You are on page 1of 23

Journal of Autoimmunity xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Journal of Autoimmunity
journal homepage: www.elsevier.com/locate/jautimm

Connecting the immune system, systemic chronic inflammation and the gut
microbiome: The role of sex
Lisa Rizzettoa,∗, Francesca Favaa, Kieran M. Tuohya, Carlo Selmib,c
a
Department of Food Quality and Nutrition, Research and Innovation Centre, Fondazione Edmund Mach, San Michele all’Adige, Trento, Italy
b
Division of Rheumatology and Clinical Immunology, Humanitas Research Hospital, Rozzano, Italy
c
BIOMETRA Department, University of Milan, Italy

A R T I C LE I N FO A B S T R A C T

Keywords: Unresolved low grade systemic inflammation represents the underlying pathological mechanism driving immune
Autoimmunity and metabolic pathways involved in autoimmune diseases (AID). Mechanistic studies in animal models of AID
Gut microbiota and observational studies in patients have found alterations in gut microbiota communities and their metabo-
Bile acids lites, suggesting a microbial contribution to the onset or progression of AID. The gut microbiota and its meta-
SCFA
bolites have been shown to influence immune functions and immune homeostasis both within the gut and
Microbial metabolism
systematically. Microbial derived-short chain fatty acid (SCFA) and bio-transformed bile acid (BA) have been
shown to influence the immune system acting as ligands specific cell signaling receptors like GPRCs, TGR5 and
FXR, or via epigenetic processes. Similarly, intestinal permeability (leaky gut) and bacterial translocation are
important contributors to chronic systemic inflammation and, without repair of the intestinal barrier, might
represent a continuous inflammatory stimulus capable of triggering autoimmune processes. Recent studies in-
dicate gender-specific differences in immunity, with the gut microbiota shaping and being concomitantly shaped
by the hormonal milieu governing differences between the sexes. A bi-directional cross-talk between microbiota
and the endocrine system is emerging with bacteria being able to produce hormones (e.g. serotonin, dopamine
and somatostatine), respond to host hormones (e.g. estrogens) and regulate host hormones' homeostasis (e.g by
inhibiting gene prolactin transcription or converting glucocorticoids to androgens). We review herein how gut
microbiota and its metabolites regulate immune function, intestinal permeability and possibly AID pathological
processes. Further, we describe the dysbiosis within the gut microbiota observed in different AID and speculate
how restoring gut microbiota composition and its regulatory metabolites by dietary intervention including
prebiotics and probiotics could help in preventing or ameliorating AID. Finally, we suggest that, given consistent
observations of microbiota dysbiosis associated with AID and the ability of SCFA and BA to regulate intestinal
permeability and inflammation, further mechanistic studies, examining how dietary microbiota modulation can
protect against AID, hold considerable potential to tackle increased incidence of AID at the population level.

1. Introduction an important contribution from environmental factors. A number of


recent studies have highlighted the influence of genetics and environ-
The incidence of autoimmune diseases (AID) has increased sig- mental factors [1], such as antibiotics [2], dietary habits [3,4] and sex
nificantly in Westernized countries over the past decades, with a higher hormones [5] in AID. The host immune system plays an important role
prevalence in women (∼80% of overall incidence). The prevailing in shaping the gut microbiota and reciprocally, host-associated micro-
hypothesis suggests unresolved systemic inflammation as a key factor in organisms significantly influence the development and function of in-
disease emergence and progression. In genetically predisposed subjects nate and adaptive immunity [6,7], by establishing a “tolerant” pheno-
unchecked systemic inflammation leads to the formation of auto- type and therefore facilitating the continuation of host-microbe co-
antibodies, physiological dysregulation and tissue damage which can existence [8]. Similarly, the gastrointestinal microbiota has recently
eventually manifest as specific AID. been highlighted as one of the major contributors to the regulation of
While AID has a strong genetic basis, increasing evidence indicates immune effector cell maturation and activity, and its dysbiosis has been


Corresponding author. Department of Food Quality and Nutrition, Research and Innovation Centre, Fondazione Edmund Mach, via Edmund Mach 1, 38010 San Michele all’Adige,
Trento, Italy.
E-mail address: lisa.rizzetto@fmach.it (L. Rizzetto).

https://doi.org/10.1016/j.jaut.2018.05.008
Received 3 April 2018; Received in revised form 18 May 2018; Accepted 21 May 2018
0896-8411/ © 2018 Elsevier Ltd. All rights reserved.

Please cite this article as: Rizzetto, L., Journal of Autoimmunity (2018), https://doi.org/10.1016/j.jaut.2018.05.008
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

shown to contribute to intestinal mucosa permeability, induction of Through binding specific receptors in immune cells, sex hormones
innate defenses, and thus a candidate environmental risk factor capable such as estrogens, can influence immune function, such as enhancing
of triggering AID [3,9–11]. In addition, mucosal immune function and the humoral responses as in the case of 17-β-estradiol and prolactin [5].
susceptibility to chronic inflammation differs between sexes [12,13]. It has been suggested that male hormones are protective in SLE or T1D
Aberrant microbiota community structure has been reported in and estrogens may contribute to disease progression by promoting B
different AID including asthma [14], inflammatory bowel disease (IBD) cell survival [12]. In lupus-prone mice, ERα deficiency leads to auto-
[15], type 1 diabetes (T1D) [16], rheumatoid arthritis [17], food al- antibody production ameliorates renal damage and prolongs survival
lergies [18], systemic lupus erythematosus (SLE) [11] and multiple compared to ERα-sufficient controls [32]. Knocking out ERα in both
sclerosis (MS) [19]. The human microbiome provides key immune and lupus-prone and control mice resulted in impaired TLR4 activation in
metabolic regulatory services to the host, which are set early in life and immune cells, indicating that estrogen and ER signaling can influence
contribute to life-long immune function. However, environmental fac- TLR4 responsiveness [31]. ERα deficient mice and lupus-prone ERα
tors like nutrition, medicines, hormonal changes and psychological deficient mice also showed a decreased TLR9-mediated inflammatory
stress can disrupt the gut microbiota and determine its influence on cytokine production by DCs [33], particularly interleukin (IL)-6, MCP-
immune homeostasis. The contribution of diet and the relevance of 1, IL-1β and IL-23, a cytokine milieu modulating the IL-23/IL-17
microbially-derived dietary metabolites are only just starting to be pathway and the priming of the T helper (Th) 17 response. Further-
appreciated [8,20]. more, decreased percentages of IL-17+ RORγt+ cells were isolated from
This review will present evidence supporting a role for diet:microbe the spleen of ERα−/−lupus prone mice [33]. The protective effects of
interactions in the emergence of AID, with a special focus on extra- estrogens, particularly 17-β-estradiol, has been described in the onset
intestinal AID, such as T1D, SLE and MS. We will highlight connections and progression of experimental autoimmune encephalomyelitis (EAE),
with the gut microbiota and its cell signaling metabolites which influ- the prototypic animal model of MS, by enhancing regulatory B cells and
ence inflammation systemically or provide a pathological pressure for M2 macrophage activities, and prevents the dysbiosis associated to EAE
the onset and maintenance of AID via continuous exposure to in- [34]. By contrast, endocrine disrupting chemicals such as bisphenol A,
flammatory triggers by compromising the intestinal barrier. The con- have been shown to be associated with risk of T1D in non-obese dia-
tribution of the estrogen-gut microbiota axis, the bi-directional mod- betic mice, a mouse model of insulitis and leukocytic infiltration of
ulation existing between gut microbiota and sex hormones, and the pancreatic islets leading to T1D [35–37] as well as with SLE patho-
gender bias in autoimmunity will be also discussed. genesis, as evidenced in MRL/lpr mice, the animal model of lupus ne-
phritis [38,39].
2. Impact of gut microbiota on gender-specific differences in These evidences suggest a link between estrogen setting and its
immunity and onset of autoimmunity impact in regulating immune function through activation of specific
microbial recognizing receptors (Fig. 1).
Males and females present some differences in the immune system
[12] and sex bias is an important aspect of many AID [12,21,22]. The 2.2. Sex hormones and their impact on microbiota
increasing number of microbiome studies is revealing a bi-directional
cross-talk between microbiota and the endocrine system is emerging Host hormones can affect bacterial gene expression, bacterial viru-
with bacteria being able to produce hormones (e.g. serotonin, dopa- lence and growth, with consequences on host physiology [24]. The sex
mine and somatostatine), respond to host hormones (e.g. estrogens) and hormones estriol and estradiol decrease bacterial virulence by in-
regulate host hormones' homeostasis by inhibiting gene transcription hibiting quorum sensing signaling [40] while progesterone has been
(e.g prolactin) or converting them (e.g. glucocorticoids to androgens) shown to promote growth of oral Bacteroides species and Prevotella in-
[23,24]. termedius [41]. Changes in expression of ER-β ERβ affects the compo-
sition of gut microbiota of female mice and that microbiota enriched
2.1. Sex differences in immunity and pathogen sensing from differential ERβ expression respond differently to changes in diet
complexity [42]. By using high-throughput sequencing, sex hormones
In general, normal healthy women show greater antigen presenting have been shown to modulate mammalian microbiota composition
activity and mitogenic responses, higher immunoglobulin levels and [22,43–49] (Fig. 1). Healthy male mice have been shown to bear higher
more enhanced antibody production than males, being more resistant abundance of Ruminococcus, Coprococcus and Dorea and lower abun-
to infection diseases and much more likely to develop AID, while dance of Porphyromonaceae and Rikenella, while female mice present
healthy male are thought to have an immune system that maintain higher abundance of Allobaculum, Anaeroplasma, and Lactobacillaceae
tolerance [25]. Generally, steroid hormones exert an opposite role on and Veilonellaceae [22,46,49]. Mueller at al. (2006) showed that
the immune response, with estrogens promoting humoral immunity by healthy male subjects had a higher abundance of Bacteroides-Prevotella
enhancing B cell survival and androgens and progesterone acting in than female while post-menopausal woman microbiota did not differed
prevention or suppression of autoimmunity [26,27]. from male one [47]. Very recently, Fransen et al. (2017) showed a
Accumulating evidence has linked Toll like receptor (TLR) function decreased relative abundance of Alistipes, Rikenella and Porphyr-
to estrogen and estrogen receptor α (ERα) and β (ERβ), suggesting a omonadaceae in female mice associated with a concomitant higher cell
possible contribution of different microbial sensing and TLR activation content in the thymus, a bigger thymus size and an increased IFN-β
to sex bias in AID, likely linked to gut microbe-associated molecular signaling than in male mice [43]. However, whether sex-dependent
patterns (MAMPs) recognition [28]. Several immune-related genes in- immunity differences are a cause or consequence of an alteration of the
cluding TLR7/8 - involved in nucleic acid recognition - and FOXP3 – the gut microbiota is still not clear. Gut microbiota analysis on 341 female
master transcriptional regulator of Treg expansion -, are present on X and 348 male mice from 89 inbred animals demonstrated that under a
chromosome [1] and incomplete inactivation of X chromosome in fe- controlled environment male and female mice show significant differ-
male can potentially lead to increased activation of those genes [21,29]. ences in gut microbiota composition, although genetic differences ob-
The balance among TLR7/8 and 9, signaling receptors detecting mi- scured sex differences in examination of the entire population [22].
crobial- and self-nucleic acids, has been shown to be crucial to SLE Furthermore, gonadectomy and hormone replacement induced a mi-
onset in animal models [30], as discussed below. Moreover, Jiang et al. crobiota shift driven by sex hormones. Those treatments led also to
found that female SLE patients possess more active monocytes with gender-specific differences in bile acid (BA) profiles between sexes
enhanced TLR4 receptiveness than male SLE patients and present in- [22]. An higher rate of BA synthesis and BA pool sizes in females than in
creased plasma levels of soluble CD14 upon in vitro LPS challenge [31]. males has been previously reported [50,51]. This is important given the

2
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Fig. 1. Schematic representation of the interconnections occurring among immune system, gut microbiota and sex hormones. Even more evidence are
indicating the establishment of a tight reciprocal regulation between immune functions, sexual hormones and gut microbiota. Gut microbiota composition contribute
to the immune homeostasis by regulating through its composition and its metabolites immune system functions and are concomitantly shaped by the immune system.
Studies in mammals have shown that sex hormones can also modulate microbiota composition and indeed could potentially contribute to the sex bias observed in
autoimmunity. Interestingly, gut microbiota metabolizes estrogens by secreted β-glucuronidase leading to estrogen activation and their translocation into the
bloodstream for reaching distal sites. The alteration of such interconnected regulation contributes to chronic inflammation and the onset of autoimmunity. Data
reported on microbial abundance are taken from mice data [22,43–47,49]. BA, bile acid; IL, interleukin; iTreg, inducible regulatory T cells; MΦ, macrophages, NK,
natural killer; regB, regulatory B cells; SCFA, short chain fatty acid; Th, T helper cells; Treg, regulatory T cells.

role of BA in immune cell signaling and the relative difference in cell binding to their receptors and inducing of their physiological activities
signaling potential between different BA species/chemical structure (Fig. 1). In particular, Clostridium scindens, one of the species involved
involved in pathways responsible for regulating inflammation via TGR5 on the bio-transformation of BA [56] has been also shown to convert
and FXR signaling (see below). A variety of tissues express estrogen glucorticoids to androgens [57].
receptors, including, gut, brain, bone and adipose tissue [52]. Gut The important role of gut microbiota in driving a gender bias in
epithelial barrier integrity can also be modified by estrogen as observed autoimmunity is supported by two recent studies which connect hor-
in mice where females are more resistant to gut injury compared to monal influences and microbiota to explain the sexual dimorphism of
their male counterparts [53]. This indication and the observation that autoimmunity [44,45]. Previous findings showed that germ free ani-
gut microbiota composition differs between males and females could mals lose the sexual dimorphism of T1D, with both females and males
potentially contribute to the sex bias observed in autoimmunity. having a high incidence of the disease. Both studies found that puberty
affects the male microbiota composition, which becomes less diverse
than the female microbiota and that the microbiota contributed to in-
2.3. Microbiota regulation of sex hormones creased levels of testosterone in the blood. Importantly, post-pubescent
female and castrated male microbiota were more similar than to that of
It appears that not only is the gut microbiota influenced by estro- post-pubescent males, revealing that sex hormones affected microbiota
gens but also the gut microbiota itself impacts on estrogen levels, and composition. Markle et al. [44] found that T1D can be suppressed by
may be an important regulator of circulating estrogen and estrogenic transferring gut microbiota from male mice into immature females and
molecules [54] (Fig. 1). Use of antibiotics has been shown to lower suggested microbiome manipulations can provoke testosterone-depen-
estrogen levels and urinary estrogen levels has been correlated with a dent protection from autoimmunity. In parallel, Yurkovetskiy et al. [45]
higher abundance of Clostridia and Ruminococcaceae [55]. The gut suggested that both hormones and microbiota were required to ensure
microbiota secretes β-glucuronidase, an enzyme that carries out de- male protection from T1D. Despite the different interpretation of the
conjugation of estrogens. The estrogen active deconjugated forms are in results, the two studies agree that the microbiota regulates the levels of
this way facilitated in entering the bloodstream and subsequently in

3
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

androgens in the blood. Interestingly, in germ free mice, male protec- females [77]. Moreover, the autoimmune regulator (AIRE), a negative
tion (but not female protection) was achieved by mono-colonization regulator of autoimmunity is differentially expressed in males and fe-
with segmented filamentous bacteria [114]. males. In human and mouse thymus, females express less AIRE (mRNA
Whether microbes contribute to sexual dimorphism in other auto- and protein) than males after puberty. AIRE expression has also been
immunity models remains to be examined. Investigation of the role of demonstrated to be dependent on sexual hormones, as male castration
the microbiota in biasing autoimmunity in humans also needs to be decreases AIRE thymic expression and estrogen treatment down-
investigated, especially taking into account the possible role of micro- regulates AIRE expression in cultured human thymic epithelial cells
bial-related signaling molecules which affect the endocrine/hormones [78]. Noteworthy, gut microbiota may alter host histone acetylation
system, such as BA or indeed estrogenic metabolites like equol, en- and methylation in different tissue such as colon, liver and white adi-
terolactone and enterodiol, produced by gut bacteria from dietary pose tissues [79,80] and microbiota-induced chromatin changes have
phytoestrogens [58,59]. been shown to be responsive to diet [80]. Moreover, AIRE expression is
reduced in germ-free mice [81]. Gut microbial metabolites, such as B
3. Microbial regulation of host immune responses vitamins, mainly produced by bifidobacteria in human gut [82], mi-
crobial bio-transformation of dietary polyphenols [83] and short chain
Microbiota can affect both innate and adaptive immunity by direct fatty acids (SCFA) [80,84], have been all shown to contribute sig-
contact to immune cells, by inducing epigenetic modification and nificantly to epigenetic processes. Indeed, restoring an aberrant mi-
through the production of signaling molecules. crobiota and microbial fermentation end products, especially SCFAs,
may be particularly important for regulation of inflammatory reactions,
3.1. Immune signaling via microbial sensing as described below.

The mammalian immune system recognizes MAMPs via pathogen 3.3. Microbial regulation of host immune responses via the synthesis of cell
recognition receptors like TLRs and NOD like receptors (NLRs) ex- signaling metabolites
pressed on immune and non-immune cells. Dendritic cells (DCs), mac-
rophages, neutrophils and natural killer (NK) cells, important players in Gut microbial biotransformation of exogenous and endogenous
the innate immunity, are necessary for the development of many types compounds generates metabolites that play an important role in im-
of autoimmunity and microorganisms can influence the activity of these mune homeostasis. Among these, microbial fermentation end-products
cells. Macrophages isolated from gut microbiota-depleted mice release SCFA and microbially transformed bile acids have been demonstrated
lower levels of TNF-α and higher levels of IL-10 when stimulated with to influence mucosal immune function, immune tolerance and regula-
inflammatory bacterial cell wall component lipopolysaccharide (LPS) tion of inflammation both within the gut and systemically [85]. In the
compared to macrophages from specific pathogen free (SPF) animals immune system-microbial cross-talk, these metabolites regulate the
[11]. Moreover, LPS activation of splenic DCs has been shown to de- differentiation and function of almost every cell type in the gut lym-
crease the expression of IL-15, IL-6, TNF-α and type I interferons in phoid compartments [86] and regulate intestinal permeability [87].
germ free mice [60]. NK function is also compromised in the absence of These metabolites can also circulate systemically to affect immune cells
the gut microbiota, possibly due to the inability of DCs to produce type I in peripheral tissues [88].
interferons in response to microbial stimuli [61]. The proper function
and activity of bone marrow-derived neutrophils is highly influenced by 3.3.1. SCFA
exposure to commensal microorganisms and their circulating cell wall SCFA, primarily acetate, butyrate and propionate are the main gut
components [62]. Inducible NK T cells, which also contribute to several microbial end products of carbohydrate fermentation (Box 1). SCFA are
AID, are also impaired in the germ free animals and require microbial known to affect several cellular processes including gene expression,
ligands for their maturation [61,63]. Germ free mice are also reported chemotaxis, differentiation, proliferation and apoptosis, all of which
to have distinct alterations in their adaptive immunity including de- play important roles in inflammation [20,89]. SCFA represents an im-
creased numbers of T and B cells, decreased levels of IgA and IgG an- portant energy source for intestinal epithelium (∼50% of energy re-
tibodies, and a strong skewing towards the CD4+ T cell helper (Th) 2 quirement) [90], act as ligands for activation of cellular signaling cas-
population. Introduction of microorganisms restores the Th1 and Th17 cades that carry their effects to districts other than the intestine and are
subsets, normally not present in germ free mice [11]. Colonization by involved in the regulation of physiological functions such as lipid me-
the segmented filamentous bacteria (SFB) Candidatus arthromitus en- tabolism, thermogenesis, fat storage in liver, muscle, and adipose tissue,
hances the expansion of intestinal Th17 cells through adhesion to gut and blood pressure regulation [91].
epithelial cells [64–66]. Germ-free animals harboring C. arthromitus The anti-inflammatory and immuno-modulatory effects of SCFAs
alone develop EAE showing that gut bacteria can affect neurologic in- involve both the activation of the specific cell receptors GPR109a (also
flammation since SFB-induced Th17 cells in the lamina propria can known as HCA2), GPR41 (also referred as FFAR3) and GPR43 (FFAR2)
recirculate to the brain causing inflammation [67]. Therefore, an [92–95], and intracellular targets via inhibiting histone deacetylase
aberrant microbiota can imbalance immune homeostasis and the in- (HDACs) activity [96–99] (Box 1). Mice deficient in GPR43 develop
duced autoreactive T cell could migrate in different organs, from gut to colitis, arthritis, and asthma, demonstrating a direct effect of SCFA on
brain as in the case of MS or to liver as in the case of autoimmune colonic T regulatory (Treg) cells through the increased gene expression
cholestatic liver disease [68,69], or kidney as in the case of SLE [70], of GPR43 [92]. Butyrate engagement of GPR109a is known to affect the
thus exacerbating specific organ inflammation. pro-inflammatory activities in colonic phagocytes, enhance Treg dif-
ferentiation and IL-10 production and reduce the colonic epithelial
3.2. Microbiota and epigenetic control secretion of IL-18, a cytokine known to regulate colonic homeostasis
[94]. SCFAs also promote inducible colonic Treg cells number and
Epigenetic dysregulation, including histone modification, DNA function [96,97] through epigenetic processes by inhibiting HDAC ac-
hypo- and hyper methylation, has been recently shown to play a pivotal tivity [96–99]. In particular, butyrate enhances histone H3 acetylation
role in autoimmunity affecting the activity of several immune cells in in the intronic conserved non-coding sequences of the FoxP3 locus,
T1D [71] and SLE [72]. Sex hormones can regulate immunity and au- consequently promoting extrathymic generation of Tregs [96]. More-
toimmunity also through epigenetic mechanisms, such as regulation of over, IL-10-producing inducible Tregs develop after TGF-β cytokine
miRNA expression [73] and DNA methylation status [74–76]. Healthy exposure in the periphery from naive CD4+ T cells [100]. Adoptive
males have a tendency to bear higher methylation levels than healthy transfer of CD4+CD45RBhi naïve T cells into Rag1−/− mice showed

4
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Box 1
SCFA pool and the gut microbiota

SCFA are produced mainly through saccharolytic fermentation of carbohydrates that escape digestion and absorption in the small intestine
and the pathways of SCFA production are relatively well understood and recently described in detail [303,304]. The major products are
acetate, propionate and butyrate. Total SCFAs in gut lumen is ∼100 mM, in blood is, portal ∼400 μM and peripheral ∼100 μM [305].
There is a strong biological gradient for each SCFA from the gut lumen to the periphery which leads to differing cell and tissue SCFA
exposure and therefore signaling potential. There is a significant reduction in butyrate, relative to acetate and propionate, across the gut
epithelium and also the significant reduction in propionate relative to acetate across the liver in humans [306–308]. Hence, SCFAs exhibit
the tissue specific physiological functions according to this concentration gradient. SCFA mainly exert their action through activation of G
protein coupled receptors (GPCRs), as well as epigenetic activities via the inhibition of histone deacetylases (HDACs), stimulation of histone
acetyltransferase activity and stabilization of the hypoxia-inducible factor (HIF) [98,309]. GPCRs show distinct patterns of expression and
they have been partially associated with the effects of the SCFAs on innate and adaptive immune cells and intestinal epithelial cells [310].
SCFA show differential abilities to activate the different signaling receptors [89]. GPR43 is mainly activated by acetate and propionate
followed by butyrate [311,312], while GPR41 si activated mainly by propionate and butyrate [311,312]. GPR109A, firstly identified as a
receptor for niacin, is also activated by β-hydroxybutryate and butyrate, but not by acetate and propionate [313]. Olfactory receptor 78
(OLFR78) was identified as SCFAs in a ligand screen for orphan GPCRs being it activated by acetate and propionate but not by butyrate
[314]. Circulating and gut SCFA pools can influence the immune system homeostasis by differently interacting with these signaling re-
ceptors at different body sites. Diverse bacteria are responsible for SCFA production. Acetate production pathways are widely distributed
among bacterial genera whereas pathways for propionate and butyrate production appear more highly conserved and substrate specific.
Overall, SCFAs, especially butyrate, seem to regulate several host functions exerting broad anti-inflammatory activities as well as affecting
cellular processes such as proliferation, activation, and apoptosis also through epigenetic effect. Butyrate has significantly greater HDAC
inhibiting potential, an important epigenetic process, compared to propionate and acetate [97,315,316]. For instance, de novo Treg cell
expansion through HDAC inhibition is promoted by butyrate and secondly by propionate, but not acetate [97]. DCs briefly exposed to
butyrate, and (to a lesser extent) propionate, but not acetate, potently induce Foxp3 expression [97]. Similarly butyrate and propionate
stimulate lipolysis, while acetate and aminobutyric acid have little or no effect [317]. Moreover, butyrate, but not propionate and acetate
are able to promote retinoic acid production via class 3 HDAC (HDAC3), but not HDAC2, inhibition [98]. Differently, it has been reported
that acetate could affect inflammation and host physiology both acting as HDAC2 inhibitor and promoting histone H3 and H4 acetylation
[318–320]. Altogether this evidence suggests that SCFAs are signaling molecules acting through different receptors and via different forms
of epigenetic regulation.

their conversion into Treg cells when mice were fed a butyrylated diet shown to contribute to mucosal homeostasis by affecting cytokine and
[100]. Propionate and acetate were similarly able to enhance periph- chemokine secretion of intestinal epithelial cells and by increasing the
eral Treg generation [96]. Indeed, it is likely that commensal bacterial production of retinoic acid through enhancement of aldehyde dehy-
species promote inducible Tregs in the gut [101] through SCFA release. drogenase expression [98]. Butyrate also reduces levels of pro-in-
SCFA anti-inflammatory effects are not limited to adaptive immune flammatory cytokines IL-6, TNF-α as well as the LPS binding receptor
response. DCs exposed to butyrate suppress the conversion of naïve T TLR4, concomitantly restoring gut mucosa integrity and potentially
cells into IFN-γ producing cells [102]. Through GPR41 binding, pro- reducing bacterial translocation in an experimental mucositis model
pionate has also been shown to affect DCs properties in the bone [113]. Attenuation of liver injury and subsequent inflammatory re-
marrow by increasing their phagocytic capacity and limiting their sponses following increased level of butyrate has also been observed
ability to prime a Th2 response in the airways [95]. SCFA have also [114]. Recently, Xia et al. reported [108] that continuous oral admin-
been shown to increase both intestinal IgA and systemic IgG responses istration of propionate during lactation, by increasing its concentration
supporting plasma B cell differentiation. Mice under low dietary fiber in the proximal colon, contributes to maintenance of mucosal barrier
regimes and consequent low SCFA production, have been shown to be function by promoting the expression of Zonulin-1, Claudin-8 and 1 and
defective in antibody homeostasis and to be greatly susceptible to in- Occludin through the expression of MAPKs in rats [108].
fection [103]. Thus, the evidence so far indicates that SCFA can strongly impact
Animal studies have related dietary induced aberrant gut micro- both the regulation of innate and adaptive immune systems, and in-
biota composition to inflammation through decreased intestinal barrier testinal permeability, all of which are commonly altered in AID. It is
function and increased intestinal permeability [104,105]. Increased noteworthy that Western-dietary habits bear about 7–10 times less
intestinal permeability has been previously shown to trigger low grade dietary fiber intake than ancient or traditional diets such as the
chronic systemic inflammation in metabolic diseases and extra-in- Mediterranean diet [115]. This lower fiber intake translates in lower
testinal inflammation [91], [106,107]. The combination of disrupted availability of substrate for microbial fermentation and therefore SCFA
barrier function and lack of dietary support for microbiota immuno- production (Box 1). Diet, and especially dietary fiber has the ability to
modulatory metabolite production could be an important contributor to influence both the source of inflammatory triggers through regulation
AID risk. SCFA can dose-dependently suppress intestinal permeability of intestinal permeability and the response to such inflammatory trig-
[98,108]. Butyrate in particular has been shown to suppress the per- gers by DC and Treg maturation.
meability of isolated colonic mucosa epithelium [109] and in Caco-
2 cell monolayers [110], by influencing tight junction proteins (TJ)
expression (i.e increasing Claudin-5, Claudin-7, Zonulin-1 and de- 3.3.2. Bile acids
creasing Claudin-2) and increasing trans-epithelial electrical resistance BA are cholesterol-derived molecules produced in the liver to fa-
in vitro. Histological studies have shown that butyrate is also able to cilitate the absorption of dietary lipids and fat-soluble vitamins and
alleviate intestinal injury following peritonitis induction in mice [111] maintain systemic cholesterol homeostasis. They also act as immune
and mitigate experimental colitis [112] by improving barrier function, regulators. BAs act as signaling molecules with hormone-like actions via
thus acting as an anti-inflammatory agent. Butyrate has also been dedicated BA receptors such as the nuclear receptor farnesoid X re-
ceptor (FXR/NR1H4) and the membrane-bound G protein-coupled

5
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Box 2
Microbial modulation of the BA pool

Primary bile acids (BAs) are synthesized in the liver from cholesterol and released into the small intestine in response to food intake to aid
lipid absorption and cholesterol catabolism. Newly-synthesized BAs are termed primary BAs, namely chenodeoxycholic acid and cholic
acid. Before active secretion, primary BAs are conjugated with the amino acids taurine or glycine, a fundamental amidation making BAs less
hydrophobic, less cytotoxic, and more readily secretable into bile. During the enterohepatic circulation, conjugated chenodeoxycholic acid
and cholic acid are then transformed by gut microbial enzymes through deconjugation and dihydroxylation reactions to generate un-
conjugated and secondary bile acids respectively [117]. In particular, the major microbial biotransformations include: hydrolysis of
conjugated bile acids to free bile acids and glycine or taurine by BSH activity; 7β-dehydroxylation of cholic acid and chenodexycholic acid,
leading to deoxycholic acid and litocholic acid production, respectively; bile acid 7β-dehydroxylation of ursodeoxycholic acid generating
LCA. Clostridia as well as other Gram positive gut bacteria have the most diverse bear BSH activity such as Enterococcus, Bifidobacterium,
and Lactobacillus [321]. So far, among the Gram-negative bacteria, a BSH was purified only from Bacteroides fragilis subsp. fragilis [322].
BSH activity by lactobacilli and bifidobacteria in the small intestine is particularly important in this process, making the intestinal BA pool
more available for transformation into secondary and thirtary structures by other members of the gut microbiota. The bile acid 7α/7β-
dehydroxylations involve a multistep biochemical pathway found only in the anaerobic gut bacteria, mainly Clostridia, C. perfringens and C.
scindens amongst others [56,321,323]. In addition, the gut microbiota are capable of the oxidation and epimerization of hydroxy groups at
the C3, C7 and C12 position of bile acids yielding isobile (β-hydroxy) acids.
Circulating BA pool is indeed strongly dependant on the profile of BA-converting bacteria. This dependency was firstly apparent in germ
free animals. Germ free mice show a strong increase in muricholic acid, levels compared to conventional animals, but not cholic acid.
Antibiotic treatment of conventional animals increases muricholic acid and reduces the levels of plasma secondary BAs indicating that
commensal bacteria modulate the size of the BA pool as well as its composition [324]. This bacterial activity is even more evident in disease
[117,282,325,326]. For instance, IBD patients, bearing an altered gut microbiota, often characterized by a reduction in Faecalibacterium
prausnitzii, present changes in BA profiles compared to healthy subjects [327]. Moreover, primary sclerosing cholangitis severity is ex-
acerbated in germ free mice and in conventional animals replete with intestinal microbiota correlates with decreasing level of urso-
deoxycholic acid, a microbial produced tertiary BA [328]. Transformation of conjugated primary BA cholic acid into the secondary BA
deoxycholic acid by the gut microbiota has been shown to determine the percentage of deoxycholic acid in blood [329]. Importantly, both
deoxycholic acid and ursodeoxycholic acid, produced by the gut microbiota, respond to changes in diet in animal models and are closely
linked to profiles of gut bacteria and their BSH and 7α/β-hydroxylation capabilities. Recent studies in animal models have shown that
probiotic BSH activities [330], polyphenols [331] and prebiotic fibers like oat β-glucans [332,333] can modulate the gut microbiota and
the profile of BA returning through the enterohepatic circulation. Dietary change has been shown to alter both fecal BA profiles and the
ability of the gut microbiota to generate secondary and tertiary BA [334].
Therefore, microbiota modulation through the diet might be exploited to modulate microbially-driven changes in BA pool both in terms
of BA chemical structures and in circulating concentrations of BA. Furthermore, BA themselves also have a direct modulatory affect upon
the gut microbiota [335], with taurocholate for example increasing the relative abundance of bacteria associated with inflammation in
animal models of IBD.
The importance of the composition of the BA pool is evident as BA show differential abilities to activate FXR and TGR5, as nicely
described in Refs. [336,337]. BAs are amphipatic molecules which structure characterized by a concave hydrophilic face (α face) and a
convex hydrophobic face (β face), which determined the different binding to the bile-acid receptors [336]. TGR5 showed a dose-dependent
activation by BA, with the following rank order of potency: lithocholic acid ≥ deoxycholic acid > chenodeoxycholic acid > cholic acid.
Similarly, the different BA show different affinity to FXR. The hydrophobic pocket of the FXR interacts with BAs mainly through the β
face. The α face contains several hydroxyl groups in the 3, 7 and/or 12α positions that impacts on the ability to activate FXR [338]. The 7α-
hydroxy group, conferred by gut microbiota activity, confers high capacity to chenodeoxycholic acid and synthetic ligands, such as fex-
aramine and obeticolic acid, to activate FXR. Conversely, the absence of a 7α-hydroxy group, in the case of deoxycholic acid and lithocholic
acid, compromises stable coactivator recruitment and results in partial agonistic properties. Finally, the FXR binding site cannot accom-
modate a 12α-hydroxy group and consequently, cholic acid and chenodeoxycholic acid have low affinity for FXR. Ursodeoxycholic acid has
been described as an antagonist, able to inhibit activation by chenodeoxycholic acid [339]. In line with these observation FXR showed a
dose-dependent activation by BA, with the following rank order of potency: chenodeoxycholic acid > deoxycholic acid = lithocholic
acid > cholic acid = chenodeoxycholic acid. However, the FXR binding site is not perfectly conserved between species. This makes, for
instance, the mouse FXR much more responsive to cholic acid than its human counterpart [340]. To note that mice do not have cheno-
deoxycholic acid, the most potent natural activator of human FXR, which in mice is converted to β-muricholic acid.

receptor TGR5 (also called GPBAR1 or M-BAR/BG37) [116] (Box 2). project CABALA_DIETH&HEALTH (CirculAting Bile Acids as bio-
Besides FXR and TGR5, BAs are able to activate other nuclear receptors markers of metabolic Health_Linking microbiotA, diet and Health,
including pregnane X receptor and vitamin D receptor. Ligand-activated http://www.cabalaproject.eu/) will provide new insight into how diet
nuclear receptors such as FXR control a broad range of metabolic shapes the gut microbiota, its ability to modulate circulating BAs and
processes including hepatic BA transport and metabolism, lipid and impact on human physiological processes through altered BA signaling.
glucose metabolism, drug disposition, hepatic regeneration, fibrosis, This will fill a current gap in knowledge since most studies thus far have
cell differentiation, tumour formation and also inflammation [116]. been conducted in animals and few studies have measured how specific
The diversity of BA pool in terms of molecular structures largely interactions between foods rich in fibers, polyphenols and probiotics
depends on gut microbiota composition and activity. Dietary influences can alter BA signaling in humans. Importantly, this project will achieve
which regulate the composition of the gut microbiota will also influence this through a combination of existing large scale studies, the RoCAV
microbial bioconversion of host BA, BA pool size and its chemical cohort [118] and the Med diet and exercise dietary intervention DI-
composition/structure (Box 2) [117]. The ongoing ERA-HDHL action RECT-PLUS in Israel (ClinicalTrials.gov Identifier: NCT03020186), and

6
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

also bespoke mechanistic studies combining chronic dietary interven- patients and animal models have indicated that impaired intestinal
tion to modulate gut microbiota composition and post-prandial food barrier function occurs before disease onset [105,143]. Studies using
challenge tests to provide new knowledge on how oats, apples and the the permeability markers lactulose and mannitol verified that Bio-
probiotic Lactobacillus reuteri NCIMB 30242 can influence circulating Breeding diabetes (BB-D) prone rats, which spontaneously develop T1D
BA profiles and their cell signaling potential (ClinicalTrials.gov Iden- due to genetic susceptibility, exhibit a highly permeable gut associated
tifier: NCT03369548). Providing the mechanistic basis by which fibers, with low levels of intestinal claudin, a major intercellular TJ protein,
polyphenols and BSH active probiotics can modulate circulating BA before the onset of T1D [144]. Intestinal myeloperoxidase activities and
profiles will provide much needed support for designing efficacious goblet cell density are also higher in the diabetes-prone rats than in
dietary interventions to prove cause and effect relationships between controls, supporting the notion of an early intestinal inflammatory re-
functional food ingestion and beneficial physiological effects observed sponse linked to increased early intestinal permeability. This is not a
in animal models, both in terms of immune related diseases and me- phenomenon occurring only in animal models of diabetes. In humans
tabolic diseases, but for which current literature in humans is contra- with a propensity to develop T1D also appear to possess an abnormal
dictory. Importantly CABALA_diet&health partners in University Col- intestinal barrier. Intestinal samples from individuals at risk of [145] or
lege Cork (Ireland) will investigate how different profiles of circulating already diagnosed with T1D [146,147] demonstrated abnormalities
BA influence activation of BA cell signaling receptors. This is particu- identified with sugar permeability tests, the clinical standard for mea-
larly important since different BA have very different ligand potentials suring intestinal permeability. A loss of the intestinal barrier function
and activation potentials, as described in Box 2. has been observed in non-celiac T1D patients with a frequent associa-
BA signaling through FXR and TGR5 binding attenuates pro-in- tion with mucosa ultra-structural alteration [148]. Moreover, the po-
flammatory innate immune response as demonstrated in several dis- tential pathogenic role that increased intestinal permeability plays in
eases, including MS [119,120], and LPS-induced hepatic damage [121], T1D appears to be dependent on high plasma levels of Zonulin-1, a
atherosclerosis [122] and IBD [123]. Activation of FXR in lamina pro- secretory protein identical to prehaptoglobin 2 regulating epithelial TJ
pria monocytes and macrophages results in repression of NF-kB binding function [149]. Its production relies on bacterial colonization [150] and
to its responsive elements [124]. Its activation reduces inflammation, increased TJ paracellular transport [151]. Reversion of intestinal bar-
downregulating IL-1β, IL-2, IL-6, TNF-α, and IFN-γ gene expression and rier dysbiosis by adding a Zonulin-1 inhibitor ameliorated T1D mani-
attenuates experimentally induced colitis [124]. BA-dependent FXR festations in disease-prone rats [152]. Recent mechanistic studies in
activation also controls bacterial overgrowth and maintains mucosal animal models show that microbial translocation, a consequence of
integrity in the small intestine under physiological conditions by in- increased intestinal permeability due to decreased intestinal epithelial
ducing the transcription of multiple genes involved in intestinal mu- integrity, contributes to T1D development [163]. Gut bacterial trans-
cosal defence against microbes [125]. location into mesenteric and pancreatic lymph nodes has been shown to
In addition to FXR, BA exert anti-inflammatory effects via TGR5 increase destruction of insulin-producing pancreatic β cells in strepto-
engagement. TGR5 signaling also suppresses NF-kB mediated pro-in- zotocin-induced T1D mice [163]. Translocated bacteria in pancreatic
flammatory gene regulation impacting on macrophage [122] and B lymph nodes triggered the nucleotide-binding oligomerization domain
cells activation in vivo [121]. Using in vitro human monocyte-derived containing 2 (NOD2) receptor activation, highly expressed by DCs and
DCs, it has been demonstrated that TGR5 activation by BAs can induce macrophages, but not NOD1, leading to NOD2-mediated induction T
differentiation of IL-12 hypo-producing DCs through TGR5-cAMP helper 1 (Th1) and Th17 responses in pancreatic lymph nodes and
pathway [126,127]. TGR5 has also been found to inhibit the in- pancreas. This unchecked inflammatory response exacerbated T1D
flammatory response associated with liver ischemia through suppres- symptoms [163]. These findings suggest that a leaky intestine in the
sion of the TLR4-NF-kB pathway [128]. Activation of this receptor also setting of T1D could allow a greater exposure of the intestinal immune
suppressed gastric inflammation [129,130] and TGR5 deficiency ex- system to antigens and a continuous source of inflammatory triggers
acerbated injury-induced colitis in mice [131], while its activation in- systemically. The recent evidence showing epigenetic modification as
hibits inflammation and prevents colitis development [123]. Consistent important trigger for T1D development [71], further suggests intestinal
with these observations, treatment of wild-type mice with oleanolic microbiota and its dietary derived-metabolites involvement in immune
acid, a triterpene molecule extracted from olive leaves acting as a TGR5 function through HDAC inhibition could exert an important role in
agonist and with a shape resembling the secondary BA litocholic acid contributing to disease onset.
[132], attenuated colitis development and monocyte infiltration into Altered microbiota composition has also been reported in T1D
the colon [131]. This indicates that the TGR5 pathway may have po- compared to healthy controls. In Non-obese diabetic mice, Bacteroides,
tential as a therapeutic target for Th1 dominant chronic inflammatory Oscillospira, Sutterella, and Bifidobacterium genera were more abundant
disorders, such as Crohn's disease and psoriasis [133] but also Th1- compared to controls [113]. Oscillospira is an important gut microbe
driven autoimmune diseases. previously shown to mediate the disruption of the intestinal epithelial
Indeed, BA activity is not merely limited to the liver but influences barrier, thus causing the translocation of bacteria from the gut in
host immune function systemically. Indication on the potential use of obesity [153]. A culture-independent analysis of the bacteria in fecal
BA as immunosuppressant is discussed below. samples collected from Biobreeding diabetes (BB-D) prone rats and BB-
D resistant rats revealed that Lactobacillus and Bifidobacterium are
4. Gut microbiota dysbiosis in extra-intestinal AID dominant genera in BB-D resistant rats, showing that these species are
negatively correlated with the onset of T1D [154]. Antibiotic treatment
Several evidence are suggesting that alterations of the gut microbial accelerated disease onset in Non-obese diabetic mice in concomitance
composition may be correlated to AID manifestation [14,134–139], with an expansion of Th1 cells and a reduction of Th17 cells in the gut
influencing both immune homeostasis and gut permeability. lymphoid tissues [155]. Fecal transplantation of Non-obese diabetic
microbiota in non-obese resistant mice resulted in insulitis induction,
4.1. Type 1 diabetes revealing that Non-obese diabetic animals bear a diabetogenic gut mi-
crobial community. This gut microbiota is characterized by increased
T1D is an organ-specific AID characterized by an autoimmune re- ileal level of Anaeroplasma and Desulfovibrio species with respect to non-
sponse against the host pancreatic β cells, leading to insufficient insulin obese resistant mice. Lactobacillus abundance was reduced. Non-obese
production from the pancreas [140]. Even though it is still unclear diabetic mice colon appeared devoid of Bacteroides acidifaciiens and
whether intestinal permeability is an initiator or accelerator of T1D or showed decreased abundance of Ruminococcus gnavus and increased
merely an outcome of disease progression [141,142], studies in T1D abundance of Prevotella [155,156].

7
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Table 1 Table 1 (continued)


Key findings supporting a role for gut dysbiosis in AID. Studies investigating the
gut microbiota of AID in animal models and in patients compared with healthy Disease Human cohort/animal Major finding (AID vs HS) Reference
studies
subjects. AID, autoimmune disease; T1D type 1 diabetes; SLE, systemic lupus
erythematosus; MS, multiple sclerosis; HS, healthy subjects.
Relapsing HS (n = 31), MS Prior vitamin D [343]
Disease Human cohort/animal Major finding (AID vs HS) Reference remitting (n = 30); 4 MS supplementation,
studies patients with pre- and ↓ Bacteroidaceae
post vitamin D ↓ Faecalibacterium,
T1D supplementation ↑ Ruminococcus
Animal study non-obese diabetic ↑ Bacteroides [341] Relapsing HS (n = 50), MS ↓ frequencies of specific [213]
mice ↑ Oscillospira remitting (n = 20) Clostridium cluster and
↑ Sutterella Bacteorides
↑ Bifidobacterium Relapsing HS (n = 36), MS ↑ Pedobacter, [216]
Animal study Biobreeding diabetes ↓ Bifidobacterium [154] remitting (n = 59) ↑ Flavobacterium,
(BB-D) prone, BB-D ↓ Lactobacillus ↑ Blautia,
resistant rats ↑ Dorea,
Human study, T1D (n = 16) vs HS ↑ Clostridium [157] ↑ Pseudomonas
pediatric (n = 16) ↑ Bacteroides ↑ Mycoplana
↑ Veillonella ↓ Adlercreutzia,
↓ Lactobacillus ↓ Collinsella,
↓ Bifidobacterium ↓ Parabacteroides,
↓ Blautia coccides- ↓ Erysipelotrichaceae,
Eubacterium rectale ↓ Lachnospiraceae,
↓ Prevotella ↓ Veillonellaceae,
Human study, Mexican T1D children ↑ Bacteroides [158] ↓ Lactobacillus,
pediatric (n = 21), 8 newly ↓ Prevotella ↓ Coprobacillus,
diagnosed, 13 ↓ Haemophilus
established T1D Relapsing HS (n = 43), MS ↑ Methanobrevibacter, [217]
Human study, European T1D newly ↑ Bacteroidetes [159] remitting (n = 60) ↑ Akkermansia
pediatric diagnosed children ↑ Streptococcus ↓ Butyricimonas
(n = 28), vs HS ↓ Clostridium cluster IV Relapsing HS (n = 71), MS ↑ Akkermansia [218]
(n = 27) and XIVa remitting (n = 71) ↑ Acinetobacter
Human study, Russian vs Finnish and ↑ Bacteroides LPS in [161] ↓ Parabacteroides
pediatric Estonia children Finnish and Estonia Relapsing Monozygotic twin, ↑ Akkermansia [219]
(n = 222) children, with high remitting discordant for MS
incidence of T1D (n = 34)
Human study, Scandinavia infants, Reduced bacterial alpha- [160] Relapsing HS (n = 43), MS No differences in alpha [214]
pediatric predisposed to T1D diversity remitting, (n = 60) diversity and beta
(n = 33) ↑ Blautia pediatric diversity
↑ Rikennellaceae subjects Early microbiota
↑ Ruminococcus alterations in MS found at
↑ Streptococcus specific taxonomic levels:
↓ Lachnopiraceae ↑ Faecalibacterium
↓ Veillonellaceae ↑ Enterobacteriaceae
Human study, Finnish children ↑ Bacteroides [163] ↑ Clostridiales
pediatric (n = 8), predisposed ↑Bifidobacterium ↑ Bacteroides
to T1D fragilis
Human study, Scandinavia children ↑ Bacteroides [162] ↓ Methanobrevibacter
pediatric with HLA-conferred ↑Bifidobacterium Relapsing HS (n = 9), MS Found microbiota [215]
susceptibility to T1D remitting, (n = 15) dysfunction
(n = 18), HS (n = 18) pediatric ↑ Desulfovibrionaceae
SLE subjects ↓ Lachnospiraceae
Animal study MRL/lpr mice Female lupus prone mice [202] ↓ Ruminococcaceae
showed
↓ Lactobacillaceae
↓ Lachnospiraceae Initial findings indicate that the gut microbiota of subjects with
Animal study SNF1 mice Decreased beta-diversity [203]
prediabetes or T1D is different to that of healthy people (Table 1). The
↓ Lactobacillacus
↓ Firmicutes/Bacteroides analysis of 16 T1D pediatric patients revealed a marked increase in the
ratio relative abundance of Clostridium, Bacteroides and Veillonella and a
Human study HS (n = 20), SLE ↓ Firmicutes/Bacteroides [198] concomitant decrease in Lactobacillus, Bifidobacterium, the Blautia coc-
(n = 20) ratio cides-Eubacterium rectale group and Prevotella compared to healthy
↓ bacterial metabolites
(homoserine lactone, N-
children [157]. Higher abundance of Bacteroides and lower abundance
acetylmuramic acid, of Prevotella characterize also newly-diagnosed Mexican pediatric T1D
ribose-1,5-bisphosphate) patients compared to healthy children [158]. A European study showed
Human study HS (n = 17), SLE ↓ metabolites associated [199] that < 3 years old patients had increased level of Bacteroidetes and
(n = 18) to pyrimidine, purine and
streptococci and reduced abundance of Clostridium cluster IV and XIVa
aminoacid metabolism
MS species. Increased microbial diversity and decreased level of butyrate
Relapsing HS (n = 31), MS Type A Clostridium [342] producing clostridia have been found in pediatric T1D who were < 3
remitting (n = 30) perfringes reduced in MS years old with respect to healthy age-matched control [159]. Com-
Type B C. perfringes found paring the gut microbiota of subjects who develop T1D with those who
in one MS patient
do not, has revealed that a reduced bacterial diversity is more evident
before the time of diabetes onset [160]. A recent study reported that
Russian children, who have a low incidence of T1D have a low relative
abundance of Bacteroides in contrast with the high relative abundance

8
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

of Bacteroides shown in children from Finland and Estonia, where the [173]. The authors showed that alterations in the composition of mucin
incidence of T1D is much higher [161]. Three human studies including degrading bacteria, with increasing Bacteroides and decreased abun-
T1D patients were carried out in Scandinavia where the incidence of dance of Akkermansia, associate with the risk of early development of
T1D is high [160,162,163]. Similar to observations in animal experi- islet auto-antibodies [173]. Thus, changes in microbial diversity and
ments, T1D patients showed an increased abundance of Bacteroidaceae differences in relative abundance, in particular of mucin-degrading and
and decrease of bifidobacteria [157,163]. Specifically, Brown et al. butyrate-producing bacterial taxa, between diabetic subjects and
[162] and De Goffau et al. [163] identified a decrease in lactate- and healthy/control, highlight that gut microbiota and associated increased
butyrate-producing bacteria being associated with development of au- gut permeability may contribute to disease onset or progression. In-
toimmunity, with a dearth of Bifidobacterium adolescentis and B. pseu- triguingly, the above study also suggests that diet-microbe interactions
docatenalatum species, whereas Kostic et al. [160] did not see similar and their influence in immune function early in life hold significant
changes in patients that developed autoimmunity during the study potential in developing dietary, population based strategies to reduce
period. However, the authors showed that T1D children bear higher the risk of T1D.
abundance of Blautia, Rikennellaceae, Ruminococcus and Streptococcus
and a lower abundance of Lachnospiraceae and Veillonellaceae [118]. 4.2. Systemic Lupus Erythematosus (SLE)
Differences in methods and study groups (patients with established T1D
vs. patients with anti-pancreatic antibodies who later develop overt SLE is an autoimmune disorder characterized by severe and per-
diabetes) may account for the differences or contradictions in the re- sistent inflammation that leads to tissue damage in multiple organs
sults. Importantly, all studies found that the alpha-diversity of T1D [176,177]. Although SLE affects both men and women, women of
patient microbiota was reduced, indicating that T1D is characterized by childbearing age are diagnosed about 9 times more often than men
a reduced bacterial diversity, possibly reduced functional diversity and [178]. It has been shown that LPS can promote SLE development and
possibly low community stability. disease progression upon penetration of the intestinal epithelium and
Activation of TLRs has been linked to the pathogenesis of T1D translocation into systemic tissues [179]. In SLE patients, either in ac-
[164]. TLR4 has been identified as the main receptor on pancreatic β tive phase or remission, the higher level of soluble CD14 compared to
cells contributing to islet cell inflammation during the course of T1D healthy subjects correlates with disease activity parameters, suggesting
development [165]. LPS is the major ligand triggering TLR4 signaling the involvement of free circulating LPS in SLE development [180]. In
activation. Controversial results have suggested a role of LPS as an addition, repeated injections of LPS into lupus-prone mice result in
etiological agent increasing the risk of developing T1D [166] or as a increased autoantibody production and development of glomerulone-
suppressive agent on the incidence and severity of the disease by in- phritis [181,182]. Activation of TLR4 also promotes lupus disease ac-
creasing Tregs [167,168]. LPS administration to Non-obese diabetic tivity in lupus prone mice, as the apolipoprotein E-deficient (ApoE−/−)
mice during the prediabetic state did not reverse insulitis but delayed mice [181,183,184]. Mice spontaneously develop lupus when TLR4
the onset of diabetes [169]. Multiple-injection, rather than single in- responsiveness is increased, whereas the exacerbated disease phenotype
jection, of LPS was found to suppress T cell proliferation, to promote can be significantly ameliorated when the commensal gut microbiota,
Tregs expansion and to induce the differentiation of tolerogenic DC as a source of LPS, is removed by antibiotics [181]. Ni et al. found
with low TLR4 expression. Explanting those cells into Non-obese dia- increased levels of serum autoantibodies and more severe lung injury
betic/SCID diabetic mice protects those animals from T1D progression when challenging ApoE−/− mice with LPS [185]. Moreover, im-
[169]. munization of non-autoimmune mice with phospholipid-binding pro-
Interestingly, Non-obese diabetic mice deficient in the innate sig- teins induced lupus-like disease and can be facilitated by the adminis-
naling molecule MyD88 are protected from the development of T1D tration of LPS [186]. Conversely, inhibition of TLR4 results in reduced
[170]. However, such protection is lost when Myd88−/− mice of the autoantibody production and lowered renal glomerular IgG deposits in
Non-obese diabetic strain are housed under germ-free conditions. The lupus-prone mice [187,188]. Taken together, these data suggest LPS
absence of MyD88 in Non-obese diabetic mice leads to over-re- stimulation and TLR4 activation or hyper-responsiveness, as factors
presentation of the bacterial phylum Bacteroidetes [170] and suppres- contributing to SLE initiation in genetically susceptible hosts.
sion of diabetes, presumably through the production of an, as yet uni- Circulating lipoteichoic acid (LTA), an inflammatory component of
dentified, immunomodulatory microbial product. MyD88 deficiency the Gram-positive bacterial cell walls, has also been linked lupus dis-
changes the composition of the distal gut microbiota, and that exposure ease. The expression of TLR2, the receptor of LTA, has been reported to
to the microbiota of specific pathogen-free MyD88−/−Non-obese dia- be increased in SLE patients [189]. In lupus-prone mice, TLR2 activa-
betic donors attenuates T1D in germ-free Non-obese diabetic recipients. tion triggers lupus nephritis whereas TLR2 knockout attenuates lupus-
Transfer of gut microbiota of Myd88−/− Non-obese diabetic to wild- like symptoms [187,190,191]. Recently, another bacterial antigen that
type Non-obese diabetic mice reduced the intensity of insulitis and may mimic self-antigens has been recognized to induce autoantibody
delayed the onset of T1D. This result has been shown to correlate with a production [192].
decreased abundance of Lactobacillaceae, an increased abundance of Since the hallmark of SLE is the presence of high levels of anti-
Lachnospiraceae and Clostridiaceae, increased level of TGF-β in the double-stranded DNA autoantibody (anti-dsDNA) in sera, the con-
lamina propria and increased number of CD8+, CD103+ and CD8 α/β T tribution the nucleic acid receptor, recognizing bacterial hypomethy-
cells [171]. Together, these findings indicate that the interaction be- lated CpG DNA sequences, TLR9, in the pathogenesis of SLE has been
tween the gut microbiota and the innate immune system could be an investigated in a transgenic mice carrying anti-dsDNA antibody trans-
important factor modifying T1D predisposition. gene and challenged the mice with TLR4-and TLR9-agonists [183].
So far, few studies have investigated the association between altered Splenocytes from these mice were found to secrete higher levels of in-
gut microbial communities and autoimmunity prior to islet autoanti- terleukin-10 (IL-10) and anti-dsDNA. Concomitant activation of extra-
body development [172,173]. A shift in the ratio of Bacteroidetes and cellular TLR4 and intracellular TLR9 affect SLE progression in vivo by
Firmicutes [159] and a higher abundance of Bacteroides dorei and accelerating the accumulation of anti-ds-DNA autoantibody [183].
Bacteroides vulgatus [174] have been associate with increased risk of Several downstream proteins in the TLR7 and 9 signaling cascade
islet autoantibody. Autoantibody positive children also show lower are highly relevant to the pathogenesis of SLE [193]. Deficiency of
relative abundance of butyrate-producing and mucin-degrading bac- MyD88 in B cells, in particular, has been shown to ameliorate lupus
teria [162]. Recently, co-occurrence analysis on the BABYDIET cohort disease in MRL/lpr mice, the classical animal model of lupus nephritis
[172,175] revealed a functional association between diet, gut micro- [30,194] by ameliorating nephritis, reducing IgG anti-nuclear anti-
biota structure and metabolism and islet auto-antibody development bodies production and activation of nephrotoxic T cells [194].

9
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Conversely, there is a paucity of data pertaining to NLRs such as environmental factors inducing loss of tolerance and activation of self-
NOD1 and NOD2 which are associated with inflammasome formation reactive T cells recognizing myelin antigens [205]. Studies on EAE
[195,196]. Interestingly, NLRP3/ASC inflammasome were found com- animal models suggest a strong influence of the gut and its microbiota
promised in New Zealand Blackmice, a lupus-prone model. Consistent in disease development and associated inflammation [67,207–209]. An
with this finding, loss of ASC in B6-Faslpr mice led to exacerbation of altered gut mucosa and a disrupted mucosal immune homeostasis have
lupus-like disease by increasing lung T cell infiltrates and nephritis as been suggested to contribute to the emergence of the disease (REF). An
well as increasing DCs and macrophages activation and expansion of T increased intestinal permeability preceding EAE onset which worsens
and B cells [197]. These results suggest opposite role for TLRs and NLRs during disease progression has been described [207]. Intestinal per-
in contributing to SLE. The evidence that TLR-mediator deficiencies meability was shown to be associated with an alteration of mucosal
ameliorate lupus symptoms while NLR-mediator deficiencies exacer- structure, as indicated by the increased crypt depth and thickness of
bate it, suggest distinct contributions for those receptors in SLE pa- jejunum and ileum mucosa, and an overexpression of Zonulin-1 [207]
thogenesis. Both receptor's family are involved in sensing bacterial as previously observed in T1D [152,210]. These changes lead to a
components, leading to hypothesize that they could differently re- higher immune cells infiltration during MS progression. Elevated Th1
cognize translocated bacteria through a leaky gut, eliciting pro-in- and Th17 pro-inflammatory responses was observed in lamina propria,
flammatory response or tolerance, respectively. Payer's patches and mesenteric lymphnodes as well as in spleen. This
Intestinal dysbiosis has been reported in SLE patients (Table 1). increase was concomitant with lower Treg cell numbers [207].
Preliminary 16S rRNA profiling showed that the fecal Firmicutes/Bac- The complete absence of microbes in gnotobiotic mice C57BL/6
teroidetes ratio was significantly lower in SLE patients compared to mice subjected to the induction of active EAE [67] and in gnotobiotic
healthy controls, even during remission [198]. Fecal metabolite con- transgenic mice [209] that develop the disease spontaneously sig-
centrations measured by liquid chromatography coupled with a time of nificantly affects the development of disease. Germ-free EAE animals
fly mass spectrometry (LC-ESI-QTOF-MS) and capillary electrophoresis produce lower levels of the pro-inflammatory cytokines IFN-γ and IL-
mass spectrometry technique showed that SLE patients exhibited re- 17A in both intestine and spinal cord but display a reciprocal increase
duced levels of homoserine lactone, N-acetylmuramic acid and ribose- in Treg cells. This was associated with a reduced ability of germ-free
1,5-bisphosphate compared to healthy subjects [199]. The same re- animal DCs to stimulate pro-inflammatory T cell responses. In germ-free
search group also described a deficiency in metabolites associated with mice demyelination was shown to initiate following colonization with
pyrimidine, purine, and amino acid metabolism [199]. The authors feces isolated from SPF mice and is caused following entry of auto-
suggested that SLE patients bear deficiencies in chemical species par- antigen specific B cells into the germinal center of cervical lymph nodes
ticipating in, for example, quorum sensing, cell-to-cell communication [209]. Furthermore, oral treatment with a mixture of non-absorbing
and cell wall biosynthesis, which are known to be important for gut antibiotics beginning one week prior to sensitization also reduce sig-
microbiota homeostasis. nificantly the severity of EAE [211]. Microbiota depletion ameliorates
In mouse models, a role for the intestinal gut microbiota in the the development of EAE by reducing the production of pro-in-
development of SLE is not completely clear since the onset and severity flammatory cytokines from the draining lymph node cells, as well as by
of lupus-like disease is not profoundly altered in the germ-free state reducing mesenteric Th17 cells in an invariant NK cells dependent
[200,201]. However, there is emerging evidence that there is dysbiosis manner [211]. Moreover, antibiotic treatment, by reducing gut bac-
and gut-pathology in different animal models of lupus. Consistent with terial populations protects against EAE development by inducing
findings of microbiota composition in human SLE patients [198], lupus- FoxP3+ Treg cells accumulation in the mesenteric and cervical
prone MRL/lpr mice have increased microbial diversity in the feces lymphnodes [208,212].
compared to non-disease controls [202], and the relative abundance of Although the cause of MS is unknown, the gut microbiota seems to
Lactobacillus and the ratio of Firmicutes/Bacteroidetes were higher in be important in the onset and/or progression of the associated systemic
mice with lower lupus severity [203]. Moreover, female lupus-prone autoimmunity. Studies conducted so far suggest the presence of a dis-
MRL/lpr mice displayed a significant reduction of Lactobacillaceae fa- rupted or altered microbiota in relapsing remitting MS patients com-
mily and increase of Lachnospiraceae both prior to disease onset and in pared to healthy people [213–217] (Table 1). Detailed faecal micro-
the late stage of disease with severe lupus symptoms [202]. biome analyses of 31 MS patients revealed that MS patients had distinct
Zhang et al. observed that there were sex-related changes in the gut microbial community profile compared to healthy subjects (n = 36)
microbiota of lupus-prone mice compared to healthy mice. Females [216]. Chen et al. observed a higher abundance of Psuedomonas, My-
were more likely to have higher Lachnospiraceae abundance, which coplasma, Haemophilus, Blautia and Dorea genera in MS patients,
correlated with the development of more severe disease [202]. Another whereas control group showed higher abundance of Parabacteroides,
study also showed that gender-related differences in the gut of lupus- Adlercreutzia and Prevotella genera. Other studies reported that MS pa-
prone mice. Female SWR x NZBF1 (SNF1), which develop more severe tients, compared to healthy subjects, bear significantly lower Clostridia
nephritis than male mice, have more plasma cells and α4β7-expressing XIVa and IV clusters, e.g. Butyricimonas [217] and Bacteroides
T cells in the Peyer's patches than male SNF1 mice [204]. Analysis of [213,214], and higher relative abundances of Methanobrevibacter and
RNA expression in the distal ileum showed that female SNF1 mice ex- Akkermansia [217]. The differences observed correlated with variation
press much higher levels of pro-inflammatory mediators including IL-6, in the gene expression of molecules involved in DC maturation, inter-
IL-9, IL-17, IL-22, IFN-α and IFN-β than male mice [204]. Although the feron and NF-kB signaling pathways in circulating T cells and mono-
gut microbiota was not analyzed in this study, it does further support a cytes [217]. In contrast, another study comparing 16S rRNA profiles
role for the gut microbiota in lupus development and in the gender bias from 60 faecal MS samples with 43 healthy donors showed no sig-
of this disease, given the regulation of these inflammatory cytokines by nificant differences in the diversity of microbial populations in MS
the gut microbiota and their metabolites. [214]. Moreover, the gut microbiota analysis of 15 pediatric MS pa-
tients revealed an inverse association of Bacteroides with Th17 immune
4.3. Multiple Sclerosis (MS) markers, while Fusobacteria correlated with Treg expression in healthy
children [215]. Very recently, Cekanaviciute et al. reported higher le-
MS is a chronic inflammatory demyelinating disease of the central vels of Akkermansia muciniphila and Acinetobacter calcoaceticus in MS
nervous system with a pathogenesis characterized by a breakdown of patients compared to healthy subjects [218]. These two bacterial spe-
the blood-brain barrier and demyelination of the central nervous cies induced a pro-inflammatory response ex vivo, while, Para-
system by infiltrating autoreactive T cells [205,206]. Epidemiological bacteroides distasonis, found reduced in MS, induced an anti-in-
and genetic studies suggests that MS is provoked by exposure to flammatory response in mice, driven by the expansion of IL-10+ Treg

10
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

cells [218]. It is noteworthy to observed that monocolonization of mice inflammatory signaling pathways [233–237].
with Acinetobacter calcoaceticus but not with Akkermansia, exacerbated Evidence mainly from animal studies suggests effects of probiotics
EAE symptoms [218]. Similarly, the gut microbial profiles of MS twins in regulation of EAE via IL-10 expansion [67,238–242]. Oral adminis-
and healthy twins were compared, evidencing an increased Akker- tration of a mixture of L. paracasei DSM 13434, L. plantarum DSM 15312
mansia relative abundance in untreated MS twins [219]. When human and DSM 15313 suppressed established EAE disease [239]. The treat-
MS microbiota was transplanted in mice, the modest dysbiosis observed ment resulted in IL-10-dependent activation of Tregs in gut-related
in MS subjects determined spontaneous autoimmune encephalomyelitis lymphoid organs as well as the central nervous system, followed by a
in recipient healthy germ free mice [219], which spleen T cells pro- concomitant reduction of immune cells promoting and inducing the
duced less IL-10 after transplantation. production of TNF-α, IFN-γ and IL-17 [239]. Treatment with Pedio-
Recently, Benedek et al. [34] showed that treatment with pregnancy coccus acidolactici R037 [241], Bifidobacterium animalis (a strain by
levels of 17-β-estradiol significantly increasing the relative abundance Numico Research, NL) [240], and Bifidobacterium animalis PTCC 1631 –
of lactobacilli and Lachnospiraceae and induces regulatory B cells and alone or in combination with L. plantarum A7 - [243] improved central
anti-inflammatory macrophage proliferation in MLN. These studies in- nervous system inflammation through the induction of Treg cells in the
deed associate changes in the relative abundance of gut species with gut mucosa by promoting the secretion of transforming growth factor-β,
phenotypic modification and balance of immune regulatory cell popu- and inducing decreased Th1/Th17 inflammatory subsets
lations. The changes in the abundance of these microbial species could [240,241,244]. Similarly, a probiotic mixture consisting of Lactobacillus
have different effects on the immune systems. For example, different casei, Lactobacillus acidophilus, Lactobacillus reuteni, Bifidobacterium bi-
profiles of gut bacteria are likely to impact on SCFA profiles, including fidum, and Streptococcus thermophilus (Korea Yakult Co) has been shown
acetate, butyrate and propionate (Box 1) and also modulate BA profile to prevent EAE, delay disease onset and reduce the infiltration of in-
and pool size (Box 2), which can directly regulated Treg expansion flammatory cells into the CNS [244]. Furthermore, oral treatment with
[95–97,100,220], oxidative stress, reinforce the intestinal barrier and capsular polysaccharide A, derived from commensal Bacteroides species,
blood-brain barrier function [221,222] as also described above. protects mice against EAE through the induction of regulatory CD39+ T
cells and Tregs [238,245]. B. fragilis and Lactobacillus rhamnosus have
5. Can microbiota modulation improve AID associated symptoms also been shown to interact directly with neurons in the enteric nervous
and chronic inflammation? system affecting their function [246]. A randomized double‐blind pla-
cebo‐controlled clinical trial analyzed the effect of the intake of a
The prospect of restoring a normal microbiota is an appealing hy- probiotic supplementation containing L. acidophilus, L. casei, L. fer-
pothesis to manage conditions associated with a gut microbiota dys- mentum and Bifidobacterium bifidum for 12 weeks in 60 MS patients. The
biosis. Restoration of the composition of gut microbiota and its com- study indicated that probiotic administration improved Expanded Dis-
munity structure conceivably leads to restoration of its function. ability Status Score, insulin resistance and decrease the levels of in-
Several ways for manipulating gut microbiota composition and function flammatory markers in MS patients [247]. T1D pathogenesis appears to
include the use of dietary intervention, prebiotics, probiotics and fecal be responsive to dietary intervention early in life in some animal
microbiota transplantation, as discussed in the following sections. models. Formula feeding during the neonatal period accelerates the
development of T1D in BBD prone rats through regulation of Treg cells
5.1. Dietary microbiota modulation and its immune-regulatory potential in and anti-inflammatory cytokines [248]. The administration of a hy-
AID drolyzed casein diet [249] or antibiotics [250,251] accelerates disease
development. Furthermore, a reduced abundance of lactobacilli and
Findings from population studies and intervention trials provide bifidobacteria has been associated with onset and progression of T1D
some evidence that adherence to a healthy diet over time reduces the [154]. Lai et al. (2009) showed that Lactobacillus strains with cinnamoyl
risk of long-term inflammation and associated diseases. The esterase activity, namely L. johnsonii N6.2 and L. reuteri TD1, were
Mediterranean diet, with beneficial fats, high polyphenols and fibers, negatively correlated with T1D development [252]. This led to the in-
and low refined sugars and low energy, has been shown to lower in- vestigation of the potential use of L. johnsonii N6.2 in the treatment of
flammatory markers in patients with metabolic disease and in the el- T1D [253]. The administration of L. johnsonii isolated from BB-D re-
derly [223–225]. sistant rats delayed or inhibited the onset of T1D in T1D-prone rats. In
Indeed, treating chronic inflammation has become one of the most particular, L. johnsonii N6.2 induced changes in ileal microbiota, de-
promising targets for reducing the disease burden, especially since it creased oxidative response of the intestinal mucosal and reduced in-
appears responsive to diet, with beneficial fats, polyphenols and fibers flammation by inhibiting IFN-γ production. L. johnsonii N6.2 adminis-
all shown to impact on inflammatory markers [223–226]. In addition, tration also resulted in higher levels of interephitelial junction proteins
such interventions also boost gut bacteria that potentially suppress in- boosting intestinal integrity [253]. In a recent study in NOD mice, oral
flammation. Human fecal butyrate-producing Clostridia clusters XIVa administration of a the probiotic VSL#3- which consists of a mixture of
and IV, as well as Bacteroides fragilis have been correlated with sup- L. casei, L. plantarum, L. acidophilus, L. delbrueckii subsp. bulgaricus), B.
pression of inflammatory conditions through the induction of Foxp3+ longum, B. breve and B. infantis and Streptococcus salivarius subsp. ther-
Tregs [227,228]. Indeed, the ultimate goal of therapies targeting the mophilus - protects mice from T1D by suppressing IL-1β expression and
gut microbiota is to the growth of anti-inflammatory commensal bac- the release of immunomodulatory indoleamine 2,3-dioxygenase (IDO)
teria or their metabolites while eliminating potential pathobionts and to and by promoting the differentiation of gut tolerogenic DCs reducing
reverse gut permeability (Fig. 2). Prebiotics and probiotics and dietary differentiation/expansion of Th1 and Th17 cells in the intestinal mu-
intervention with certain whole foods have shown some promise for cosa and at the sites of autoimmunity [254]. Moreover, the effect of
this respect [229]. early exposure of probiotics on T1D onset and progression has also been
studied in humans [255]. Use of probiotics supplemented though diet
5.1.1. Probiotic administration to reverse chronic systemic inflammation or taken in infant formula has been recorded in the large “The en-
associated to AID vironmental Determinants of Diabetes in the Young” (TEDDY) cohort,
Probiotics are “live microorganisms which when administered in ade- encompassing 8676 children at increased risk of T1D. Early probiotic
quate amounts confer a health benefit on the host” [230]. Probiotic spe- supplementation, within the first 27 days of life, was associated with a
cies, mainly lactobacilli and bifidobacteria, have been shown to possess decreased risk of islet autoimmunity when compared with probiotic
properties which protect the intestinal barrier by targeting different supplementation after 27 days or no supplementation [255].
components of the mucosa [231,232] and by inhibiting pro- In a lupus-like animal model, restoring Lactobacillus spp. by

11
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Fig. 2. Proposed intervention for gut dysbiosis recovery and schematic model on how diet could contribute to chronic inflammation. Diet induces changes
to gut microbiota, and reduces the production of BAs and SCFAs, leading to impaired signaling activation, epigenetic transcriptional changes and resulting in
alteration of gut homeostasis, Treg activity and immune homeostasis. Therapeutic targeting gut dysbiosis can be approached by a direct manipulation of the
microbiota through antibiotic administration or faecal transplantation or by promoting growth of beneficial bacteria by dietary administration of prebiotic or
probiotic or by consuming high fiber diet.

administration of retinoic acid improved lupus symptoms, and sug- in inulin-fed rats has also been reported [262,263]. Similarly, increased
gesting the use of lactobacilli as a probiotic to diminish inflammation in SCFA production from fermentation of various fibers has been shown to
patients with SLE [202]. In this context, a 3-weeks oral administration increase mucin production in the colon of laboratory animals [264].
of a mixture of 5 lactobacilli, namely L. oris F043, L. rhamnosus LMS201, Mucin is important in the intestine because it fortifies the intestinal
L. reuteri CF48-3A, L. johnsonii 135-1-CHN and L. gasseri JV-V03, has barrier and also provides carbohydrate to members of the gut micro-
been shown to promote Treg proliferation in the kidney of lupus-prone biota, and also supports community stability and resilience trough
MRL/lpr mice [256]. To note, the probiotic administration also im- inter-species metabolic cross-feeding. Some of these microorganisms
proved gut epithelium integrity, and decreased inflammation by de- especially Akkermansia muciniphila have been shown to impact on im-
creasing gut IL-6 levels and concomitantly increase both gut and cir- mune function by promoting intestinal barrier and by increasing bu-
culating IL-10 levels. Modulating the gut microbiota revealed that lupus tyrate levels thus increasing its immuno-modulatory activity
nephritis is influenced by gut microbiota in a sex-dependent fashion as [265–267].
probiotics has been shown to be effective in female and castrated male Prebiotics B galactooligosaccharide [268] and short chain fructoo-
mice but not in intact males [256]. ligosaccharides [269] have been shown to inhibit production of pro-
Further investigations are warranted focusing on understanding the inflammatory cytokines and enhance production of anti-inflammatory
molecular and metabolic mechanisms underlying the health effect of cytokines in animal models. These effects appear, at least in part, to be
probiotics in ameliorating the chronic systemic inflammation asso- mediated by SCFA produced upon prebiotic fermentation.
ciated to AID. Particular focus should be paid on the ability of pro- A study on the TEDDY cohort showed that increasing dietary soluble
biotics to reduce intestinal permeability and whether this reduces the fiber intake in early childhood did not decrease the risk of developing
risk of AID by reducing chronic systemic inflammation in humans. T1D [270]. However, Mariňo et al. (2017) recently showed that spe-
cialized diets containing acetylated or butyrilated starch provided
5.1.2. Prebiotic administration to reverse chronic systemic inflammation protection from T1D in the non-obese diabetic mouse model [271].
associated to AID Compared to controls, both diets promoted a significant reduction of
Prebiotics are “non-digestible food ingredient that beneficially affects insulitis and diabetes development, as well as an increase in gut in-
the host by selectively stimulating the growth and/or activity of one or a tegrity. The protective effect was augmented when SCFA were provided
limited number of bacteria in the colon” [257]. Most common prebiotics combined in the diet. Acetate was particularly effective at decreasing
are dietary fibers which specifically target the gut microbiota. They the proliferation of autoreactive T cells by modulating splenic B cells
show promise, at least in animal studies, to modulate inflammatory expansion, while butyrate was superior in promoting the number and
processes in AID. Several in vitro and in vivo approaches have shown function of Treg cells [271]. Downregulation of the expression of co-
that prebiotics can modulate the gut microbial composition towards a stimulatory molecules and MHC class I on B cells of non-obese diabetic
potentially healthier community structure, especially increasing bifi- mice fed the acetate rich diet correlated with a low frequency of
dobacteria levels. Prebiotics, especially the inulin-type fructans, have IGRP+CD8+ T cells and diminished the proliferation of autoreactive
also been shown to improve the barrier integrity by stimulating the NOD8.3 T cells in vivo. The authors also found that diabetes protection
growth of protective bacteria and the intestinal production of SCFA, of MyD88-deficient non-obese diabetic mice housed in SPF condition is
that upregulate epithelial defense mechanisms that protect against in- associated with high level concentration of circulating SCFA [271].
testinal inflammation. These bacteria or SCFA restore intestinal barrier Very recently Needell et al. (2017) investigated the ability of maternal
function by enhancing TJ between intestinal epithelial cells and by therapy with SCFA to modulate the intestinal microbiota and reverse
increasing mucus production [258]. Prebiotics have also been shown to virus-induced T1D in LEW. WR1 rat offspring [272]. Administration of
decrease the levels of circulating LPS [259,260], whilst increasing GLP- SCFA through drinking water prior to pregnancy and further treatment
1 secretion and TJ expression, and reducing chronic low grade in- of the offspring with SCFA not only decreased the pro-inflammatory
flammatory state associated with metabolic endotoxemia. Other dietary milieu induced by Kilham Rat Virus but also prevented islet auto-
fibers, such as inulin, arabinoxylan, guar gum and citrus fibers, have immunity. SCFA administration determined also a modulation of the
been shown to directly increase mucin production, resulting in reduced gut microbiota, with an increase in the relative abundance of Bacter-
intestinal permeability [261]. An increase in sulphomucin production oidetes and a decrease in Firmicutes level. SCFA treatment resulted also

12
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

in a reduced abundance of Clostridium and Bifidobacterium compared the proportion of genes associated with activation of macrophage to-
with rats infected by the virus only [272]. The effect of SCFAs admin- wards an anti-inflammatory phenotype.
istration in MS has been recently investigated in the EAE mice model In summary, the broad immuno-metabolic actions of BA-activated
[220,273]. Oral administration of SCFA suppressed EAE progression by receptors hold considerable promise in the development of pharma-
inhibiting lymphocyte-mediated systemic inflammation [273]. In par- ceuticals or nutraceuticals capable of tackling inflammation in an au-
ticular butyrate treatment promoted Treg expansion and decreased Th1 toimmune setting (Fig. 2). While BA are synthesized within the liver,
cells, even though a slight increase of Th17 was observed as previously the significant diversity of the bile acid pool is ultimately generated by
reported [220,274]. Similarly, oral administration of propionate re- the gut microbiota (Box 2). Thus, alteration of the gut microbiota will
sulted in an increase of Treg and IL-10 production [220]. influence the signaling properties of the BA pool in the host, de-
Overall, by changing microbiota profiles and impacting on immune termining health or disease.
function through increased/changed levels of SCFA [275], prebiotics
hold the promise as a not invasive strategy for treating AID. Further 5.1.4. Diet and sex-dependent effects
investigations should follow for translating the evidence collected in Recent data show sex-specific impact of diet on the gut microbiota,
animal studies in controlled randomized human trial. and the apparent sex-specific nature of gut microbiota modulation of
host health [46]. In particular, the impact of diet on the gut microbiota
5.1.3. BA as potential immunosuppressant in AID therapy has been shown to be different in males versus females. For example,
Circulating BA pool is also strongly dependant on the profile of BA- Lactobacillus, Alistipes, Lachnospiraceae and Clostridium were more
converting bacteria. Therefore, microbiota modulation through the diet abundant in males fed a high-fat rather than chow diet, whereas in
might be exploited to modulate microbially-driven changes in BA pool females these genera were less abundant in high-fat diets [46]. Kar-
both in terms of BA chemical structures and in circulating concentra- unasena et al. [283] monitored the response to the probiotic Lactoba-
tions of BA (Box 2). cillus animalis NP-51 in male and female mice infected with the pa-
BA-based therapies are efficacious to treat chronic diseases with a thogen Mycobacterium avium sub. paratuberculosis, a model of IBD. The
strong inflammatory component and have so far focused on hydro- authors discovered that probiotic was able to repress Mycobacterium-
philic, less toxic BA or synthetic potent agonists, some of which acting driven inflammation in female mice, but not in male mice [283]. They
as dual FXR and TGR5 ligands. Among the natural ligands, urso- showed that, regardless of treatment, differences in the pro-in-
deoxycholic acid is regarded as a minor secondary BA as it is formed by flammatory cytokines level existed between male and female animals
7β-epimerization of chenodeoxycholic acid by the intestinal micro- and also at a metabolite level [284]. M. avium sub. paratubercolosis lead
biota. The portion of ursodeoxycholic acid in the human total BA pool is to an increase in SCFA, especially butyrate and acetate, and xylose in
less than 3% [276]. Chenodeoxycholic acid also has im- males and greater increases in o-phosphocholine or histidine in female
munosuppressive properties, inhibiting lymphocyte proliferation to colon tissues. This was accompanied by an increased abundance of
mitogens, suppressing IL-1β, IL-6, TNF-α and IFN-γ production from Staphylococcus and Roseburia in female mice, even though these two
mononuclear cells [277,278]. genera were found over-represented in females regardless of treatment.
In the context of AID, the immunosuppressive effects of urso- Interestingly, animals maintained on a diet with the probiotics de-
deoxycholic acid and chenodeoxycholic acid were investigated in F1 monstrated no difference in Lactobacillus species, while after probiotic
mouse model of SLE [279]. Chenodeoxycholic acid, but not urso- and synbiotic treatments, Lactobacillus species were found in greater
deoxycholic acid, delayed development of autoimmune glomerulone- numbers in female colonic tissues [284]. Indeed, it appears that sex and
phritis and prolonged the life span of lupus-prone mice. chenodeoxy- diet interactively control microbiota composition, highlighting the need
cholic acid treated mice showed also a significantly lower serum IFN-γ to identify and explain how this interplay occurs, how they relate to the
concentration, revealing an immuno-suppressive activity of this bile immune response and the underlying molecular basis of this response.
acid [279]. Ho and Steinman treated orally EAE mice with obeticholic
acid (OCA, also referred as 6-ethyl-chenodeoxycholic acid, 6-ECDCA) - 5.2. Fecal microbiota transplantation for restoring gut microbiota
asynthetic FXR agonist having more affinity to FXR than its natural functionality
ligand chenodeoxycholic acid [280] - known to improve lamina propria
permeability and chronic inflammation associated with IBD [123] and Restoring microbial health by transplanting a foreign gut microbiota
attenuate intestinal ischemia reperfusion injury in rats [281], and with has received much medical attention drawn by the success in treating
the natural FXR ligand, the BA CDCA. OCA inhibited both active and recurrent Clostridium difficile associated diarrhea [285,286]. Whole
passive EAE more effectively than chenodeoxycholic acid, by reducing fecal microbiota transplantation (FMT, Fig. 2) is being investigated for
IFN-γ production and modulating both T and B cell trafficking and the management of both intestinal and extra-intestinal disorders asso-
checkpoint inhibitors. Nevertheless, human patients receiving experi- ciated with a microbiota dysbiosis. Theoretically, FMT increases gut
mental OCA therapy have been observed to exhibit side effects, devel- microbiota diversity, altered microbial metabolism, i.e. SCFA produc-
oping severe skin itching and showing an increase in total cholesterol, tion and bile acid biotransformation, thus limiting gut permeability and
LDL cholesterol and insulin, conferring an increased risk of develop- reducing local and systemic inflammation [287]. FMT has also been
ment of atherosclerosis. Thus, treatment of MS through the BA-FXR demonstrated to be durable, mostly safe and carry relative few side
interaction needs further investigation before transfer to clinical prac- effects [288,289]. FMT has shown some effects in IBD patients
tice [282]. Interfering with the BA-TGR5 interaction has also been [290–292] and other gastro-intestinal disorders, such as irritable bowel
shown to be effective in amelioration of MS-associated inflammation syndrome and pouchitis as recently reviewed [293,294]. Currently the
[119]. Furthermore, Lewis at al. showed that the TGR5 synthetic ago- studies investigating the potential of FMT therapy include case reports,
nist BIX02694 was able to alter cytokine production in response to LPS case series, but few controlled randomized trial and cohort studies. One
activation in vivo and reduced EAE severity by decreasing monocytes Crohn's disease patient, previously failing immunosuppressant therapy,
and microglia activity and reducing the numbers of central nervous showed remission of colitis after a single FMT infusion [295]. A ran-
system infiltrating monocytes and T cells. This results in the EAE model domized controlled clinical trial showed that 6-weeks FMT treatment
are consistent with a report that oleanolic acid, a TGR5/FXR agonist, (once a week) induces remission in a significantly greater percentage of
can reduce disease severity in EAE by impacting myeloid cell activation patients with active UC than placebo (9 patients vs 2 patients, out of 70
and pro-inflammatory cytokine and chemokine production [120]. Si- completing the study) [296].
milarly, McMahan et al. [127] demonstrated that a dual TGR5/FXR Some evidence exists on the efficacy of FMT in the treatment of
agonist increased the proportion of Ly6C-low monocytes and increased extra-intestinal disorders [212,293,297]. A case report study on a

13
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

hepatic encephalopathy patient showed that after a 4 weeks FMT within the gut and at other body sites. Boosting beneficial bacterial
treatment his mental status returned to baseline, suggesting the efficacy metabolites, including SCFAs and BA, using prebiotics and/or probio-
of FMT in ameliorating hepatic encephalopathy [298]. A recent clinical tics or by incorporating them into the diet seems to be a very attractive
trial [299] evaluated a modified FMT protocol, termed Microbiota strategy. At least in animal studies such approaches have been shown to
Transfer Therapy (MTT), to resolve symptoms associated with autism in downregulate pro-inflammatory mediators and increase the number of
children. A 7–8 weeks MTT treatment resulted in 82% reduction in Treg cells, which is particularly important for T-cell mediated AID.
gastrointestinal symptoms, e.g. constipation, diarrhea, indigestion, ab- However, although most studies suggest that SCFAs and BA receptor
dominal pain. This effect was maintained after eight weeks of cessation agonisms ameliorate chronic inflammation associated to AID, caution is
of therapy. needed, as such metabolites could also exacerbate some autoimmune
Only few studies have so far demonstrated the impact of FMT in disorders, as observed in mice antibody induced-arthritis, where SCFA
ameliorating AID associated symptoms. A case series study showed an administration increased cellular infiltration and cartilage and bone
improvement in the neurological symptoms and quality of life of three destruction [273]. Dietary microbiota modulation might also account
MS patients after FMT for chronic constipation [300]. Disease pro- for why diet apparently affects male and female differently, potentially
gression was also halted in one of those patients [300]. A phase I/II requiring dietary intervention studies to be powered to be able to
clinical trial evaluating the efficacy and safety of FMT in treating measure sex effects. In conclusion, several lines of evidence, the ma-
cholestatic liver disease in ongoing [301]. FMT has been shown success jority of which have been retrieved from animal studies, indicate an
in two men affected by alopecia areata, an AID causing hair loss [302]. important contribution of the gut microbiota and its metabolites to
The effect of FMT on T1D development was investigated in Non obese immune homeostasis and regulation of intestinal permeability, two
diabetic mice showing that short-term oral transfer of fecal bacteria, physiological processes thought to be involved in the onset of extra-
from different diabetes resistant mouse strains into diabetes-prone mice intestinal AID. The potential of their manipulation through tailored
delayed the onset of T1D [156]. dietary interventions as a strategy in reducing the risk of AID at a po-
Thus, despite preliminary indications of the potential efficacy of pulation level may therefore prove a promising preventative strategy.
FMT in improving AID associated with gut dysbiosis, randomized Use of dietary interventions to treat AID is still in its infancy however,
controlled trials on large population should be performed to determine with few studies showing positive effects.
the effectiveness of the therapy. Furthermore, FMT approaches are
probably unsuitable at present for AID prevention trials because of the Conflicts of interest
problems of standardizing stool composition that could include com-
municable pathogens as well as current clinical investigation highlight The authors declare that no conflicts of interest exist.
the importance of choosing the right donor to have the expected results.
Funding
6. Conclusions
This study was funded in part by Accordo di Programma between
AID is a constellation of chronic, debilitating, often severe and fatal Fondazione Edmund Mach (San Michele all’Adige, Italy) and
immune related diseases affecting about 10% of the population and is Autonomous Province of Trento (Trento, Italy); in part by the ERA-
rapidly increasing in incidence wherever Western diet and modern life- HDHL action project “CABALA_DIETH&HEALTH. CirculAting Bile Acids
style prevail. For many patients, treatments are ineffective and they as biomarkers of metabolic Health_Linking microbiotA, diet and
often present with serious co-morbidities. Genetic and environmental Health” (http://www.healthydietforhealthylife.eu/index.php/64-open-
triggers have long been considered as the main drivers of auto- calls/309-cabala-diet-health).
immunity. Increasing evidence in recent years suggests that microbial
translocation, and the consequent immunological homeostasis disrup- References
tion are important environmental factors involved in AID manifestation
and development. [1] J.-M. Anaya, C. Ramirez-Santana, M.A. Alzate, N. Molano-Gonzalez, A. Rojas-
The clinical impact of microbiota dysbiosis is not completely un- Villarraga, The autoimmune ecology, Front. Immunol. 7 (2016), http://dx.doi.
org/10.3389/fimmu.2016.00139.
derstood and whether dysbiosis precedes AID manifestation or whether [2] A.E. Livanos, T.U. Greiner, P. Vangay, W. Pathmasiri, D. Stewart, S. McRitchie,
the disease itself alters microbiota composition and function, remains to H. Li, J. Chung, J. Sohn, S. Kim, Z. Gao, C. Barber, J. Kim, S. Ng, A.B. Rogers,
be fully elucidated. The discovery of specific activating and inhibitory S. Sumner, X.-S. Zhang, K. Cadwell, D. Knights, A. Alekseyenko, F. Bäckhed,
M.J. Blaser, Antibiotic-mediated gut microbiome perturbation accelerates devel-
receptors triggered by the microbiota and its metabolites and engaged opment of type 1 diabetes in mice, Nat. Microbiol. 1 (2016) 16140, http://dx.doi.
in autoimmunity also suggests a strong microbiota-autoimmunity con- org/10.1038/nmicrobiol.2016.140.
nection. This microbiota connection has been strengthened by the re- [3] A.N. Thorburn, L. Macia, C.R. Mackay, Diet, metabolites, and “Western-Lifestyle”
inflammatory diseases, Immunity 40 (2014) 833–842, http://dx.doi.org/10.1016/
cent suggestions of a tight reciprocal regulation between immune j.immuni.2014.05.014.
functions, sexual hormones and microbiota composition. Future ad- [4] F. Colotta, B. Jansson, F. Bonelli, Modulation of inflammatory and immune re-
vances could be achieved by specific efforts to go beyond mere corre- sponses by vitamin D, J. Autoimmun. 85 (2017) 78–97, http://dx.doi.org/10.
1016/j.jaut.2017.07.007.
lation and description of changes in microbiota composition at phylum,
[5] E. Ortona, M. Pierdominici, A. Maselli, C. Veroni, F. Aloisi, Y. Shoenfeld, Sex-based
genus or even species levels, towards metabolic functionality. Gut mi- differences in autoimmune diseases, Ann. Ist. Super Sanita 52 (2016) 205–212,
crobiota members in fact, do not function as single isolated individuals http://dx.doi.org/10.4415/ANN_16_02_12.
and gut microbiota activities are exerted in concert and synergistically [6] L.V. Hooper, D.R. Littman, A.J. Macpherson, Interactions between the microbiota
and the immune system, Science 336 (2012) 1268–1273, http://dx.doi.org/10.
with a high degree of cooperation between different bacterial compo- 1126/science.1223490.
nents therein, as a whole (Box 1 and 2). Given the observed connections [7] B. Chen, L. Sun, X. Zhang, Integration of microbiome and epigenome to decipher
between gut microbiota composition and metabolic output and AID, the the pathogenesis of autoimmune diseases, J. Autoimmun. 83 (2017) 31–42,
http://dx.doi.org/10.1016/j.jaut.2017.03.009.
potential use of dietary interventions designed to correct the im- [8] G. Goverse, R. Molenaar, L. Macia, J. Tan, M.N. Erkelens, T. Konijn,
munological defects contributing to AID progression may hold promise. M. Knippenberg, E.C.L. Cook, D. Hanekamp, M. Veldhoen, A. Hartog, G. Roeselers,
Dietary manipulation of microbial composition on relative abundance C.R. Mackay, R.E. Mebius, Diet-derived short chain fatty acids stimulate intestinal
epithelial cells to induce mucosal tolerogenic dendritic cells, J. Immunol. 198
within the gut has the potential to radically resolve chronic unchecked (2017) 2172–2181, http://dx.doi.org/10.4049/jimmunol.1600165.
inflammation by changing the metabolic activity of the gut microbiota. [9] L. Yurkovetskiy, J.M. Pickard, A.V. Chervonsky, Microbiota and Autoimmunity:
SCFA and BA have been shown to influence host metabolic function, exploring new avenues, Cell Host Microbe 17 (2015) 548–552, http://dx.doi.org/
10.1016/j.chom.2015.04.010.
integrity of the intestinal epithelium and also the immune system both

14
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

[10] P.C. Dorrestein, S.K. Mazmanian, R. Knight, Finding the missing links among term bisphenol A exposure accelerates insulitis development in diabetes-prone
metabolites, microbes, and the host, Immunity 40 (2014) 824–832, http://dx.doi. NOD mice, Immunopharmacol. Immunotoxicol. 35 (2013) 349–358, http://dx.
org/10.1016/j.immuni.2014.05.015. doi.org/10.3109/08923973.2013.772195.
[11] E.C. Rosser, C. Mauri, A clinical update on the significance of the gut microbiota in [37] J. Bodin, A.K. Bølling, R. Becher, F. Kuper, M. Løvik, U.C. Nygaard, Transmaternal
systemic autoimmunity, J. Autoimmun. 74 (2016) 85–93, http://dx.doi.org/10. bisphenol A exposure accelerates diabetes type 1 development in NOD mice,
1016/j.jaut.2016.06.009. Toxicol. Sci. Off. J. Soc. Toxicol 137 (2014) 311–323, http://dx.doi.org/10.1093/
[12] J.G. Markle, E.N. Fish, SeXX matters in immunity, Trends Immunol. 35 (2014) toxsci/kft242.
97–104, http://dx.doi.org/10.1016/j.it.2013.10.006. [38] K.M. Pollard, D.M. Cauvi, C.B. Toomey, K.V. Morris, D.H. Kono, Interferon-γ and
[13] J. Ji, J. Sundquist, K. Sundquist, Gender-specific incidence of autoimmune dis- systemic autoimmunity, Discov. Med. 16 (2013) 123–131.
eases from national registers, J. Autoimmun. 69 (2016) 102–106, http://dx.doi. [39] M.R. Edwards, R. Dai, B. Heid, T.E. Cecere, D. Khan, Q. Mu, C. Cowan, X.M. Luo,
org/10.1016/j.jaut.2016.03.003. S.A. Ahmed, Commercial rodent diets differentially regulate autoimmune glo-
[14] M. Hauptmann, U.E. Schaible, Linking microbiota and respiratory disease, FEBS merulonephritis, epigenetics and microbiota in MRL/lpr mice, Int. Immunol. 29
Lett. 590 (2016) 3721–3738, http://dx.doi.org/10.1002/1873-3468.12421. (2017) 263–276, http://dx.doi.org/10.1093/intimm/dxx033.
[15] M. Prosberg, F. Bendtsen, I. Vind, A.M. Petersen, L.L. Gluud, The association be- [40] A. Beury-Cirou, M. Tannières, C. Minard, L. Soulère, T. Rasamiravaka, R.H. Dodd,
tween the gut microbiota and the inflammatory bowel disease activity: a sys- Y. Queneau, Y. Dessaux, C. Guillou, O.M. Vandeputte, D. Faure, At a supra-phy-
tematic review and meta-analysis, Scand. J. Gastroenterol. 51 (2016) 1407–1415, siological concentration, human sexual hormones act as quorum-sensing in-
http://dx.doi.org/10.1080/00365521.2016.1216587. hibitors, PLoS One 8 (2013) e83564, , http://dx.doi.org/10.1371/journal.pone.
[16] E. Gianchecchi, A. Fierabracci, On the pathogenesis of insulin-dependent diabetes 0083564.
mellitus: the role of microbiota, Immunol. Res. 65 (2017) 242–256, http://dx.doi. [41] K.S. Kornman, W.J. Loesche, Effects of estradiol and progesterone on Bacteroides
org/10.1007/s12026-016-8832-8. melaninogenicus and Bacteroides gingivalis, Infect. Immun. 35 (1982) 256–263.
[17] S. Abdollahi-Roodsaz, S.B. Abramson, J.U. Scher, The metabolic role of the gut [42] R. Menon, S.E. Watson, L.N. Thomas, C.D. Allred, A. Dabney, M.A. Azcarate-Peril,
microbiota in health and rheumatic disease: mechanisms and interventions, Nat. J.M. Sturino, Diet complexity and estrogen receptor β status affect the composition
Rev. Rheumatol. 12 (2016) 446–455, http://dx.doi.org/10.1038/nrrheum. of the murine intestinal microbiota, Appl. Environ. Microbiol. 79 (2013)
2016.68. 5763–5773, http://dx.doi.org/10.1128/AEM.01182-13.
[18] A.B. Blázquez, M.C. Berin, Microbiome and food allergy, Transl. Res. J. Lab. Clin. [43] F. Fransen, A.A. van Beek, T. Borghuis, B. Meijer, F. Hugenholtz, C. van der Gaast-
Med 179 (2017) 199–203, http://dx.doi.org/10.1016/j.trsl.2016.09.003. de Jongh, H.F. Savelkoul, M.I. de Jonge, M.M. Faas, M.V. Boekschoten, H. Smidt,
[19] J. Ochoa-Repáraz, K. Magori, L.H. Kasper, The chicken or the egg dilemma: in- S. El Aidy, P. de Vos, The impact of gut microbiota on gender-specific differences
testinal dysbiosis in multiple sclerosis, Ann. Transl. Med. 5 (2017), http://dx.doi. in immunity, Front. Immunol. 8 (2017) 754, http://dx.doi.org/10.3389/fimmu.
org/10.21037/atm.2017.01.18. 2017.00754.
[20] M. Spiljar, D. Merkler, M. Trajkovski, The immune system bridges the gut mi- [44] J.G.M. Markle, D.N. Frank, S. Mortin-Toth, C.E. Robertson, L.M. Feazel, U. Rolle-
crobiota with systemic energy homeostasis: focus on TLRs, mucosal barrier, and Kampczyk, M. von Bergen, K.D. McCoy, A.J. Macpherson, J.S. Danska, Sex dif-
SCFAs, Front. Immunol. 8 (2017), http://dx.doi.org/10.3389/fimmu.2017.01353. ferences in the gut microbiome drive hormone-dependent regulation of auto-
[21] D.L. Smith-Bouvier, A.A. Divekar, M. Sasidhar, S. Du, S.K. Tiwari-Woodruff, immunity, Science 339 (2013) 1084–1088, http://dx.doi.org/10.1126/science.
J.K. King, A.P. Arnold, R.R. Singh, R.R. Voskuhl, A role for sex chromosome 1233521.
complement in the female bias in autoimmune disease, J. Exp. Med. 205 (2008) [45] L. Yurkovetskiy, M. Burrows, A.A. Khan, L. Graham, P. Volchkov, L. Becker,
1099–1108, http://dx.doi.org/10.1084/jem.20070850. D. Antonopoulos, Y. Umesaki, A.V. Chervonsky, Gender bias in autoimmunity is
[22] E. Org, M. Mehrabian, B.W. Parks, P. Shipkova, X. Liu, T.A. Drake, A.J. Lusis, Sex influenced by microbiota, Immunity 39 (2013) 400–412, http://dx.doi.org/10.
differences and hormonal effects on gut microbiota composition in mice, Gut 1016/j.immuni.2013.08.013.
Microb. 7 (2016) 313–322, http://dx.doi.org/10.1080/19490976.2016.1203502. [46] D.I. Bolnick, L.K. Snowberg, P.E. Hirsch, C.L. Lauber, E. Org, B. Parks, A.J. Lusis,
[23] M. Edwards, R. Dai, S.A. Ahmed, Our environment shapes Us: the importance of R. Knight, J.G. Caporaso, R. Svanbäck, Individual diet has sex-dependent effects
environment and sex differences in regulation of autoantibody production, Front. on vertebrate gut microbiota, Nat. Commun. 5 (2014) 4500, http://dx.doi.org/10.
Immunol. 9 (2018), http://dx.doi.org/10.3389/fimmu.2018.00478. 1038/ncomms5500.
[24] H. Neuman, J.W. Debelius, R. Knight, O. Koren, Microbial endocrinology: the [47] S. Mueller, K. Saunier, C. Hanisch, E. Norin, L. Alm, T. Midtvedt, A. Cresci, S. Silvi,
interplay between the microbiota and the endocrine system, FEMS Microbiol. Rev. C. Orpianesi, M.C. Verdenelli, T. Clavel, C. Koebnick, H.-J.F. Zunft, J. Doré,
39 (2015) 509–521, http://dx.doi.org/10.1093/femsre/fuu010. M. Blaut, Differences in fecal microbiota in different European study populations
[25] E.N. Fish, The X-files in immunity: sex-based differences predispose immune re- in relation to age, gender, and country: a cross-sectional study, Appl. Environ.
sponses, Nat. Rev. Immunol. 8 (2008) 737–744, http://dx.doi.org/10.1038/ Microbiol. 72 (2006) 1027–1033, http://dx.doi.org/10.1128/AEM.72.2.1027-
nri2394. 1033.2006.
[26] S.T. Ngo, F.J. Steyn, P.A. McCombe, Gender differences in autoimmune disease, [48] O. Koren, J.K. Goodrich, T.C. Cullender, A. Spor, K. Laitinen, H.K. Bäckhed,
Front. Neuroendocrinol. 35 (2014) 347–369, http://dx.doi.org/10.1016/j.yfrne. A. Gonzalez, J.J. Werner, L.T. Angenent, R. Knight, F. Bäckhed, E. Isolauri,
2014.04.004. S. Salminen, R.E. Ley, Host remodeling of the gut microbiome and metabolic
[27] J. Asaba, M. Bandyopadhyay, M. Kindy, S. Dasgupta, Estrogen receptor signal in changes during pregnancy, Cell 150 (2012) 470–480, http://dx.doi.org/10.1016/
regulation of B cell activation during diverse immune responses, Int. J. Biochem. j.cell.2012.07.008.
Cell Biol. 68 (2015) 42–47, http://dx.doi.org/10.1016/j.biocel.2015.08.012. [49] B. Chassaing, O. Koren, J.K. Goodrich, A.C. Poole, S. Srinivasan, R.E. Ley,
[28] M. Cunningham, G. Gilkeson, Estrogen receptors in immunity and autoimmunity, A.T. Gewirtz, Dietary emulsifiers impact the mouse gut microbiota promoting
Clin. Rev. Allergy Immunol. 40 (2011) 66–73, http://dx.doi.org/10.1007/s12016- colitis and metabolic syndrome, Nature 519 (2015) 92–96, http://dx.doi.org/10.
010-8203-5. 1038/nature14232.
[29] J. Wang, C.M. Syrett, M.C. Kramer, A. Basu, M.L. Atchison, M.C. Anguera, Unusual [50] Z.D. Fu, I.L. Csanaky, C.D. Klaassen, Gender-divergent profile of bile acid home-
maintenance of X chromosome inactivation predisposes female lymphocytes for ostasis during aging of mice, PLoS One 7 (2012) e32551, , http://dx.doi.org/10.
increased expression from the inactive X, Proc. Natl. Acad. Sci. U. S. A 113 (2016) 1371/journal.pone.0032551.
E2029–E2038, http://dx.doi.org/10.1073/pnas.1520113113. [51] L. Sheng, P.K. Jena, H.-X. Liu, K.M. Kalanetra, F.J. Gonzalez, S.W. French,
[30] K.M. Nickerson, S.R. Christensen, J. Shupe, M. Kashgarian, D. Kim, K. Elkon, V.V. Krishnan, D.A. Mills, Y.-J.Y. Wan, Gender differences in bile acids and mi-
M.J. Shlomchik, TLR9 regulates TLR7- and MyD88-dependent autoantibody pro- crobiota in relationship with gender dissimilarity in steatosis induced by diet and
duction and disease in a murine model of lupus, J. Immunol. Baltim. Md 1950 184 FXR inactivation, Sci. Rep. 7 (2017) 1748, http://dx.doi.org/10.1038/s41598-
(2010) 1840–1848, http://dx.doi.org/10.4049/jimmunol.0902592. 017-01576-9.
[31] W. Jiang, G. Gilkeson, Sex Differences in monocytes and TLR4 associated immune [52] K.M. Eyster, The estrogen receptors: an overview from different perspectives,
responses; implications for systemic lupus erythematosus (SLE), J. Immunother. Estrogen Recept, Humana Press, New York, NY, 2016, pp. 1–10, , http://dx.doi.
Appl. 1 (2014) 1, http://dx.doi.org/10.7243/2055-2394-1-1. org/10.1007/978-1-4939-3127-9_1.
[32] J.L. Svenson, J. EuDaly, P. Ruiz, K.S. Korach, G.S. Gilkeson, Impact of estrogen [53] H. Homma, E. Hoy, D.-Z. Xu, Q. Lu, R. Feinman, E.A. Deitch, The female intestine
receptor deficiency on disease expression in the NZM2410 lupus prone mouse, is more resistant than the male intestine to gut injury and inflammation when
Clin. Immunol. Orlando Fla 128 (2008) 259–268, http://dx.doi.org/10.1016/j. subjected to conditions associated with shock states, Am. J. Physiol. Gastrointest.
clim.2008.03.508. Liver Physiol. 288 (2005) G466–G472, http://dx.doi.org/10.1152/ajpgi.00036.
[33] M.A. Cunningham, O.S. Naga, J.G. Eudaly, J.L. Scott, G.S. Gilkeson, Estrogen re- 2004.
ceptor alpha modulates Toll-like receptor signaling in murine lupus, Clin. [54] J.M. Baker, L. Al-Nakkash, M.M. Herbst-Kralovetz, Estrogen-gut microbiome axis:
Immunol. Orlando Fla 144 (2012) 1–12, http://dx.doi.org/10.1016/j.clim.2012. physiological and clinical implications, Maturitas 103 (2017) 45–53, http://dx.
04.001. doi.org/10.1016/j.maturitas.2017.06.025.
[34] G. Benedek, J. Zhang, H. Nguyen, G. Kent, H.A. Seifert, S. Davin, P. Stauffer, [55] H. Adlercreutz, M.O. Pulkkinen, E.K. Hämäläinen, J.T. Korpela, Studies on the role
A.A. Vandenbark, L. Karstens, M. Asquith, H. Offner, Estrogen protection against of intestinal bacteria in metabolism of synthetic and natural steroid hormones, J.
EAE modulates the microbiota and mucosal-associated regulatory cells, J. Steroid Biochem. 20 (1984) 217–229.
Neuroimmunol. 310 (2017) 51–59, http://dx.doi.org/10.1016/j.jneuroim.2017. [56] D.-J. Kang, J.M. Ridlon, D.R. Moore, S. Barnes, P.B. Hylemon, Clostridium scin-
06.007. dens baiCD and baiH genes encode stereo-specific 7α/7β-hydroxy-3-oxo-Δ4-cho-
[35] J. Bodin, A. Kocbach Bølling, A. Wendt, L. Eliasson, R. Becher, F. Kuper, M. Løvik, lenoic acid oxidoreductases, Biochim. Biophys. Acta 1781 (2008) 16–25, http://
U.C. Nygaard, Exposure to bisphenol A, but not phthalates, increases spontaneous dx.doi.org/10.1016/j.bbalip.2007.10.008.
diabetes type 1 development in NOD mice, Toxicol. Rep 2 (2015) 99–110, http:// [57] J.M. Ridlon, S. Ikegawa, J.M.P. Alves, B. Zhou, A. Kobayashi, T. Iida, K. Mitamura,
dx.doi.org/10.1016/j.toxrep.2015.02.010. G. Tanabe, M. Serrano, A. De Guzman, P. Cooper, G.A. Buck, P.B. Hylemon,
[36] J. Bodin, A.K. Bølling, M. Samuelsen, R. Becher, M. Løvik, U.C. Nygaard, Long- Clostridium scindens: a human gut microbe with a high potential to convert

15
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

glucocorticoids into androgens, J. Lipid Res. 54 (2013) 2437–2449, http://dx.doi. nu9090962.


org/10.1194/jlr.M038869. [80] K.A. Krautkramer, J.H. Kreznar, K.A. Romano, E.I. Vivas, G.A. Barrett-Wilt,
[58] M. Baptissart, E. Martinot, A. Vega, L. Sédes, B. Rouaisnel, A. de Haze, S. Baron, M.E. Rabaglia, M.P. Keller, A.D. Attie, F.E. Rey, J.M. Denu, Diet-microbiota in-
K. Schoonjans, F. Caira, D.H. Volle, Bile acid-FXRα pathways regulate male sexual teractions mediate global epigenetic programming in multiple host tissues, Mol.
maturation in mice, Oncotarget 7 (2016) 19468–19482, http://dx.doi.org/10. Cell. 64 (2016) 982–992, http://dx.doi.org/10.1016/j.molcel.2016.10.025.
18632/oncotarget.7153. [81] A. Nakajima, N. Negishi, H. Tsurui, N. Kadowaki-Ohtsuji, K. Maeda, M. Nanno,
[59] J.M. Ridlon, J.S. Bajaj, The human gut sterolbiome: bile acid-microbiome endo- Y. Yamaguchi, N. Shimizu, H. Yagita, K. Okumura, S. Habu, Commensal bacteria
crine aspects and therapeutics, Acta Pharm. Sin. B. 5 (2015) 99–105, http://dx. regulate thymic aire expression, PLoS One 9 (2014), http://dx.doi.org/10.1371/
doi.org/10.1016/j.apsb.2015.01.006. journal.pone.0105904.
[60] D.G. Souza, A.T. Vieira, A.C. Soares, V. Pinho, J.R. Nicoli, L.Q. Vieira, [82] B.A. Shenderov, Gut indigenous microbiota and epigenetics, Microb. Ecol. Health
M.M. Teixeira, The essential role of the intestinal microbiota in facilitating acute Dis. 23 (2012), http://dx.doi.org/10.3402/mehd.v23i0.17195.
inflammatory responses, J. Immunol. Baltim. Md 1950 173 (2004) 4137–4146. [83] K.A. Krautkramer, F.E. Rey, J.M. Denu, Chemical signaling between gut micro-
[61] T. Selvanantham, Q. Lin, C.X. Guo, A. Surendra, S. Fieve, N.K. Escalante, biota and host chromatin: what is your gut really saying? J. Biol. Chem. (2017),
D.S. Guttman, C.J. Streutker, S.J. Robertson, D.J. Philpott, T. Mallevaey, NKT cell- http://dx.doi.org/10.1074/jbc.R116.761577 jbc.R116.761577.
deficient mice harbor an altered microbiota that fuels intestinal inflammation [84] B. Hippe, M. Remely, E. Aumueller, A. Pointner, A.G. Haslberger, SCFA producing
during chemically induced colitis, J. Immunol. Baltim. Md 1950 197 (2016) gut microbiota and its effects on the epigenetic regulation of inflammation, Benef.
4464–4472, http://dx.doi.org/10.4049/jimmunol.1601410. Microorg. Med. Health Appl., Springer, Cham (2015) 181–197, http://dx.doi.org/
[62] T.B. Clarke, K.M. Davis, E.S. Lysenko, A.Y. Zhou, Y. Yu, J.N. Weiser, Recognition 10.1007/978-3-319-23213-3_9.
of peptidoglycan from the microbiota by Nod1 enhances systemic innate im- [85] J. Park, M. Kim, S.G. Kang, A.H. Jannasch, B. Cooper, J. Patterson, C.H. Kim,
munity, Nat. Med. 16 (2010) 228–231, http://dx.doi.org/10.1038/nm.2087. Short-chain fatty acids induce both effector and regulatory T cells by suppression
[63] D. An, S.F. Oh, T. Olszak, J.F. Neves, F.Y. Avci, D. Erturk-Hasdemir, X. Lu, of histone deacetylases and regulation of the mTOR-S6K pathway, Mucosal
S. Zeissig, R.S. Blumberg, D.L. Kasper, Sphingolipids from a symbiotic microbe Immunol. 8 (2015) 80–93, http://dx.doi.org/10.1038/mi.2014.44.
regulate homeostasis of host intestinal natural killer T cells, Cell 156 (2014) [86] C.H. Kim, J. Park, M. Kim, Gut microbiota-derived short-chain fatty acids, T cells,
123–133, http://dx.doi.org/10.1016/j.cell.2013.11.042. and inflammation, Immune Netw 14 (2014) 277–288, http://dx.doi.org/10.4110/
[64] K. Atarashi, T. Tanoue, M. Ando, N. Kamada, Y. Nagano, S. Narushima, W. Suda, in.2014.14.6.277.
A. Imaoka, H. Setoyama, T. Nagamori, E. Ishikawa, T. Shima, T. Hara, S. Kado, [87] J. Tan, C. McKenzie, M. Potamitis, A.N. Thorburn, C.R. Mackay, L. Macia, The role
T. Jinnohara, H. Ohno, T. Kondo, K. Toyooka, E. Watanabe, S.-I. Yokoyama, of short-chain fatty acids in health and disease, Adv. Immunol. 121 (2014)
S. Tokoro, H. Mori, Y. Noguchi, H. Morita, I.I. Ivanov, T. Sugiyama, G. Nuñez, 91–119, http://dx.doi.org/10.1016/B978-0-12-800100-4.00003-9.
J.G. Camp, M. Hattori, Y. Umesaki, K. Honda, Th17 cell induction by adhesion of [88] K.A. Kuhn, T.S. Stappenbeck, Peripheral education of the immune system by the
microbes to intestinal epithelial cells, Cell 163 (2015) 367–380, http://dx.doi.org/ colonic microbiota, Semin. Immunol. 25 (2013) 364–369, http://dx.doi.org/10.
10.1016/j.cell.2015.08.058. 1016/j.smim.2013.10.002.
[65] A.M. Farkas, C. Panea, Y. Goto, G. Nakato, M. Galan-Diez, S. Narushima, K. Honda, [89] M. Kasubuchi, S. Hasegawa, T. Hiramatsu, A. Ichimura, I. Kimura, Dietary gut
I.I. Ivanov, Colonization and induction of Th17 cells by segmented filamentous microbial metabolites, short-chain fatty acids, and host metabolic regulation,
bacteria in the murine intestine, J. Immunol. Meth. 421 (2015) 104–111, http:// Nutrients 7 (2015) 2839–2849, http://dx.doi.org/10.3390/nu7042839.
dx.doi.org/10.1016/j.jim.2015.03.020. [90] J.G. LeBlanc, F. Chain, R. Martín, L.G. Bermúdez-Humarán, S. Courau, P. Langella,
[66] I.I. Ivanov, K. Atarashi, N. Manel, E.L. Brodie, T. Shima, U. Karaoz, D. Wei, Beneficial effects on host energy metabolism of short-chain fatty acids and vita-
K.C. Goldfarb, C.A. Santee, S.V. Lynch, T. Tanoue, A. Imaoka, K. Itoh, K. Takeda, mins produced by commensal and probiotic bacteria, Microb. Cell Factories 16
Y. Umesaki, K. Honda, D.R. Littman, Induction of intestinal Th17 cells by seg- (2017), http://dx.doi.org/10.1186/s12934-017-0691-z.
mented filamentous bacteria, Cell 139 (2009) 485–498, http://dx.doi.org/10. [91] C.N. Heiss, L.E. Olofsson, Gut microbiota-dependent modulation of energy meta-
1016/j.cell.2009.09.033. bolism, J. Innate Immun (2017), http://dx.doi.org/10.1159/000481519.
[67] Y.K. Lee, J.S. Menezes, Y. Umesaki, S.K. Mazmanian, Proinflammatory T-cell re- [92] W. Wu, M. Sun, F. Chen, A.T. Cao, H. Liu, Y. Zhao, X. Huang, Y. Xiao, S. Yao,
sponses to gut microbiota promote experimental autoimmune encephalomyelitis, Q. Zhao, Z. Liu, Y. Cong, Microbiota metabolite short chain fatty acid acetate
Proc. Natl. Acad. Sci. U. S. A 108 (2011) 4615–4622, http://dx.doi.org/10.1073/ promotes intestinal IgA response to microbiota which is mediated by GPR43,
pnas.1000082107. Mucosal Immunol. 10 (2017) 946–956, http://dx.doi.org/10.1038/mi.2016.114.
[68] X. Wu, Z. Tian, Gut-liver axis: gut microbiota in shaping hepatic innate immunity, [93] N. Singh, A. Gurav, S. Sivaprakasam, E. Brady, R. Padia, H. Shi, M. Thangaraju,
Sci. China Life Sci. 60 (2017) 1191–1196, http://dx.doi.org/10.1007/s11427-017- P.D. Prasad, S. Manicassamy, D.H. Munn, J.R. Lee, S. Offermanns, V. Ganapathy,
9128-3. Activation of the receptor (Gpr109a) for niacin and the commensal metabolite
[69] H.-D. Ma, Y.-H. Wang, C. Chang, M.E. Gershwin, Z.-X. Lian, The intestinal mi- butyrate suppresses colonic inflammation and carcinogenesis, Immunity 40 (2014)
crobiota and microenvironment in liver, Autoimmun. Rev. 14 (2015) 183–191, 128–139, http://dx.doi.org/10.1016/j.immuni.2013.12.007.
http://dx.doi.org/10.1016/j.autrev.2014.10.013. [94] E. Elinav, T. Strowig, A.L. Kau, J. Henao-Mejia, C.A. Thaiss, C.J. Booth,
[70] A. Okamoto, K. Fujio, N.H. Tsuno, K. Takahashi, K. Yamamoto, Kidney-infiltrating D.R. Peaper, J. Bertin, S.C. Eisenbarth, J.I. Gordon, R.A. Flavell, NLRP6 in-
CD4+ T-cell clones promote nephritis in lupus-prone mice, Kidney Int. 82 (2012) flammasome regulates colonic microbial ecology and risk for colitis, Cell 145
969–979, http://dx.doi.org/10.1038/ki.2012.242. (2011) 745–757, http://dx.doi.org/10.1016/j.cell.2011.04.022.
[71] S.T. Jerram, M.N. Dang, R.D. Leslie, The role of epigenetics in type 1 diabetes, [95] A. Trompette, E.S. Gollwitzer, K. Yadava, A.K. Sichelstiel, N. Sprenger, C. Ngom-
Curr. Diabetes Rep. 17 (2017), http://dx.doi.org/10.1007/s11892-017-0916-x. Bru, C. Blanchard, T. Junt, L.P. Nicod, N.L. Harris, B.J. Marsland, Gut microbiota
[72] H. Long, H. Yin, L. Wang, M.E. Gershwin, Q. Lu, The critical role of epigenetics in metabolism of dietary fiber influences allergic airway disease and hematopoiesis,
systemic lupus erythematosus and autoimmunity, J. Autoimmun. 74 (2016) Nat. Med. 20 (2014) 159–166, http://dx.doi.org/10.1038/nm.3444.
118–138, http://dx.doi.org/10.1016/j.jaut.2016.06.020. [96] P.M. Smith, M.R. Howitt, N. Panikov, M. Michaud, C.A. Gallini, M. Bohlooly-Y,
[73] R. Dai, S.A. Ahmed, Sexual dimorphism of miRNA expression: a new perspective in J.N. Glickman, W.S. Garrett, The microbial metabolites, short-chain fatty acids,
understanding the sex bias of autoimmune diseases, Therapeut. Clin. Risk Manag. regulate colonic Treg cell homeostasis, Science 341 (2013) 569–573, http://dx.
10 (2014) 151–163, http://dx.doi.org/10.2147/TCRM.S33517. doi.org/10.1126/science.1241165.
[74] L. Li, K.-M. Lee, W. Han, J.-Y. Choi, J.-Y. Lee, G.H. Kang, S.K. Park, D.-Y. Noh, K.- [97] N. Arpaia, C. Campbell, X. Fan, S. Dikiy, J. van der Veeken, P. deRoos, H. Liu,
Y. Yoo, D. Kang, Estrogen and progesterone receptor status affect genome-wide J.R. Cross, K. Pfeffer, P.J. Coffer, A.Y. Rudensky, Metabolites produced by com-
DNA methylation profile in breast cancer, Hum. Mol. Genet. 19 (2010) mensal bacteria promote peripheral regulatory T-cell generation, Nature 504
4273–4277, http://dx.doi.org/10.1093/hmg/ddq351. (2013) 451–455, http://dx.doi.org/10.1038/nature12726.
[75] Y. Xu, L. Chao, J. Wang, Y. Sun, miRNA-148a regulates the expression of the es- [98] R. Schilderink, C. Verseijden, J. Seppen, V. Muncan, G.R. van den Brink,
trogen receptor through DNMT1-mediated DNA methylation in breast cancer cells, T.T. Lambers, E.A. van Tol, W.J. de Jonge, The SCFA butyrate stimulates the
Oncol. Lett. 14 (2017) 4736–4740, http://dx.doi.org/10.3892/ol.2017.6803. epithelial production of retinoic acid via inhibition of epithelial HDAC, Am. J.
[76] A. Stone, E. Zotenko, W.J. Locke, D. Korbie, E.K.A. Millar, R. Pidsley, C. Stirzaker, Physiol. Gastrointest, Liver Physiol 310 (2016) G1138–G1146, http://dx.doi.org/
P. Graham, M. Trau, E.A. Musgrove, R.I. Nicholson, J.M.W. Gee, S.J. Clark, DNA 10.1152/ajpgi.00411.2015.
methylation of oestrogen-regulated enhancers defines endocrine sensitivity in [99] Y. Zhang, J. Maksimovic, G. Naselli, J. Qian, M. Chopin, M.E. Blewitt, A. Oshlack,
breast cancer, Nat. Commun. 6 (2015) 7758, http://dx.doi.org/10.1038/ L.C. Harrison, Genome-wide DNA methylation analysis identifies hypomethylated
ncomms8758. genes regulated by FOXP3 in human regulatory T cells, Blood 122 (2013)
[77] O. El-Maarri, T. Becker, J. Junen, S.S. Manzoor, A. Diaz-Lacava, R. Schwaab, 2823–2836, http://dx.doi.org/10.1182/blood-2013-02-481788.
T. Wienker, J. Oldenburg, Gender specific differences in levels of DNA methylation [100] Y. Furusawa, Y. Obata, S. Fukuda, T.A. Endo, G. Nakato, D. Takahashi,
at selected loci from human total blood: a tendency toward higher methylation Y. Nakanishi, C. Uetake, K. Kato, T. Kato, M. Takahashi, N.N. Fukuda,
levels in males, Hum. Genet. 122 (2007) 505–514, http://dx.doi.org/10.1007/ S. Murakami, E. Miyauchi, S. Hino, K. Atarashi, S. Onawa, Y. Fujimura, T. Lockett,
s00439-007-0430-3. J.M. Clarke, D.L. Topping, M. Tomita, S. Hori, O. Ohara, T. Morita, H. Koseki,
[78] N. Dragin, J. Bismuth, G. Cizeron-Clairac, M.G. Biferi, C. Berthault, A. Serraf, J. Kikuchi, K. Honda, K. Hase, H. Ohno, Commensal microbe-derived butyrate
R. Nottin, D. Klatzmann, A. Cumano, M. Barkats, R. Le Panse, S. Berrih-Aknin, induces the differentiation of colonic regulatory T cells, Nature 504 (2013)
Estrogen-mediated downregulation of AIRE influences sexual dimorphism in au- 446–450, http://dx.doi.org/10.1038/nature12721.
toimmune diseases, J. Clin. Invest. 126 (2016) 1525–1537, http://dx.doi.org/10. [101] M.B. Geuking, J. Cahenzli, M.A.E. Lawson, D.C.K. Ng, E. Slack, S. Hapfelmeier,
1172/JCI81894. K.D. McCoy, A.J. Macpherson, Intestinal bacterial colonization induces mutualistic
[79] K. Aleksandrova, B. Romero-Mosquera, V. Hernandez, Diet, gut microbiome and regulatory T cell responses, Immunity 34 (2011) 794–806, http://dx.doi.org/10.
epigenetics: emerging links with inflammatory bowel diseases and prospects for 1016/j.immuni.2011.03.021.
management and prevention, Nutrients 9 (2017), http://dx.doi.org/10.3390/ [102] A. Gurav, S. Sivaprakasam, Y.D. Bhutia, T. Boettger, N. Singh, V. Ganapathy,

16
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Slc5a8, a Na+-coupled high-affinity transporter for short-chain fatty acids, is a 2011.11.006.


conditional tumor suppressor in colon that protects against colitis and colon [123] R.M. Gadaleta, K.J. van Erpecum, B. Oldenburg, E.C.L. Willemsen, W. Renooij,
cancer under low-fiber dietary conditions, Biochem. J. 469 (2015) 267–278, S. Murzilli, L.W.J. Klomp, P.D. Siersema, M.E.I. Schipper, S. Danese, G. Penna,
http://dx.doi.org/10.1042/BJ20150242. G. Laverny, L. Adorini, A. Moschetta, S.W.C. van Mil, Farnesoid X receptor acti-
[103] M. Kim, Y. Qie, J. Park, C.H. Kim, Gut microbial metabolites fuel host antibody vation inhibits inflammation and preserves the intestinal barrier in inflammatory
responses, Cell Host Microbe 20 (2016) 202–214, http://dx.doi.org/10.1016/j. bowel disease, Gut 60 (2011) 463–472, http://dx.doi.org/10.1136/gut.2010.
chom.2016.07.001. 212159.
[104] Y.-Y. Li, J.A. Pearson, C. Chao, J. Peng, X. Zhang, Z. Zhou, Y. Liu, F.S. Wong, [124] P. Vavassori, A. Mencarelli, B. Renga, E. Distrutti, S. Fiorucci, The bile acid re-
L. Wen, Nucleotide-binding oligomerization domain-containing protein 2 (Nod2) ceptor FXR is a modulator of intestinal innate immunity, J. Immunol. Baltim. Md
modulates T1DM susceptibility by gut microbiota, J. Autoimmun. 82 (2017) 1950 183 (2009) 6251–6261, http://dx.doi.org/10.4049/jimmunol.0803978.
85–95, http://dx.doi.org/10.1016/j.jaut.2017.05.007. [125] L. Verbeke, R. Farre, B. Verbinnen, K. Covens, T. Vanuytsel, J. Verhaegen,
[105] A. Paun, C. Yau, J.S. Danska, Immune recognition and response to the intestinal M. Komuta, T. Roskams, S. Chatterjee, P. Annaert, I. Vander Elst, P. Windmolders,
microbiome in type 1 diabetes, J. Autoimmun. 71 (2016) 10–18, http://dx.doi. J. Trebicka, F. Nevens, W. Laleman, The FXR agonist obeticholic acid prevents gut
org/10.1016/j.jaut.2016.02.004. barrier dysfunction and bacterial translocation in cholestatic rats, Am. J. Pathol.
[106] J.R. Kelly, P.J. Kennedy, J.F. Cryan, T.G. Dinan, G. Clarke, N.P. Hyland, Breaking 185 (2015) 409–419, http://dx.doi.org/10.1016/j.ajpath.2014.10.009.
down the barriers: the gut microbiome, intestinal permeability and stress-related [126] T. Kida, Y. Tsubosaka, M. Hori, H. Ozaki, T. Murata, Bile acid receptor TGR5
psychiatric disorders, Front. Cell. Neurosci. 9 (2015), http://dx.doi.org/10.3389/ agonism induces NO production and reduces monocyte adhesion in vascular en-
fncel.2015.00392. dothelial cells, Arterioscler. Thromb. Vasc. Biol. 33 (2013) 1663–1669, http://dx.
[107] L.W. van den Elsen, H.C. Poyntz, L.S. Weyrich, W. Young, E.E. Forbes-Blom, doi.org/10.1161/ATVBAHA.113.301565.
Embracing the gut microbiota: the new frontier for inflammatory and infectious [127] R.H. McMahan, X.X. Wang, L.L. Cheng, T. Krisko, M. Smith, K.E. Kasmi,
diseases, Clin. Transl. Immunol 6 (2017) e125, http://dx.doi.org/10.1038/cti. M. Pruzanski, L. Adorini, L. Golden-Mason, M. Levi, H.R. Rosen, Bile acid receptor
2016.91. activation modulates hepatic monocyte activity and improves nonalcoholic fatty
[108] Z. Xia, Y. Han, K. Wang, S. Guo, D. Wu, X. Huang, Z. Li, L. Zhu, Oral adminis- liver disease, J. Biol. Chem. 288 (2013) 11761–11770, http://dx.doi.org/10.
tration of propionic acid during lactation enhances the colonic barrier function, 1074/jbc.M112.446575.
Lipids Health Dis. 16 (2017) 62, http://dx.doi.org/10.1186/s12944-017-0452-3. [128] H. Yang, H. Zhou, L. Zhuang, J. Auwerx, K. Schoonjans, X. Wang, C. Feng, L. Lu,
[109] L. Zheng, C.J. Kelly, K.D. Battista, R. Schaefer, J.M. Lanis, E.E. Alexeev, R.X. Wang, Plasma membrane-bound G protein-coupled bile acid receptor attenuates liver
J.C. Onyiah, D.J. Kominsky, S.P. Colgan, Microbial-derived butyrate promotes ischemia/reperfusion injury via the inhibition of toll-like receptor 4 signaling in
epithelial barrier function through IL-10 receptor-dependent repression of mice, Liver Transplant. Off. Publ. Am. Assoc. Study Liver Dis. Int. Liver Transplant.
Claudin-2, J. Immunol. Baltim. Md 1950 199 (2017) 2976–2984, http://dx.doi. Soc. 23 (2017) 63–74, http://dx.doi.org/10.1002/lt.24628.
org/10.4049/jimmunol.1700105. [129] C. Guo, H. Qi, Y. Yu, Q. Zhang, J. Su, D. Yu, W. Huang, W.-D. Chen, Y.-D. Wang,
[110] U.-H. Jin, Y. Cheng, H. Park, L.A. Davidson, E.S. Callaway, R.S. Chapkin, The g-protein-coupled bile acid receptor Gpbar1 (TGR5) inhibits gastric in-
A. Jayaraman, A. Asante, C. Allred, E.A. Weaver, S. Safe, Short chain fatty acids flammation through antagonizing NF-κb signaling pathway, Front. Pharmacol. 6
enhance aryl hydrocarbon (ah) responsiveness in mouse colonocytes and Caco-2 (2015) 287, http://dx.doi.org/10.3389/fphar.2015.00287.
human colon cancer cells, Sci. Rep. 7 (2017) 10163, http://dx.doi.org/10.1038/ [130] C. Guo, J. Su, Z. Li, R. Xiao, J. Wen, Y. Li, M. Zhang, X. Zhang, D. Yu, W. Huang,
s41598-017-10824-x. W.-D. Chen, Y.-D. Wang, The G-protein-coupled bile acid receptor Gpbar1 (TGR5)
[111] X. Han, H. Song, Y. Wang, Y. Sheng, J. Chen, Sodium butyrate protects the in- suppresses gastric cancer cell proliferation and migration through antagonizing
testinal barrier function in peritonitic mice, Int. J. Clin. Exp. Med. 8 (2015) STAT3 signaling pathway, Oncotarget 6 (2015) 34402–34413, http://dx.doi.org/
4000–4007. 10.18632/oncotarget.5353.
[112] M. Kespohl, N. Vachharajani, M. Luu, H. Harb, S. Pautz, S. Wolff, N. Sillner, [131] S. Cipriani, A. Mencarelli, M.G. Chini, E. Distrutti, B. Renga, G. Bifulco, F. Baldelli,
A. Walker, P. Schmitt-Kopplin, T. Boettger, H. Renz, S. Offermanns, U. Steinhoff, A. Donini, S. Fiorucci, The bile acid receptor GPBAR-1 (TGR5) modulates integrity
A. Visekruna, The microbial metabolite butyrate induces expression of Th1- of intestinal barrier and immune response to experimental colitis, PLoS One 6
Associated factors in CD4+ T cells, Front. Immunol. 8 (2017) 1036, http://dx.doi. (2011) e25637, , http://dx.doi.org/10.1371/journal.pone.0025637.
org/10.3389/fimmu.2017.01036. [132] H. Sato, C. Genet, A. Strehle, C. Thomas, A. Lobstein, A. Wagner, C. Mioskowski,
[113] T.M. Ferreira, A.J. Leonel, M.A. Melo, R.R.G. Santos, D.C. Cara, V.N. Cardoso, J. Auwerx, R. Saladin, Anti-hyperglycemic activity of a TGR5 agonist isolated from
M.I.T.D. Correia, J.I. Alvarez-Leite, Oral supplementation of butyrate reduces Olea europaea, Biochem. Biophys. Res. Commun. 362 (2007) 793–798, http://dx.
mucositis and intestinal permeability associated with 5-Fluorouracil administra- doi.org/10.1016/j.bbrc.2007.06.130.
tion, Lipids 47 (2012) 669–678, http://dx.doi.org/10.1007/s11745-012-3680-3. [133] R. Ichikawa, T. Takayama, K. Yoneno, N. Kamada, M.T. Kitazume, H. Higuchi,
[114] M. Miyoshi, H. Sakaki, M. Usami, N. Iizuka, K. Shuno, M. Aoyama, Y. Usami, Oral K. Matsuoka, M. Watanabe, H. Itoh, T. Kanai, T. Hisamatsu, T. Hibi, Bile acids
administration of tributyrin increases concentration of butyrate in the portal vein induce monocyte differentiation toward interleukin-12 hypo-producing dendritic
and prevents lipopolysaccharide-induced liver injury in rats, Clin. Nutr. Edinb. cells via a TGR5-dependent pathway, Immunology 136 (2012) 153–162, http://
Scotl 30 (2011) 252–258, http://dx.doi.org/10.1016/j.clnu.2010.09.012. dx.doi.org/10.1111/j.1365-2567.2012.03554.x.
[115] F. Del Chierico, P. Vernocchi, B. Dallapiccola, L. Putignani, Mediterranean diet [134] M. Prosberg, F. Bendtsen, I. Vind, A.M. Petersen, L.L. Gluud, The association be-
and health: food effects on gut microbiota and disease control, Int. J. Mol. Sci. 15 tween the gut microbiota and the inflammatory bowel disease activity: a sys-
(2014) 11678–11699, http://dx.doi.org/10.3390/ijms150711678. tematic review and meta-analysis, Scand. J. Gastroenterol. 51 (2016) 1407–1415,
[116] T.W.H. Pols, L.G. Noriega, M. Nomura, J. Auwerx, K. Schoonjans, The bile acid http://dx.doi.org/10.1080/00365521.2016.1216587.
membrane receptor TGR5 as an emerging target in metabolism and inflammation, [135] E. Gianchecchi, A. Fierabracci, On the pathogenesis of insulin-dependent diabetes
J. Hepatol. 54 (2011) 1263–1272, http://dx.doi.org/10.1016/j.jhep.2010.12.004. mellitus: the role of microbiota, Immunol. Res. 65 (2017) 242–256, http://dx.doi.
[117] S.L. Long, C.G.M. Gahan, S.A. Joyce, Interactions between gut bacteria and bile in org/10.1007/s12026-016-8832-8.
health and disease, Mol. Aspect. Med. 56 (2017) 54–65, http://dx.doi.org/10. [136] S. Abdollahi-Roodsaz, S.B. Abramson, J.U. Scher, The metabolic role of the gut
1016/j.mam.2017.06.002. microbiota in health and rheumatic disease: mechanisms and interventions, Nat.
[118] F. Gianfagna, G. Veronesi, M. Tozzi, A. Tarallo, R. Borchini, M.M. Ferrario, Rev. Rheumatol. 12 (2016) 446–455, http://dx.doi.org/10.1038/nrrheum.
L. Bertù, A. Montonati, P. Castelli, RoCAV (risk of cardiovascular diseases and 2016.68.
abdominal aortic aneurysm in varese) project investigators, prevalence of ab- [137] A.B. Blázquez, M.C. Berin, Microbiome and food allergy, Transl. Res. J. Lab. Clin.
dominal aortic aneurysms in the general population and in subgroups at high Med 179 (2017) 199–203, http://dx.doi.org/10.1016/j.trsl.2016.09.003.
cardiovascular risk in Italy. Results of the RoCAV population based study, Eur. J. [138] E.C. Rosser, C. Mauri, A clinical update on the significance of the gut microbiota in
Vasc. Endovasc. Surg. Off. J. Eur. Soc. Vasc. Surg 55 (2018) 633–639, http://dx. systemic autoimmunity, J. Autoimmun. 74 (2016) 85–93, http://dx.doi.org/10.
doi.org/10.1016/j.ejvs.2018.01.008. 1016/j.jaut.2016.06.009.
[119] N.D. Lewis, L.A. Patnaude, J. Pelletier, D.J. Souza, S.M. Lukas, F.J. King, J.D. Hill, [139] J. Ochoa-Repáraz, K. Magori, L.H. Kasper, The chicken or the egg dilemma: in-
D.E. Stefanopoulos, K. Ryan, S. Desai, D. Skow, S.G. Kauschke, A. Broermann, testinal dysbiosis in multiple sclerosis, Ann. Transl. Med. 5 (2017), http://dx.doi.
D. Kuzmich, C. Harcken, E.R. Hickey, L.K. Modis, A GPBAR1 (TGR5) small mo- org/10.21037/atm.2017.01.18.
lecule agonist shows specific inhibitory effects on myeloid cell activation in vitro [140] Z. Wang, Z. Xie, Q. Lu, C. Chang, Z. Zhou, Beyond genetics: what causes type 1
and reduces experimental autoimmune encephalitis (EAE) in vivo, PLoS One 9 diabetes, Clin. Rev. Allergy Immunol. 52 (2017) 273–286, http://dx.doi.org/10.
(2014) e100883, , http://dx.doi.org/10.1371/journal.pone.0100883. 1007/s12016-016-8592-1.
[120] R. Martín, J. Carvalho-Tavares, M. Hernández, M. Arnés, V. Ruiz-Gutiérrez, [141] M.A. Odenwald, J.R. Turner, Intestinal permeability defects: is it time to treat?
M.L. Nieto, Beneficial actions of oleanolic acid in an experimental model of Clin. Gastroenterol. Hepatol. Off. Clin. Pract. J. Am. Gastroenterol. Assoc. 11
multiple sclerosis: a potential therapeutic role, Biochem. Pharmacol. 79 (2010) (2013) 1075–1083, http://dx.doi.org/10.1016/j.cgh.2013.07.001.
198–208, http://dx.doi.org/10.1016/j.bcp.2009.08.002. [142] J.H. Buckner, C.J. Greenbaum, Stacking the deck: studies of patients with multiple
[121] Y.-D. Wang, W.-D. Chen, D. Yu, B.M. Forman, W. Huang, The G-Protein-coupled autoimmune diseases propelled our understanding of type 1 diabetes as an auto-
bile acid receptor, Gpbar1 (TGR5), negatively regulates hepatic inflammatory immune disease, J. Immunol. Baltim. Md 1950 199 (2017) 3011–3013, http://dx.
response through antagonizing nuclear factor kappa light-chain enhancer of ac- doi.org/10.4049/jimmunol.1701299.
tivated B cells (NF-κB) in mice, Hepatology 54 (2011) 1421–1432, http://dx.doi. [143] C. Kuhn, A. Besançon, S. Lemoine, S. You, C. Marquet, S. Candon, L. Chatenoud,
org/10.1002/hep.24525. Regulatory mechanisms of immune tolerance in type 1 diabetes and their failures,
[122] T.W.H. Pols, M. Nomura, T. Harach, G. Lo Sasso, M.H. Oosterveer, C. Thomas, J. Autoimmun. 71 (2016) 69–77, http://dx.doi.org/10.1016/j.jaut.2016.05.002.
G. Rizzo, A. Gioiello, L. Adorini, R. Pellicciari, J. Auwerx, K. Schoonjans, TGR5 [144] J. Neu, C.M. Reverte, A.D. Mackey, K. Liboni, L.M. Tuhacek-Tenace, M. Hatch,
activation inhibits atherosclerosis by reducing macrophage inflammation and lipid N. Li, R.A. Caicedo, D.A. Schatz, M. Atkinson, Changes in intestinal morphology
loading, Cell Metabol. 14 (2011) 747–757, http://dx.doi.org/10.1016/j.cmet. and permeability in the biobreeding rat before the onset of type 1 diabetes, J.

17
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

Pediatr. Gastroenterol. Nutr. 40 (2005) 589–595. 1703–1716, http://dx.doi.org/10.1517/14712598.2010.534080.


[145] E. Bosi, L. Molteni, M.G. Radaelli, L. Folini, I. Fermo, E. Bazzigaluppi, L. Piemonti, [165] M. Li, L. Song, X. Gao, W. Chang, X. Qin, Toll-like receptor 4 on islet β cells senses
M.R. Pastore, R. Paroni, Increased intestinal permeability precedes clinical onset expression changes in high-mobility group box 1 and contributes to the initiation
of type 1 diabetes, Diabetologia 49 (2006) 2824–2827, http://dx.doi.org/10. of type 1 diabetes, Exp. Mol. Med. 44 (2012) 260–267, http://dx.doi.org/10.
1007/s00125-006-0465-3. 3858/emm.2012.44.4.021.
[146] M. Kuitunen, T. Saukkonen, J. Ilonen, H.K. Akerblom, E. Savilahti, Intestinal [166] B. Balasa, K. Van Gunst, N. Sarvetnick, The microbial product lipopolysaccharide
permeability to mannitol and lactulose in children with type 1 diabetes with the confers diabetogenic potential on the T cell repertoire of BDC2.5/NOD mice: im-
HLA-DQB1*02 allele, Autoimmunity 35 (2002) 365–368. plications for the etiology of autoimmune diabetes, Clin. Immunol. Orlando Fla 95
[147] A. Sapone, L. de Magistris, M. Pietzak, M.G. Clemente, A. Tripathi, F. Cucca, (2000) 93–98, http://dx.doi.org/10.1006/clim.2000.4855.
R. Lampis, D. Kryszak, M. Cartenì, M. Generoso, D. Iafusco, F. Prisco, F. Laghi, [167] I. Caramalho, L. Rodrigues-Duarte, A. Perez, S. Zelenay, C. Penha-Gonçalves,
G. Riegler, R. Carratu, D. Counts, A. Fasano, Zonulin upregulation is associated J. Demengeot, Regulatory T cells contribute to diabetes protection in lipopoly-
with increased gut permeability in subjects with type 1 diabetes and their re- saccharide-treated non-obese diabetic mice, Scand. J. Immunol. 74 (2011)
latives, Diabetes 55 (2006) 1443–1449. 585–595, http://dx.doi.org/10.1111/j.1365-3083.2011.02627.x.
[148] M. Secondulfo, D. Iafusco, R. Carratù, L. deMagistris, A. Sapone, M. Generoso, [168] A. Aumeunier, F. Grela, A. Ramadan, L. Pham Van, E. Bardel, A. Gomez Alcala,
A. Mezzogiomo, F.C. Sasso, M. Cartenì, R. De Rosa, F. Prisco, V. Esposito, P. Jeannin, S. Akira, J.-F. Bach, N. Thieblemont, Systemic Toll-like receptor sti-
Ultrastructural mucosal alterations and increased intestinal permeability in non- mulation suppresses experimental allergic asthma and autoimmune diabetes in
celiac, type I diabetic patients, Dig. Liver Dis. Off. J. Ital. Soc. Gastroenterol. Ital. NOD mice, PLoS One 5 (2010) e11484, , http://dx.doi.org/10.1371/journal.pone.
Assoc. Study Liver 36 (2004) 35–45. 0011484.
[149] A. Fasano, Zonulin, regulation of tight junctions, and autoimmune diseases, Ann. [169] J. Wang, H. Cao, H. Wang, G. Yin, J. Du, F. Xia, J. Lu, M. Xiang, Multiple me-
N. Y. Acad. Sci. 1258 (2012) 25–33, http://dx.doi.org/10.1111/j.1749-6632. chanisms involved in diabetes protection by lipopolysaccharide in non-obese
2012.06538.x. diabetic mice, Toxicol. Appl. Pharmacol. 285 (2015) 149–158, http://dx.doi.org/
[150] R. El Asmar, P. Panigrahi, P. Bamford, I. Berti, T. Not, G.V. Coppa, C. Catassi, 10.1016/j.taap.2015.04.006.
A. Fasano, R. El Asmar, Host-dependent zonulin secretion causes the impairment [170] L. Wen, R.E. Ley, P.Y. Volchkov, P.B. Stranges, L. Avanesyan, A.C. Stonebraker,
of the small intestine barrier function after bacterial exposure, Gastroenterology C. Hu, F.S. Wong, G.L. Szot, J.A. Bluestone, J.I. Gordon, A.V. Chervonsky, Innate
123 (2002) 1607–1615. immunity and intestinal microbiota in the development of Type 1 diabetes, Nature
[151] A. Fasano, Zonulin and its regulation of intestinal barrier function: the biological 455 (2008) 1109–1113, http://dx.doi.org/10.1038/nature07336.
door to inflammation, autoimmunity, and cancer, Physiol. Rev. 91 (2011) [171] C. King, N. Sarvetnick, The incidence of type-1 diabetes in NOD mice is modulated
151–175, http://dx.doi.org/10.1152/physrev.00003.2008. by restricted flora not germ-free conditions, PLoS One 6 (2011) e17049, , http://
[152] T. Watts, I. Berti, A. Sapone, T. Gerarduzzi, T. Not, R. Zielke, A. Fasano, Role of the dx.doi.org/10.1371/journal.pone.0017049.
intestinal tight junction modulator zonulin in the pathogenesis of type I diabetes in [172] D. Endesfelder, W. zu Castell, A. Ardissone, A.G. Davis-Richardson, P. Achenbach,
BB diabetic-prone rats, Proc. Natl. Acad. Sci. U. S. A 102 (2005) 2916–2921, M. Hagen, M. Pflueger, K.A. Gano, J.R. Fagen, J.C. Drew, C.T. Brown,
http://dx.doi.org/10.1073/pnas.0500178102. B. Kolaczkowski, M. Atkinson, D. Schatz, E. Bonifacio, E.W. Triplett, A.-G. Ziegler,
[153] Y.Y. Lam, C.W.Y. Ha, C.R. Campbell, A.J. Mitchell, A. Dinudom, J. Oscarsson, Compromised gut microbiota networks in children with anti-islet cell auto-
D.I. Cook, N.H. Hunt, I.D. Caterson, A.J. Holmes, L.H. Storlien, Increased gut immunity, Diabetes 63 (2014) 2006–2014, http://dx.doi.org/10.2337/db13-
permeability and microbiota change associate with mesenteric fat inflammation 1676.
and metabolic dysfunction in diet-induced obese mice, PLoS One 7 (2012) e34233, [173] D. Endesfelder, M. Engel, A.G. Davis-Richardson, A.N. Ardissone, P. Achenbach,
, http://dx.doi.org/10.1371/journal.pone.0034233. S. Hummel, C. Winkler, M. Atkinson, D. Schatz, E. Triplett, A.-G. Ziegler, W. zu
[154] L.F. Roesch, G.L. Lorca, G. Casella, A. Giongo, A. Naranjo, A.M. Pionzio, N. Li, Castell, Towards a functional hypothesis relating anti-islet cell autoimmunity to
V. Mai, C.H. Wasserfall, D. Schatz, M.A. Atkinson, J. Neu, E.W. Triplett, Culture- the dietary impact on microbial communities and butyrate production,
independent identification of gut bacteria correlated with the onset of diabetes in a Microbiome 4 (2016) 17, http://dx.doi.org/10.1186/s40168-016-0163-4.
rat model, ISME J. 3 (2009) 536–548, http://dx.doi.org/10.1038/ismej.2009.5. [174] A.G. Davis-Richardson, A.N. Ardissone, R. Dias, V. Simell, M.T. Leonard,
[155] K. Brown, A. Godovannyi, C. Ma, Y. Zhang, Z. Ahmadi-Vand, C. Dai, K.M. Kemppainen, J.C. Drew, D. Schatz, M.A. Atkinson, B. Kolaczkowski,
M.A. Gorzelak, Y. Chan, J.M. Chan, A. Lochner, J.P. Dutz, B.A. Vallance, J. Ilonen, M. Knip, J. Toppari, N. Nurminen, H. Hyöty, R. Veijola, T. Simell,
D.L. Gibson, Prolonged antibiotic treatment induces a diabetogenic intestinal J. Mykkänen, O. Simell, E.W. Triplett, Bacteroides dorei dominates gut micro-
microbiome that accelerates diabetes in NOD mice, ISME J. 10 (2016) 321–332, biome prior to autoimmunity in Finnish children at high risk for type 1 diabetes,
http://dx.doi.org/10.1038/ismej.2015.114. Front. Microbiol. 5 (2014) 678, http://dx.doi.org/10.3389/fmicb.2014.00678.
[156] J. Peng, S. Narasimhan, J.R. Marchesi, A. Benson, F.S. Wong, L. Wen, Long term [175] S. Hummel, M. Pflüger, M. Hummel, E. Bonifacio, A.-G. Ziegler, Primary dietary
effect of gut microbiota transfer on diabetes development, J. Autoimmun. 53 intervention study to reduce the risk of islet autoimmunity in children at increased
(2014) 85–94, http://dx.doi.org/10.1016/j.jaut.2014.03.005. risk for type 1 diabetes: the BABYDIET study, Diabetes Care 34 (2011) 1301–1305,
[157] M. Murri, I. Leiva, J.M. Gomez-Zumaquero, F.J. Tinahones, F. Cardona, http://dx.doi.org/10.2337/dc10-2456.
F. Soriguer, M.I. Queipo-Ortuño, Gut microbiota in children with type 1 diabetes [176] M. Teruel, M.E. Alarcón-Riquelme, The genetic basis of systemic lupus er-
differs from that in healthy children: a case-control study, BMC Med. 11 (2013) 46, ythematosus: what are the risk factors and what have we learned, J. Autoimmun.
http://dx.doi.org/10.1186/1741-7015-11-46. 74 (2016) 161–175, http://dx.doi.org/10.1016/j.jaut.2016.08.001.
[158] M.E. Mejía-León, J.F. Petrosino, N.J. Ajami, M.G. Domínguez-Bello, A.M.C. de la [177] M. Giannelou, C.P. Mavragani, Cardiovascular disease in systemic lupus er-
Barca, Fecal microbiota imbalance in Mexican children with type 1 diabetes, Sci. ythematosus: a comprehensive update, J. Autoimmun. 82 (2017) 1–12, http://dx.
Rep. 4 (2014) 3814, http://dx.doi.org/10.1038/srep03814. doi.org/10.1016/j.jaut.2017.05.008.
[159] M.C. de Goffau, S. Fuentes, B. van den Bogert, H. Honkanen, W.M. de Vos, [178] P.K. Bundhun, M.Z.S. Soogund, F. Huang, Impact of systemic lupus erythematosus
G.W. Welling, H. Hyöty, H.J.M. Harmsen, Aberrant gut microbiota composition at on maternal and fetal outcomes following pregnancy: a meta-analysis of studies
the onset of type 1 diabetes in young children, Diabetologia 57 (2014) 1569–1577, published between years 2001-2016, J. Autoimmun. 79 (2017) 17–27, http://dx.
http://dx.doi.org/10.1007/s00125-014-3274-0. doi.org/10.1016/j.jaut.2017.02.009.
[160] A.D. Kostic, D. Gevers, H. Siljander, T. Vatanen, T. Hyötyläinen, A.- [179] Q. Mu, H. Zhang, X.M. Luo, SLE: another autoimmune disorder influenced by
M. Hämäläinen, A. Peet, V. Tillmann, P. Pöhö, I. Mattila, H. Lähdesmäki, microbes and diet? Front. Immunol. 6 (2015) 608, http://dx.doi.org/10.3389/
E.A. Franzosa, O. Vaarala, M. de Goffau, H. Harmsen, J. Ilonen, S.M. Virtanen, fimmu.2015.00608.
C.B. Clish, M. Orešič, C. Huttenhower, M. Knip, DIABIMMUNE Study Group, [180] W.A. Nockher, R. Wigand, W. Schoeppe, J.E. Scherberich, Elevated levels of so-
R.J. Xavier, The dynamics of the human infant gut microbiome in development luble CD14 in serum of patients with systemic lupus erythematosus, Clin. Exp.
and in progression toward type 1 diabetes, Cell Host Microbe 17 (2015) 260–273, Immunol. 96 (1994) 15–19.
http://dx.doi.org/10.1016/j.chom.2015.01.001. [181] B. Liu, Y. Yang, J. Dai, R. Medzhitov, M.A. Freudenberg, P.L. Zhang, Z. Li, TLR4
[161] T. Vatanen, A.D. Kostic, E. d'Hennezel, H. Siljander, E.A. Franzosa, M. Yassour, up-regulation at protein or gene level is pathogenic for lupus-like autoimmune
R. Kolde, H. Vlamakis, T.D. Arthur, A.-M. Hämäläinen, A. Peet, V. Tillmann, disease, J. Immunol. Baltim. Md 1950 177 (2006) 6880–6888.
R. Uibo, S. Mokurov, N. Dorshakova, J. Ilonen, S.M. Virtanen, S.J. Szabo, [182] J.-X. Zhai, Z.-X. Zhang, Y.-J. Feng, S.-S. Ding, X.-H. Wang, L.-W. Zou, D.-Q. Ye,
J.A. Porter, H. Lähdesmäki, C. Huttenhower, D. Gevers, T.W. Cullen, M. Knip, PDTC attenuate LPS-induced kidney injury in systemic lupus erythematosus-prone
R.J. Xavier, Variation in microbiome LPS immunogenicity contributes to auto- MRL/lpr mice, Mol. Biol. Rep. 39 (2012) 6763–6771, http://dx.doi.org/10.1007/
immunity in humans, Cell 165 (2016) 842–853, http://dx.doi.org/10.1016/j.cell. s11033-012-1501-7.
2016.04.007. [183] T.-P. Lee, J.C. Huang, C.-J. Liu, H.-J. Chen, Y.-H. Chen, Y.-T. Tsai, W. Yang, K.-
[162] C.T. Brown, A.G. Davis-Richardson, A. Giongo, K.A. Gano, D.B. Crabb, H. Sun, Interactions of surface-expressed TLR-4 and endosomal TLR-9 accelerate
N. Mukherjee, G. Casella, J.C. Drew, J. Ilonen, M. Knip, H. Hyöty, R. Veijola, lupus progression in anti-dsDNA antibody transgenic mice, Exp. Biol. Med.
T. Simell, O. Simell, J. Neu, C.H. Wasserfall, D. Schatz, M.A. Atkinson, Maywood NJ 239 (2014) 715–723, http://dx.doi.org/10.1177/
E.W. Triplett, Gut microbiome metagenomics analysis suggests a functional model 1535370214525299.
for the development of autoimmunity for type 1 diabetes, PLoS One 6 (2011) [184] T.-P. Lee, S.-J. Tang, M.-F. Wu, Y.-C. Song, C.-L. Yu, K.-H. Sun, Transgenic over-
e25792, , http://dx.doi.org/10.1371/journal.pone.0025792. expression of anti-double-stranded DNA autoantibody and activation of Toll-like
[163] M.C. de Goffau, K. Luopajärvi, M. Knip, J. Ilonen, T. Ruohtula, T. Härkönen, receptor 4 in mice induce severe systemic lupus erythematosus syndromes, J.
L. Orivuori, S. Hakala, G.W. Welling, H.J. Harmsen, O. Vaarala, Fecal microbiota Autoimmun. 35 (2010) 358–367, http://dx.doi.org/10.1016/j.jaut.2010.07.007.
composition differs between children with β-cell autoimmunity and those without, [185] J. Ni, Q. Ouyang, L. Lin, Z. Huang, H. Lu, X. Chen, H. Lin, Z. Wang, D. Xu,
Diabetes 62 (2013) 1238–1244, http://dx.doi.org/10.2337/db12-0526. Y. Zhang, Role of toll-like receptor 4 on lupus lung injury and atherosclerosis in
[164] F.I.L. Clanchy, S.M. Sacre, Modulation of toll-like receptor function has ther- LPS-challenge ApoE−/− mice, Clin. Dev. Immunol. 2013 (2013) 476856, , http://
apeutic potential in autoimmune disease, Expet Opin. Biol. Ther. 10 (2010) dx.doi.org/10.1155/2013/476856.

18
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

[186] J.S. Levine, R. Subang, S. Setty, J. Cabrera, P. Laplante, M.J. Fritzler, J. Rauch, 1950 183 (2009) 6041–6050, http://dx.doi.org/10.4049/jimmunol.0900747.
Phospholipid-binding proteins differ in their capacity to induce autoantibodies [209] K. Berer, M. Mues, M. Koutrolos, Z.A. Rasbi, M. Boziki, C. Johner, H. Wekerle,
and murine systemic lupus erythematosus, Lupus 23 (2014) 752–768, http://dx. G. Krishnamoorthy, Commensal microbiota and myelin autoantigen cooperate to
doi.org/10.1177/0961203314525676. trigger autoimmune demyelination, Nature 479 (2011) 538–541, http://dx.doi.
[187] A. Lartigue, N. Colliou, S. Calbo, A. François, S. Jacquot, C. Arnoult, F. Tron, org/10.1038/nature10554.
D. Gilbert, P. Musette, Critical role of TLR2 and TLR4 in autoantibody production [210] J. Visser, J. Rozing, A. Sapone, K. Lammers, A. Fasano, Tight junctions, intestinal
and glomerulonephritis in lpr mutation-induced mouse lupus, J. Immunol. Baltim. permeability, and autoimmunity: celiac disease and type 1 diabetes paradigms,
Md 1950 183 (2009) 6207–6216, http://dx.doi.org/10.4049/jimmunol.0803219. Ann. N. Y. Acad. Sci. 1165 (2009) 195–205, http://dx.doi.org/10.1111/j.1749-
[188] S.A. Summers, A. Hoi, O.M. Steinmetz, K.M. O'Sullivan, J.D. Ooi, D. Odobasic, 6632.2009.04037.x.
S. Akira, A.R. Kitching, S.R. Holdsworth, TLR9 and TLR4 are required for the [211] H. Yokote, S. Miyake, J.L. Croxford, S. Oki, H. Mizusawa, T. Yamamura, NKT cell-
development of autoimmunity and lupus nephritis in pristane nephropathy, J. dependent amelioration of a mouse model of multiple sclerosis by altering gut
Autoimmun. 35 (2010) 291–298, http://dx.doi.org/10.1016/j.jaut.2010.05.004. flora, Am. J. Pathol. 173 (2008) 1714–1723, http://dx.doi.org/10.2353/ajpath.
[189] Y. Liu, J. Liao, M. Zhao, H. Wu, S. Yung, T.M. Chan, A. Yoshimura, Q. Lu, 2008.080622.
Increased expression of TLR2 in CD4(+) T cells from SLE patients enhances im- [212] M.-Q. Xu, H.-L. Cao, W.-Q. Wang, S. Wang, X.-C. Cao, F. Yan, B.-M. Wang, Fecal
mune reactivity and promotes IL-17 expression through histone modifications, microbiota transplantation broadening its application beyond intestinal disorders,
Eur. J. Immunol. 45 (2015) 2683–2693, http://dx.doi.org/10.1002/eji. World J. Gastroenterol. 21 (2015) 102–111, http://dx.doi.org/10.3748/wjg.v21.
201445219. i1.102.
[190] R.D. Pawar, L. Castrezana-Lopez, R. Allam, O.P. Kulkarni, S. Segerer, E. Radomska, [213] S. Miyake, S. Kim, W. Suda, K. Oshima, M. Nakamura, T. Matsuoka, N. Chihara,
T.N. Meyer, C.-M. Schwesinger, N. Akis, H.-J. Gröne, H.-J. Anders, Bacterial li- A. Tomita, W. Sato, S.-W. Kim, H. Morita, M. Hattori, T. Yamamura, Dysbiosis in
popeptide triggers massive albuminuria in murine lupus nephritis by activating the gut microbiota of patients with multiple sclerosis, with a striking depletion of
Toll-like receptor 2 at the glomerular filtration barrier, Immunology 128 (2009) species belonging to clostridia XIVa and IV clusters, PLoS One 10 (2015)
e206–221, http://dx.doi.org/10.1111/j.1365-2567.2008.02948.x. e0137429, , http://dx.doi.org/10.1371/journal.pone.0137429.
[191] V. Urbonaviciute, C. Starke, W. Pirschel, S. Pohle, S. Frey, C. Daniel, K. Amann, [214] H. Tremlett, D.W. Fadrosh, A.A. Faruqi, J. Hart, S. Roalstad, J. Graves,
G. Schett, M. Herrmann, R.E. Voll, Toll-like receptor 2 is required for autoantibody C.M. Spencer, S.V. Lynch, S.S. Zamvil, E. Waubant, US Network of Pediatric MS
production and development of renal disease in pristane-induced lupus, Arthritis Centers, Associations between the gut microbiota and host immune markers in
Rheum. 65 (2013) 1612–1623, http://dx.doi.org/10.1002/art.37914. pediatric multiple sclerosis and controls, BMC Neurol. 16 (2016) 182, http://dx.
[192] P.M. Gallo, G.J. Rapsinski, R.P. Wilson, G.O. Oppong, U. Sriram, M. Goulian, doi.org/10.1186/s12883-016-0703-3.
B. Buttaro, R. Caricchio, S. Gallucci, Ç. Tükel, Amyloid-dna composites of bacterial [215] H. Tremlett, D.W. Fadrosh, A.A. Faruqi, F. Zhu, J. Hart, S. Roalstad, J. Graves,
biofilms stimulate autoimmunity, Immunity 42 (2015) 1171–1184, http://dx.doi. S. Lynch, E. Waubant, US Network of Pediatric MS Centers, Gut microbiota in
org/10.1016/j.immuni.2015.06.002. early pediatric multiple sclerosis: a case-control study, Eur. J. Neurol. 23 (2016)
[193] Y. Wu, W. Tang, J. Zuo, Toll-like receptors: potential targets for lupus treatment, 1308–1321, http://dx.doi.org/10.1111/ene.13026.
Acta Pharmacol. Sin. 36 (2015) 1395–1407, http://dx.doi.org/10.1038/aps. [216] J. Chen, N. Chia, K.R. Kalari, J.Z. Yao, M. Novotna, M.M.P. Soldan, D.H. Luckey,
2015.91. E.V. Marietta, P.R. Jeraldo, X. Chen, B.G. Weinshenker, M. Rodriguez,
[194] L.L. Teichmann, D. Schenten, R. Medzhitov, M. Kashgarian, M.J. Shlomchik, O.H. Kantarci, H. Nelson, J.A. Murray, A.K. Mangalam, Multiple sclerosis patients
Signals via the adaptor MyD88 in B cells and DCs make distinct and synergistic have a distinct gut microbiota compared to healthy controls, Sci. Rep. 6 (2016)
contributions to immune activation and tissue damage in lupus, Immunity 38 28484, http://dx.doi.org/10.1038/srep28484.
(2013) 528–540, http://dx.doi.org/10.1016/j.immuni.2012.11.017. [217] S. Jangi, R. Gandhi, L.M. Cox, N. Li, F. von Glehn, R. Yan, B. Patel, M.A. Mazzola,
[195] R. Karki, S.M. Man, T.-D. Kanneganti, Inflammasomes and cancer, Cancer S. Liu, B.L. Glanz, S. Cook, S. Tankou, F. Stuart, K. Melo, P. Nejad, K. Smith,
Immunol. Res. 5 (2017) 94–99, http://dx.doi.org/10.1158/2326-6066.CIR-16- B.D. Topçuolu, J. Holden, P. Kivisäkk, T. Chitnis, P.L. De Jager, F.J. Quintana,
0269. G.K. Gerber, L. Bry, H.L. Weiner, Alterations of the human gut microbiome in
[196] R. Karki, S.M. Man, R.K.S. Malireddi, S. Kesavardhana, Q. Zhu, A.R. Burton, multiple sclerosis, Nat. Commun. 7 (2016) 12015, http://dx.doi.org/10.1038/
B.R. Sharma, X. Qi, S. Pelletier, P. Vogel, P. Rosenstiel, T.-D. Kanneganti, NLRC3 is ncomms12015.
an inhibitory sensor of PI3K-mTOR pathways in cancer, Nature (2016), http://dx. [218] E. Cekanaviciute, B.B. Yoo, T.F. Runia, J.W. Debelius, S. Singh, C.A. Nelson,
doi.org/10.1038/nature20597. R. Kanner, Y. Bencosme, Y.K. Lee, S.L. Hauser, E. Crabtree-Hartman, I.K. Sand,
[197] M. Lech, G. Lorenz, O.P. Kulkarni, M.O.O. Grosser, N. Stigrot, M.N. Darisipudi, M. Gacias, Y. Zhu, P. Casaccia, B.A.C. Cree, R. Knight, S.K. Mazmanian,
R. Günthner, M.W.M. Wintergerst, D. Anz, H.E. Susanti, H.-J. Anders, NLRP3 and S.E. Baranzini, Gut bacteria from multiple sclerosis patients modulate human T
ASC suppress lupus-like autoimmunity by driving the immunosuppressive effects cells and exacerbate symptoms in mouse models, Proc. Natl. Acad. Sci. 114 (2017)
of TGF-β receptor signalling, Ann. Rheum. Dis. 74 (2015) 2224–2235, http://dx. 10713–10718, http://dx.doi.org/10.1073/pnas.1711235114.
doi.org/10.1136/annrheumdis-2014-205496. [219] K. Berer, L.A. Gerdes, E. Cekanaviciute, X. Jia, L. Xiao, Z. Xia, C. Liu, L. Klotz,
[198] A. Hevia, C. Milani, P. López, A. Cuervo, S. Arboleya, S. Duranti, F. Turroni, U. Stauffer, S.E. Baranzini, T. Kümpfel, R. Hohlfeld, G. Krishnamoorthy,
S. González, A. Suárez, M. Gueimonde, M. Ventura, B. Sánchez, A. Margolles, H. Wekerle, Gut microbiota from multiple sclerosis patients enables spontaneous
Intestinal dysbiosis associated with systemic lupus erythematosus, mBio 5 (2014), autoimmune encephalomyelitis in mice, Proc. Natl. Acad. Sci. 114 (2017)
http://dx.doi.org/10.1128/mBio.01548-14 e01548–01514. 10719–10724, http://dx.doi.org/10.1073/pnas.1711233114.
[199] D. Rojo, A. Hevia, R. Bargiela, P. López, A. Cuervo, S. González, A. Suárez, [220] A. Haghikia, S. Jörg, A. Duscha, J. Berg, A. Manzel, A. Waschbisch, A. Hammer,
B. Sánchez, M. Martínez-Martínez, C. Milani, M. Ventura, C. Barbas, A. Moya, D.-H. Lee, C. May, N. Wilck, A. Balogh, A.I. Ostermann, N.H. Schebb, D.A. Akkad,
A. Suárez, A. Margolles, M. Ferrer, Ranking the impact of human health disorders D.A. Grohme, M. Kleinewietfeld, S. Kempa, J. Thöne, S. Demir, D.N. Müller,
on gut metabolism: systemic lupus erythematosus and obesity as study cases, Sci. R. Gold, R.A. Linker, Dietary fatty acids directly impact central nervous system
Rep. 5 (2015) 8310, http://dx.doi.org/10.1038/srep08310. autoimmunity via the small intestine, Immunity 43 (2015) 817–829, http://dx.
[200] M.A. Maldonado, V. Kakkanaiah, G.C. MacDonald, F. Chen, E.A. Reap, E. Balish, doi.org/10.1016/j.immuni.2015.09.007.
W.R. Farkas, J.C. Jennette, M.P. Madaio, B.L. Kotzin, P.L. Cohen, R.A. Eisenberg, [221] V. Braniste, M. Al-Asmakh, C. Kowal, F. Anuar, A. Abbaspour, M. Tóth, A. Korecka,
The role of environmental antigens in the spontaneous development of auto- N. Bakocevic, L.G. Ng, N.L. Guan, P. Kundu, B. Gulyás, C. Halldin, K. Hultenby,
immunity in MRL-lpr mice, J. Immunol. 162 (1999) 6322–6330. H. Nilsson, H. Hebert, B.T. Volpe, B. Diamond, S. Pettersson, The gut microbiota
[201] J. East, M. Branca, Autoimmune reactions and malignant changes in germ-free influences blood-brain barrier permeability in mice, Sci. Transl. Med. 6 (2014)
New Zealand Black mice, Clin. Exp. Immunol. 4 (1969) 621–635. 263ra158, http://dx.doi.org/10.1126/scitranslmed.3009759.
[202] H. Zhang, X. Liao, J.B. Sparks, X.M. Luo, Dynamics of gut microbiota in auto- [222] H. Wang, P. Shi, L. Zuo, J. Dong, J. Zhao, Q. Liu, W. Zhu, Dietary non-digestible
immune lupus, Appl. Environ. Microbiol. 80 (2014) 7551–7560, http://dx.doi. polysaccharides ameliorate intestinal epithelial barrier dysfunction in IL-10
org/10.1128/AEM.02676-14. knockout mice, J. Crohns Colitis. 10 (2016) 1076–1086, http://dx.doi.org/10.
[203] B.M. Johnson, M.-C. Gaudreau, M.M. Al-Gadban, R. Gudi, C. Vasu, Impact of 1093/ecco-jcc/jjw065.
dietary deviation on disease progression and gut microbiome composition in [223] C. Richard, P. Couture, S. Desroches, B. Lamarche, Effect of the Mediterranean diet
lupus-prone SNF1 mice, Clin. Exp. Immunol. 181 (2015) 323–337, http://dx.doi. with and without weight loss on markers of inflammation in men with metabolic
org/10.1111/cei.12609. syndrome, Obes. Silver Spring Md 21 (2013) 51–57, http://dx.doi.org/10.1002/
[204] M.-C. Gaudreau, B.M. Johnson, R. Gudi, M.M. Al-Gadban, C. Vasu, Gender bias in oby.20239.
lupus: does immune response initiated in the gut mucosa have a role? Clin. Exp. [224] R. Casas, E. Sacanella, M. Urpí-Sardà, D. Corella, O. Castañer, R.-M. Lamuela-
Immunol. 180 (2015) 393–407, http://dx.doi.org/10.1111/cei.12587. Raventos, J. Salas-Salvadó, M.-A. Martínez-González, E. Ros, R. Estruch, Long-
[205] J. Goverman, Autoimmune T cell responses in the central nervous system, Nat. Term Immunomodulatory Effects of a Mediterranean Diet in Adults at High Risk of
Rev. Immunol. 9 (2009) 393–407, http://dx.doi.org/10.1038/nri2550. Cardiovascular Disease in the PREvención con DIeta MEDiterránea (PREDIMED)
[206] C. Selmi, J.G. Barin, N.R. Rose, Current trends in autoimmunity and the nervous Randomized Controlled Trial, J. Nutr. 146 (2016) 1684–1693, http://dx.doi.org/
system, J. Autoimmun. 75 (2016) 20–29, http://dx.doi.org/10.1016/j.jaut.2016. 10.3945/jn.115.229476.
08.005. [225] E.P. Neale, M.J. Batterham, L.C. Tapsell, Consumption of a healthy dietary pattern
[207] M. Nouri, A. Bredberg, B. Weström, S. Lavasani, Intestinal barrier dysfunction results in significant reductions in C-reactive protein levels in adults: a meta-
develops at the onset of experimental autoimmune encephalomyelitis, and can be analysis, Nutr. Res. N. Y. N 36 (2016) 391–401, http://dx.doi.org/10.1016/j.
induced by adoptive transfer of auto-reactive T cells, PLoS One 9 (2014) e106335, nutres.2016.02.009.
, http://dx.doi.org/10.1371/journal.pone.0106335. [226] M. Kolehmainen, S.M. Ulven, J. Paananen, V. de Mello, U. Schwab, C. Carlberg,
[208] J. Ochoa-Repáraz, D.W. Mielcarz, L.E. Ditrio, A.R. Burroughs, D.M. Foureau, M. Myhrstad, J. Pihlajamäki, E. Dungner, E. Sjölin, I. Gunnarsdottir, L. Cloetens,
S. Haque-Begum, L.H. Kasper, Role of gut commensal microflora in the develop- M. Landin-Olsson, B. Akesson, F. Rosqvist, J. Hukkanen, K.-H. Herzig,
ment of experimental autoimmune encephalomyelitis, J. Immunol. Baltim. Md L.O. Dragsted, M.J. Savolainen, L. Brader, K. Hermansen, U. Risérus, I. Thorsdottir,

19
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

K.S. Poutanen, M. Uusitupa, P. Arner, I. Dahlman, Healthy Nordic diet down- (2017) 1535–1548, http://dx.doi.org/10.1016/j.biopha.2017.08.117.
regulates the expression of genes involved in inflammation in subcutaneous adi- [244] H.-K. Kwon, G.-C. Kim, Y. Kim, W. Hwang, A. Jash, A. Sahoo, J.-E. Kim, J.H. Nam,
pose tissue in individuals with features of the metabolic syndrome, Am. J. Clin. S.-H. Im, Amelioration of experimental autoimmune encephalomyelitis by pro-
Nutr. 101 (2015) 228–239, http://dx.doi.org/10.3945/ajcn.114.092783. biotic mixture is mediated by a shift in T helper cell immune response, Clin.
[227] K. Atarashi, T. Tanoue, T. Shima, A. Imaoka, T. Kuwahara, Y. Momose, G. Cheng, Immunol. Orlando Fla 146 (2013) 217–227, http://dx.doi.org/10.1016/j.clim.
S. Yamasaki, T. Saito, Y. Ohba, T. Taniguchi, K. Takeda, S. Hori, I.I. Ivanov, 2013.01.001.
Y. Umesaki, K. Itoh, K. Honda, Induction of colonic regulatory T cells by in- [245] J. Ochoa-Repáraz, D.W. Mielcarz, Y. Wang, S. Begum-Haque, S. Dasgupta,
digenous Clostridium species, Science 331 (2011) 337–341, http://dx.doi.org/10. D.L. Kasper, L.H. Kasper, A polysaccharide from the human commensal
1126/science.1198469. Bacteroides fragilis protects against CNS demyelinating disease, Mucosal
[228] Y. Furusawa, Y. Obata, S. Fukuda, T.A. Endo, G. Nakato, D. Takahashi, Immunol. 3 (2010) 487–495, http://dx.doi.org/10.1038/mi.2010.29.
Y. Nakanishi, C. Uetake, K. Kato, T. Kato, M. Takahashi, N.N. Fukuda, [246] Y.-K. Mao, D.L. Kasper, B. Wang, P. Forsythe, J. Bienenstock, W.A. Kunze,
S. Murakami, E. Miyauchi, S. Hino, K. Atarashi, S. Onawa, Y. Fujimura, T. Lockett, Bacteroides fragilis polysaccharide A is necessary and sufficient for acute activa-
J.M. Clarke, D.L. Topping, M. Tomita, S. Hori, O. Ohara, T. Morita, H. Koseki, tion of intestinal sensory neurons, Nat. Commun. 4 (2013) 1465, http://dx.doi.
J. Kikuchi, K. Honda, K. Hase, H. Ohno, Commensal microbe-derived butyrate org/10.1038/ncomms2478.
induces the differentiation of colonic regulatory T cells, Nature 504 (2013) [247] E. Kouchaki, O.R. Tamtaji, M. Salami, F. Bahmani, R.D. Kakhaki, E. Akbari,
446–450, http://dx.doi.org/10.1038/nature12721. M. Tajabadi-Ebrahimi, P. Jafari, Z. Asemi, Clinical and metabolic response to
[229] M. Lamprecht, S. Bogner, G. Schippinger, K. Steinbauer, F. Fankhauser, probiotic supplementation in patients with multiple sclerosis: a randomized,
S. Hallstroem, B. Schuetz, J.F. Greilberger, Probiotic supplementation affects double-blind, placebo-controlled trial, Clin. Nutr. 36 (2017) 1245–1249, http://
markers of intestinal barrier, oxidation, and inflammation in trained men; a ran- dx.doi.org/10.1016/j.clnu.2016.08.015.
domized, double-blinded, placebo-controlled trial, J. Int. Soc. Sports Nutr 9 (2012) [248] R.A. Caicedo, N. Li, C. Des Robert, P.O. Scumpia, C.P. Hubsher, C.H. Wasserfall,
45, http://dx.doi.org/10.1186/1550-2783-9-45. D.A. Schatz, M.A. Atkinson, J. Neu, Neonatal formula feeding leads to im-
[230] C. Hill, F. Guarner, G. Reid, G.R. Gibson, D.J. Merenstein, B. Pot, L. Morelli, munological alterations in an animal model of type 1 diabetes, Pediatr. Res. 63
R.B. Canani, H.J. Flint, S. Salminen, P.C. Calder, M.E. Sanders, Expert consensus (2008) 303–307, http://dx.doi.org/10.1203/PDR.0b013e31815ed662.
document. The International Scientific Association for Probiotics and Prebiotics [249] J. Visser, S. Brugman, F. Klatter, L. Vis, H. Groen, J. Strubbe, J. Rozing, Short-term
consensus statement on the scope and appropriate use of the term probiotic, Nat. dietary adjustment with a hydrolyzed casein–based diet postpones diabetes de-
Rev. Gastroenterol. Hepatol. 11 (2014) 506–514, http://dx.doi.org/10.1038/ velopment in the diabetes-prone BB rat, Metab. Clin. Exp. 52 (2003) 333–337,
nrgastro.2014.66. http://dx.doi.org/10.1053/meta.2003.50052.
[231] R. Toumi, K. Abdelouhab, H. Rafa, I. Soufli, D. Raissi-Kerboua, Z. Djeraba, [250] S. Brugman, F.A. Klatter, J.T.J. Visser, A.C.M. Wildeboer-Veloo, H.J.M. Harmsen,
C. Touil-Boukoffa, Beneficial role of the probiotic mixture Ultrabiotique on J. Rozing, N.A. Bos, Antibiotic treatment partially protects against type 1 diabetes
maintaining the integrity of intestinal mucosal barrier in DSS-induced experi- in the Bio-Breeding diabetes-prone rat. Is the gut flora involved in the develop-
mental colitis, Immunopharmacol. Immunotoxicol. 35 (2013) 403–409, http://dx. ment of type 1 diabetes? Diabetologia 49 (2006) 2105–2108, http://dx.doi.org/
doi.org/10.3109/08923973.2013.790413. 10.1007/s00125-006-0334-0.
[232] B.P. Abraham, E.M.M. Quigley, Probiotics in inflammatory bowel disease, [251] R.F. Schwartz, J. Neu, D. Schatz, M.A. Atkinson, C. Wasserfall, Antibiotic treat-
Gastroenterol. Clin. N. Am. 46 (2017) 769–782, http://dx.doi.org/10.1016/j.gtc. ment partially protects against type 1 diabetes in the Bio-Breeding diabetes-prone
2017.08.003. rat. Is the gut flora involved in the development of type 1 diabetes? Diabetologia
[233] R. Toumi, I. Soufli, H. Rafa, M. Belkhelfa, A. Biad, C. Touil-Boukoffa, Probiotic 49:2105-2108, Comment on: Brugman S et al, Diabetologia 50 (2007) (2006)
bacteria Lactobacillus and Bifidobacterium attenuate inflammation in dextran 220–221, http://dx.doi.org/10.1007/s00125-006-0526-7.
sulfate sodium-induced experimental colitis in mice, Int. J. Immunopathol. [252] K.K. Lai, G.L. Lorca, C.F. Gonzalez, Biochemical properties of two cinnamoyl es-
Pharmacol. 27 (2014) 615–627, http://dx.doi.org/10.1177/ terases purified from a Lactobacillus johnsonii strain isolated from stool samples of
039463201402700418. diabetes-resistant rats, Appl. Environ. Microbiol. 75 (2009) 5018–5024, http://dx.
[234] C.-U. Riedel, F. Foata, D. Philippe, O. Adolfsson, B.-J. Eikmanns, S. Blum, Anti- doi.org/10.1128/AEM.02837-08.
inflammatory effects of bifidobacteria by inhibition of LPS-induced NF-kappaB [253] R. Valladares, D. Sankar, N. Li, E. Williams, K.-K. Lai, A.S. Abdelgeliel,
activation, World J. Gastroenterol. 12 (2006) 3729–3735. C.F. Gonzalez, C.H. Wasserfall, J. Larkin, D. Schatz, M.A. Atkinson, E.W. Triplett,
[235] A.M. O'Hara, P. O'Regan, Á. Fanning, C. O'Mahony, J. MacSharry, A. Lyons, J. Neu, G.L. Lorca, Lactobacillus johnsonii N6.2 mitigates the development of type
J. Bienenstock, L. O'Mahony, F. Shanahan, Functional modulation of human in- 1 diabetes in BB-DP rats, PLoS One 5 (2010) e10507, , http://dx.doi.org/10.1371/
testinal epithelial cell responses by Bifidobacterium infantis and Lactobacillus journal.pone.0010507.
salivarius, Immunology 118 (2006) 202–215, http://dx.doi.org/10.1111/j.1365- [254] J. Dolpady, C. Sorini, C. Di Pietro, I. Cosorich, R. Ferrarese, D. Saita, M. Clementi,
2567.2006.02358.x. F. Canducci, M. Falcone, Oral probiotic VSL#3 prevents autoimmune diabetes by
[236] E.V. Khokhlova, V.V. Smeianov, B.A. Efimov, L.I. Kafarskaia, S.I. Pavlova, modulating microbiota and promoting indoleamine 2,3-dioxygenase-enriched
A.N. Shkoporov, Anti-inflammatory properties of intestinal Bifidobacterium tolerogenic intestinal environment, J. Diabetes Res. 2016 (2016) 7569431, ,
strains isolated from healthy infants, Microbiol. Immunol. 56 (2012) 27–39, http://dx.doi.org/10.1155/2016/7569431.
http://dx.doi.org/10.1111/j.1348-0421.2011.00398.x. [255] U. Uusitalo, X. Liu, J. Yang, C.A. Aronsson, S. Hummel, M. Butterworth,
[237] M. Roselli, A. Finamore, M.S. Britti, E. Mengheri, Probiotic bacteria Å. Lernmark, M. Rewers, W. Hagopian, J.-X. She, O. Simell, J. Toppari,
Bifidobacterium animalis MB5 and Lactobacillus rhamnosus GG protect intestinal A.G. Ziegler, B. Akolkar, J. Krischer, J.M. Norris, S.M. Virtanen, Association of
Caco-2 cells from the inflammation-associated response induced by en- early exposure of probiotics and islet autoimmunity in the TEDDY study, JAMA
terotoxigenic Escherichia coli K88, Br. J. Nutr. 95 (2006) 1177–1184. Pediatr 170 (2016) 20–28, http://dx.doi.org/10.1001/jamapediatrics.2015.2757.
[238] Y. Wang, K.M. Telesford, J. Ochoa-Repáraz, S. Haque-Begum, M. Christy, [256] Q. Mu, H. Zhang, X. Liao, K. Lin, H. Liu, M.R. Edwards, S.A. Ahmed, R. Yuan, L. Li,
E.J. Kasper, L. Wang, Y. Wu, S.C. Robson, D.L. Kasper, L.H. Kasper, An intestinal T.E. Cecere, D.B. Branson, J.L. Kirby, P. Goswami, C.M. Leeth, K.A. Read,
commensal symbiosis factor controls neuroinflammation via TLR2-mediated CD39 K.J. Oestreich, M.D. Vieson, C.M. Reilly, X.M. Luo, Control of lupus nephritis by
signalling, Nat. Commun. 5 (2014) 4432, http://dx.doi.org/10.1038/ changes of gut microbiota, Microbiome 5 (2017), http://dx.doi.org/10.1186/
ncomms5432. s40168-017-0300-8.
[239] S. Lavasani, B. Dzhambazov, M. Nouri, F. Fåk, S. Buske, G. Molin, H. Thorlacius, [257] M. Roberfroid, G.R. Gibson, L. Hoyles, A.L. McCartney, R. Rastall, I. Rowland,
J. Alenfall, B. Jeppsson, B. Weström, A novel probiotic mixture exerts a ther- D. Wolvers, B. Watzl, H. Szajewska, B. Stahl, F. Guarner, F. Respondek, K. Whelan,
apeutic effect on experimental autoimmune encephalomyelitis mediated by IL-10 V. Coxam, M.-J. Davicco, L. Léotoing, Y. Wittrant, N.M. Delzenne, P.D. Cani,
producing regulatory T cells, PLoS One 5 (2010) e9009, http://dx.doi.org/10. A.M. Neyrinck, A. Meheust, Prebiotic effects: metabolic and health benefits, Br. J.
1371/journal.pone.0009009. Nutr. 104 (Suppl 2) (2010) S1–S63, http://dx.doi.org/10.1017/
[240] J. Ezendam, A. de Klerk, E.R. Gremmer, H. van Loveren, Effects of Bifidobacterium S0007114510003363.
animalis administered during lactation on allergic and autoimmune responses in [258] J.O. Lindsay, K. Whelan, A.J. Stagg, P. Gobin, H.O. Al-Hassi, N. Rayment,
rodents, Clin. Exp. Immunol. 154 (2008) 424–431, http://dx.doi.org/10.1111/j. M.A. Kamm, S.C. Knight, A. Forbes, Clinical, microbiological, and immunological
1365-2249.2008.03788.x. effects of fructo-oligosaccharide in patients with Crohn's disease, Gut 55 (2006)
[241] K. Takata, M. Kinoshita, T. Okuno, M. Moriya, T. Kohda, J.A. Honorat, 348–355, http://dx.doi.org/10.1136/gut.2005.074971.
T. Sugimoto, A. Kumanogoh, H. Kayama, K. Takeda, S. Sakoda, Y. Nakatsuji, The [259] E.M. Dewulf, P.D. Cani, S.P. Claus, S. Fuentes, P.G. Puylaert, A.M. Neyrinck,
lactic acid bacterium Pediococcus acidilactici suppresses autoimmune en- L.B. Bindels, W.M. de Vos, G.R. Gibson, J.-P. Thissen, N.M. Delzenne, Insight into
cephalomyelitis by inducing IL-10-producing regulatory T cells, PLoS One 6 the prebiotic concept: lessons from an exploratory, double blind intervention study
(2011) e27644, , http://dx.doi.org/10.1371/journal.pone.0027644. with inulin-type fructans in obese women, Gut 62 (2013) 1112–1121, http://dx.
[242] R.M. Rezende, R.P. Oliveira, S.R. Medeiros, A.C. Gomes-Santos, A.C. Alves, doi.org/10.1136/gutjnl-2012-303304.
F.G. Loli, M.A.F. Guimarães, S.S. Amaral, A.P. da Cunha, H.L. Weiner, V. Azevedo, [260] P.D. Cani, S. Possemiers, T.V. de Wiele, Y. Guiot, A. Everard, O. Rottier, L. Geurts,
A. Miyoshi, A.M.C. Faria, Hsp65-producing Lactococcus lactis prevents experi- D. Naslain, A. Neyrinck, D.M. Lambert, G.G. Muccioli, N.M. Delzenne, Changes in
mental autoimmune encephalomyelitis in mice by inducing CD4+LAP+ reg- gut microbiota control inflammation in obese mice through a mechanism invol-
ulatory T cells, J. Autoimmun. 40 (2013) 45–57, http://dx.doi.org/10.1016/j.jaut. ving GLP-2-driven improvement of gut permeability, Gut 58 (2009) 1091–1103,
2012.07.012. http://dx.doi.org/10.1136/gut.2008.165886.
[243] Z. Salehipour, D. Haghmorad, M. Sankian, M. Rastin, R. Nosratabadi, M.M. Soltan [261] S. Satchithanandam, M. Vargofcak-Apker, R.J. Calvert, A.R. Leeds, M.M. Cassidy,
Dallal, N. Tabasi, M. Khazaee, L.R. Nasiraii, M. Mahmoudi, Bifidobacterium ani- Alteration of gastrointestinal mucin by fiber feeding in rats, J. Nutr. 120 (1990)
malis in combination with human origin of Lactobacillus plantarum ameliorate 1179–1184.
neuroinflammation in experimental model of multiple sclerosis by altering CD4+ [262] N. Fontaine, J.C. Meslin, S. Lory, C. Andrieux, Intestinal mucin distribution in the
T cell subset balance, Biomed. Pharmacother. Biomedecine Pharmacother 95 germ-free rat and in the heteroxenic rat harbouring a human bacterial flora: effect

20
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

of inulin in the diet, Br. J. Nutr. 75 (1996) 881–892. 1524890113.


[263] P. Van den Abbeele, P. Gérard, S. Rabot, A. Bruneau, S. El Aidy, M. Derrien, [283] E. Karunasena, K.W. McMahon, P.C. Kurkure, M.M. Brashears, A comparison of
M. Kleerebezem, E.G. Zoetendal, H. Smidt, W. Verstraete, T. Van de Wiele, cell mediators and serum cytokines transcript expression between male and female
S. Possemiers, Arabinoxylans and inulin differentially modulate the mucosal and mice infected with Mycobacterium avium subspecies paratuberculosis and/or
luminal gut microbiota and mucin-degradation in humanized rats, Environ. consuming probiotics, Pathog. Dis. 72 (2014) 104–110, http://dx.doi.org/10.
Microbiol. 13 (2011) 2667–2680, http://dx.doi.org/10.1111/j.1462-2920.2011. 1111/2049-632X.12193.
02533.x. [284] E. Karunasena, K.W. McMahon, D. Chang, M.M. Brashears, Host responses to the
[264] A. Barcelo, J. Claustre, F. Moro, J.A. Chayvialle, J.C. Cuber, P. Plaisancié, Mucin pathogen Mycobacterium avium subsp. paratuberculosis and beneficial microbes
secretion is modulated by luminal factors in the isolated vascularly perfused rat exhibit host sex specificity, Appl. Environ. Microbiol. 80 (2014) 4481–4490,
colon, Gut 46 (2000) 218–224. http://dx.doi.org/10.1128/AEM.01229-14.
[265] M. Derrien, P. Van Baarlen, G. Hooiveld, E. Norin, M. Müller, W.M. de Vos, [285] L.J. Brandt, O.C. Aroniadis, M. Mellow, A. Kanatzar, C. Kelly, T. Park, N. Stollman,
Modulation of mucosal immune response, tolerance, and proliferation in mice F. Rohlke, C. Surawicz, Long-term follow-up of colonoscopic fecal microbiota
colonized by the mucin-degrader Akkermansia muciniphila, Front. Microbiol. 2 transplant for recurrent Clostridium difficile infection, Am. J. Gastroenterol. 107
(2011), http://dx.doi.org/10.3389/fmicb.2011.00166. (2012) 1079–1087, http://dx.doi.org/10.1038/ajg.2012.60.
[266] M. Derrien, C. Belzer, W.M. de Vos, Akkermansia muciniphila and its role in [286] E. Mattila, R. Uusitalo-Seppälä, M. Wuorela, L. Lehtola, H. Nurmi, M. Ristikankare,
regulating host functions, Microb. Pathog. 106 (2017) 171–181, http://dx.doi. V. Moilanen, K. Salminen, M. Seppälä, P.S. Mattila, V.-J. Anttila, P. Arkkila, Fecal
org/10.1016/j.micpath.2016.02.005. transplantation, through colonoscopy, is effective therapy for recurrent
[267] A. Everard, C. Belzer, L. Geurts, J.P. Ouwerkerk, C. Druart, L.B. Bindels, Y. Guiot, Clostridium difficile infection, Gastroenterology 142 (2012) 490–496, http://dx.
M. Derrien, G.G. Muccioli, N.M. Delzenne, W.M. de Vos, P.D. Cani, Cross-talk doi.org/10.1053/j.gastro.2011.11.037.
between Akkermansia muciniphila and intestinal epithelium controls diet-induced [287] A. Khoruts, M.J. Sadowsky, Understanding the mechanisms of faecal microbiota
obesity, Proc. Natl. Acad. Sci. 110 (2013) 9066–9071, http://dx.doi.org/10.1073/ transplantation, Nat. Rev. Gastroenterol. Hepatol. 13 (2016) 508–516, http://dx.
pnas.1219451110. doi.org/10.1038/nrgastro.2016.98.
[268] J. Vulevic, A. Drakoularakou, P. Yaqoob, G. Tzortzis, G.R. Gibson, Modulation of [288] A. Khoruts, J. Dicksved, J.K. Jansson, M.J. Sadowsky, Changes in the composition
the fecal microflora profile and immune function by a novel trans-galactooligo- of the human fecal microbiome after bacteriotherapy for recurrent Clostridium
saccharide mixture (B-GOS) in healthy elderly volunteers, Am. J. Clin. Nutr. 88 difficile-associated diarrhea, J. Clin. Gastroenterol. 44 (2010) 354–360, http://dx.
(2008) 1438–1446. doi.org/10.1097/MCG.0b013e3181c87e02.
[269] Y. Liu, G.R. Gibson, G.E. Walton, An in vitro approach to study effects of prebiotics [289] C.R. Kelly, S. Kahn, P. Kashyap, L. Laine, D. Rubin, A. Atreja, T. Moore, G. Wu,
and probiotics on the faecal microbiota and selected immune parameters relevant Update on fecal microbiota transplantation 2015: indications, methodologies,
to the elderly, PLoS One 11 (2016) e0162604, , http://dx.doi.org/10.1371/ mechanisms, and outlook, Gastroenterology 149 (2015) 223–237, http://dx.doi.
journal.pone.0162604. org/10.1053/j.gastro.2015.05.008.
[270] A. Beyerlein, X. Liu, U.M. Uusitalo, M. Harsunen, J.M. Norris, K. Foterek, [290] M. Fischer, D. Kao, C. Kelly, A. Kuchipudi, S.-M. Jafri, M. Blumenkehl, D. Rex,
S.M. Virtanen, M.J. Rewers, J.-X. She, O. Simell, Å. Lernmark, W. Hagopian, M. Mellow, N. Kaur, H. Sokol, G. Cook, M.J. Hamilton, E. Phelps, B. Sipe, H. Xu,
B. Akolkar, A.-G. Ziegler, J.P. Krischer, S. Hummel, Dietary intake of soluble fiber J.R. Allegretti, Fecal microbiota transplantation is safe and efficacious for re-
and risk of islet autoimmunity by 5 y of age: results from the TEDDY study12, Am, current or refractory Clostridium difficile infection in patients with inflammatory
J. Clin. Nutr. 102 (2015) 345–352, http://dx.doi.org/10.3945/ajcn.115.108159. bowel disease, Inflamm. Bowel Dis. 22 (2016) 2402–2409, http://dx.doi.org/10.
[271] E. Mariño, J.L. Richards, K.H. McLeod, D. Stanley, Y.A. Yap, J. Knight, 1097/MIB.0000000000000908.
C. McKenzie, J. Kranich, A.C. Oliveira, F.J. Rossello, B. Krishnamurthy, [291] J.L. Anderson, R.J. Edney, K. Whelan, Systematic review: faecal microbiota
C.M. Nefzger, L. Macia, A. Thorburn, A.G. Baxter, G. Morahan, L.H. Wong, transplantation in the management of inflammatory bowel disease, Aliment.
J.M. Polo, R.J. Moore, T.J. Lockett, J.M. Clarke, D.L. Topping, L.C. Harrison, Pharmacol. Ther. 36 (2012) 503–516, http://dx.doi.org/10.1111/j.1365-2036.
C.R. Mackay, Gut microbial metabolites limit the frequency of autoimmune T cells 2012.05220.x.
and protect against type 1 diabetes, Nat. Immunol. 18 (2017) 552–562, http://dx. [292] R.J. Colman, D.T. Rubin, Fecal microbiota transplantation as therapy for in-
doi.org/10.1038/ni.3713. flammatory bowel disease: a systematic review and meta-analysis, J. Crohns
[272] J.C. Needell, D. Ir, C.E. Robertson, M.E. Kroehl, D.N. Frank, D. Zipris, Maternal Colitis 8 (2014) 1569–1581, http://dx.doi.org/10.1016/j.crohns.2014.08.006.
treatment with short-chain fatty acids modulates the intestinal microbiota and [293] N.A. Cohen, N. Maharshak, Novel indications for fecal microbial transplantation:
immunity and ameliorates type 1 diabetes in the offspring, PLoS One 12 (2017), update and review of the literature, Dig. Dis. Sci. 62 (2017) 1131–1145, http://dx.
http://dx.doi.org/10.1371/journal.pone.0183786. doi.org/10.1007/s10620-017-4535-9.
[273] M. Mizuno, D. Noto, N. Kaga, A. Chiba, S. Miyake, The dual role of short fatty acid [294] S. Khanna, Microbiota replacement therapies: innovation in gastrointestinal care,
chains in the pathogenesis of autoimmune disease models, PLoS One 12 (2017) Clin. Pharmacol. Ther. 103 (2018) 102–111, http://dx.doi.org/10.1002/cpt.923.
e0173032, , http://dx.doi.org/10.1371/journal.pone.0173032. [295] D. Kao, N. Hotte, P. Gillevet, K. Madsen, Fecal microbiota transplantation inducing
[274] J. Park, C.J. Goergen, H. HogenEsch, C.H. Kim, Chronically elevated levels of remission in Crohn's colitis and the associated changes in fecal microbial profile, J.
short-chain fatty acids induce T cell-mediated ureteritis and hydronephrosis, J. Clin. Gastroenterol. 48 (2014) 625–628, http://dx.doi.org/10.1097/MCG.
Immunol. Baltim. Md 1950 196 (2016) 2388–2400, http://dx.doi.org/10.4049/ 0000000000000131.
jimmunol.1502046. [296] P. Moayyedi, M.G. Surette, P.T. Kim, J. Libertucci, M. Wolfe, C. Onischi,
[275] E. Catry, L.B. Bindels, A. Tailleux, S. Lestavel, A.M. Neyrinck, J.-F. Goossens, D. Armstrong, J.K. Marshall, Z. Kassam, W. Reinisch, C.H. Lee, Fecal microbiota
I. Lobysheva, H. Plovier, A. Essaghir, J.-B. Demoulin, C. Bouzin, B.D. Pachikian, transplantation induces remission in patients with active ulcerative colitis in a
P.D. Cani, B. Staels, C. Dessy, N.M. Delzenne, Targeting the gut microbiota with randomized controlled trial, Gastroenterology 149 (2015), http://dx.doi.org/10.
inulin-type fructans: preclinical demonstration of a novel approach in the man- 1053/j.gastro.2015.04.001 102–109.e6.
agement of endothelial dysfunction, Gut (2017), http://dx.doi.org/10.1136/ [297] D. Kim, S.-A. Yoo, W.-U. Kim, Gut microbiota in autoimmunity: potential for
gutjnl-2016-313316. clinical applications, Arch. Pharm. Res. 39 (2016) 1565–1576, http://dx.doi.org/
[276] K. Dilger, S. Hohenester, U. Winkler-Budenhofer, B.A.J. Bastiaansen, F.G. Schaap, 10.1007/s12272-016-0796-7.
C. Rust, U. Beuers, Effect of ursodeoxycholic acid on bile acid profiles and in- [298] D. Kao, B. Roach, H. Park, N. Hotte, K. Madsen, V. Bain, P. Tandon, Fecal mi-
testinal detoxification machinery in primary biliary cirrhosis and health, J. crobiota transplantation in the management of hepatic encephalopathy, Hepatol.
Hepatol. 57 (2012) 133–140, http://dx.doi.org/10.1016/j.jhep.2012.02.014. Baltim. Md 63 (2016) 339–340, http://dx.doi.org/10.1002/hep.28121.
[277] Y. Calmus, J. Guechot, P. Podevin, M.T. Bonnefis, J. Giboudeau, R. Poupon, [299] D.-W. Kang, J.B. Adams, A.C. Gregory, T. Borody, L. Chittick, A. Fasano,
Differential effects of chenodeoxycholic and ursodeoxycholic acids on interleukin A. Khoruts, E. Geis, J. Maldonado, S. McDonough-Means, E.L. Pollard, S. Roux,
1, interleukin 6 and tumor necrosis factor-alpha production by monocytes, M.J. Sadowsky, K.S. Lipson, M.B. Sullivan, J.G. Caporaso, R. Krajmalnik-Brown,
Hepatol. Baltim. Md 16 (1992) 719–723. Microbiota Transfer Therapy alters gut ecosystem and improves gastrointestinal
[278] G. Filaci, N. Pelli, T. Sacco, P. Contini, L. Lanza, A. Picciotto, M. Scudeletti, and autism symptoms: an open-label study, Microbiome 5 (2017) 10, http://dx.
F. Puppo, G. Castiglioni, F. Indiveri, S-adenosil-L-methionine is able to reverse the doi.org/10.1186/s40168-016-0225-7.
immunosuppressive effects of chenodeoxycholic acid in vitro, Int. J. [300] T.J. Borody, A. Khoruts, Fecal microbiota transplantation and emerging applica-
Immunopharm. 19 (1997) 157–165. tions, Nat. Rev. Gastroenterol. Hepatol. 9 (2011) 88–96, http://dx.doi.org/10.
[279] S. Suwannaroj, A. Lagoo, R.W. McMurray, Suppression of renal disease and 1038/nrgastro.2011.244.
mortality in the female NZB x NZW F1 mouse model of systemic lupus er- [301] A.H. Ali, E.J. Carey, K.D. Lindor, The microbiome and primary sclerosing cho-
ythematosus (SLE) by chenodeoxycholic acid, Lupus 10 (2001) 562–567, http:// langitis, Semin. Liver Dis. 36 (2016) 340–348, http://dx.doi.org/10.1055/s-0036-
dx.doi.org/10.1191/096120301701549697. 1594007.
[280] V. Sepe, C. Festa, B. Renga, A. Carino, S. Cipriani, C. Finamore, D. Masullo, F. del [302] D. Rebello, E. Wang, E. Yen, P.A. Lio, C.R. Kelly, Hair growth in two alopecia
Gaudio, M.C. Monti, S. Fiorucci, A. Zampella, Insights on FXR selective modula- patients after fecal microbiota transplant, ACG Case Rep. J. 4 (2017) e107, http://
tion. Speculation on bile acid chemical space in the discovery of potent and se- dx.doi.org/10.14309/crj.2017.107.
lective agonists, Sci. Rep. 6 (2016) 19008, http://dx.doi.org/10.1038/srep19008. [303] L.-L. Wang, H.-H. Guo, S. Huang, C.-L. Feng, Y.-X. Han, J.-D. Jiang,
[281] L.J. Ceulemans, L. Verbeke, J.-P. Decuypere, R. Farré, G. De Hertogh, K. Lenaerts, Comprehensive evaluation of SCFA production in the intestinal bacteria regulated
I. Jochmans, D. Monbaliu, F. Nevens, J. Tack, W. Laleman, J. Pirenne, Farnesoid X by berberine using gas-chromatography combined with polymerase chain reac-
receptor activation attenuates intestinal ischemia reperfusion injury in rats, PLoS tion, J. Chromatogr. B 1057 (2017) 70–80, http://dx.doi.org/10.1016/j.jchromb.
One 12 (2017) e0169331, , http://dx.doi.org/10.1371/journal.pone.0169331. 2017.05.004.
[282] P.P. Ho, L. Steinman, Obeticholic acid, a synthetic bile acid agonist of the farne- [304] D.J. Morrison, T. Preston, Formation of short chain fatty acids by the gut micro-
soid X receptor, attenuates experimental autoimmune encephalomyelitis, Proc. biota and their impact on human metabolism, Gut Microb. 7 (2016) 189–200,
Natl. Acad. Sci. 113 (2016) 1600–1605, http://dx.doi.org/10.1073/pnas. http://dx.doi.org/10.1080/19490976.2015.1134082.

21
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

[305] J.H. Cummings, E.W. Pomare, W.J. Branch, C.P. Naylor, G.T. Macfarlane, Short 1002/hep.24537.
chain fatty acids in human large intestine, portal, hepatic and venous blood, Gut [326] H. Zhang, Y. Liu, Z. Bian, S. Huang, X. Han, Z. You, Q. Wang, D. Qiu, Q. Miao,
28 (1987) 1221–1227. Y. Peng, X. Li, P. Invernizzi, X. Ma, The critical role of myeloid-derived suppressor
[306] G.T. Macfarlane, G.R. Gibson, J.H. Cummings, Comparison of fermentation reac- cells and FXR activation in immune-mediated liver injury, J. Autoimmun. 53
tions in different regions of the human colon, J. Appl. Bacteriol. 72 (1992) 57–64. (2014) 55–66, http://dx.doi.org/10.1016/j.jaut.2014.02.010.
[307] C.M. van der Beek, J.G. Bloemen, M.A. van den Broek, K. Lenaerts, K. Venema, [327] H. Duboc, S. Rajca, D. Rainteau, D. Benarous, M.-A. Maubert, E. Quervain,
W.A. Buurman, C.H. Dejong, Hepatic uptake of rectally administered butyrate G. Thomas, V. Barbu, L. Humbert, G. Despras, C. Bridonneau, F. Dumetz, J.-
prevents an increase in systemic butyrate concentrations in humans, J. Nutr. 145 P. Grill, J. Masliah, L. Beaugerie, J. Cosnes, O. Chazouillères, R. Poupon, C. Wolf,
(2015) 2019–2024, http://dx.doi.org/10.3945/jn.115.211193. J.-M. Mallet, P. Langella, G. Trugnan, H. Sokol, P. Seksik, Connecting dysbiosis,
[308] J.G. Bloemen, K. Venema, M.C. van de Poll, S.W. Olde Damink, W.A. Buurman, bile-acid dysmetabolism and gut inflammation in inflammatory bowel diseases,
C.H. Dejong, Short chain fatty acids exchange across the gut and liver in humans Gut 62 (2013) 531–539, http://dx.doi.org/10.1136/gutjnl-2012-302578.
measured at surgery, Clin. Nutr. Edinb. Scotl 28 (2009) 657–661, http://dx.doi. [328] J.H. Tabibian, S.P. O'Hara, C.E. Trussoni, P.S. Tietz, P.L. Splinter, T. Mounajjed,
org/10.1016/j.clnu.2009.05.011. L.R. Hagey, N.F. LaRusso, Absence of the intestinal microbiota exacerbates he-
[309] C.J. Kelly, L. Zheng, E.L. Campbell, B. Saeedi, C.C. Scholz, A.J. Bayless, patobiliary disease in a murine model of primary sclerosing cholangitis, Hepatol.
K.E. Wilson, L.E. Glover, D.J. Kominsky, A. Magnuson, T.L. Weir, S.F. Ehrentraut, Baltim. Md 63 (2016) 185–196, http://dx.doi.org/10.1002/hep.27927.
C. Pickel, K.A. Kuhn, J.M. Lanis, V. Nguyen, C.T. Taylor, S.P. Colgan, Crosstalk [329] L.A. Thomas, M.J. Veysey, G. French, P.B. Hylemon, G.M. Murphy, R.H. Dowling,
between microbiota-derived short-chain fatty acids and intestinal epithelial HIF Bile acid metabolism by fresh human colonic contents: a comparison of caecal
augments tissue barrier function, Cell Host Microbe 17 (2015) 662–671, http://dx. versus faecal samples, Gut 49 (2001) 835–842.
doi.org/10.1016/j.chom.2015.03.005. [330] S.A. Joyce, J. MacSharry, P.G. Casey, M. Kinsella, E.F. Murphy, F. Shanahan,
[310] R. Corrêa-Oliveira, J.L. Fachi, A. Vieira, F.T. Sato, M.A.R. Vinolo, Regulation of C. Hill, C.G.M. Gahan, Regulation of host weight gain and lipid metabolism by
immune cell function by short-chain fatty acids, Clin. Transl. Immunol 5 (2016) bacterial bile acid modification in the gut, Proc. Natl. Acad. Sci. U. S. A 111 (2014)
e73, http://dx.doi.org/10.1038/cti.2016.17. 7421–7426, http://dx.doi.org/10.1073/pnas.1323599111.
[311] A.J. Brown, S.M. Goldsworthy, A.A. Barnes, M.M. Eilert, L. Tcheang, D. Daniels, [331] M. Chen, L. Yi, Y. Zhang, X. Zhou, L. Ran, J. Yang, J. Zhu, Q. Zhang, M. Mi,
A.I. Muir, M.J. Wigglesworth, I. Kinghorn, N.J. Fraser, N.B. Pike, J.C. Strum, Resveratrol attenuates trimethylamine-N-oxide (TMAO)-Induced atherosclerosis
K.M. Steplewski, P.R. Murdock, J.C. Holder, F.H. Marshall, P.G. Szekeres, by regulating TMAO synthesis and bile acid metabolism via remodeling of the gut
S. Wilson, D.M. Ignar, S.M. Foord, A. Wise, S.J. Dowell, The Orphan G protein- microbiota, mBio 7 (2016), http://dx.doi.org/10.1128/mBio.02210-15.
coupled receptors GPR41 and GPR43 are activated by propionate and other short [332] T. Arora, R.L. Loo, J. Anastasovska, G.R. Gibson, K.M. Tuohy, R.K. Sharma,
chain carboxylic acids, J. Biol. Chem. 278 (2003) 11312–11319, http://dx.doi. J.R. Swann, E.R. Deaville, M.L. Sleeth, E.L. Thomas, E. Holmes, J.D. Bell, G. Frost,
org/10.1074/jbc.M211609200. Differential effects of two fermentable carbohydrates on central appetite regula-
[312] E. Le Poul, C. Loison, S. Struyf, J.-Y. Springael, V. Lannoy, M.-E. Decobecq, tion and body composition, PLoS One 7 (2012) e43263, , http://dx.doi.org/10.
S. Brezillon, V. Dupriez, G. Vassart, J. Van Damme, M. Parmentier, M. Detheux, 1371/journal.pone.0043263.
Functional characterization of human receptors for short chain fatty acids and [333] B. Drzikova, G. Dongowski, E. Gebhardt, Dietary fibre-rich oat-based products
their role in polymorphonuclear cell activation, J. Biol. Chem. 278 (2003) affect serum lipids, microbiota, formation of short-chain fatty acids and steroids in
25481–25489, http://dx.doi.org/10.1074/jbc.M301403200. rats, Br. J. Nutr. 94 (2005) 1012–1025.
[313] K. Ahmed, S. Tunaru, S. Offermanns, GPR109A, GPR109B and GPR81, a family of [334] S.J.D. O'Keefe, J.V. Li, L. Lahti, J. Ou, F. Carbonero, K. Mohammed, J.M. Posma,
hydroxy-carboxylic acid receptors, Trends Pharmacol. Sci. 30 (2009) 557–562, J. Kinross, E. Wahl, E. Ruder, K. Vipperla, V. Naidoo, L. Mtshali, S. Tims,
http://dx.doi.org/10.1016/j.tips.2009.09.001. P.G.B. Puylaert, J. DeLany, A. Krasinskas, A.C. Benefiel, H.O. Kaseb, K. Newton,
[314] J.L. Pluznick, R.J. Protzko, H. Gevorgyan, Z. Peterlin, A. Sipos, J. Han, I. Brunet, J.K. Nicholson, W.M. de Vos, H.R. Gaskins, E.G. Zoetendal, Fat, fibre and cancer
L.-X. Wan, F. Rey, T. Wang, S.J. Firestein, M. Yanagisawa, J.I. Gordon, risk in African Americans and rural Africans, Nat. Commun. 6 (2015) 6342, http://
A. Eichmann, J. Peti-Peterdi, M.J. Caplan, Olfactory receptor responding to gut dx.doi.org/10.1038/ncomms7342.
microbiota-derived signals plays a role in renin secretion and blood pressure [335] S. Devkota, Y. Wang, M.W. Musch, V. Leone, H. Fehlner-Peach, A. Nadimpalli,
regulation, Proc. Natl. Acad. Sci. U. S. A 110 (2013) 4410–4415, http://dx.doi. D.A. Antonopoulos, B. Jabri, E.B. Chang, Dietary-fat-induced taurocholic acid
org/10.1073/pnas.1215927110. promotes pathobiont expansion and colitis in Il10-/- mice, Nature 487 (2012)
[315] M. Waldecker, T. Kautenburger, H. Daumann, C. Busch, D. Schrenk, Inhibition of 104–108, http://dx.doi.org/10.1038/nature11225.
histone-deacetylase activity by short-chain fatty acids and some polyphenol me- [336] S. Modica, R.M. Gadaleta, A. Moschetta, Deciphering the nuclear bile acid receptor
tabolites formed in the colon, J. Nutr. Biochem. 19 (2008) 587–593, http://dx.doi. FXR paradigm, Nucl. Recept. Signal. 8 (2010), http://dx.doi.org/10.1621/nrs.
org/10.1016/j.jnutbio.2007.08.002. 08005.
[316] H.N. Sanchez, T. Shen, J.R. Taylor, K. Munoz, H. Zan, P. Casali, Short-chain fatty [337] Y. Li, R. Tang, P.S.C. Leung, M.E. Gershwin, X. Ma, Bile acids and intestinal mi-
acid (SCFA) histone deacetylase inhibitors produced by gut microbiota regulate crobiota in autoimmune cholestatic liver diseases, Autoimmun. Rev. 16 (2017)
selected microRNAs to modulate local and systemic antibody responses, J. 885–896, http://dx.doi.org/10.1016/j.autrev.2017.07.002.
Immunol. 196 (2016) 127.6-127.6. [338] T. Fujino, M. Une, T. Imanaka, K. Inoue, T. Nishimaki-Mogami, Structure-activity
[317] J.M. Rumberger, J.R.S. Arch, A. Green, Butyrate and other short-chain fatty acids relationship of bile acids and bile acid analogs in regard to FXR activation, J. Lipid
increase the rate of lipolysis in 3T3-L1 adipocytes, PeerJ 2 (2014) e611, http://dx. Res. 45 (2004) 132–138, http://dx.doi.org/10.1194/jlr.M300215-JLR200.
doi.org/10.7717/peerj.611. [339] G. Campana, P. Pasini, A. Roda, S. Spampinato, Regulation of ileal bile acid-
[318] X. Gao, S.-H. Lin, F. Ren, J.-T. Li, J.-J. Chen, C.-B. Yao, H.-B. Yang, S.-X. Jiang, G.- binding protein expression in Caco-2 cells by ursodeoxycholic acid: role of the
Q. Yan, D. Wang, Y. Wang, Y. Liu, Z. Cai, Y.-Y. Xu, J. Chen, W. Yu, P.-Y. Yang, Q.- farnesoid X receptor, Biochem. Pharmacol. 69 (2005) 1755–1763, http://dx.doi.
Y. Lei, Acetate functions as an epigenetic metabolite to promote lipid synthesis org/10.1016/j.bcp.2005.03.019.
under hypoxia, Nat. Commun. 7 (2016) 11960, http://dx.doi.org/10.1038/ [340] J. Cui, L. Huang, A. Zhao, J.-L. Lew, J. Yu, S. Sahoo, P.T. Meinke, I. Royo,
ncomms11960. F. Pelaez, S.D. Wright, Guggulsterone is a farnesoid X receptor antagonist in
[319] M.L. Soliman, T.A. Rosenberger, Acetate supplementation increases brain histone coactivator association assays but acts to enhance transcription of bile salt export
acetylation and inhibits histone deacetylase activity and expression, Mol. Cell. pump, J. Biol. Chem. 278 (2003) 10214–10220, http://dx.doi.org/10.1074/jbc.
Biochem. 352 (2011) 173–180, http://dx.doi.org/10.1007/s11010-011-0751-3. M209323200.
[320] M.L. Soliman, M.D. Smith, H.M. Houdek, T.A. Rosenberger, Acetate supple- [341] F.R.C. Costa, M.C.S. Françozo, G.G. de Oliveira, A. Ignacio, A. Castoldi,
mentation modulates brain histone acetylation and decreases interleukin-1β ex- D.S. Zamboni, S.G. Ramos, N.O. Câmara, M.R. de Zoete, N.W. Palm, R.A. Flavell,
pression in a rat model of neuroinflammation, J. Neuroinflammation 9 (2012) 51, J.S. Silva, D. Carlos, Gut microbiota translocation to the pancreatic lymph nodes
http://dx.doi.org/10.1186/1742-2094-9-51. triggers NOD2 activation and contributes to T1D onset, J. Exp. Med. 213 (2016)
[321] J.M. Ridlon, D.-J. Kang, P.B. Hylemon, Bile salt biotransformations by human 1223–1239, http://dx.doi.org/10.1084/jem.20150744.
intestinal bacteria, J. Lipid Res. 47 (2006) 241–259, http://dx.doi.org/10.1194/ [342] K.R. Rumah, J. Linden, V.A. Fischetti, T. Vartanian, Isolation of Clostridium per-
jlr.R500013-JLR200. fringens type B in an individual at first clinical presentation of multiple sclerosis
[322] E.J. Stellwag, P.B. Hylemon, Purification and characterization of bile salt hydro- provides clues for environmental triggers of the disease, PLoS One 8 (2013)
lase from Bacteroides fragilis subsp. fragilis, Biochim. Biophys. Acta 452 (1976) e76359, , http://dx.doi.org/10.1371/journal.pone.0076359.
165–176. [343] B.L. Cantarel, E. Waubant, C. Chehoud, J. Kuczynski, T.Z. DeSantis, J. Warrington,
[323] R. Edenharder, J. Schneider, 12 beta-dehydrogenation of bile acids by Clostridium A. Venkatesan, C.M. Fraser, E.M. Mowry, Gut microbiota in multiple sclerosis:
paraputrificum, C. tertium, and C. difficile and epimerization at carbon-12 of possible influence of immunomodulators, J. Investig. Med. Off. Publ. Am. Fed.
deoxycholic acid by cocultivation with 12 alpha-dehydrogenating Eubacterium Clin. Res. 63 (2015) 729–734, http://dx.doi.org/10.1097/JIM.
lentum, Appl. Environ. Microbiol. 49 (1985) 964–968. 0000000000000192.
[324] S.I. Sayin, A. Wahlström, J. Felin, S. Jäntti, H.-U. Marschall, K. Bamberg,
B. Angelin, T. Hyötyläinen, M. Orešič, F. Bäckhed, Gut microbiota regulates bile
acid metabolism by reducing the levels of tauro-beta-muricholic acid, a naturally List of abbreviations
occurring FXR antagonist, Cell Metabol. 17 (2013) 225–235, http://dx.doi.org/10.
1016/j.cmet.2013.01.003. 6-ECDCA: 6-ethyl chenodeoxycholic acid
[325] A. Baghdasaryan, T. Claudel, J. Gumhold, D. Silbert, L. Adorini, A. Roda, AID: autoimmune disease
S. Vecchiotti, F.J. Gonzalez, K. Schoonjans, M. Strazzabosco, P. Fickert, BA: bile acid
M. Trauner, Dual farnesoid X receptor/TGR5 agonist INT-767 reduces liver injury BB-D: Biobreeding diabetes
in the Mdr2-/- (Abcb4-/-) mouse cholangiopathy model by promoting biliary BSH: bile salt hydrolase
HCO−₃ output, Hepatol. Baltim. Md 54 (2011) 1303–1312, http://dx.doi.org/10. DC: dendritic cell

22
L. Rizzetto et al. Journal of Autoimmunity xxx (xxxx) xxx–xxx

EAE: experimental autoimmune encephalomyelitis NK: natural killer


ERα: estrogen receptor α NOD: nucleotide-binding oligomerization domain containing
IBD: inflammatory bowel disease OCA: obeticholic acid
HDAC: histone deacetylase regB: regulatory B cells
Ig: immunoglobulin SCFA: short chain fatty acids
IL: interleukin SFB: segmented filamentous bacteria
FMT: fecal microbiota transplantation SLE: systemic lupus erythematosus
FXR: farnesoid X receptor SPF: specific pathogen free
GPCR: G protein coupled receptor T1D: type 1 diabetes
LPS: lipopolysaccharide Th: T helper cells
MAMPs: microbe-associated molecular patterns TJ: tight junction
MDP: muramyl dipeptide TLR: Toll like receptors
MS: multiple sclerosis Treg: regulatory T cells
NLR: NOD like receptor

23

You might also like