formation of DNA adducts require gut micro-
bial β-glucuronidase (165) that converts inactive MAM-glucuronide back into active methyla-
zoxymethanol. Colon carcinogenesis induced by AOM is prevented by prebiotics that reduce the
number of β-glucuronidase-expressing bacteria, or by inhibitors of the enzyme (166, 167). Sys-
temic effects of the microbiota can also be inflammatory. In mice where gut dysbiosis was induced
by fungal Candida sp. and by T. gondii infection, the lung macrophages and bone marrow mono-
cyte precursors, respectively, were polarized into cells with antiallergic and anti-inflammatory
characteristics (51, 168).
There are many examples of systemic regulation of carcinogenesis by intestinal microbes in
experimental mouse cancer models. Colonic infection with Helicobacter hepaticus has contrasting
effects on carcinogenesis in the intestine and at distant organs. Colonic H. hepaticus infection
increases ileal and colonic carcinogenesis in APCmin/+mice and in AOM-treated Rag2/mice by
www.annualreviews.org Microbes and Cancer 211
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
inducing IL-22 production by innate lymphoid cells (155, 169). It also promotes mammary and
prostate carcinoma in APCmin/+/Rag2/mice (124, 170) and enhances chemical and viral liver
carcinogenesis (124, 171). However, in Rag2-sufficient animals H. hepaticus infection protects
against intestinal and mammary carcinogenesis by inducing IL-10-producing Tregs (172). In a
different model of mammary carcinogenesis based on expression of SV40 T antigen under the
control of sex steroid hormones, infection with H. hepaticus increases tumor multiplicity, affecting
migration of neutrophils into the mammary gland (173).
Variation in intestinal microbes between different animal facilities or as a consequence of
experimental perturbations profoundly affects incidence of lymphoma and survival of Atm (ataxia
telangiectasia mutated)-deficient mice (174). The gut microbiota contributes to the incidence of
thymic lymphoma in these animals by modulating the TNF-regulated inflammatory tone and by
inducing oxidative stress as well as genotoxicity in leukocytes and epithelial cells (170, 171, 174,
175). In mice TLR5-mediated recognition of commensal microbiota induces production of IL-6
and recruitment of myeloid suppressor cells, enhancing progression of malignant tumors at distal
extramucosal locations (176).
BACTERIA AS ANTICANCER TREATMENT
Following the personal and historical observations of tumor regression associated with acute bac-
terial infections, at the end of the nineteenth century, the New York surgeon William Coley suc-
cessfully treated soft tissue sarcoma patients with “Coley’s toxins,” a combination of Streptococcus
pyogenes and gram-negative Bacillus prodigiosus (Serratia marcescens), providing evidence that a severe
localized infection may induce a systemic antitumor immune response (177). Subsequently, several
bacteria or bacterial preparations, such as Corynebacterium parvum and the streptococcal prepa-
ration OK-432, were tested in cancer therapy; local treatment with bacille Calmette Gu´
erin, an
attenuated strain of Mycobacterium bovis, is still a first-line therapy for superficial bladder carcinoma
(178). Many genera of bacteria, including Salmonella,Escherichia,andClostridium, preferentially
accumulate in tumors when delivered systemically, and they have been tested as anticancer agents
(179). A major limitation of cancer therapies is that large tumors contain hypoxic, necrotic, and
quiescent regions in which the tumor cells do not proliferate. These regions are either inaccessible
to drugs or, because of hypoxia, poorly susceptible to DNA damage induced by chemotherapy or
radiation (179). Because obligate anaerobic bacteria such as Clostridium spp. can only proliferate
in hypoxic regions of the tumors, when they are delivered as spores they germinate and prolifer-
ate with antitumor effects only in the hypoxic tissues (180). Although the use of spores could be
considered for large hypoxic tumors, small tumors or metastases are better oxygenated and might
be better treated with facultative anaerobes, such as Salmonella and Escherichia, that are attracted
by small molecules released by tumors (181). The ideal bacterial therapeutic would have the fol-
lowing properties: toxic for tumor cells but not normal cells, selective for the tumor (particularly
regions not targeted by conventional therapies), susceptible to immunological clearance but not
immediately destroyed by the immune response, proliferative, and genetically modifiable (181).
Furthermore, the therapeutic bacteria should be mobile and allow genetic alteration to reduce
systemic toxicity and localization in the tumors (182, 183). For example, deletion in Salmonella of
the Trg gene that encodes a sugar-sensing transmembrane receptor allows the bacteria to deeply
penetrate within poorly vascularized and nutrient-poor regions of the tumor (183). Bacterial
therapeutics can be used for delivery of antitumor molecules such as toxin, cytokines, antigens,
and antibodies, as well as for transfer of genes encoding molecules with antitumor activity (179,
182). The advantages of bacterial cancer therapies include not only their efficacy when other
212 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
therapies fail but also their tumor tropism and ability to proliferate and maintain favorable phar-
macokinetics for an extended period of time (179).
CANCER CHEMOTHERAPY AND THE MICROBIOTA
The microbiota regulates the response to different types of cancer chemotherapy by affecting their
pharmacokinetics, mechanism of action, and toxicity (184, 185) (Figure 3). Intestinal host and
bacterial enzymes control the bioavailability of many oral drugs (186). Exposure to xenobiotics,
including chemotherapy agents, alters the composition, physiology, and gene expression of the
human gut microbiota (187). The gut microbiota also metabolizes injected drugs after biliary excre-
tion and reabsorption (188). For example, tissue carboxylesterase transforms Irinotecan (CPT-11;
an intravenous topoisomerase I inhibitor used for colorectal cancer treatment) into its active form,
SN-38, which is then detoxified in the liver by UDP-glucuronosyltransferases, becoming inactive
SN-38-G before being secreted in the gut (188). When bacterial β-glucuronidase reconverts SN-
38-G in the gut into SN-38, intestinal toxicity and diarrhea are observed (189). β-Glucuronidase
activity in the human gut is mostly associated with abundance of Firmicutes, particularly within
clostridial clusters XIVa and IV (190). Intestinal inflammation induced by CPT-11 therapy can
be successfully treated with antibiotics to decrease the abundance of β-glucuronidase-positive
bacteria or with bacterial β-glucuronidase-specific inhibitors (191).
Platinum-based anticancer drugs inhibit DNA replication by forming intrastrand platinum-
DNA adducts and double-stranded breaks (192). In addition to inducing tumor cell cytotoxicity
and apoptosis, platinum drugs cause severe intestinal toxicity, nephrotoxicity, and peripheral neu-
ropathy that compromise the quality of life (193–195). In germfree mice or in mice depleted of
gut commensals using nonabsorbable broad-spectrum antibiotics, the antitumor efficacy of ox-
aliplatin or cisplatin is dramatically decreased (53). In the absence of commensal microbiota the
drugs still reach the tumor and form platinum-DNA adducts, but DNA damage is severely atten-
uated (53). The microbiota is necessary for training tumor-infiltrating myeloid cells to produce
reactive oxygen species (ROS) via NADPH oxidase 2 (NOX2), which is necessary for platinum
compound–induced DNA damage (53). Mice deficient for the Cybb gene encoding the gp91phox
chain of NOX2 and mice treated with antibodies depleting myeloid cells are poorly responsive
to oxaliplatin (53). Although ROS were known to be involved in the induction of DNA damage
and apoptosis of the tumor cells, the source of ROS was expected to be autocrine (196). However,
observations in microbiota-depleted mice are more compatible with paracrine production of ROS
by NOX2-expressing and ROS-producing tumor infiltrating myeloid cells (53). Administration
of Lactobacillus acidophilus to antibiotics-treated mice restores cisplatin antitumor activity (197).
How precisely the gut commensals and L. acidophilus prime myeloid cells for ROS production in
response to platinum drugs is still unclear. Interestingly, L. acidophilus also attenuates intestinal
toxicity in patients treated with both radiotherapy and cisplatin, suggesting that this probiotic
may enhance the antitumor effect while preventing adverse reactions (197, 198). Acetovanillone,
an inhibitor of NADPH oxidases, protects mice from cisplatin nephrotoxicity by preventing tox-
icity of both the ROS produced by kidney tubular cells soon after treatment and the late ROS
produced by infiltrating myeloid cells (199). Not surprisingly, these observations indicate that the
antitumor effect and the toxicity of platinum compounds are modulated in a similar way by gut
commensals. In devising procedures targeting the microbiota to improve therapeutic efficacy while
limiting toxicity, it will be important to determine how the drugs and the microbiota intersect in
mediating toxicity in tumors and organs and to identify the roles and mechanisms of action of dif-
ferent commensals. Treatment with prebiotics, probiotics, and symbiotics could prevent dysbiosis
www.annualreviews.org Microbes and Cancer 213
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
after chemotherapy and reduce inflammation in the gut and liver, although data from stringent
clinical trials are needed (200).
Antitumor treatment with the alkylating agent cyclophosphamide (CTX) rapidly damages
the mouse gut epithelial barrier by increasing mucosal permeability, resulting in transmucosal
translocation of gram-positive gut bacteria, such as Lactobacillus johnsonii,Lactobacillus murinus,
and Enterococcus hirae, into the mesenteric lymph nodes (201). Within a week, CTX treatment
also alters the composition of commensals in the small intestine of mice, similar to what ob-
served in cancer patients (201, 202). CTX treatment selectively reduces Firmicutes (Clostridium
cluster XIVa,Roseburia,Coprococcus and unclassified Lachnospiraceae) and Spirochaetes (particularly
Untreated CpG-ODN/
Anti-IL-10R Anti-PD-L1
CTX
Anti-CTLA-4 TBI/
Adoptive T cell
transfer
TREATMENTS Cisplatin
Oxaliplatin
Hemorrhagic
necrosis Promoting
antitumor
immunity
Licensing of
T cell–
mediated
killing
Genotoxicity
and cell
death
T cell–
mediated
tumor killing
Immuno-
genic cell
death and
T cell–
mediated
tumor cell
clearance
CELLULAR
RESPONSE TO
ANTICANCER
THERAPY
INTESTINAL
EPITHELIUM
GUT
MICROBIOTA
TUMORS
CTLA-4
T cells
T cells ROS
Transferred
CTLs
PD1/PD-L1
TNF
Monocyte
Dendritic cell
Macrophage
Neutrophils DNA damage
TLR4
TLR4
MyD88
MyD88
Anti-CTLA-4
Anti-CTLA-4
CpG-ODN/
Anti-IL-10R
CpG-ODN/
Anti-IL-10R
Anti-PD-L1
Anti-PD-L1
Total body
irradiation
Total body
irradiation
Cisplatin
Oxaliplatin
Cisplatin
Oxaliplatin
CTX
CTX
No
treatment
No
treatment
L. acidophilus
L. acidophilus
Gram-negative
bacteria
Gram-negative
bacteria
L. johnsonii
L. murinus
E. hirae
B. intestinihominis
L. reuteri
Ruminococcus
A. shahii
L. murinum
L. intestinalis
L. fermentum
Burkholderiales
B. thetaiotaomicron
B. fragilis
Burkholderiales
B. thetaiotaomicron
B. fragilis
B. breve
B. longum
B. adolescentis
TLR9
TLR4
TLR4
THERAPEUTIC
OUTCOME
12
3
4
5
67
pTh17 cells
CTLs
Th1 cells
pTh17 cells
CTLs
Th1 cells
214 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
Treponema genus) phyla in the small intestinal mucosa while it enhances gram-positive bacteria,
mainly L. johnsonii,L. murinus,E. hirae,andL. reuteri, some of which translocate into mesen-
teric lymph nodes (201). CTX induces immunogenic tumor cell death that, in concert with the
translocated gram-positive bacteria and the accumulation of the gram-negative Barnesiella intes-
tinihominis in the colon, activates pathogenic T helper 17 (pTh17) cells and memory Th1 immune
responses that mediate an antitumor adaptive immune response (201, 203). The pTh17 responses
and the antitumor effect of CTX treatment are reduced in germfree mice and in mice depleted
of gram-positive bacteria by antibiotics (201). The antitumor effect of CTX can be restored in
microbiota-depleted mice by adoptive transfer of pTh17 cells (201).
IMMUNOTHERAPY AND THE GUT MICROBIOTA
Immunotherapy is one of the greatest successes of cancer research (204). Enduring, complete
responses have been obtained in patients with metastatic melanoma and lung cancer even when
previous treatments failed. However, the response in different patients and cell types has varied.
New evidence that the composition of the gut microbiota modulates the response to immunother-
apy opens new possibilities of improving outcomes (53, 205, 206).
An early study showed that in mice preconditioned with total body irradiation (TBI) before
receiving adoptive T cell therapy, the antitumor effect was dependent on the presence of gut
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
Figure 3
Gut microbiota in cancer therapy. Gut commensal microorganisms regulate complex cellular networks,
enhancing (red names) or attenuating (blue names) the efficacy of cancer treatments. Some of the mechanisms
by which the microbiota affects the various anticancer therapies (bottom) are depicted here:
() An intact, healthy intestinal epithelial cell layer and mucous layer are an efficient barrier for
luminal commensal bacteria. Together with well-regulated interactions between the intestinal innate
and adaptive immune system, they maintain mucosal homeostasis. Some but not all cancer therapies
damage mucosal integrity and allow transmucosal translocation of commensal bacteria.
() Intratumoral treatment with the immunostimulating TLR9 agonist CpG-ODN combined with in-
hibition of IL-10 signaling (anti-IL10R antibodies) induces rapid tumor hemorrhagic necrosis mediated
by TLR9-induced TNF production. The gut microbiota primes (via a TLR4-dependent mechanism)
tumor-infiltrating myeloid cells to produce TNF in response to CpG-ODN.
() Immunotherapy with anti-CTLA-4 induces mucosal damage and translocation of Burkholderi-
ales and Bacteriodales, which promote anticommensal immunity that acts as an adjuvant for antitumor
immunity and is required for inducing a positive response to therapy.
() The antitumor effect of anti-PD-L1 therapy, which does not damage the gut epithelia, requires
preexisting antitumor immunity that is particularly effective in mice harboring intestinal Bifidobacterium
spp.
() Preconditioning total body irradiation (TBI) enhances the efficacy of adoptive T cell therapy
by inducing mucosal damage, which allows translocation of gram-negative commensals that activate
dendritic cells via TLR4 signaling, augmenting proliferation and cytotoxic functions of the transferred
T cells in the tumor microenvironment.
() Platinum-based drugs, including cisplatin and oxaliplatin, cause DNA damage in tumor cells that
is dependent on both the formation of platinum-DNA adducts and the production of NADPH-oxidase
dependent reactive oxygen species (ROS) by tumor-infiltrating myeloid cells that have been primed in
a MyD88-dependent way by components of the commensal microbiota.
() Cyclophosphamide (CTX) therapy induces immunologic cell death of tumor cells, which elicits
the generation of antitumor pathogenic Th17 (pTh17) cells, Th1 cells, and cytotoxic T lymphocytes
(CTLs), leading to tumor destruction. Optimal generation of this antitumor immune response requires
the activation of tumor-antigen-presenting dendritic cells by components of the intestinal microbiota
that translocate following CTX-induced mucosal damage.
www.annualreviews.org Microbes and Cancer 215
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
commensals (207). TBI damaged the intestinal mucosa, inducing translocation of bacteria and
activation of dendritic cells via TLR4 activation. This improved the proliferation and antitumor
effect of the transferred CD8+T cells (207). The beneficial effect of TBI on therapy efficacy was lost
in Tlr4-deficient mice and in antibiotic-treated mice (207). Administration of the TLR4 agonist
lipopolysaccharide to nonirradiated animals enhanced the efficacy of transferred T cells (207).
Intratumoral administration of the TLR9 agonist CpG oligonucleotide (CpG-ODN) along
with IL-10R-blocking antibodies induces a strong inflammatory antitumor response, character-
ized by secretion of proinflammatory cytokines such as TNF and IL-12 that was potentiated by
antagonizing the immunosuppressive role of IL-10 produced by Tregs and myeloid cells (208–
210). CpG-ODN/anti-IL-10R treatment induces rapid hemorrhagic necrosis and repolarization
of tumor-infiltrating dendritic cells and macrophages to a proinflammatory state. These then pro-
mote T cell–mediated antitumor immunity to permanently eliminate the tumors in most mice
(208). In germfree mice or mice whose gut microbiota has been depleted by antibiotic treatment,
tumor-infiltrating myeloid cell production of proinflammatory cytokines in response to CpG-
ODN is poor, preventing the induction of TNF-dependent necrosis and antitumor adaptive im-
munity (53). Tumors of microbiota-depleted mice have only minor alterations in the number and
differentiation of tumor-infiltrating myeloid cells, mostly derived by recruited Ly6C+circulating
inflammatory monocytes. Tumor-infiltrating myeloid cell subsets in microbiota-depleted mice
show modest alterations in their gene expression profile compared with conventional mice. This
observation contrasts with the major gene expression differences between microbiota-depleted and
conventionally raised mice that are observed after CpG-ODN treatment. Myeloid cells from Tlr4-
deficient tumor-bearing mice are also partially unresponsive to CpG-ODN, and administration of
lipopolysaccharide by gavage to microbiota-depleted mice reconstitutes the myeloid cell response
to CpG, suggesting that activation of TLR4 by products of commensal bacteria primes tumor
myeloid cells for responsiveness to TLR9 (53). These results are reminiscent of the reported in-
ability of mononuclear phagocytes in nonmucosal lymphoid organs of germfree mice to respond to
microbial stimulation with transcription of inflammatory genes due to epigenetically closed chro-
matin conformation at those loci (34). Thus, colonizing the mice postnatally introduces chromatin
changes in the myeloid cells, poising them for rapid responses to inflammatory stimuli, similar
to the phenomenon of trained innate resistance recently reported for various innate effector cell
types (34, 211). However, this epigenetic conformation can be reversed by antibiotic-induced de-
pletion of the commensal microbiota for two to three weeks, suggesting either that this chromatin
conformation is not stable and requires continued instructions from the commensal microbiota
or that inflammatory monocytes are newly generated in the bone marrow and require training
(53). The ability of the tumor-infiltrating myeloid cells from animals with altered microbiota to
produce TNF in response to CpG-ODN intratumoral treatment correlated with the abundance of
specific bacterial genera. For example, TNF production was positively correlated with the frequen-
cies of gram-negative Alistipes spp. and gram-positive Ruminococcus spp. in the fecal microbiota.
Frequencies of Lactobacillus species that are considered to be probiotics with anti-inflammatory
properties, including L. murinum,L. intestinalis,andL. fermentum, negatively correlated with TNF
production (53). Alistipes shahii administration by gavage to mice previously exposed to antibiotics
restored TNF production, whereas L. fermentum impaired TNF production in conventionally
raised mice (53). Thus, the training of myeloid cells for response to CpG-ODN is reversed when
the commensal microbiota is depleted, and individual bacterial strains can either increase or at-
tenuate priming for TNF production. Selective antibiotics, prebiotics, or probiotics that modify
microbiota composition could be considered for optimizing anticancer immunotherapy.
Immunity plays a dual role in cancer. As part of immunosurveillance, it may prevent or de-
lay the growth of the tumor by destroying tumor cells or creating a hostile microenvironment.
216 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
On the other hand, it can also promote tumor initiation and progression by activating an anti-
inflammatory microenvironment or selecting cells that are able to evade the antitumor immune
response (212). In patients with progressive tumors, anticancer immunity is dormant or sup-
pressed, unable to prevent tumor growth. However, in many patients, it can be reactivated by
releasing the immunological brakes responsible for tumor escape (204, 213). Immune checkpoint
inhibitors include antibodies against cytotoxic T lymphocyte–associated protein 4 (CLTA-4), ex-
pressed on activated T effector cells and Tregs, and programmed cell death protein 1 (PD1) or its
ligand PDL-1. These inhibitors have had enduring clinical efficacy in many cancer patients (214).
Anti-CTLA-4 antibodies enhance T cell immune responses and proliferation by preventing the
T cell–suppressive interaction of CTLA-4 with its CD80 and CD86 ligands (214). Anti-PD1 and
anti-PD-L1 antibodies avoid T cell exhaustion and maintain T cell effector functions by prevent-
ing the binding of PD1 on activated T cells with PD-L1 that is expressed on the tumor cells and
other stromal cells (214). As with many effective anticancer therapies, checkpoint inhibitors may
have immune-related adverse effects. Onset of colitis and hypophysitis is mostly observed in re-
sponse to anti-CTLA-4 antibodies, whereas blocking PD1/PD-L1 interactions can lead to thyroid
dysfunctions and pneumonitis (215). Variability in patient response and susceptibility to adverse
reactions has been primarily attributed to the genetic characteristics of the tumors (including the
number of mutations and possibly the number of neoantigens); however, the immune status of the
patient and the possible effect of the microbiota in modulating it are additional important factors
(213, 214).
Two recent studies have identified the important role of the microbiota in modulating the
efficacy of anti-CTLA-4 and anti-PD-L1 therapy (205, 206). The anti-CTLA-4 treatment was
found to be ineffective in antibiotic-treated mice and in germfree mice (205). Treatment of mice
with anti-CTLA-4 antibodies consistently induces T cell–mediated mucosal damage in the ileum
and colon and alters the proportion of bacterial species both in the intestine and in the feces (205).
Bacteroides thetaiotaomicron or B. fragilis, when orally administered to microbiota-depleted mice,
corrected the deficient response to anti-CTLA-4 by activating intratumoral dendritic cells and
inducing a Th1 response in the tumor-draining lymph nodes (205). Of particular interest, feeding
with B. fragilis and Burkholderia cepacia combined not only enhanced the anticancer response but
also prevented the intestinal inflammation and colitis induced by anti-CTLA-4 (205). The correla-
tion between adverse effects and microbiota composition was confirmed by clinical trials in which
the increased frequency of the Bacteroidetes phylum was found to be correlated with resistance to
colitis, whereas underrepresented genetic pathways involved in polyamine transport and B vitamin
biosynthesis were associated with increased risk (216). In mice, selective antibiotics alter the com-
position of the gut microbiota affecting the anti-CTLA-4 response: e.g., vancomycin prevents loss
of Bacteroidales and Burkholderiales and enhances the efficacy of anti-CTLA-4 therapy (205). The
melanoma patients treated with anti-CTLA-4 clustered into three fecal microbiota enterotypes:
Cluster A was dominated by Prevotella spp. and clusters B and C by Bacteroides spp. After treatment,
some of the patients in cluster B acquired the cluster C enterotype. When germfree mice were
colonized with the patients’ fecal microbiota, the mice colonized with cluster C enterotype showed
expansion of B. thetaiotaomicron or B. fragilis and an ability to respond to anti-CTLA-4 treatment
(205). Thus, anti-CTLA-4 treatment may, in some cases, alter the composition of the gut micro-
biota in a direction that favors its antitumor activity. Anti-CTLA-4 elicits, both in mice and in the
humans, a Th1 response specific for either B. thetaiotaomicron or B. fragilis. Microbiota-depleted
mice recover their ability to respond to anti-CTLA-4 therapy when B. fragilis–specific T cells are
adoptively transferred (205). A similar observation was reported for mucosal damage induced by
infection that also induces a persistent memory Th1 cell response against commensal bacteria that,
upon reinfection at the same anatomical site, prime for a polarized Th1 response (14). However,
www.annualreviews.org Microbes and Cancer 217
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
in the case of the anti-CTLA-4 treatment, the mucosal response to commensals and the antitumor
response are at different anatomical sites and the mechanism regulating migration and tropism of
the T cells involved is not yet understood.
Unlike anti-CTLA4, anti-PD-L1 treatment in mice does not induce intestinal damage, and
its anticancer effect does not have an absolute requirement for the presence of gut commensals
(206). However, B16 melanoma grew faster in C57BL/6 mice purchased from Taconic (TAC)
than in those purchased from Jackson Laboratory ( JAX) (206). Anti-PD-L1 almost completely
arrested growth of the smoldering tumors of JAX mice, whereas it only slowed the progression
of the faster-growing tumors of TAC mice (206). More CD8+T cells infiltrated within tumors
in JAX mice compared to TAC mice, suggesting that the slower tumor growth and better re-
sponsiveness to anti-PD-L1 in JAX mice was due to a more robust anticancer immune response
(206). TAC mice cohoused with JAX mice acquired the same rate of tumor growth, antitumor
resistance, and responsiveness to anti-PD-L1 observed in JAX mice (206). The responsiveness of
JAX mice to anti-PD-L1 correlated with the fecal abundancy of the Bifidobacterium species, in-
cluding B. breve,B. longum,andB. adolescentis (206). A commercially available probiotic cocktail of
Bifidobacterium species (including B. breve and B. longum), administered alone or with anti-PD-L1,
to TAC mice enhanced CD8+T cell–induced antitumor activity (206). Treatment of TAC mice
with Bifidobacterium spp. or anti-PD-L1 showed equivalent therapeutic efficacy (206). Thus, unlike
anti-CTL-4, anti-PD-L1 treatment does not require an inflammatory immune activation that is
supported by the presence of the commensal microbiota. However, the presence of Bifidobacterium
spp. in the gut microbiota fosters a more robust antitumor immune response that, when enhanced
by anti-PD-L1, prevents tumor progression.
CONCLUSIONS
Recent technological advances have revolutionized our understanding of human metaorganism
physiology. This has opened new possibilities for basic and clinical science, including precision
medicine. Although novel sequencing technologies have contributed most to the success of mi-
crobiology studies, progress in other areas has been critical as well: improved bacterial culture
methodology, gnotobiotic technology, bioinformatics, computational science (including an expo-
nentially growing number of reference data sets), microscopy, and systems biology. Indeed, we
might safely predict that microbiota research will benefit from both experimental and computa-
tional advances. For example, there is a need for software that interprets larger metagenomic data
sets and presents hierarchically organized functional data such as novel microorganism-centered
annotated pathways and other gene function interpretations. Metagenomic and metatranscrip-
tomic analyses will be improved by new reference genomes from human and animal microbiota.
This will be aided by new single-molecule sequencing methods that generate reads tens of kilobases
long, easing genome assembly of even uncultivable microbes. Single-cell sequencing technologies
will also allow us to sequence microbial genomes and transcriptomes without the need of culturing
the microorganisms. New computational methods will be developed to better understand the ki-
netics of microbial communities and to identify keystone species that have greater influence than
their abundance may imply. Advances in bacterial culture methods and artificial gut development
[e.g., “robogut” technology (217) and “gut on a chip” (218)] will serve to validate computational
predictions and identify novel microbe-microbe and host-microbe interactions. Finally, the def-
inition of microbiota will be broadened to include not just prokaryotes, unicellular eukaryotes,
and viruses but also small multicellular animals residing in the gut (e.g., helminths) and in the skin
(e.g., Demodex mites). Our internal ecosystem is much more diverse than previously realized.
218 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
The studies reviewed here have led us to conclude that many of the mechanisms of microbe-
dependent host physiology and tissue homeostasis are derived from ancestral mechanisms stem-
ming from the earliest interactions between prokaryotes and eukaryotes. Phagocytosis began as a
method of obtaining nutrients and probably led to the genesis of mitochondria. Phagocytes and
myeloid cells, in particular, communicate with commensal microbiota to dictate responses in-
volving immune defense, tissue homeostasis, repair, cancer development, and response to cancer
therapy (219). As phagocytes act as bridges between all the leukocyte subsets involved in resistance
to infection and tissue damage, the cross talk between the host and microbiota transverses both
innate and adaptive immunity.
The composition and functional status of different members of the microbial community can
modulate or even control cancer initiation and progression, comorbidity, and response to ther-
apy. We are at a critical point when the accumulated knowledge of host-microbe interactions
can be used to design new therapeutic strategies. The commensal microbiota is a target for in-
terventions to prevent cancer, control its progression, induce its destruction via genetically en-
gineered bacteria, and prevent cancer-associated comorbidities such as cachexia that can affect
treatment success and quality of life. Indeed, recent studies demonstrating that microbes can
modulate anticancer therapies offer the prospect that targeting the gut microbiota may improve
therapeutic efficacy and attenuate toxicity and adverse reactions. Currently, however, few ther-
apeutic approaches target microbes; clinical applications are still from the dark ages. Probiotics
and prebiotics might be a more precise tool for manipulating the gut microbiota than antibi-
otics. However, until now, there has been little consistent or sufficient evidence of their clinical
efficacy in many microbiota-related diseases. An exception is Clostridium difficile infection, where
crude community-altering therapies, such as fecal transplantation, have been successful (220).
Progress in bacterial ecology and in understanding the role of bile acids in controlling C. dif-
ficile infection has identified bacterial species that for treating C. difficile infection (221) as an
alternative to fecal transplantation. Therapeutic use of bacterial species presents the challenge of
formulation that will achieve efficient colonization in the intestine after oral delivery. Whereas
preparations of facultative anaerobic bacteria may be relatively easy to formulate and adminis-
ter, obligate anaerobes may present more obstacles. Fortunately, many obligate anaerobes, like
Clostridium spp., form spores that germinate only in hypoxic environments. Spores are durable
and survive dry, oxygen-rich environments and therefore could be used as a vector for micro-
bial delivery. These characteristics of the spores also make them attractive for use in anticancer
treatments, since they would germinate in the hypoxic environment of large tumors. Looking
at the road ahead there is hope that we will overcome the many challenges of using microbes
as anticancer therapeutics. In addition to identifying therapeutic microbes, we have to address
the temporal and geographical variability of the microbiota in human populations. Through the
accumulation of clinical data, we need to identify key metabolic processes of the healthy micro-
biota and microbial taxa that promote resistance to disease or improve therapeutic outcomes in
different clinical situations. As we identify beneficial microbes and metabolic processes that pro-
mote resistance to disease or improve efficacy of available treatments, microbiology will likely
become an important component of precision and personalized medicine for cancer and other
diseases.
DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.
www.annualreviews.org Microbes and Cancer 219
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
LITERATURE CITED
1. Ley RE, Lozupone CA, Hamady M, Knight R, Gordon JI. 2008. Worlds within worlds: Evolution of
the vertebrate gut microbiota. Nat. Rev. Microbiol. 6:776–88
2. Yutin N, Wolf MY, Wolf YI, Koonin EV. 2009. The origins of phagocytosis and eukaryogenesis. Biol.
Direct 4:9
3. McInerney JO, O’Connell MJ, Pisani D. 2014. The hybrid nature of the Eukaryota and a consilient view
of life on Earth. Nat. Rev. Microbiol. 12:449–55
4. Chovatiya R, Medzhitov R. 2014. Stress, inflammation, and defense of homeostasis. Mol. Cell 54:281–88
5. Chu H, Mazmanian SK. 2013. Innate immune recognition of the microbiota promotes host-microbial
symbiosis. Nat. Immunol. 14:668–75
6. Hum. Microbiome Proj. Consort. 2012. Structure, function and diversity of the healthy human micro-
biome. Nature 486:207–14
7. Sender R, Fuchs S, Milo R. 2016. Are we really vastly outnumbered? revisiting the ratio of bacterial to
host cells in humans. Cell 164:337–40
8. Bianconi E, Piovesan A, Facchin F, Beraudi A, Casadei R, et al. 2013. An estimation of the number of
cells in the human body. Ann. Hum. Biol. 40:463–71
9. Bosch TC, McFall-Ngai MJ. 2011. Metaorganisms as the new frontier. Zoology 114:185–90
10. Costello EK, Stagaman K, Dethlefsen L, Bohannan BJ, Relman DA. 2012. The application of ecological
theory toward an understanding of the human microbiome. Science 336:1255–62
11. McFall-Ngai M. 2008. Are biologists in ‘future shock’? Symbiosis integrates biology across domains.
Nat. Rev. Microbiol. 6:789–92
12. Candela M, Biagi E, Maccaferri S, Turroni S, Brigidi P. 2012. Intestinal microbiota is a plastic factor
responding to environmental changes. Trends Microbiol. 20:385–91
13. Zeng MY, Cisalpino D, Varadarajan S, Hellman J, Warren HS, et al. 2016. Gut microbiota-induced
immunoglobulin G controls systemic infection by symbiotic bacteria and pathogens. Immunity 44:647–58
14. Hand TW, Dos Santos LM, Bouladoux N, Molloy MJ, Pagan AJ, et al. 2012. Acute gastrointestinal
infection induces long-lived microbiota-specific T cell responses. Science 337:1553–56
15. Stewart EJ. 2012. Growing unculturable bacteria. J. Bacteriol. 194:4151–60
16. Goodman AL, Kallstrom G, Faith JJ, Reyes A, Moore A, et al. 2011. Extensive personal human gut
microbiota culture collections characterized and manipulated in gnotobiotic mice. PNAS 108:6252–57
17. Tringe SG, Hugenholtz P. 2008. A renaissance for the pioneering 16S rRNA gene. Curr. Opin. Microbiol.
11:442–46
18. Schoch CL, Seifert KA, Huhndorf S, Robert V, Spouge JL, et al. 2012. Nuclear ribosomal internal
transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. PNAS 109:6241–46
19. Parfrey LW, Walters WA, Lauber CL, Clemente JC, Berg-Lyons D, et al. 2014. Communities of
microbial eukaryotes in the mammalian gut within the context of environmental eukaryotic diversity.
Front. Microbiol. 5:298
20. Morgun A, Dzutsev A, Dong X, Greer RL, Sexton DJ, et al. 2015. Uncovering effects of antibiotics on
the host and microbiota using transkingdom gene networks. Gut 64:1732–43
21. Zeevi D, Korem T, Zmora N, Israeli D, Rothschild D, et al. 2015. Personalized nutrition by prediction
of glycemic responses. Cell 163:1079–94
22. Belkaid Y, Hand TW. 2014. Role of the microbiota in immunity and inflammation. Cell 157:121–41
23. Cryan JF, Dinan TG. 2012. Mind-altering microorganisms: the impact of the gut microbiota on brain
and behaviour. Nat. Rev. Neurosci. 13:701–12
24. Qin J, Li R, Raes J, Arumugam M, Burgdorf KS, et al. 2010. A human gut microbial gene catalogue
established by metagenomic sequencing. Nature 464:59–65
25. Tsai YC, Conlan S, Deming C, Segre JA, Kong HH, et al. 2016. Resolving the complexity of human
skin metagenomes using single-molecule sequencing. mBio 7:e01948-15
26. Blekhman R, Goodrich JK, Huang K, Sun Q, Bukowski R, et al. 2015. Host genetic variation impacts
microbiome composition across human body sites. Genome Biol. 16:191
27. Mueller NT, Bakacs E, Combellick J, Grigoryan Z, Dominguez-Bello MG. 2015. The infant microbiome
development: Mom matters. Trends Mol. Med. 21:109–17
220 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
28. Yatsunenko T, Rey FE, Manary MJ, Trehan I, Dominguez-Bello MG, et al. 2012. Human gut micro-
biome viewed across age and geography. Nature 486:222–27
29. Jackson MA, Jeffery IB, Beaumont M, Bell JT, Clark AG, et al. 2016. Signatures of early frailty in the
gut microbiota. Genome Med.8:8
30. O’Toole PW, Jeffery IB. 2015. Gut microbiota and aging. Science 350:1214–15
31. Brestoff JR, Artis D. 2013. Commensal bacteria at the interface of host metabolism and the immune
system. Nat. Immunol. 14:676–84
32. Belkaid Y, Naik S. 2013. Compartmentalized and systemic control of tissue immunity by commensals.
Nat. Immunol. 14:646–53
33. Ichinohe T, Pang IK, Kumamoto Y, Peaper DR, Ho JH, et al. 2011. Microbiota regulates immune
defense against respiratory tract influenza A virus infection. PNAS 108:5354–59
34. Ganal SC, Sanos SL, Kallfass C, Oberle K, Johner C, et al. 2012. Priming of natural killer cells by
nonmucosal mononuclear phagocytes requires instructive signals from commensal microbiota. Immunity
37:171–86
35. Abt MC. 2016. Acute gastroenteritis leaves a lasting impression. Cell Host Microbe 19:3–5
36. Chervonsky AV. 2013. Microbiota and autoimmunity. Cold Spring Harb. Perspect. Biol. 5:a007294
37. Naik S, Bouladoux N, Wilhelm C, Molloy MJ, Salcedo R, et al. 2012. Compartmentalized control of
skin immunity by resident commensals. Science 337:1115–19
38. Vaishnava S, Behrendt CL, Ismail AS, Eckmann L, Hooper LV. 2008. Paneth cells directly sense gut
commensals and maintain homeostasis at the intestinal host-microbial interface. PNAS 105:20858
39. Shen W, Li W, Hixon JA, Bouladoux N, Belkaid Y, et al. 2014. Adaptive immunity to murine skin
commensals. PNAS 111:E2977–86
40. Khosravi A, Yanez A, Price JG, Chow A, Merad M, et al. 2014. Gut microbiota promote hematopoiesis
to control bacterial infection. Cell Host Microbe. 15:374–81
41. Balmer ML, Schurch CM, Saito Y, Geuking MB, Li H, et al. 2014. Microbiota-derived compounds drive
steady-state granulopoiesis via MyD88/TICAM signaling. J. Immunol. 193:5273–83
42. Trompette A, Gollwitzer ES, Yadava K, Sichelstiel AK, Sprenger N, et al. 2014. Gut microbiota
metabolism of dietary fiber influences allergic airway disease and hematopoiesis. Nat. Med. 20:159–66
43. Erny D, Hrab ˇ
e de Angelis AL, Jaitin D, Wieghofer P, Staszewski O, et al. 2015. Host microbiota
constantly control maturation and function of microglia in the CNS. Nat. Neurosci. 18:965–77
44. Zhang D, Chen G, Manwani D, Mortha A, Xu C, et al. 2015. Neutrophil ageing is regulated by the
microbiome. Nature 525:528–32
45. Mukherji A, Kobiita A, Ye T, Chambon P. 2013. Homeostasis in intestinal epithelium is orchestrated
by the circadian clock and microbiota cues transduced by TLRs. Cell 153:812–27
46. Nguyen KD, Fentress SJ, Qiu Y, Yun K, Cox JS, Chawla A. 2013. Circadian gene Bmal1 regulates diurnal
oscillations of Ly6Chi inflammatory monocytes. Science 341:1483–88
47. Thaiss CA, Zeevi D, Levy M, Zilberman-Schapira G, Suez J, et al. 2014. Transkingdom control of
microbiota diurnal oscillations promotes metabolic homeostasis. Cell 159:514–29
48. Perussia B, Dayton ET, Fanning V, Thiagarajan P, Hoxie J, Trinchieri G. 1983. Immune interferon
and leukocyte-conditioned medium induce normal and leukemic myeloid cells to differentiate along the
monocytic pathway. J. Exp. Med. 158:2058–80
49. Colotta F, Re F, Polentarutti N, Sozzani S, Mantovani A. 1992. Modulation of granulocyte survival and
programmed cell death by cytokines and bacterial products. Blood 80:2012–20
50. de Oliveira S, Rosowski EE, Huttenlocher A. 2016. Neutrophil migration in infection and wound repair:
going forward in reverse. Nat. Rev. Immunol. 16:378–91
51. Askenase MH, Han S-J, Byrd AL, Morais da Fonseca D, Bouladoux N, et al. 2015. Bone-marrow-resident
NK cells prime monocytes for regulatory function during infection. Immunity 42:1130–42
52. Gottschalk RA, Martins AJ, Angermann BR, Dutta B, Ng CE, et al. 2016. Distinct NF-κBandMAPK
activation thresholds uncouple steady-state microbe sensing from anti-pathogen inflammatory responses.
Cell Syst. 2:378–90
53. Iida N, Dzutsev A, Stewart CA, Smith L, Bouladoux N, et al. 2013. Commensal bacteria control cancer
response to therapy by modulating the tumor microenvironment. Science 342:967–70
www.annualreviews.org Microbes and Cancer 221
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
54. Ozmen L, Pericin M, Hakimi J, Chizzonite RA, Wysocka M, et al. 1994. Interleukin 12, interferon γ,
and tumor necrosis factor αare the key cytokines of the generalized Shwartzman reaction. J. Exp. Med.
180:907–15
55. Hanahan D, Weinberg RA. 2000. The hallmarks of cancer. Cell 100:57–70
56. Fidler IJ. 2003. The pathogenesis of cancer metastasis: the ‘seed and soil’ hypothesis revisited. Nat. Rev.
Cancer 3:453–58
57. Dolberg DS, Bissell MJ. 1984. Inability of Rous sarcoma virus to cause sarcomas in the avian embryo.
Nature 309:552–56
58. Martincorena I, Roshan A, Gerstung M, Ellis P, Van Loo P, et al. 2015. Tumor evolution: high burden
and pervasive positive selection of somatic mutations in normal human skin. Science 348:880–86
59. Bissell MJ, Hines WC. 2011. Why don’t we get more cancer? A proposed role of the microenvironment
in restraining cancer progression. Nat. Med. 17:320–29
60. Adami J, Gabel H, Lindelof B, Ekstrom K, Rydh B, et al. 2003. Cancer risk following organ transplan-
tation: a nationwide cohort study in Sweden. Br. J. Cancer 89:1221–27
61. Vogtmann E, Goedert JJ. 2016. Epidemiologic studies of the human microbiome and cancer. Br. J.
Cancer 114:237–42
62. IARC Work. Group Evaluation Carcinog. Risks Hum. 1994. Schistosomes, Liver Flukes and Helicobacter
pylori. IARC Monogr. Evaluation Carcinog. Risks Hum. Vol. 61. Geneva, Switz.: World Health Organ.
63. Atherton JC, Blaser MJ. 2009. Coadaptation of Helicobacter pylori and humans: ancient history, modern
implications. J. Clin. Investig. 119:2475–87
64. Kostic AD, Chun E, Robertson L, Glickman JN, Gallini CA, et al. 2013. Fusobacterium nucleatum poten-
tiates intestinal tumorigenesis and modulates the tumor-immune microenvironment. Cell Host Microbe
14:207–15
65. Sears CL, Garrett WS. 2014. Microbes, microbiota, and colon cancer. Cell Host Microbe 15:317–28
66. Morgan XC, Tickle TL, Sokol H, Gevers D, Devaney KL, et al. 2012. Dysfunction of the intestinal
microbiome in inflammatory bowel disease and treatment. Genome Biol. 13:R79
67. Arthur JC, Perez-Chanona E, Muhlbauer M, Tomkovich S, Uronis JM, et al. 2012. Intestinal inflam-
mation targets cancer-inducing activity of the microbiota. Science 338:120–23
68. Thomas RM, Jobin C. 2015. The microbiome and cancer: Is the ‘oncobiome’ mirage real? Trends Cancer
1:24–35
69. Poutahidis T, Varian BJ, Levkovich T, Lakritz JR, Mirabal S, et al. 2015. Dietary microbes modulate
transgenerational cancer risk. Cancer Res. 75:1197–204
70. Bhaskaran K, Douglas I, Forbes H, dos-Santos-Silva I, Leon DA, Smeeth L. 2014. Body-mass index and
risk of 22 specific cancers: a population-based cohort study of 5.24 million UK adults. Lancet 384:755–65
71. Conroy MJ, Dunne MR, Donohoe CL, Reynolds JV. 2016. Obesity-associated cancer: an immunological
perspective. Proc. Nutr. Soc. 75:125–38
72. Beyaz S, Mana MD, Roper J, Kedrin D, Saadatpour A, et al. 2016. High-fat diet enhances stemness and
tumorigenicity of intestinal progenitors. Nature 531:53–58
73. Ohtani N, Yoshimoto S, Hara E. 2014. Obesity and cancer: a gut microbial connection. Cancer Res.
74:1885–89
74. Abnet CC, Corley DA, Freedman ND, Kamangar F. 2015. Diet and upper gastrointestinal malignancies.
Gastroenterology 148:1234–43.e4
75. Doll R, Peto R. 1981. The causes of cancer: Quantitative estimates of avoidable risks of cancer in the
United States today. J. Natl. Cancer Inst. 66:1191–308
76. Song M, Garrett WS, Chan AT. 2015. Nutrients, foods, and colorectal cancer prevention. Gastroenterol-
ogy 148:1244–60.e16
77. Wei W, Sun W, Yu S, Yang Y, Ai L. 2016. Butyrate production from high-fiber diet protects against
lymphoma tumor. Leuk. Lymphoma 57:2401–8
78. Masuyama H, Mitsui T, Nobumoto E, Hiramatsu Y. 2015. The effects of high-fat diet exposure in utero
on the obesogenic and diabetogenic traits through epigenetic changes in adiponectin and leptin gene
expression for multiple generations in female mice. Endocrinology 156:2482–91
79. Modi SR, Collins JJ, Relman DA. 2014. Antibiotics and the gut microbiota. J. Clin. Investig. 124:4212–18
222 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
80. Cho I, Yamanishi S, Cox L, Methe BA, Zavadil J, et al. 2012. Antibiotics in early life alter the murine
colonic microbiome and adiposity. Nature 488:621–26
81. Scott FI, Horton DB, Mamtani R, Haynes K, Goldberg DS, et al. 2016. Administration of antibiotics to
children before age 2 years increases risk for childhood obesity. Gastroenterology 151:120–129.e5
82. Hviid A, Svanstrom H, Frisch M. 2011. Antibiotic use and inflammatory bowel diseases in childhood.
Gut 60:49–54
83. Sergentanis TN, Zagouri F, Zografos GC. 2010. Is antibiotic use a risk factor for breast cancer? A
meta-analysis. Pharmacoepidemiol. Drug Saf. 19:1101–7
84. Xuan C, Shamonki JM, Chung A, Dinome ML, Chung M, et al. 2014. Microbial dysbiosis is associated
with human breast cancer. PLOS ONE 9:e83744
85. Lee YY, Mahendra Raj S, Graham DY. 2013. Helicobacter pylori infection—a boon or a bane: lessons
from studies in a low-prevalence population. Helicobacter 18:338–46
86. Levkovich T, Poutahidis T, Cappelle K, Smith MB, Perrotta A, et al. 2014. ‘Hygienic’ lymphocytes
convey increased cancer risk. J. Anal. Oncol. 3:113–21
87. Fonseca DM, Hand TW, Han S-J, Gerner MY, Glatman Zaretsky A, et al. 2015. Microbiota-dependent
sequelae of acute infection compromise tissue-specific immunity. Cell 163:354–66
88. Chen JA, Splenser A, Guillory B, Luo J, Mendiratta M, et al. 2015. Ghrelin prevents tumour- and
cisplatin-induced muscle wasting: characterization of multiple mechanisms involved. J. Cachexia Sar-
copenia Muscle 6:132–43
89. de Matos-Neto EM, Lima JDCC, de Pereira WO, Figuer ˆ
edo RG, Riccardi DMDR, et al. 2015. Systemic
inflammation in cachexia—is tumor cytokine expression profile the culprit? Front. Immunol. 6:629
90. Kir S, White JP, Kleiner S, Kazak L, Cohen P, et al. 2014. Tumour-derived PTH-related protein triggers
adipose tissue browning and cancer cachexia. Nature 513:100–4
91. Armougom F, Henry M, Vialettes B, Raccah D, Raoult D. 2009. Monitoring bacterial community of
human gut microbiota reveals an increase in Lactobacillus in obese patients and Methanogens in anorexic
patients. PLOS ONE 4:e7125
92. Varian BJ, Goureshetti S, Poutahidis T, Lakritz JR, Levkovich T, et al. 2016. Beneficial bacteria inhibit
cachexia. Oncotarget 7:11803–16
93. Bindels LB, Neyrinck AM, Claus SP, Le Roy CI, Grangette C, et al. 2016. Synbiotic approach restores
intestinal homeostasis and prolongs survival in leukaemic mice with cachexia. ISME J. 10:1456–70
94. Schieber AMP, Lee YM, Chang MW, Leblanc M, Collins B, et al. 2015. Disease tolerance mediated by
microbiome E. coli involves inflammasome and IGF-1 signaling. Science 350:558–63
95. Klein GL, Petschow BW, Shaw AL, Weaver E. 2013. Gut barrier dysfunction and microbial translocation
in cancer cachexia: a new therapeutic target. Curr. Opin. Support. Palliat. Care 7:361–67
96. de Martel C, Ferlay J, Franceschi S, Vignat J, Bray F, et al. 2012. Global burden of cancers attributable
to infections in 2008: a review and synthetic analysis. Lancet Oncol. 13:607–15
97. IARC Work. Group Evaluation Carcinog. Risks Hum. 2012. A Review of Human Carcinogens: Part B;
Biological agents. IARC Monogr. Evaluation Carcinog. Risks Hum. Vol. 100B. Geneva, Switz.: World
Health Organ.
98. Zhang X, Zhang Z, Zheng B, He Z, Winberg G, Ernberg I. 2013. An update on viral association of
human cancers. Arch. Virol. 158:1433–43
99. Ng J, Wu J. 2012. Hepatitis B- and hepatitis C-related hepatocellular carcinomas in the United States:
similarities and differences. Hepat. Mon. 12:e7635
100. Psyrri A, Rampias T, Vermorken JB. 2014. The current and future impact of human papillomavirus on
treatment of squamous cell carcinoma of the head and neck. Ann. Oncol. 25:2101–15
101. Tommasino M. 2014. The human papillomavirus family and its role in carcinogenesis. Semin. Cancer
Biol. 26:13–21
102. Ciminale V, Rende F, Bertazzoni U, Romanelli MG. 2014. HTLV-1 and HTLV-2: highly similar viruses
with distinct oncogenic properties. Front. Microbiol. 5:398
103. Feng H, Shuda M, Chang Y, Moore PS. 2008. Clonal integration of a polyomavirus in human Merkel
cell carcinoma. Science 319:1096–100
104. Zheng ZM. 2010. Viral oncogenes, noncoding RNAs, and RNA splicing in human tumor viruses. Int.
J. Biol. Sci. 6:730–55
www.annualreviews.org Microbes and Cancer 223
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
105. Mitra A, MacIntyre DA, Lee YS, Smith A, Marchesi JR, et al. 2015. Cervical intraepithelial neoplasia
disease progression is associated with increased vaginal microbiome diversity. Sci. Rep. 5:16865
106. Rehermann B, Nascimbeni M. 2005. Immunology of hepatitis B virus and hepatitis C virus infection.
Nature Rev. Immunol. 5:215–29
107. Bender S, Reuter A, Eberle F, Einhorn E, Binder M, Bartenschlager R. 2015. Activation of type I and III
interferon response by mitochondrial and peroxisomal MAVS and inhibition by hepatitis C virus. PLOS
Pathog. 11:e1005264
108. Chen Y, Chen Z, Guo R, Chen N, Lu H, et al. 2011. Correlation between gastrointestinal fungi and
varying degrees of chronic hepatitis B virus infection. Diagn. Microbiol. Infect. Dis. 70:492–98
109. Xu M, Wang B, Fu Y, Chen Y, Yang F, et al. 2012. Changes of fecal Bifidobacterium species in adult
patients with hepatitis B virus-induced chronic liver disease. Microb. Ecol. 63:304–13
110. Henao-Mejia J, Elinav E, Thaiss CA, Licona-Limon P, Flavell RA. 2013. Role of the intestinal micro-
biome in liver disease. J. Autoimmun. 46:66–73
111. Chou HH, Chien WH, Wu LL, Cheng CH, Chung CH, et al. 2015. Age-related immune clearance of
hepatitis B virus infection requires the establishment of gut microbiota. PNAS 112:2175–80
112. Moore PS, Chang Y. 2010. Why do viruses cause cancer? Highlights of the first century of human
tumour virology. Nat. Rev. Cancer 10:878–89
113. Ki MR, Hwang M, Kim AY, Lee EM, Lee EJ, et al. 2014. Role of vacuolating cytotoxin VacA and
cytotoxin-associated antigen CagA of Helicobacter pylori in the progression of gastric cancer. Mol. Cell
Biochem. 396:23–32
114. Khosravi Y, Bunte RM, Chiow KH, Tan TL, Wong WY, et al. 2016. Helicobacter pylori and gut microbiota
modulate energy homeostasis prior to inducing histopathological changes in mice. Gut Microbes 7:48–53
115. Perry S, de Jong BC, Solnick JV, de la Luz Sanchez M, Yang S, et al. 2010. Infection with Helicobacter
pylori is associated with protection against tuberculosis. PLOS ONE 5:e8804
116. Koch KN, Hartung ML, Urban S, Kyburz A, Bahlmann AS, et al. 2015. Helicobacter urease-induced
activation of the TLR2/NLRP3/IL-18 axis protects against asthma. J. Clin. Investig. 125:3297–302
117. Kim DJ, Park JH, Franchi L, Backert S, Nunez G. 2013. The Cag pathogenicity island and interaction
between TLR2/NOD2 and NLRP3 regulate IL-1βproduction in Helicobacter pylori infected dendritic
cells. Eur. J. Immunol. 43:2650–58
118. Peek RM Jr., Blaser MJ. 2002. Helicobacter pylori and gastrointestinal tract adenocarcinomas. Nat. Rev.
Cancer 2:28–37
119. Plottel CS, Blaser MJ. 2011. Microbiome and malignancy. Cell Host Microbe 10:324–35
120. Yap TW, Gan HM, Lee YP, Leow AH, Azmi AN, et al. 2016. Helicobacter pylori eradication causes
perturbation of the human gut microbiome in young adults. PLOS ONE 11:e0151893
121. Belkaid Y, Segre JA. 2014. Dialogue between skin microbiota and immunity. Science 346:954–59
122. Goldszmid RS, Trinchieri G. 2012. The price of immunity. Nat. Immunol. 13:932–38
123. Jobin C. 2012. Colorectal cancer: CRC—all about microbial products and barrier function? Nat. Rev.
Gastroenterol. Hepatol. 9:694–96
124. Rao VP, Poutahidis T, Ge Z, Nambiar PR, Boussahmain C, et al. 2006. Innate immune inflammatory
response against enteric bacteria Helicobacter hepaticus induces mammary adenocarcinoma in mice. Cancer
Res. 66:7395–400
125. Attene-Ramos MS, Wagner ED, Plewa MJ, Gaskins HR. 2006. Evidence that hydrogen sulfide is a
genotoxic agent. Mol. Cancer Res. 4:9–14
126. Andriamihaja M, Lan A, Beaumont M, Audebert M, Wong X, et al. 2015. The deleterious metabolic
and genotoxic effects of the bacterial metabolite p-cresol on colonic epithelial cells. Free Radic. Biol. Med.
85:219–27
127. Garrett WS, Lord GM, Punit S, Lugo-Villarino G, Mazmanian SK. 2007. Communicable ulcerative
colitis induced by T-bet deficiency in the innate immune system. Cell 131:33
128. Elinav E, Strowig T, Kau AL, Henao-Mejia J, Thaiss CA, et al. 2011. NLRP6 inflammasome regulates
colonic microbial ecology and risk for colitis. Cell 145:745–57
129. Rubinstein MR, Wang X, Liu W, Hao Y, Cai G, Han YW. 2013. Fusobacterium nucleatum promotes
colorectal carcinogenesis by modulating E-cadherin/β-catenin signaling via its FadA adhesin. Cell Host
Microbe 14:195–206
224 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
130. Zackular JP, Baxter NT, Iverson KD, Sadler WD, Petrosino JF, et al. 2013. The gut microbiome
modulates colon tumorigenesis. mBio 4:e00692-13
131. Mira-Pascual L, Cabrera-Rubio R, Ocon S, Costales P, Parra A, et al. 2015. Microbial mucosal colonic
shifts associated with the development of colorectal cancer reveal the presence of different bacterial and
archaeal biomarkers. J. Gastroenterol. 50:167–79
132. Okuda T, Kokubu E, Kawana T, Saito A, Okuda K, Ishihara K. 2012. Synergy in biofilm formation
between Fusobacterium nucleatum and Prevotella species. Anaerobe 18:110–16
133. Abed J, Emgard JE, Zamir G, Faroja M, Almogy G, et al. 2016. Fap2 mediates Fusobacterium nucleatum
colorectal adenocarcinoma enrichment by binding to tumor-expressed Gal-GalNAc. Cell Host Microbe
20:215–25
134. Gur C, Ibrahim Y, Isaacson B, Yamin R, Abed J, et al. 2015. Binding of the Fap2 protein of Fusobacterium
nucleatum to human inhibitory receptor TIGIT protects tumors from immune cell attack. Immunity
42:344–55
135. Dejea CM, Wick EC, Hechenbleikner EM, White JR, Mark Welch JL, et al. 2014. Microbiota organi-
zation is a distinct feature of proximal colorectal cancers. PNAS 111:18321–26
136. Johnson CH, Dejea CM, Edler D, Hoang LT, Santidrian AF, et al. 2015. Metabolism links bacterial
biofilms and colon carcinogenesis. Cell Metab. 21:891–97
137. Dalmasso G, Cougnoux A, Delmas J, Darfeuille-Michaud A, Bonnet R. 2014. The bacterial genotoxin
colibactin promotes colon tumor growth by modifying the tumor microenvironment. Gut Microbes 5:675–
80
138. Nesic D, Hsu Y, Stebbins CE. 2004. Assembly and function of a bacterial genotoxin. Nature 429:429–33
139. Kim JJ, Tao H, Carloni E, Leung WK, Graham DY, Sepulveda AR. 2002. Helicobacter pylori impairs
DNA mismatch repair in gastric epithelial cells. Gastroenterology 123:542–53
140. Maddocks OD, Short AJ, Donnenberg MS, Bader S, Harrison DJ. 2009. Attaching and effacingEscherichia
coli downregulate DNA mismatch repair protein in vitro and are associated with colorectal adenocarci-
nomas in humans. PLOS ONE 4:e5517
141. Yao Q, Zhang L, Wan X, Chen J, Hu L, et al. 2014. Structure and specificity of the bacterial cysteine
methyltransferase effector NleE suggests a novel substrate in human DNA repair pathway. PLOS Pathog.
10:e1004522
142. Boleij A, Hechenbleikner EM, Goodwin AC, Badani R, Stein EM, et al. 2015. The Bacteroides fragilis
toxin gene is prevalent in the colon mucosa of colorectal cancer patients. Clin. Infect. Dis. 60:208–15
143. Wick EC, Rabizadeh S, Albesiano E, Wu X, Wu S, et al. 2014. Stat3 activation in murine colitis induced
by enterotoxigenic Bacteroides fragilis.Inflamm. Bowel Dis. 20:821–34
144. Geis AL, Fan H, Wu X, Wu S, Huso DL, et al. 2015. Regulatory T-cell response to enterotoxigenic
Bacteroides fragilis colonization triggers IL17-dependent colon carcinogenesis. Cancer Discov. 5:1098–109
145. Grivennikov SI, Wang K, Mucida D, Stewart CA, Schnabl B, et al. 2012. Adenoma-linked barrier defects
and microbial products drive IL-23/IL-17-mediated tumour growth. Nature 491:254–58
146. Ernst M, Putoczki TL. 2013. Targeting IL-11 signaling in colon cancer. Oncotarget 4:1860–61
147. Salcedo R, Worschech A, Cardone M, Jones Y, Gyulai Z, et al. 2010. MyD88-mediated signaling prevents
development of adenocarcinomas of the colon: role of interleukin 18. J. Exp. Med. 207:1625–36
148. Henao-Mejia J, Elinav E, Jin C, Hao L, Mehal WZ, et al. 2012. Inflammasome-mediated dysbiosis
regulates progression of NAFLD and obesity. Nature 482:179–85
149. Macia L, Tan J, Vieira AT, Leach K, Stanley D, et al. 2015. Metabolite-sensing receptors GPR43 and
GPR109A facilitate dietary fibre-induced gut homeostasis through regulation of the inflammasome. Nat.
Commun. 6:6734
150. Chudnovskiy A, Mortha A, Kana V, Kennard A, Ramirez JD, et al. 2016. Host-protozoan interactions
protect from mucosal infections through activation of the inflammasome. Cell 167:444–56
151. Smith PM, Howitt MR, Panikov N, Michaud M, Gallini CA, et al. 2013. The microbial metabolites,
short-chain fatty acids, regulate colonic Treg cell homeostasis. Science 341:569–73
152. Singh N, Gurav A, Sivaprakasam S, Brady E, Padia R, et al. 2014. Activation of Gpr109a, receptor
for niacin and the commensal metabolite butyrate, suppresses colonic inflammation and carcinogenesis.
Immunity 40:128–39
www.annualreviews.org Microbes and Cancer 225
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
153. Saleh M, Trinchieri G. 2011. Innate immune mechanisms of colitis and colitis-associated colorectal
cancer. Nat. Rev. Immunol. 11:9–20
154. Huber S, Gagliani N, Zenewicz LA, Huber FJ, Bosurgi L, et al. 2012. IL-22BP is regulated by the
inflammasome and modulates tumorigenesis in the intestine. Nature 491:259–63
155. Kirchberger S, Royston DJ, Boulard O, Thornton E, Franchini F, et al. 2013. Innate lymphoid cells
sustain colon cancer through production of interleukin-22 in a mouse model. J. Exp. Med. 210:917–31
156. Munoz M, Eidenschenk C, Ota N, Wong K, Lohmann U, et al. 2015. Interleukin-22 induces interleukin-
18 expression from epithelial cells during intestinal infection. Immunity 42:321–31
157. Nowarski R, Jackson R, Gagliani N, de Zoete MR, Palm NW, et al. 2015. Epithelial IL-18 equilibrium
controls barrier function in colitis. Cell 163:1444–56
158. Elangovan S, Pathania R, Ramachandran S, Ananth S, Padia RN, et al. 2014. The niacin/butyrate receptor
GPR109A suppresses mammary tumorigenesis by inhibiting cell survival. Cancer Res. 74:1166–78
159. Van den Abbeele P, Belzer C, Goossens M, Kleerebezem M, De Vos WM, et al. 2013. Butyrate-producing
Clostridium cluster XIVa species specifically colonize mucins in an in vitro gut model. ISME J. 7:949–61
160. Tan J, McKenzie C, Potamitis M, Thorburn AN, Mackay CR, Macia L. 2014. The role of short-chain
fatty acids in health and disease. Adv. Immunol. 121:91–119
161. Belcheva A, Irrazabal T, Robertson SJ, Streutker C, Maughan H, et al. 2014. Gut microbial metabolism
drives transformation of Msh2-deficient colon epithelial cells. Cell 158:288–99
162. Uittamo J, Siikala E, Kaihovaara P, Salaspuro M, Rautemaa R. 2009. Chronic candidosis and oral cancer
in APECED-patients: production of carcinogenic acetaldehyde from glucose and ethanol by Candida
albicans.Int. J. Cancer 124:754–56
163. Howitt MR, Lavoie S, Michaud M, Blum AM, Tran SV, et al. 2016. Tuft cells, taste-chemosensory cells,
orchestrate parasite type 2 immunity in the gut. Science 351:1329–33
164. Nowak A, Czyzowska A, Huben K, Sojka M, Kuberski S, et al. 2016. Prebiotics and age, but not probiotics
affect the transformation of 2-amino-3-methyl-3H-imidazo[4,5-f]quinoline (IQ) by fecal microbiota—an
in vitro study. Anaerobe 39:124–35
165. Megaraj V, Ding X, Fang C, Kovalchuk N, Zhu Y, Zhang QY. 2014. Role of hepatic and intestinal p450
enzymes in the metabolic activation of the colon carcinogen azoxymethane in mice. Chem. Res. Toxicol.
27:656–62
166. Morita N, Walaszek Z, Kinjo T, Nishimaki T, Hanausek M, et al. 2008. Effects of synthetic and natural
in vivo inhibitors of β-glucuronidase on azoxymethane-induced colon carcinogenesis in rats. Mol. Med.
Rep. 1:741–46
167. Vergara-Castaneda HA, Guevara-Gonzalez RG, Ramos-Gomez M, Reynoso-Camacho R, Guzman-
Maldonado H, et al. 2010. Non-digestible fraction of cooked bean (Phaseolus vulgaris L.) cultivar Bayo
Madero suppresses colonic aberrant crypt foci in azoxymethane-induced rats. Food Funct. 1:294–300
168. Kim YG, Udayanga KG, Totsuka N, Weinberg JB, Nunez G, Shibuya A. 2014. Gut dysbiosis promotes
M2 macrophage polarization and allergic airway inflammation via fungi-induced PGE2. Cell Host Microbe.
15:95–102
169. Nagamine CM, Sohn JJ, Rickman BH, Rogers AB, Fox JG, Schauer DB. 2008. Helicobacter hepati-
cus infection promotes colon tumorigenesis in the BALB/c-Rag2/ApcMin/+mouse. Infect. Immun.
76:2758–66
170. Poutahidis T, Cappelle K, Levkovich T, Lee CW, Doulberis M, et al. 2013. Pathogenic intestinal
bacteria enhance prostate cancer development via systemic activation of immune cells in mice. PLOS
ONE 8:e73933
171. Fox JG, Feng Y, Theve EJ, Raczynski AR, Fiala JL, et al. 2010. Gut microbes define liver cancer risk in
mice exposed to chemical and viral transgenic hepatocarcinogens. Gut 59:88–97
172. Erdman SE, Poutahidis T, Tomczak M, Rogers AB, Cormier K, et al. 2003. CD4+CD25+regulatory T
lymphocytes inhibit microbially induced colon cancer in Rag2-deficient mice. Am.J.Pathol.162:691–702
173. Lakritz JR, Poutahidis T, Mirabal S, Varian BJ, Levkovich T, et al. 2015. Gut bacteria require neutrophils
to promote mammary tumorigenesis. Oncotarget 6:9387–96
174. Yamamoto ML, Maier I, Dang AT, Berry D, Liu J, et al. 2013. Intestinal bacteria modify lymphoma inci-
dence and latency by affecting systemic inflammatory state, oxidative stress, and leukocyte genotoxicity.
Cancer Res. 73:4222–32
226 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
175. Westbrook AM, Wei B, Hacke K, Xia M, Braun J, Schiestl RH. 2012. The role of tumour necrosis
factor-αand tumour necrosis factor receptor signalling in inflammation-associated systemic genotoxicity.
Mutagenesis 27:77–86
176. Rutkowski MR, Stephen TL, Svoronos N, Allegrezza MJ, Tesone AJ, et al. 2015. Microbially driven
TLR5-dependent signaling governs distal malignant progression through tumor-promoting inflamma-
tion. Cancer Cell 27:27–40
177. Coley WB. 1893. Treatment of malignant tumors by repeated inoculation of erysipelas: with a report of
10 cases. Am. J. Med. Sci. 105:487–564
178. Tang DH, Chang SS. 2015. Management of carcinoma in situ of the bladder: best practice and recent
developments. Adv. Urol. 7:351–64
179. Forbes NS. 2010. Engineering the perfect (bacterial) cancer therapy. Nat. Rev. Cancer 10:785–94
180. Bettegowda C, Huang X, Lin J, Cheong I, Kohli M, et al. 2006. The genome and transcriptomes of the
anti-tumor agent Clostridium novyi-NT.Nat. Biotechnol. 24:1573–80
181. Forbes NS. 2006. Profile of a bacterial tumor killer. Nat. Biotechnol. 24:1484–85
182. Mlynarczyk GS, Berg CA, Withrock IC, Fick ME, Anderson SJ, et al. 2014. Salmonella as a biological
“Trojan horse” for neoplasia: future possibilities including brain cancer. Med. Hypotheses 83:343–45
183. Zhang M, Forbes NS. 2015. Trg-deficient Salmonella colonize quiescent tumor regions by exclusively
penetrating or proliferating. J. Control Release 199:180–89
184. Spanogiannopoulos P, Bess EN, Carmody RN, Turnbaugh PJ. 2016. The microbial pharmacists within
us: a metagenomic view of xenobiotic metabolism. Nat. Rev. Microbiol. 14:273–87
185. Dzutsev A, Goldszmid RS, Viaud S, Zitvogel L, Trinchieri G. 2015. The role of the microbiota in
inflammation, carcinogenesis, and cancer therapy. Eur. J. Immunol. 45:17–31
186. Li H, Jia W. 2013. Cometabolism of microbes and host: implications for drug metabolism and drug-
induced toxicity. Clin. Pharmacol. Ther. 94:574–81
187. Maurice CF, Haiser HJ, Turnbaugh PJ. 2013. Xenobiotics shape the physiology and gene expression of
the active human gut microbiome. Cell 152:39–50
188. Fujita K, Sparreboom A. 2010. Pharmacogenetics of irinotecan disposition and toxicity: a review. Curr.
Clin. Pharmacol. 5:209–17
189. Stringer AM, Gibson RJ, Logan RM, Bowen JM, Yeoh AS, Keefe DM. 2008. Faecal microflora and β-
glucuronidase expression are altered in an irinotecan-induced diarrhea model in rats. Cancer Biol. Ther.
7:1919–25
190. McIntosh FM, Maison N, Holtrop G, Young P, Stevens VJ, et al. 2012. Phylogenetic distribution of
genes encoding β-glucuronidase activity in human colonic bacteria and the impact of diet on faecal
glycosidase activities. Environ. Microbiol. 14:1876–87
191. Wallace BD, Wang H, Lane KT, Scott JE, Orans J, et al. 2010. Alleviating cancer drug toxicity by
inhibiting a bacterial enzyme. Science 330:831–35
192. Galluzzi L, Vitale I, Michels J, Brenner C, Szabadkai G, et al. 2014. Systems biology of cisplatin resistance:
past, present and future. Cell Death Dis. 5:e1257
193. Roy S, Ryals MM, Van den Bruele AB, Fitzgerald TS, Cunningham LL. 2013. Sound preconditioning
therapy inhibits ototoxic hearing loss in mice. J. Clin. Investig. 123:4945–49
194. Zhu S, Pabla N, Tang C, He L, Dong Z. 2015. DNA damage response in cisplatin-induced nephrotox-
icity. Arch. Toxicol. 89:2197–205
195. Park SB, Goldstein D, Krishnan AV, Lin CSY, Friedlander ML, et al. 2013. Chemotherapy-induced
peripheral neurotoxicity: a critical analysis. CA Cancer J. Clin. 63:419–37
196. Laurent A, Nicco C, Chereau C, Goulvestre C, Alexandre J, et al. 2005. Controlling tumor growth by
modulating endogenous production of reactive oxygen species. Cancer Res. 65:948–56
197. Gui QF, Lu HF, Zhang CX, Xu ZR, Yang YH. 2015. Well-balanced commensal microbiota contributes
to anti-cancer response in a lung cancer mouse model. Genet. Mol. Res. 14:5642–51
198. Chitapanarux I, Chitapanarux T, Traisathit P, Kudumpee S, Tharavichitkul E, Lorvidhaya V. 2010.
Randomized controlled trial of live Lactobacillus acidophilus plus Bifidobacterium bifidum in prophylaxis of
diarrhea during radiotherapy in cervical cancer patients. Radiat. Oncol. 5:31
www.annualreviews.org Microbes and Cancer 227
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
IY35CH08-Trinchieri ARI 18 January 2017 13:30
199. Wang Y, Luo X, Pan H, Huang W, Wang X, et al. 2015. Pharmacological inhibition of NADPH oxidase
protects against cisplatin induced nephrotoxicity in mice by two step mechanism. Food Chem. Toxicol.
83:251–60
200. Dhiman RK. 2013. Gut microbiota and hepatic encephalopathy. Metab. Brain Dis. 28:321–26
201. Viaud S, Saccheri F, Mignot G, Yamazaki T, Daillere R, et al. 2013. The intestinal microbiota modulates
the anticancer immune effects of cyclophosphamide. Science 342:971–6
202. Zwielehner J, Lassl C, Hippe B, Pointner A, Switzeny OJ, et al. 2011. Changes in human fecal microbiota
due to chemotherapy analyzed by TaqMan-PCR, 454 sequencing and PCR-DGGE fingerprinting. PLOS
ONE 6:e28654
203. Daillere R, Vetizou M, Waldschmitt N, Yamazaki T, Isnard C, et al. 2016. Enterococcus hirae and
Barnesiella intestinihominis facilitate cyclophosphamide-induced therapeutic immunomodulatory effects.
Immunity 45:931–43
204. Couzin-Frankel J. 2013. Breakthrough of the year 2013: cancer immunotherapy. Science 342:1432–33
205. Vetizou M, Pitt JM, Daillere R, Lepage P, Waldschmitt N, et al. 2015. Anticancer immunotherapy by
CTLA-4 blockade relies on the gut microbiota. Science 350:1079–84
206. Sivan A, Corrales L, Hubert N, Williams JB, Aquino-Michaels K, et al. 2015. Commensal Bifidobacterium
promotes antitumor immunity and facilitates anti-PD-L1 efficacy. Science 350:1084–89
207. Paulos CM, Wrzesinski C, Kaiser A, Hinrichs CS, Chieppa M, et al. 2007. Microbial translocation
augments the function of adoptively transferred self/tumor-specific CD8+T cells via TLR4 signaling.
J. Clin. Investig. 117:2197–204
208. Guiducci C, Vicari AP, Sangaletti S, Trinchieri G, Colombo MP. 2005. Redirecting in vivo elicited
tumor infiltrating macrophages and dendritic cells towards tumor rejection. Cancer Res. 65:3437–46
209. Vicari AP, Chiodoni C, Vaure C, Ait-Yahia S, Dercamp C, et al. 2002. Reversal of tumor-induced
dendritic cell paralysis by CpG immunostimulatory oligonucleotide and anti-interleukin 10 receptor
antibody. J. Exp. Med. 196:541–49
210. Stewart CA, Metheny H, Iida N, Smith L, Hanson M, et al. 2013. Interferon-dependent IL-10 production
by Tregs limits tumor Th17 inflammation. J. Clin. Investig. 123:4859–74
211. Netea MG, Joosten LA, Latz E, Mills KH, Natoli G, et al. 2016. Trained immunity: a program of innate
immune memory in health and disease. Science 352:aaf1098
212. Schreiber RD, Old LJ, Smyth MJ. 2011. Cancer immunoediting: integrating immunity’s roles in cancer
suppression and promotion. Science 331:1565–70
213. Page DB, Postow MA, Callahan MK, Allison JP, Wolchok JD. 2014. Immune modulation in cancer with
antibodies. Annu. Rev. Med. 65:185–202
214. Sharma P, Allison JP. 2015. The future of immune checkpoint therapy. Science 348:56–61
215. Teply BA, Lipson EJ. 2014. Identification and management of toxicities from immune checkpoint-
blocking drugs. Oncology 28(Suppl. 3):30–38
216. Dubin K, Callahan MK, Ren B, Khanin R, Viale A, et al. 2016. Intestinal microbiome analyses identify
melanoma patients at risk for checkpoint-blockade-induced colitis. Nat. Commun. 7:10391
217. Santiago-Rodriguez TM, Ly M, Daigneault MC, Brown IH, McDonald JA, et al. 2015. Chemostat
culture systems support diverse bacteriophage communities from human feces. Microbiome 3:58
218. Kim HJ, Li H, Collins JJ, Ingber DE. 2016. Contributions of microbiome and mechanical deformation
to intestinal bacterial overgrowth and inflammation in a human gut-on-a-chip. PNAS 113:E7–15
219. Goldszmid RS, Dzutsev A, Viaud S, Zitvogel L, Restifo NP, Trinchieri G. 2015. Microbiota modulation
of myeloid cells in cancer therapy. Cancer Immunol. Res. 3:103–9
220. Kelly CR, Ihunnah C, Fischer M, Khoruts A, Surawicz C, et al. 2014. Fecal microbiota transplant
for treatment of Clostridium difficile infection in immunocompromised patients. Am. J. Gastroenterol.
109:1065–71
221. Buffie CG, Bucci V, Stein RR, McKenney PT, Ling L, et al. 2015. Precision microbiome reconstitution
restores bile acid mediated resistance to Clostridium difficile.Nature 517:205–8
228 Dzutsev et al.
Changes may still occur before final publication online and in print
Annu. Rev. Immunol. 2017.35. Downloaded from www.annualreviews.org
Access provided by National Institutes of Health Library (NIH) on 02/13/17. For personal use only.
Article
Full-text available
Simple Summary Discovering new oncoviruses is a main goal of virology research. In addition to its deleterious role in immunocompromised patients and during pregnancy leading to birth defects, the human cytomegalovirus’s (HCMV) potential role as an oncogenic agent has garnered significant attention recently. This perspective article focuses on the transforming potential of HCMV based on recently unveiled molecular and cellular characteristics of HCMV-infected cells. Abstract Epstein–Barr virus (EBV), Kaposi sarcoma human virus (KSHV), human papillomavirus (HPV), hepatitis B and C viruses (HBV, HCV), human T-lymphotropic virus-1 (HTLV-1), and Merkel cell polyomavirus (MCPyV) are the seven human oncoviruses reported so far. While traditionally viewed as a benign virus causing mild symptoms in healthy individuals, human cytomegalovirus (HCMV) has been recently implicated in the pathogenesis of various cancers, spanning a wide range of tissue types and malignancies. This perspective article defines the biological criteria that characterize the oncogenic role of HCMV and based on new findings underlines a critical role for HCMV in cellular transformation and modeling the tumor microenvironment as already reported for the other human oncoviruses.
Article
Full-text available
Recent advances in cancer research have unveiled a significant yet previously underappreciated aspect of oncology: the presence and role of intratumoral microbiota. These microbial residents, encompassing bacteria, fungi, and viruses within tumor tissues, have been found to exert considerable influence on tumor development, progression, and the efficacy of therapeutic interventions. This review aims to synthesize these groundbreaking discoveries, providing an integrated overview of the identification, characterization, and functional roles of intratumoral microbiota in cancer biology. We focus on elucidating the complex interactions between these microorganisms and the tumor microenvironment, highlighting their potential as novel biomarkers and therapeutic targets. The purpose of this review is to offer a comprehensive understanding of the microbial dimension in cancer, paving the way for innovative approaches in cancer diagnosis and treatment.
Chapter
During the past decade, significant technological and computational advances have provided unprecedented opportunities for gaining a better understanding of cancer biology and translating this knowledge into meaningful and concrete clinical benefits. Research has made considerable progress in identifying genes and molecular changes that promote cancer initiation and progression. Large-scale genomic and transcriptomic studies involving thousands of patients have prompted the development of sophisticated computational approaches and tools that enable identifying key driver alterations and characterizing their impact on tumor development. This chapter provides an overview of the basic principles that describe the complexity of cancer, which are known as cancer hallmarks, as well as the most relevant computational techniques and tools that have been designed to investigate the genes and genomic changes that contribute to these hallmarks.
Article
Full-text available
Gut microbiota regulates various aspects of human physiology by producing metabolites, metabolizing enzymes, and toxins. Many studies have linked microbiota with human health and altered microbiome configurations with the occurrence of several diseases, including cancer. Accumulating evidence suggests that the microbiome can influence the initiation and progression of several cancers. Moreover, some microbiotas of the gut and oral cavity have been reported to infect tumors, initiate metastasis, and promote the spread of cancer to distant organs, thereby influencing the clinical outcome of cancer patients. The gut microbiome has recently been reported to interact with environmental factors such as diet and exposure to environmental toxicants. Exposure to environmental pollutants such as polycyclic aromatic hydrocarbons (PAHs) induces a shift in the gut microbiome metabolic pathways, favoring a proinflammatory microenvironment. In addition, other studies have also correlated cancer incidence with exposure to PAHs. PAHs are known to induce organ carcinogenesis through activating a ligand-activated transcriptional factor termed the aryl hydrocarbon receptor (AhR), which metabolizes PAHs to highly reactive carcinogenic intermediates. However, the crosstalk between AhR and the microbiome in mediating carcinogenesis is poorly reviewed. This review aims to discuss the role of exposure to environmental pollutants and activation of AhR on microbiome-associated cancer progression and explore the underlying molecular mechanisms involved in cancer development.
Article
Full-text available
The effectiveness of cancer immunization is largely dependent on the tumor’s microenvironment, especially the tumor immune microenvironment. Emerging studies say microbes exist in tumor cells and immune cells, suggesting that these microbes can affect the state of the immune microenvironment of the tumor. Our comprehensive review navigates the intricate nexus between intratumoral microorganisms and their role in tumor biology and immune modulation. Beginning with an exploration of the historical acknowledgment of microorganisms within tumors, the article underscores the evolution of the tumor microenvironment (TME) and its subsequent implications. Using findings from recent studies, we delve into the unique bacterial compositions across different tumor types and their influence on tumor growth, DNA damage, and immune regulation. Furthermore, we illuminate the potential therapeutic implications of targeting these intratumoral microorganisms, emphasizing their multifaceted roles from drug delivery agents to immunotherapy enhancers. As advancements in next-generation sequencing (NGS) technology redefine our understanding of the tumor microbiome, the article underscores the importance of discerning their precise role in tumor progression and tailoring therapeutic interventions. The review culminates by emphasizing ongoing challenges and the pressing need for further research to harness the potential of intratumoral microorganisms in cancer care.
Article
Full-text available
The aim of this study was to assess the impact of WWTP effluents on the sediment microbial communities throughout the Mures , River. This study shows the existence of an ecological equilibrium between the WWTP effluent disruptors and the resilience of the Mures , River sediment microbiomes, a fact that suggests the river's stable/balanced ecological status in this regard, partly due to the microbial communities' resilience to the local impact of WWTP effluents. High-throughput 16S bacterial metabarcoding was used to evaluate the bacterial communities in the sediment. Due to the lotic system's sediment microbial communities' sensitivity to environmental changes, we assumed the dependency of these community structures and functions on environmental abiotic and abiotic parameters. The study results show that, although bacterial communities are equally diverse in the three locations (upstream WWTP, WWTP effluents, and downstream WWTP), there is a difference in community structure between the upstream samples and the WWTP samples, while the downstream samples contain a mixture of the upstream and WWTP effluent communities. Just downstream of the WWTP sediment, microbial communities are influenced by the specific input from the WWTP effluents; nevertheless, the river sediment microbiome is resilient and able to further recover its natural microbial composition, as evidenced by the similarity in bacterial community structures at all upstream river locations. This study demonstrates the ecological equilibrium between the WWTP effluent disruptors and the resilience capacity of the Mures , River sediment microbiomes, a fact that indicates the river's stable/balanced ecological status, in part due to the microbial communities' resilience to the local impact of WWTP effluents. Based on these findings, a monitoring system should be implemented here in the future.
Article
Full-text available
It is nowadays well accepted that chronic inflammation plays a pivotal role in tumor initiation and progression. Under this aspect, the oral cavity is predestined to examine this connection because periodontitis is a highly prevalent chronic inflammatory disease and oral squamous cell carcinomas are the most common oral malignant lesions. In this review, we describe how particular molecules of the human innate host defense system may participate as molecular links between these two important chronic noncommunicable diseases (NCDs). Specific focus is directed toward antimicrobial polypeptides, such as the cathelicidin LL‐37 and human defensins, as well as S100 proteins and alarmins. We report in which way these peptides and proteins are able to initiate and support oral tumorigenesis, showing direct mechanisms by binding to growth‐stimulating cell surface receptors and/or indirect effects, for example, inducing tumor‐promoting genes. Finally, bacterial challenges with impact on oral cancerogenesis are briefly addressed.
Article
Full-text available
Muscle wasting, known as cachexia, is a debilitating condition associated with chronic inflammation such as during cancer. Beneficial microbes have been shown to optimize systemic inflammatory tone during good health; however, interactions between microbes and host immunity in the context of cachexia are incompletely understood. Here we use mouse models to test roles for bacteria in muscle wasting syndromes. We find that feeding of a human commensal microbe, Lactobacillus reuteri, to mice is sufficient to lower systemic indices of inflammation and inhibit cachexia. Further, the microbial muscle-building phenomenon extends to normal aging as wild type animals exhibited increased growth hormone levels and up-regulation of transcription factor Forkhead Box N1 [FoxN1] associated with thymus gland retention and longevity. Interestingly, mice with a defective FoxN1 gene (athymic nude) fail to inhibit sarcopenia after L. reuteri therapy, indicating a FoxN1-mediated mechanism. In conclusion, symbiotic bacteria may serve to stimulate FoxN1 and thymic functions that regulate inflammation, offering possible alternatives for cachexia prevention and novel insights into roles for microbiota in mammalian ontogeny and phylogeny.
Article
Mature circulating polymorphonuclear cells (PMN) have the shortest half- life among leukocytes and undergo rapid programmed cell death in vitro. In this study, we have examined the possibility that inflammatory signals (cytokines and bacterial products) can regulate PMN survival. PMN in culture were found to rapidly die, with percentages of survival at 24, 48, 72, and 96 hours of 97.3% +/- 1.9%, 36.8% +/- 5.3%, 14.5% +/- 3.1%, and 4.2% +/- 2.9%, respectively (mean +/- SE of 20 different donors). PMN incubated with interleukin-1 beta (IL-1 beta), tumor necrosis factor, granulocyte-macrophage colony-stimulating factor (CSF), granulocyte-CSF, and interferon-gamma (IFN-gamma), but not with prototypic chemoattractants (fMLP, recombinant C5a, and IL-8), showed a marked increase in survival, with values ranging at 72 hours of incubation from 89.5% +/- 5.8% for IL-1 beta to 47.6% +/- 6.4% for IFN- gamma. The calculated half-life was 35 hours for untreated and 115 hours for IL-1-treated PMN. PMN activated with lipopolysaccharide (LPS) or inactivated streptococci also showed a longer survival compared with untreated cells (94.4% +/- 3.2% and 95.5% +/- 2.4%, respectively, at 72 hours). PMN surviving in response to LPS or IL-1 beta retained the capacity to produce superoxide anion when treated with phorbol esters or fMLP. All inducers of PMN survival protect these cells from programmed cell death because they reduced cells with morphologic features of apoptosis and the fragmentation of DNA in multiples of 180 bp. Thus, certain cytokines and bacterial products can prolong PMN survival by interfering with the physiologic process of apoptosis. Prolongation of survival may be important for the regulation of host resistance and inflammation, and may represent a crucial permissive step for certain cytokines and microbial products that activate gene expression and function in PMN.
Article
The efficacy of the anti-cancer immunomodulatory agent cyclophosphamide (CTX) relies on intestinal bacteria. How and which relevant bacterial species are involved in tumor immunosurveillance, and their mechanism of action are unclear. Here, we identified two bacterial species, Enterococcus hirae and Barnesiella intestinihominis that are involved during CTX therapy. Whereas E. hirae translocated from the small intestine to secondary lymphoid organs and increased the intratumoral CD8/Treg ratio, B. intestinihominis accumulated in the colon and promoted the infiltration of IFN-γ-producing γδT cells in cancer lesions. The immune sensor, NOD2, limited CTX-induced cancer immunosurveillance and the bioactivity of these microbes. Finally, E. hirae and B. intestinihominis specific-memory Th1 cell immune responses selectively predicted longer progression-free survival in advanced lung and ovarian cancer patients treated with chemo-immunotherapy. Altogether, E. hirae and B. intestinihominis represent valuable “oncomicrobiotics” ameliorating the efficacy of the most common alkylating immunomodulatory compound.
Article
While conventional pathogenic protists have been extensively studied, there is an underappreciated constitutive protist microbiota that is an integral part of the vertebrate microbiome. The impact of these species on the host and their potential contributions to mucosal immune homeostasis remain poorly studied. Here, we show that the protozoan Tritrichomonas musculis activates the host epithelial inflammasome to induce IL-18 release. Epithelial-derived IL-18 promotes dendritic cell-driven Th1 and Th17 immunity and confers dramatic protection from mucosal bacterial infections. Along with its role as a “protistic” antibiotic, colonization with T. musculis exacerbates the development of T-cell-driven colitis and sporadic colorectal tumors. Our findings demonstrate a novel mutualistic host-protozoan interaction that increases mucosal host defenses at the cost of an increased risk of inflammatory disease.
Article
Fusobacterium nucleatum is associated with colorectal cancer and promotes colonic tumor formation in preclinical models. However, fusobacteria are core members of the human oral microbiome and less prevalent in the healthy gut, raising questions about how fusobacteria localize to CRC. We identify a host polysaccharide and fusobacterial lectin that explicates fusobacteria abundance in CRC. Gal-GalNAc, which is overexpressed in CRC, is recognized by fusobacterial Fap2, which functions as a Gal-GalNAc lectin. F. nucleatum binding to clinical adenocarcinomas correlates with Gal-GalNAc expression and is reduced upon O-glycanase treatment. Clinical fusobacteria strains naturally lacking Fap2 or inactivated Fap2 mutants show reduced binding to Gal-GalNAc-expressing CRC cells and established CRCs in mice. Additionally, intravenously injected F. nucleatum localizes to mouse tumor tissues in a Fap2-dependent manner, suggesting that fusobacteria use a hematogenous route to reach colon adenocarcinomas. Thus, targeting F. nucleatum Fap2 or host epithelial Gal-GalNAc may reduce fusobacteria potentiation of CRC.
Article
The innate immune system distinguishes low-level homeostatic microbial stimuli from those of invasive pathogens, yet we lack understanding of how qualitatively similar microbial products yield context-specific macrophage functional responses. Using quantitative approaches, we found that NF-κB and MAPK signaling was activated at different concentrations of a stimulatory TLR4 ligand in both mouse and human macrophages. Above a threshold of ligand, MAPK were activated in a switch-like manner, facilitating production of inflammatory mediators. At ligand concentrations below this threshold, NF-κB signaling occurred, promoting expression of a restricted set of genes and macrophage priming. Among TLR-induced genes, we observed an inverse correlation between MAPK dependence and ligand sensitivity, highlighting the role of this signaling dichotomy in partitioning innate responses downstream of a single receptor. Our study reveals an evolutionarily conserved innate immune response system in which danger discrimination is enforced by distinct thresholds for NF-κB and MAPK activation, which provide sequential barriers to inflammatory mediator production.
Article
Neutrophil migration and its role during inflammation has been the focus of increased interest in the past decade. Advances in live imaging and the use of new model systems have helped to uncover the behaviour of neutrophils in injured and infected tissues. Although neutrophils were considered to be short-lived effector cells that undergo apoptosis in damaged tissues, recent evidence suggests that neutrophil behaviour is more complex and, in some settings, neutrophils might leave sites of tissue injury and migrate back into the vasculature. The role of reverse migration and its contribution to resolution of inflammation remains unclear. In this Review, we discuss the different cues within tissues that mediate neutrophil forward and reverse migration in response to injury or infection and the implications of these mechanisms to human disease.
Article
Training immune cells to remember Classical immunological memory, carried out by T and B lymphocytes, ensures that we feel the ill effects of many pathogens only once. Netea et al. review how cells of the innate immune system, which lack the antigen specificity, clonality, and longevity of T cell and B cells, have some capacity to remember, too. Termed “trained immunity,” the property allows macrophages, monocytes, and natural killer cells to show enhanced responsiveness when they reencounter pathogens. Epigenetic changes largely drive trained immunity, which is shorter lived and less specific than classical memory but probably still gives us a leg up during many infections. Science , this issue p. 10.1126/science.aaf1098