Sie sind auf Seite 1von 13

F O C U S O N H o m e os tat i c I m m un e R e s p ons e s

REVIEWS
Intestinal epithelial cells: regulators
of barrier function and immune
homeostasis
Lance W. Peterson1 and David Artis1,2
Abstract | The abundance of innate and adaptive immune cells that reside together with
trillions of beneficial commensal microorganisms in the mammalian gastrointestinal
tract requires barrier and regulatory mechanisms that conserve host–microbial
interactions and tissue homeostasis. This homeostasis depends on the diverse functions
of intestinal epithelial cells (IECs), which include the physical segregation of commensal
bacteria and the integration of microbial signals. Hence, IECs are crucial mediators of
intestinal homeostasis that enable the establishment of an immunological environment
permissive to colonization by commensal bacteria. In this Review, we provide a
comprehensive overview of how IECs maintain host–commensal microbial relationships
and immune cell homeostasis in the intestine.

Inflammatory bowel disease


Specialized epithelial cells constitute barrier surfaces that function contributes to systemic immune activation,
(IBD). A chronic condition of separate mammalian hosts from the external environ- which promotes the progression of chronic viral infec-
the intestine characterized ment. The gastrointestinal tract is the largest of these tions, including infection with HIV and hepatitis virus2,3,
by severe inflammation barriers and is specially adapted to colonization by and metabolic disease4,5. Furthermore, host–microbial
and mucosal destruction.
commensal bacteria that aid in digestion and markedly interactions that occur at the IEC barrier contribute to a
The most common forms of
IBD in humans are ulcerative influence the development and function of the mucosal broad range of extra-intestinal autoimmune and inflam-
colitis and Crohn’s disease, immune system. However, microbial colonization car- matory diseases, including type 1 diabetes, rheumatoid
which have both distinct and ries with it the risk of infection and inflammation if epi- arthritis and multiple sclerosis6–9. Hence, a comprehen-
overlapping pathological and thelial or immune cell homeostasis is disrupted. Key to sive understanding of the barrier and immunoregulatory
clinical characteristics.
the coexistence of commensal microbial communities properties of IECs could aid in the development of new
Mucins and mucosal immune cells is the capacity to maintain the strategies to prevent and treat multiple human infectious,
Heavily glycosylated proteins segregation between host and microorganism. The intes- inflammatory and metabolic diseases.
that are the major component tinal epithelium accomplishes this by forming a physical The topics of commensal bacterial diversity, microbial
of the mucus that coats
and biochemical barrier to commensal and pathogenic regulation of immune cell development and host–viral
epithelial barrier surfaces.
microorganisms. Furthermore, intestinal epithelial cells interactions in the intestine have been reviewed exten-
(IECs) can sense and respond to microbial stimuli to sively elsewhere10–14. Therefore, in this Review, we dis-
reinforce their barrier function and to participate in the cuss the role of IECs in promoting intestinal homeostasis
1
Department of Microbiology coordination of appropriate immune responses, ranging through the segregation and regulation of commensal
and Institute for Immunology, from tolerance to anti-pathogen immunity. Thus, IECs microorganisms and the host immune system. Recent
Perelman School of Medicine, maintain a fundamental immuno­regulatory function advances in the understanding of the barrier, microbial-
University of Pennsylvania.
that influences the development and homeostasis of sensing and immunoregulatory functions of IECs are
2
Department of Pathobiology,
School of Veterinary mucosal immune cells. reviewed, with a particular focus on their relationship to
Medicine, University of The association between increased bacterial trans­ intestinal health and disease. We discuss the barrier func-
Pennsylvania, Philadelphia, location and risk of developing inflammatory bowel disease tion maintained by IEC-derived mucins and anti­microbial
Pennsylvania 19104, USA. (IBD) suggests a central role for dysregulated epithelial proteins, the pathways through which IECs regulate
e‑mails: lancep@mail.med.
upenn.edu;
barrier function in either the aetiology or the pathol- innate and adaptive immune cells present at the intes-
dartis@mail.med.upenn.edu ogy of intestinal inflammation and IBD1. Increasing tinal barrier and the contribution of IEC recognition of
doi:10.1038/nri3608 evidence also indicates that the loss of intestinal barrier microbial colonization to IEC function and homeostasis.

NATURE REVIEWS | IMMUNOLOGY VOLUME 14 | MARCH 2014 | 141

© 2014 Macmillan Publishers Limited. All rights reserved


R E V IE W S

Small intestine Follicle-associated epithelium Colon

Apoptotic
Commensal IECs
bacteria

Mucus

Second-
layer
TFF3 mucus
sIgA

AMPs
M cell
Mucus
Enteroendocrine
Goblet cell
cell
Enterocyte

Stromal B cell
cell

Lymphoid
follicle

Paneth
cell IESC
Macrophage DC

Figure 1 | The IEC barrier.  Intestinal epithelial cells (IECs) form a biochemical and physical barrier that maintains
segregation between luminal microbial communities and the mucosal immune system. The intestinal
Nature Reviewsepithelial stem
| Immunology
cell (IESC) niche, containing epithelial, stromal and haematopoietic cells, controls the continuous renewal of the
epithelial cell layer by crypt-resident stem cells. Differentiated IECs — with the exception of Paneth cells — migrate
up the crypt–villus axis, as indicated by the dashed arrows. Secretory goblet cells and Paneth cells secrete mucus and
Crypts antimicrobial proteins (AMPs) to promote the exclusion of bacteria from the epithelial surface. The transcytosis and
Tubular invaginations of the luminal release of secretory IgA (sIgA) further contribute to this barrier function. Microfold cells (M cells) and goblet
intestinal epithelium. Lining cells mediate transport of luminal antigens and live bacteria across the epithelial barrier to dendritic cells (DCs),
the base of the crypts are and intestine-resident macrophages sample the lumen through transepithelial dendrites. TFF3, trefoil factor 3.
small intestinal Paneth cells,
which produce numerous
antimicrobial proteins, and
stem cells, which continuously
divide to give rise to the entire
IEC regulation of barrier function digestive function. The luminal secretion of mucins
intestinal epithelium. The intestinal epithelium is the largest of the body’s and antimicrobial proteins (AMPs) by goblet cells and
mucosal surfaces, covering ~400 m2 of surface area Paneth cells, respectively, establishes a physical and
Villi with a single layer of cells organized into crypts and biochemical barrier to microbial contact with the epi-
Projections of the intestinal
villi (FIG. 1) . This surface is continually renewed by thelial surface and underlying immune cells17,18 (FIG. 1).
epithelium into the lumen
of the small intestine that pluripotent intestinal epithelial stem cells (pluripotent Collectively, the diverse functions of IECs result in
have an outer layer IESCs) that reside in the base of crypts, where the pro- a dynamic barrier to the environment, which protects
consisting of mature, liferation, differentiation and functional potential of the host from infection and continuous exposure to
absorptive enterocytes, epithelial cell progenitors is regulated by the local stem potentially inflammatory stimuli.
mucus-secreting goblet cells
and enteroendocrine cells.
cell niche15,16 (BOX 1). Although the majority of cells
bordering the intestinal lumen are absorptive entero- Keeping the bugs at bay — IEC secretory defences. The
Pluripotent intestinal cytes, which are adapted for metabolic and digestive secretion of highly glycosylated mucins into the intesti-
epithelial stem cells function, the diversity of functions that the intestinal nal lumen by goblet cells creates the first line of defence
(Pluripotent IESCs).
epithelium carries out is reflected by the presence of against microbial encroachment. The most abundant
Tissue-resident stem cells
that undergo continuous additional specialized IEC lineages. of these mucins, mucin 2 (MUC2), plays an essential
self-renewal and are Secretory IECs, including enteroendocrine cells, part in the organization of the intestinal mucous layers
responsible for regenerating goblet cells and Paneth cells, are specialized for main- at the epithelial surface of the colon19. The importance
all lineages of mature taining the digestive or barrier function of the epithe- of mucin production by goblet cells is emphasized by
intestinal epithelial cells,
including enterocytes,
lium. Enteroendocrine cells represent a link between the the spontaneous development of colitis and the predis-
enteroendocrine cells, central and enteric neuroendocrine systems through position to inflammation-induced colorectal cancers
goblet cells and Paneth cells. the secretion of numerous hormone regulators of observed in MUC2‑deficient mice20,21. Additional goblet

142 | MARCH 2014 | VOLUME 14 www.nature.com/reviews/immunol

© 2014 Macmillan Publishers Limited. All rights reserved


F O C U S O N H o m e os tat i c I m m un e R e sRpEons eS
V IE W s

Box 1 | The IESC niche


Paneth cell- and enterocyte-derived REGIIIγ has
recently been described as a mediator of host–microbial
Along the crypt–villus axis of the epithelium, pluripotent intestinal epithelial segregation in the gut 29. Similar to the function and
stem cells (IESCs) residing in the base of crypts give rise to a transit-amplifying regulation of MUC2 in the colon, REGIIIγ acts to
population of cells that undergo rapid proliferation and differentiation into the exclude bacteria from the epithelial surface of the
various intestinal epithelial cell (IEC) subsets. Terminally differentiated cells — with
small intestine, and its production is dependent on
the exception of Paneth cells — migrate up the crypt–villus axis until they are lost
from the epithelial layer. For this process to be maintained, epithelial stem cells
IEC-intrinsic recognition of commensal microbial sig-
must be able to undergo repeated rounds of replication and possess the capacity nals29. Interactions between AMPs, including REGIIIγ,
for continuous self-renewal16. Recent advances in stem cell biology have identified and mucins lead to concentrated antimicrobial activity
markers of IESCs that have contributed to the understanding of epithelial at the epithelial surface30. Thus, the combined func-
self-renewal and differentiation16,183–185. tions of secretory IECs seem to limit the quantity and
The patterning and distribution of proliferating crypt units in the intestine depend diversity of live bacteria that can reach the epithe-
on paracrine signalling between the epithelium and the underlying mesenchyme. lial surface or interact with the underlying mucosal
A balance between bone morphogenetic protein signals and antagonists, such as immune system.
noggin and gremlin, provides a niche for proliferating stem cells while limiting ectopic The importance of maintaining the health of secre-
crypt formation15. IESCs further rely on signalling through both the WNT–β‑catenin
tory IECs is reflected in human IBD and models of
and the Notch pathways for promoting self-renewal and directing differentiation
towards secretory versus non-secretory lineage IEC fates16.
murine intestinal inflammation, in which genetic
The responsiveness of epithelial progenitors to external regulation in settings of defects in autophagy and the unfolded protein response
inflammation or infection remains less well understood. In particular, how immune (UPR) disrupt the function of Paneth and goblet cells
system-mediated signalling integrates into the homeostatic pathways described and promote disease susceptibility 31–37. Autophagy
above or acts through alternative pathways for altering stem cell function is poorly in IECs has been shown to act in an innate immune
defined. However, several recent studies have given insight into the regulation of capacity to limit the dissemination of invasive bacteria
WNT–β‑catenin signalling by the pro-inflammatory cytokines interferon‑γ and tumour passing through the epithelium38, but it also supports
necrosis factor, offering an example of how immune signalling and homeostatic the packaging and exocytosis of Paneth cell granules33.
pathways for regulating the stem cell niche can converge186,187. Furthermore, When autophagy is disrupted in mice they become sus-
cell-intrinsic mechanisms of integrating host–commensal microorganism interactions
ceptible to a form of experimental colitis33. Notably, this
into IEC homeostasis have been recently described188.
susceptibility is dependent on exposure to a common
strain of an enteric virus (murine norovirus), providing
cell-derived products, such as trefoil factor 3 (TFF3) and an example of the compound genetic and environmen-
resistin-like molecule‑β (RELMβ), further contribute to tal interactions that contribute to disease pathogene-
the regulation of a physical barrier in the intestine. TFF3 sis36. Disruption of UPR genes results in endoplasmic
Autophagy
A cellular process by which
provides structural integrity to mucus through mucin reticulum stress in secretory cells and spontaneous
cytoplasmic organelles and crosslinking and acts as a signal that promotes epithe- intestinal inflammation34. Notably, disruption of either
macromolecular complexes lial repair, migration of IECs and resistance to apopto- autophagy or the UPR leads to the compensatory
are engulfed by double sis22,23. RELMβ functions to promote MUC2 secretion, engagement of the other, supporting a model in which
membrane-bound vesicles
regulate macrophage and adaptive T cell responses the two are interrelated39. Furthermore, the engage-
for delivery to lysosomes
and subsequent degradation. during inflammation and, in the setting of nematode ment of these pathways by Paneth cells is required for
This process is involved in infection, directly inhibit parasite chemotaxis24,25. maintaining intestinal homeostasis in mice, and their
constitutive turnover of Intestinal barrier function is further reinforced by combined absence leads to the development of a spon-
proteins and organelles and is the secretion of AMPs by IECs. Enterocytes are capa- taneous disease resembling human Crohn’s disease39.
central to cellular activities that
maintain a balance between
ble of producing some AMPs, such as the C‑type lec- These findings, coupled with genetic evidence from
the synthesis and breakdown tin regenerating islet-derived protein IIIγ (REGIIIγ), patients with IBD for the role of autophagy and the
of various proteins. throughout the small intestine and colon. By contrast, UPR in disease pathogenesis31,32,34,37, support an impor-
Paneth cells are uniquely adapted for the secretion of tant link between the disruption of Paneth cell function
Unfolded protein response
many additional AMPs, including defensins (cryptdins and the potential origins of intestinal inflammation.
(UPR). A response that
increases the ability of the in mice), cathelicidins and lysozyme, in the crypts of Finally, IECs directly transport secretory immu-
endoplasmic reticulum to the small intestine18,26. These AMPs disrupt highly con- noglobulins across the epithelial barrier. Following
fold and translocate proteins, served and essential features of bacterial biology, such as their production by plasma cells in the lamina propria,
decreases the synthesis of surface membranes, which are targeted by pore-forming dimeric IgA complexes are bound by the polymeric
proteins, causes the arrest of
the cell cycle and promotes
defensins and cathelicidins, and Gram-positive cell wall immunoglobulin receptor (pIgR) on the basolateral
apoptosis. peptidoglycans, which are targeted by C‑type lectins18,27. membrane of IECs and actively transcytosed into
This strategy enables the broad regulation of both com- the intestinal lumen 40. The collaboration between
Plasma cells mensal and pathogenic bacteria and limits resistance IgA-secreting B cells and IECs provides an adaptive
Terminally differentiated cells
of bacteria to antimicrobial responses. Regional vari- immune component to the epithelial barrier that reg-
of the B cell lineage that
secrete large amounts of ation in AMP production exists along the longitudinal ulates commensal bacterial populations to maintain
antibodies. axis of the intestinal tract 28. Although further analysis IEC and immune cell homeostasis41–43. Future studies
is required, this distribution may reflect anatomically to better understand how mucus, AMP and secretory
Lamina propria restricted host–commensal bacteria interactions that immunoglobulin dynamics can be regulated to sup-
Connective tissue that
underlies the epithelium of the
drive the differential regulation of IEC responses or port barrier function will enable the development of
mucosa and contains stromal serve to shape heterogeneity in the composition and therapeutic interventions for preserving intestinal
and haematopoietic cells. localization of microbial communities. homeostasis.

NATURE REVIEWS | IMMUNOLOGY VOLUME 14 | MARCH 2014 | 143

© 2014 Macmillan Publishers Limited. All rights reserved


R E V IE W S

Box 2 | IEC tight junctions and turnover


In addition to luminal antigen sampling by IECs,
subepithelial mononuclear phagocytes, through inter-
Below the mucous layers, intestinal epithelial cells (IECs) form a continuous physical actions with IECs, sample luminal contents through
barrier. Tight junctions connect adjacent IECs and are associated with cytoplasmic transepithelial dendrites 48,49 (discussed below). The
actin and myosin networks that regulate intestinal permeability. In the setting of adaptation of the epithelial barrier for the sampling of
inflammatory bowel disease (IBD), dysregulation of these interactions, mediated by
luminal contents accommodates limited and controlled
tumour necrosis factor signalling and by myosin light chain kinase activity, leads
to IEC cytoskeletal rearrangements that disrupt tight junctions and increase
bacterial and antigen translocation to direct appropriate
permeability189,190. These findings suggest that IEC tight junctions could be important tolerogenic or anti-pathogen responses. The influence
targets for enhancing the integrity of the intestinal barrier in IBD. of these transport pathways on the immune response is
As the IEC barrier is continuously renewed, the turnover of IECs provides an additional not well understood, but distinct pathways of acquiring
challenge to the maintenance of epithelial continuity. Recent studies have described antigens may influence the context in which immune
pathways by which adjacent cells seal potential voids created during the extrusion of cells interpret microbial signals50. Furthermore, trans-
either apoptotic or live cells from the single-cell layer191,192. As dysregulated epithelial port through these pathways may alter bacteria and
cell turnover and apoptosis are associated with intestinal inflammation, the contribution antigens to enable controlled transport of antigens to
of these mechanisms to the limiting of barrier breaches and further inflammation is of be differentiated from dysregulated bacterial transloca-
relevance to our understanding of epithelial cells as an efficient physical barrier.
tion50. Harnessing the functions of IECs in this sampling
Although increased intestinal permeability has been correlated with IBD1,193,194,
it remains unclear whether the loss of barrier function is a cause or a consequence of
process holds promise for the development of mucosal
intestinal inflammation in human disease. Evidence from mouse models with genetic vaccines and the regulation of intestinal inflammation.
defects in tight-junction-associated proteins suggests that disruption of barrier
function alone is not always sufficient to cause disease195,196. Notably, in mice with a IECs — sentinels in intestinal homeostasis
deletion of the tight-junction protein junctional adhesion molecule A, the secretion Central to the capacity of IECs to maintain barrier
of commensal bacteria-specific IgA can compensate for the loss of barrier function and immunoregulatory functions is their ability to act
and limit disease severity following chemically induced colitis195. Thus, compensatory as frontline sensors for microbial encounters and to
immune mechanisms can act to protect against the development of colitis, even in the integrate commensal bacteria-derived signals into anti­
setting of barrier disruption, supporting a multi-hit model of disease susceptibility195. microbial and immunoregulatory responses (FIG. 2). IECs
express pattern-recognition receptors (PRRs) that enable
them to act as dynamic sensors of the microbial envi-
Sampling of luminal contents by IECs. Despite the bar- ronment and as active participants in the directing of
rier function supported by IECs (BOX 2), the intestinal mucosal immune cell responses (see Supplementary
epithelium includes specialized adaptations that conflict information S1 (table)). Members of the Toll-like recep-
with the concept of complete segregation between host tor (TLR)51, NOD-like receptor (NLR)52,53 and RIG‑I‑like
immune cells and microorganisms. Specialized IECs, receptor (RLR) families54,55 provide distinct pathways for
called microfold cells (M cells), mediate the sampling the recognition of microbial ligands or endogenous sig-
Peyer’s patches of luminal antigens and intact microorganisms for pres- nals associated with pathogenesis. Unlike sterile sites in
Groups of lymphoid aggregates entation to the underlying mucosal immune system44. the body where recognition of foreign microorganisms
located in the submucosa of
These specialized IECs are concentrated in the follicle- initiates highly inflammatory cascades, the abundance
the small intestine that contain
many immune cells, including associated epithelium overlaying the luminal surface of of symbiotic commensals in the intestine necessitates
B cells, T cells and dendritic intestinal lymphoid structures, including Peyer’s patches that IECs maintain a state of altered responsiveness
cells. They have a luminal and isolated lymphoid follicles44,45. (discussed below). Although the study of PRR pathways
barrier consisting of specialized Although nonspecific uptake and transcytosis of anti- in haematopoietic cells has mostly focused on their
epithelial cells, called microfold
cells, which sample the lumen
gens represents a well-established mechanism of sampling pro-inflammatory properties in antigen-presenting and
and transport antigens. by M cells, it has recently been demonstrated that more effector immune cell populations, their role in regulating
efficient mechanisms of receptor-mediated transport also tissue homeostasis and immune tolerance has emerged
Pattern-recognition exist. The surface glycoprotein GP2 acts as a receptor for as a major component of their function in IECs (see
receptors
the bacterial pilus protein FimH, and the M cell-mediated Supplementary information S1 (table)).
(PRRs). Receptors that
recognize structures shared transport of the pathogen Salmonella enterica across the
by foreign microorganisms epithelial barrier depends on GP2–FimH interaction46. The homeostatic role of microbial recognition by IECs.
or endogenous molecules This suggests that M cells are capable of both specific Evidence of a role for PRRs in the protection against
associated with pathogenesis. receptor-mediated microbial uptake and nonspecific intestinal inflammation and repair of epithelial dam-
Signalling through these
receptors promotes
antigen uptake from the intestinal lumen. Although the age emerged from studies of mice deficient in TLRs
tissue-specific innate immune active transport of luminal contents across the epithelial and signalling adaptors or depleted of key commensal
responses including the barrier was thought to be a unique function of M cells, it microorganisms56. Landmark work by Medzhitov and
production of cytokines. was recently shown that small-intestinal goblet cells also colleagues demonstrated, through the use of TLR- and
contribute to this process through the delivery of soluble MYD88‑deficient and broad-spectrum antibiotic-treated
Toll-like receptor
(TLR). An evolutionarily luminal antigens to subepithelial dendritic cells (DCs)47. mice, that commensal bacteria-derived signals contrib-
conserved pattern-recognition Although both M cells and goblet cells seem to be capable ute to epithelial homeostasis and repair in a model of
receptor located at the cell of antigen delivery to the lamina propria, the functional chemically induced colitis using dextran sodium sulphate
surface or at intracellular importance and contribution of these two pathways to (DSS)56. This and other studies defined beneficial roles of
membranes. The natural
ligands of TLRs are conserved
the development of anti-pathogen responses or to the IEC-intrinsic TLR signalling that include the expression
molecular structures found in maintenance of immune tolerance remains incompletely of cytoprotective heat-shock proteins, epidermal growth
bacteria, viruses and fungi. understood. factor receptor ligands56,57, and TFF3 (REF. 58), and the

144 | MARCH 2014 | VOLUME 14 www.nature.com/reviews/immunol

© 2014 Macmillan Publishers Limited. All rights reserved


F O C U S O N H o m e os tat i c I m m un e R e sRpEons eS
V IE W s

a Commensal bacteria
b AMPs Mucin

TFF3

TLR9
(apical) ROS

Endosome TLR3, Tight


TLR7, junction
TLR8

MYD88 TRIF
Ub Heat-shock
Ub proteins
Ub
Iκ B IκB IKKγ
IKKα IKKβ
p50 p65
NLRP3,
NLRP6, NF-κB
NLRC4

Pro-caspase 1 NF-κB NF-κB


p50 p65 p50 p65

NOD1,
Inflammasome NOD2
RIP2
FRMPD2

IL-1β EGFR
and IL-18 TLR2, TLR4, APRIL, BAFF, IL-25,
TLR5 or TLR9 retinoic acid, TGFβ EGFR ligands
and TSLP

Figure 2 | Microbial recognition promotes IEC health and function.  a | Pattern-recognition receptors (PRRs),
including intestinal epithelial cell (IEC)-expressed Toll-like receptors (TLRs) and NOD-like receptors (NLRs), recognize
NOD-like receptor conserved microbial-associated molecular motifs and pathogen-specific virulence properties. Nature
TLRsReviews | Immunology
recruit the signalling
(NLR). A pattern-recognition
adaptors MYD88 and TIR-domain-containing adaptor protein inducing interferon‑β (TRIF) on ligation to signal molecules
receptor located in the cytosol.
NLRs recognize a wide range of
via nuclear factor‑κB (NF‑κB), p50 and p65 subunit activation and the mitogen-activated protein kinase (MAPK) pathway
foreign structures and patterns (not shown). Nucleotide-binding oligomerization domain 1 (NOD1) and NOD2 signal through receptor-interacting
associated with pathogenesis. protein 2 (RIP2) to activate NF‑κB and MAPKs, whereas other IEC-expressed NLRs, including NOD-, LRR- and pyrin
Some members of this family domain-containing 3 (NLRP3), NLRP6 and NOD-, LRR- and CARD-containing 4 (NLRC4), form inflammasome complexes
form multiprotein complexes with pro-caspase 1 for the cleavage and activation of interleukin‑1β (IL‑1β) and IL‑18. Polarized expression of PRRs by IECs
known as inflammasomes, at either the apical or basolateral membrane may contribute to the discrimination between commensal and pathogen
which regulate the processing microbial signals. For example, signalling through surface or endosomal TLR9 at the apical pole of IECs promotes the
and secretion of inhibition of NF-κB signalling, whereas TLR signalling from the basolateral pole promotes NF-κB activation. b | Microbial
pro-inflammatory cytokines.
recognition is integrated by IECs. This promotes cell survival and repair (mediated by trefoil factor 3 (TFF3), heat-shock
RIG‑I‑like receptor
proteins and epidermal growth factor receptor (EGFR) ligand expression), barrier function (mediated by increased mucin
(RLR). A pattern-recognition and antimicrobial peptide (AMP) producton) and immunoregulatory responses (mediated by a proliferation-inducing
receptor located in the cytosol ligand (APRIL), B cell-activating factor (BAFF), IL-25, retinoic acid, transforming growth factor‑β (TGFβ) and thymic
that responds to viral RNA. stromal lymphopoietin (TSLP)), FRMPD2, FERM and PDZ domain-containing 2; IκB, inhibitor of NF‑κB; IKK, IκB kinase;
ROS, reactive oxygen species; Ub, ubiquitin.
Dextran sodium sulphate
(DSS). A large polysaccharide
that causes epithelial injury
and inflammation in the enhanced integrity of apical tight-junction complexes59 (IκB) kinase (IKK) complex or NF‑κB essential modula-
intestinal tract and is (FIG. 2). Furthermore, IEC-specific deletion of elements tor (NEMO), results in enhanced DSS-induced or spon-
commonly used in models of
experimentally induced colitis
necessary for the activation of the transcription factor taneous colitis60,61. These studies establish an essential
for studying the response to complex nuclear factor-κB (NF‑κB) downstream of TLR role for TLRs, in addition to other NF‑κB signalling
intestinal injury. signalling in mice, including the inhibitor of NF‑κB pathways, in epithelial homeostasis and repair.

NATURE REVIEWS | IMMUNOLOGY VOLUME 14 | MARCH 2014 | 145

© 2014 Macmillan Publishers Limited. All rights reserved


R E V IE W S

As additional families of TLRs, such as the NLRs and delicate nature of the balance that exists between homeo-
RLRs, have been ascribed roles in regulating inflamma- stasis and inflammation and its importance in maintain-
tory immune cell responses, they have also been shown ing healthy host–microorganism symbiosis.
to be important in IECs for the regulation of intestinal
homeostasis52–55. The identification of nucleotide-bind- Specialized regulation of PRR pathways in IECs. The
ing oligomerization domain 2 (NOD2), an NLR family proximity of IECs to an abundance of luminal microbial
member that recognizes bacterial muramyl dipeptide signals necessitates specialized mechanisms for maintain-
(MDP), as the first genetic susceptibility locus for ing altered or hyporesponsive PRR signalling in response
Crohn’s disease has fuelled interest in the role of this to commensal bacteria-dependent stimuli83,84. In support
PRR and the related protein NOD1 in both immune of this, IECs express negative regulators of PRR-dependent
cells and the intestinal epithelium37,62,63. Moreover, pro-inflammatory signalling 75,83,85,86 (see Supplementary
inflammasomes formed by caspase 1 and NLRs, includ- information S1 (table)). The disruption of these regula-
ing IEC-expressed NOD-, LRR- and pyrin domain- tory pathways or constitutive activation of NF‑κB pre-
containing 3 (NLRP3), NLRP6, NLRP12 and NOD-, dispose mice to dysregulated epithelial homeostasis and
LRR- and CARD-containing 4 (NLRC4), have a complex exaggerated inflammation72,75,85,87. Furthermore, it has
influence over inflammation and epithelial repair, as dem- been appreciated that commensal bacteria-dependent
onstrated by both pathological and protective roles in con- production of ROS by IECs can attenuate the activation
stitutive knockout mouse models52,53,64 (see Supplementary of NF‑κB, broadly tolerizing IECs to microbial stimula-
information S1 (table)). As NLRs are expressed by sev- tion through PRR signalling 88,89. Although additional
eral cell populations in the intestine, conditional knock- mechanisms exist for the negative regulation of PRR sig-
out models will be required to elucidate the precise nalling pathways90, in most cases the extent to which they
haematopoietic, epithelial and stromal contributions of are active in IECs and their contributions to intestinal
these PRRs during inflammation and repair. homeostasis remain to be determined.
Finally, reactive oxygen species (ROS) produced in In addition to maintaining the hyporesponsiveness
response to commensal or pathogenic bacteria have of IECs, innate immune pathways must differentiate
a role in IEC-intrinsic signalling that acts to promote between signals derived from commensal and patho-
epithelial repair, independently of their microbicidal genic microorganisms for the scaling of an appropriate
effects65,66. Through the inactivation of cellular redox- inflammatory response91. The polarized nature of the
sensitive tyrosine phosphatases, ROS promote the form­ intestinal epithelium allows for the anatomical segrega-
ation by IECs of focal adhesions, which are necessary tion of PRRs (FIG. 2). In vitro and in vivo models demon-
for cell migration and wound healing 65,66. Strikingly, strate differential responsiveness of IECs to apical versus
Nuclear factor-κB
these findings show remarkable symmetry with studies basolateral stimulation with multiple TLR ligands92–94.
(NF‑κB). A family of in Drosophila melanogaster, in which ROS also promote For example, although basolateral exposure of IECs to
transcription factors important epithelial homeostasis, suggesting an evolutionarily TLR9 ligands results in canonical activation and nuclear
for pro-inflammatory and conserved role for ROS in mediating protective effects translocation of NF‑κB, apical exposure results in a
anti-apoptotic responses that
of commensal microorganism-dependent cellular net inhibitory effect through the stabilization of IκB94.
are activated by the ubiquitin-
dependent degradation of responses67–69. This apical signal induces tolerance to subsequent TLR
their respective inhibitors, The protective effects of microbial recognition by stimulation, demonstrating a unique adaptation for the
members of the inhibitor of IECs may come at a cost. Although commensal micro- cross-tolerance of microbial recognition pathways and
NF‑κB (IκB) family. This process bial signals are protective in settings of tissue damage or a differential response to microbial signals based on
is mediated by the kinases, IκB
kinase 1 (IKK1) and IKK2.
infection, they can drive tumorigenesis and cancer when anatomical location94 (FIG. 2).
homeostatic responses become dysregulated60,70. Epithelial This concept of subcellular segregation and polarized
Inflammasomes cell-intrinsic TLR, MYD88 and NF‑κB signalling have all distribution of TLRs has been translated to the regula-
Multiprotein complexes that been implicated in promoting tumour development and tion of additional PRR pathways95,96. Through a series
contain a member of the
progression in multiple genetic70–73 and inflammation- of elegant genetic screens, FERM and PDZ domain-
NOD-like receptor family,
adaptor proteins and the induced60,72,74,75 mouse models of colorectal cancer. The containing 2 (FRMPD2) — which is a positive regula-
protease caspase 1. These convergence between PRR signalling and pro-oncogenic tor of NOD2‑mediated NF‑κB activation in response to
complexes regulate the signalling pathways could partly explain the tumorigenic MDP recognition — was recently identified to act as a
catalytic processing and effects of microbial stimulation. The stabilization of key scaffold protein that promotes basolateral membrane
secretion of pro-inflammatory
cytokines, including
oncogenic proteins, such as MYC, has been shown to be localization and selective basolateral activation through
interleukin‑1β (IL‑1β) and promoted by MYD88 signalling 71. Furthermore, NF‑κB interactions with the leucine-rich repeat (LRR) domain
IL‑18. can enhance WNT signalling in terminally differentiated of NOD2 (REF. 96) (FIG. 2). Common Crohn’s disease-
IECs to promote their dedifferentiation into stem cell‑like associated variants of NOD2 contain mutations in this
Reactive oxygen species
tumour initiators73. LRR domain. These NOD2-mutant proteins were shown
(ROS). Chemically reactive
molecules containing oxygen Paradoxically, some NLR signalling pathways protect in vitro to lack the ability to interact with FRMPD2, to
that, when produced in against tumorigenesis, partly through the regulation of colocalize at the basolateral membrane of epithelial cells
large amounts, have cell death and proliferation in damaged or transformed and to respond to stimulation with NOD2 ligands62,63,96.
pro-inflammatory and IECs76–80 and through the regulation of tissue repair These studies give insight into the mechanism of NOD2
antimicrobial effects.
Physiological levels of ROS
responses mediated by interleukin‑18 (IL‑18) signal- dysfunction associated with IBD and how IECs may
have been shown to regulate ling 53,81,82. The complexity of the multiple roles of micro- spatially regulate the activation of PRR signals at the
cellular signalling pathways. bial recognition by IECs serves to further highlight the intestinal barrier.

146 | MARCH 2014 | VOLUME 14 www.nature.com/reviews/immunol

© 2014 Macmillan Publishers Limited. All rights reserved


F O C U S O N H o m e os tat i c I m m un e R e sRpEons eS
V IE W s

Finally, mechanisms by which IECs may break their a proliferation-inducing ligand (APRIL; also known
relative tolerance to microbial signals in settings of path- as TNFSF13) and B cell-activating factor (BAFF; also
ogen infection are poorly defined. In contrast to sterile known as TNFSF13B)108,109 by IECs is promoted by com-
sites in the body, control of inflammation in the intes- mensal bacteria via PRR signalling (FIG. 2). We discuss
tine may be more adapted to relying on the recognition below the immunoregulatory functions of IECs, describ-
of ‘danger’ signals associated with pathogenesis, rather ing their contribution to the priming of adaptive immune
than on the presence of microbial signals alone97. The cell responses, regulation of innate effector responses and
recognition of danger has been proposed to be medi- homeostasis of adaptive immune cell function in the
ated through the detection of properties associated with intestinal environment.
microbial viability, termed viability-associated PAMPs
(vita-PAMPs), that distinguish living pathogens from Mononuclear phagocytes and antigen presentation. IECs
inert microbial debris, as well as through the detection exert their influence over the priming of both cellular and
of conserved virulence factors of pathogens, such as bac- humoral adaptive immune responses via a continuous
terial secretion systems and toxins that penetrate into the dialogue with antigen-presenting mononuclear phago-
cellular cytosol91,98. Although these mechanisms for scal- cytes (FIG. 3). IEC-derived TSLP, TGFβ and retinoic acid,
ing microbial threats have been studied and identified in produced in response to commensal bacteria-derived sig-
phagocytes and antigen-presenting cells, their function nals, promote the development of DCs and macrophages
and relevance in IECs are less well understood. with tolerogenic properties, including the production
of IL‑10 and retinoic acid103,104,110. Considerable hetero­
Commensal microorganism-dependent regulation of geneity exists among intestinal mononuclear phagocytes,
barrier function. In addition to the homeostatic role of the classification of which has been previously compli-
microbial recognition by IECs, the intestinal epithelium cated by conflicting nomenclature, as well as phenotypical
acts as an essential integrator of environmental signals for and functional plasticity in settings of inflammation111.
the regulation of microbial colonization, barrier function However, two distinct populations that have been
and mucosal immune responses. As previously discussed, characterized are the pre‑DC‑derived CD11c+CD103+
the production of an apical mucous layer, the secretion DCs and monocyte-derived CD11clowF4/80+CX3CR1hi
of broadly targeted AMPs and the transcytosis of secre- intestine‑resident macrophages112–115.
tory IgA contribute to epithelial barrier function. Reduced CD103+ DCs act as migratory antigen-presenting cells
mucous layer thickness in germ-free mice can be reversed and upon activation traffic to secondary lymphoid tis-
by treatment with TLR ligands, indicating that commen- sues, including the mesenteric lymph nodes and Peyer’s
sal bacteria-dependent signals regulate mucus produc- patches, carrying with them antigenic material and live
tion by goblet cells17. Similarly, the expression of many bacteria for presentation to adaptive immune cells116,117.
epithelial cell-derived AMPs is enhanced by, or dependent Influenced by their previous interactions with IECs at the
on, the presence of commensal microbial signals18,29,99–101. intestinal barrier, these migratory DCs promote immune
As cells with specialized antimicrobial function, Paneth tolerance through the differentiation of forkhead box P3
cells play a particularly important part in the regulation (FOXP3+) regulatory T cells by a TGFβ- and retinoic
of AMP production through cell-intrinsic expression of acid‑dependent mechanism116,118,119. Furthermore, the
MYD88 and NOD2 (REFS 100,101). production of retinoic acid by IEC-conditioned CD103+
The transport of IgA across the epithelial barrier DCs is responsible for the imprinting of gut-homing
is regulated, in part, by the expression of pIgR on the properties on T cells, allowing for the targeting of recircu-
basolateral membrane of IECs, which is promoted by lating mature cells to the original site of antigen encoun-
MYD88- and NF‑κB‑dependent signalling in response ter in the intestinal lamina propria120–124. Thus, in addition
to commensal microbial signals40,102. Finally, the integ- to promoting naive T cell maturation based on antigen
rity of tight junctions and transepithelial permeability specificity, CD103+ DCs relay the original context of
are regulated by commensal microbial signals, including antigenic encounter at the intestinal epithelial barrier.
TLR2‑dependent redistribution of the tight-junction pro- In contrast to CD103+ DCs, CX3CR1hi intestine-
teins to apical cell–cell contacts59. Thus, the ability of IECs resident macrophages lack migratory properties in the
to sense their microbial surroundings has an integral role steady state and instead persist in close physical contact
in regulating their barrier function. with IECs, where they act as avid phagocytes to medi-
ate clearance of pathogens and commensal bacteria
Viability-associated PAMPs Regulation of immune cells by IECs that traverse the epithelial barrier 116,125. Their expres-
(Vita-PAMPs). Members of a IECs produce numerous immunoregulatory signals that sion of tight-junction proteins allows the formation of
special class of pathogen- are necessary for tolerizing immune cells, limiting steady- trans­epithelial dendrites that penetrate into the lumen
associated molecular patterns
state inflammation and directing appropriate innate and of the intestine for sampling of exogenous antigens48,125.
recognized by the innate
immune system to signify adaptive immune cell responses against pathogens and Reflecting the functional dependence of these cells on
microbial life. These patterns commensal bacteria. Many of these responses depend the epithelium, the extension of these trans­epithelial
differentiate dead and living on the translation of commensal bacteria-derived sig- dendrites is initiated by TLR signalling, not in myeloid
microorganisms to allow for nals by IECs to mucosal immune cells. The produc- cells themselves, but in IECs49. The CX3CR1hi intestine‑
scaling of appropriate immune
responses based on the level of
tion of the cytokines thymic stromal lymphopoietin resident macrophage population has also been impli-
threat the microbial signals (TSLP)103–105, transforming growth factor‑β (TGFβ)104,106 cated in the maintenance of mucosal tolerance, as they
represents. and IL‑25 (REF. 107) and the B cell-stimulating factors have been shown to promote the survival and local

NATURE REVIEWS | IMMUNOLOGY VOLUME 14 | MARCH 2014 | 147

© 2014 Macmillan Publishers Limited. All rights reserved


R E V IE W S

Innate immune regulation Adaptive immune regulation sIgA


Commensal
bacterium TLA SEMA7A

IL-25, TSLP
TSLP,
IL-33, IL-25 TGFβ, APRIL,
IFNγ, IEL IL-7,
TSLP RA BAFF
TNF IL-15
IL-13, IL-1β,
amphiregulin IL-17, IL-23 IL-12
IL-22
IL-10

IgA+ plasma cell


IL-25
ILC2 ILC3 ILC1

TSLP TReg cell


DC
Lamina propria
TCR RA,
MHC TGFβ

Macrophage Monocyte
Naive T cell TReg cell
Basophil Type 2
progenitor MPP
IL-10, Direct IEC effect
Mast cell Basophil B cell RA,
TGFβ Indirect IEC effect
Peyer’s patch Immune response
or mesenteric Differentiation
lymph node Basophil

Figure 3 | IECs regulate innate and adaptive immunity.  Intestinal epithelial cell (IEC)-derived cytokines interleukin‑25
Nature progenitors
(IL‑25) and thymic stromal lymphopoietin (TSLP) elicit the expansion and differentiation of basophil Reviews | Immunology
and
multipotent progenitor type 2 (type 2 MPP) cells, respectively. IL‑25, IL‑33 and TSLP stimulate group 2 innate lymphoid
cells (ILC2s), whereas IL‑25 suppresses innate lymphoid cell subset 1 (ILC1) and ILC3 function by limiting macrophage
production of pro-inflammatory cytokines IL‑1β, IL‑12 and IL‑23. IECs condition dendritic cells (DCs) and macrophages
towards a tolerogenic phenotype through the production of TSLP, transforming growth factor‑β (TGFβ) and retinoic acid
(RA). These DCs promote the differentiation of naive CD4+ T cells into regulatory T (TReg) cells and the maturation of B cells
into IgA-secreting plasma cells. Mucosal cell-derived DCs also imprint a gut-homing phenotype on primed B cells and
T cells through the production of RA. After trafficking to the intestine, TReg cells are expanded in number by macrophages
that are conditioned to produce IL‑10 by TSLP-mediated stimulation and through contact-dependent interactions with
IEC-expressed semaphorin 7A (SEMA7A). The production of a proliferation-inducing ligand (APRIL) and B cell-activating
factor (BAFF) by IECs and by TSLP-stimulated macrophages and DCs promotes class-switch recombination and
the production of IgA by B cells in the intestinal lamina propria. IEL, intra-epithelial lymphocyte; IFNγ, interferon‑γ;
sIgA, secretory IgA; TCR, T cell receptor; TLA, thymus leukaemia antigen; TNF, tumour necrosis factor.

expansion of previously primed regulatory T cells 126. and myeloid cell phenotypes that promote the develop-
CX3CR1hi macrophages promote tolerance in the intes- ment of type 2 cytokine responses at mucosal sites130–133.
tinal lamina propria through the production of IL‑10, These cells include a distinct population of basophil pro-
which leads to suppression of inflammatory cytokine genitors and a population of multipotent progenitor cells,
production by colitogenic T cells and promotion of which undergo extramedullary haematopoiesis and rep-
regulatory T cell function127,128. IECs maintain this resent an innate link between IEC-derived signals and the
Innate lymphoid cells tolerogenic function through their production of solu- polarization of TH2 cell immune responses to helminths
(ILCs). A group of innate ble factors, such as TSLP, TGFβ and retinoic acid103,104,110, and allergens132,133.
immune cells that are as well as through contact-dependent interactions
lymphoid in morphology and
involving IEC expression of the integrin ligand sema- Innate lymphocyte function. In addition to the myeloid
developmental origin, but lack
properties of adaptive B cells phorin 7A, which induces IL‑10 expression by CX3CR1hi cell and granulocyte populations, a recently identified
and T cells such as recombined macrophages and promotes intestinal homeostasis129. innate immune cell population of innate lymphoid cells
antigen-specific receptors. IECs also play an important part in the induction of (ILCs) plays a crucial part in intestinal immune homeo-
They function in the T helper 2 (TH2) cell responses during helminth infec- stasis. ILCs lack properties of adaptive lymphocytes, such
regulation of immunity, tissue
homeostasis and inflammation
tion. In this setting, the IEC-derived cytokines TSLP as recombined antigen-specific receptors134. They are
in response to cytokine and IL‑25 promote the expansion and differentiation of found at barrier surfaces, including mouse and human
stimulation. haematopoietic progenitor cells towards mononuclear lung 135, skin136 and intestine137, where they function

148 | MARCH 2014 | VOLUME 14 www.nature.com/reviews/immunol

© 2014 Macmillan Publishers Limited. All rights reserved


F O C U S O N H o m e os tat i c I m m un e R e sRpEons eS
V IE W s

as regulators of tissue homeostasis, inflammation and IECs play an indirect part in the regulation of ILC3s
early innate response to infection. ILCs are regulated, in response to commensal bacteria-derived signals. For
in part, by epithelial cell-derived immunoregulatory example, IEC-derived IL‑25 leads to the suppression of
signals (FIG. 3). ILCs display phenotypical and functional IL‑23 production by macrophages and decreased IL‑22
heterogeneity, which has been reviewed extensively production by ILC3s161. By contrast, commensal bacteria-
elsewhere134,138–140. ILCs are characterized by their devel- dependent signals have also been shown to stimulate the
opmental requirements and differential cytokine expres- production of IL‑7 by IECs162, which supports the pro-
sion into group 1, group 2 and group 3 ILCs, which share duction of IL‑22 by ILC3s through the stabilization of the
functional similarities with the adaptive CD4+ TH1, TH2 transcription factor retinoid-related orphan receptor-γt
and TH17 cell populations, respectively. (RORγt; encoded by RORC)162,163. These seemingly con-
Group 1 ILCs include classical natural killer cells flicting roles for IECs in regulating ILC3 function in
(NK cells) and innate lymphoid cell subset 1 (ILC1) response to commensal bacterial stimulation may be
cells, and are characterized by the production of the explained by heterogeneity among intestinal ILCs and by
TH1 cell‑associated cytokines interferon-γ (IFNγ) and differential targeting of cell types capable of producing
tumour necrosis factor (TNF) in response to IL‑12 pro-inflammatory versus tissue-protective cytokines139.
and/or IL‑15 (REF. 140). Although NK cells can directly Although the function of ILCs has been appreciated
kill target cells through cytotoxic activity, other ILC1s are in numerous mouse models, the importance and relative
limited to cytokine production in response to stimula- contribution of these cells to inflammation in settings
tion. Although these ILC1s have a less well-understood of human disease remain incompletely defined. Future
function than NK cells, several recent reports suggest work in this field will be required to further characterize
a possible role in mediating intestinal inflammation in the heterogeneity and tissue-specific functions of these
murine colitis models and human IBD141,142. cells, elaborate our understanding of their contribu-
Group 2 ILCs (collectively termed ILC2s) produce the tions to human disease and develop means of clinically
TH2 cell‑associated cytokines IL‑5 and IL‑13 (REF. 140). targeting their protective or detrimental functions.
These factors contribute to an early innate response to
intestinal helminth infection and invoke a protective Tissue-resident T cells. Following priming by intestine-
epithelial response, including goblet cell hyperplasia and derived antigen-presenting cells in secondary lymphoid
enhanced mucus secretion143–145. Furthermore, ILC2s pre- tissues, conventional effector T cells recirculate through
sent in the lung promote airway hyperresponsiveness or the body before settling in the intestine, where they exert
tissue repair in mouse models of allergy and influenza their tolerogenic or inflammatory effect on the local
virus infection146–149. This suggests that ILC2s may have environment (FIG. 3). Here, mature T cells are subject to
analogous functions in the intestine, perhaps during food the direct influence of IECs for their functional main-
allergy or wound repair; however, evidence for these roles tenance and survival in the lamina propria. Specialized
has yet to be described. The proliferation and activation cells known as intraepithelial lymphocytes (IELs) exist
of ILC2s is supported by the predominantly epithelial in intimate contact with the IEC layer, and bidirectional
cell-derived cytokines IL‑25, IL‑33 and TSLP143–145,150. interactions between IELs and IECs maintain immune
The contribution of microbial stimulation to these sig- homeostasis at the intestinal barrier 164–166. IELs display an
nals reinforces the idea of the epithelium as an integrator activated phenotype and include conventional T cells, as
of environmental signals for the regulation of immune well as subsets of cells expressing a restricted repertoire
cell function103,104,107. of T cell receptor specificities and specialized properties,
Finally, group 3 ILCs produce TH17 and TH22 cell- including γδ T cells and NKT cells165,167. Recent studies
associated cytokines, including IL‑17A and IL‑22, in have advanced the understanding of the developmental
response to stimulation by IL‑23 (REF. 140). This group origin of these cells and the functions that they have at
includes ILC3s, as well as lymphoid tissue inducer (LTi) the intestinal barrier 165. These include the demonstration
cells, which have a well-established role in secondary that committed CD4+ T cells can undergo transcriptional
lymphoid tissue organogenesis, mediated by interactions reprogramming when they become IELs to develop a
with stromal cells during embryonic development 151. distinct phenotype resembling that of CD8+ cytotoxic
IL‑22 has an important role in protecting the intestinal T cells168. Although the influence of the local environ-
epithelium following injury or infection by bacterial ment in promoting this developmental change has not
pathogens152,153. In addition, ILC3‑derived IL‑22 sup- been explored, the intimate interactions that these cells
ports the anatomical containment of gut-associated have with IECs suggest that epithelial cell-derived signals
lymphoid tissue-resident commensal bacteria and the may promote their maintenance and function.
protection of IESCs in models of graft-versus-host dis- Tissue-resident conventional T cells primed to act
Natural killer cells
ease154–156. These tissue-protective functions of IL‑22 are as rapidly responsive effectors are important during
(NK cells). A subset of innate
lymphoid cells originally balanced by detrimental effects in certain inflammatory on­going inflammation and infection, as well as for the
defined on the basis of their settings and in the initiation of inflammation-induced protection of the mucosal barrier against future chal-
cytolytic activity against cancer 82,156,157. Collectively, these studies demonstrate lenge. This is thought to be particularly important
tumour targets but now the context-dependent nature of IL‑22 function. By con- in CD8+ T cell-dependent memory responses169. As such,
recognized to serve a broader
role in host defence and
trast, ILC3‑derived IL‑17 is thought to have a primarily CD8+ T cells with a tissue-resident memory phenotype
inflammation through the pro-inflammatory effect in the intestine and has been are uniquely enriched among αβ T cells present in the
production of cytokines. implicated in both mouse colitis and human IBD158–160. intestinal IEL compartment of mice and humans169,170.

NATURE REVIEWS | IMMUNOLOGY VOLUME 14 | MARCH 2014 | 149

© 2014 Macmillan Publishers Limited. All rights reserved


R E V IE W S

Class-switch recombination These tissue-resident memory T (TRM) cells interact In the presence of a cognate CD4+ T cell response, T cell
(CSR). The process by which with IECs through CD103 (also known as αEβ7 integ- expression of CD40L acts as a necessary signal for B cell
proliferating B cells rearrange rin), which binds the adhesion molecule E-cadherin on CSR. In the absence of help from T cells, CSR can occur
their DNA to switch from IECs171,172. This may promote retention of these and other through the stimulation of B cells by APRIL and BAFF,
expressing IgM (or another
class of immunoglobulin) to
cells at the intestinal epithelium. and signalling through transmembrane activator and
expressing a different Mouse IECs were recently demonstrated to contrib- CAML interactor (TACI) and BAFF receptor (BAFFR;
immunoglobulin heavy-chain ute to the refinement of the CD8+ TRM cell pool in favour also known as TNFRSF13C)108,109,179,180. This process is
constant region, thereby of high-affinity precursors that allow for a more efficient directly supported by IECs through the production of
producing antibody with
memory response to secondary mucosal challenge173. APRIL and BAFF in response to commensal bacteria-
different effector functions.
This occurs through the contact-dependent selective induced NF‑κB signalling 108,109. Moreover, IECs induce
expansion and survival of high-affinity or high-avidity APRIL and BAFF production by mucosal DCs through
CD8+ T cell populations expressing homodimers of the TSLP signalling, which acts to amplify the effect on B cell
co-receptor subunit CD8α (known as CD8αα+ IELs), stimulation108,109. This pathway is of clinical relevance to
which interact with the IEC-expressed MHC class I‑like the most prevalent human primary immunodeficiencies,
molecule, thymus leukaemia antigen (TLA) 173 . common variable immunodeficiency and IgA deficiency,
Understanding how such memory cell populations are in which a subset of patients have mutations in the gene
maintained is of particular interest in the design of effi- encoding the TACI receptor that lead to defects in CSR
cient vaccines against pathogens that invade mucosal and IgA production181,182.
surfaces. Strategies have been explored for generating
CD8+ TRM cells with protective effects at extra-intestinal Concluding remarks
sites174,175. Through an improved understanding of how Collectively, the studies highlighted in this Review dem-
IELs are maintained within the intestinal epithelium, onstrate the diverse and multifaceted roles that IECs
we can hope to improve vaccine strategies for prevent- have in the continuous maintenance of intestinal home-
ing infections with pathogens such as HIV and enteric ostasis. Through secretory epithelial cell responses and
viruses176. the maintenance of a continuous cell layer, IECs effec-
tively sustain a physical and biochemical barrier between
IgA-secreting plasma cells. The maturation of naive hosts and their environment. As cells forming a uniquely
B cells into mature IgA-secreting plasma cells through adapted barrier surface, IECs actively respond to their
heavy chain class-switch recombination (CSR) depends local environment through regulatory mechanisms that
on priming by mucosal DCs carrying antigen and live earn IECs recognition as central mediators of microbial
bacteria from the intestinal epithelium117,177. Similar to and immune homeostasis in the intestine. As much of
the priming of a T cell mucosal phenotype, these DCs what is understood of IEC function has been derived
are conditioned by IEC-derived signals to promote IgA from studies using mouse models, a future challenge
class switching and a gut-homing phenotype through lies in the translation of this understanding into human
the production of nitric oxide (NO), IL‑10 and retinoic systems and the development of novel therapeutics for
acid, in conjunction with TGFβ signalling 117,124,178 (FIG. 3). targeting the pathways that contribute to human health.

1. Mankertz, J. & Schulzke, J.‑D. Altered permeability 10. Kamada, N., Seo, S.‑U., Chen, G. Y. & Núñez, G. 20. Velcich, A. et al. Colorectal cancer in mice
in inflammatory bowel disease: pathophysiology and Role of the gut microbiota in immunity and genetically deficient in the mucin Muc2. Science 295,
clinical implications. Curr. Opin. Gastroenterol. inflammatory disease. Nature Rev. Immunol. 13, 1726–1729 (2002).
23, 379–383 (2007). 321–335 (2013). 21. Van der Sluis, M. et al. Muc2‑deficient mice
2. Brenchley, J. M. et al. Microbial translocation is a 11. Hooper, L. V., Littman, D. R. & Macpherson, A. J. spontaneously develop colitis, indicating that MUC2 is
cause of systemic immune activation in chronic HIV Interactions between the microbiota and the immune critical for colonic protection. Gastroenterology 131,
infection. Nature Med. 12, 1365–1371 (2006). system. Science 336, 1268–1273 (2012). 117–129 (2006).
3. Sandler, N. G. et al. Host response to translocated 12. Honda, K. & Littman, D. R. The microbiome in 22. Taupin, D. R., Kinoshita, K. & Podolsky, D. K.
microbial products predicts outcomes of patients infectious disease and inflammation. Annu. Rev. Intestinal trefoil factor confers colonic epithelial
with HBV or HCV infection. Gastroenterology 141, Immunol. 30, 759–795 (2012). resistance to apoptosis. Proc. Natl Acad. Sci. USA 97,
1220–1230 (2011). 13. Maynard, C. L., Elson, C. O., Hatton, R. D. & 799–804 (2000).
4. Cani, P. D. et al. Metabolic endotoxemia initiates Weaver, C. T. Reciprocal interactions of the 23. Dignass, A., Lynch-Devaney, K., Kindon, H., Thim, L. &
obesity and insulin resistance. Diabetes 56, intestinal microbiota and immune system. Podolsky, D. K. Trefoil peptides promote epithelial
1761–1772 (2007). Nature 489, 231–241 (2012). migration through a transforming growth factor
5. Amar, J. et al. Intestinal mucosal adherence and 14. Moon, C. & Stappenbeck, T. S. Viral interactions β-independent pathway. J. Clin. Invest. 94, 376–383
translocation of commensal bacteria at the early with the host and microbiota in the intestine. (1994).
onset of type 2 diabetes: molecular mechanisms and Curr. Opin. Immunol. 24, 405–410 (2012). 24. Artis, D. et al. RELMβ/FIZZ2 is a goblet cell-specific
probiotic treatment. EMBO Mol. Med. 3, 559–572 15. Crosnier, C., Stamataki, D. & Lewis, J. Organizing immune-effector molecule in the gastrointestinal tract.
(2011). cell renewal in the intestine: stem cells, signals Proc. Natl Acad. Sci. USA 101, 13596–13600 (2004).
6. Wen, L. et al. Innate immunity and intestinal and combinatorial control. Nature Rev. Genet. 7, 25. Nair, M. G. et al. Goblet cell-derived resistin-like
microbiota in the development of type 1 diabetes. 349–359 (2006). molecule β augments CD4+ T cell production of
Nature 455, 1109–1113 (2008). 16. Van der Flier, L. G. & Clevers, H. Stem cells, IFN‑γ and infection-induced intestinal inflammation.
7. Lee, Y. K., Menezes, J. S., Umesaki, Y. & self‑renewal, and differentiation in the intestinal J. Immunol. 181, 4709–4715 (2008).
Mazmanian, S. K. Colloquium paper: epithelium. Annu. Rev. Physiol. 71, 241–260 (2009). 26. Bevins, C. L. & Salzman, N. H. Paneth cells,
Proinflammatory T‑cell responses to gut 17. Kim, Y. S. & Ho, S. B. Intestinal goblet cells and mucins antimicrobial peptides and maintenance of intestinal
microbiota promote experimental autoimmune in health and disease: recent insights and progress. homeostasis. Nature Rev. Microbiol. 9, 356–368
encephalomyelitis. Proc. Natl Acad. Sci. USA 108, Curr. Gastroenterol. Rep. 12, 319–330 (2010). (2011).
4615–4622 (2010). 18. Gallo, R. L. & Hooper, L. V. Epithelial antimicrobial 27. Mukherjee, S. et al. Antibacterial membrane
8. Berer, K. et al. Commensal microbiota and myelin defence of the skin and intestine. Nature Rev. attack by a pore-forming intestinal C‑type lectin.
autoantigen cooperate to trigger autoimmune Immunol. 12, 503–516 (2012). Nature 505, 103–107 (2014).
demyelination. Nature 479, 538–541 (2011). 19. Johansson, M. E. V. et al. The inner of the two 28. Darmoul, D. & Ouellette, A. J. Positional specificity
9. Wu, H.‑J. et al. Gut-residing segmented filamentous Muc2 mucin-dependent mucus layers in colon is of defensin gene expression reveals Paneth cell
bacteria drive autoimmune arthritis via T helper 17 devoid of bacteria. Proc. Natl Acad. Sci. USA 105, heterogeneity in mouse small intestine.
cells. Immunity 32, 815–827 (2010). 15064–15069 (2008). Am. J. Physiol. 271, G68–G74 (1996).

150 | MARCH 2014 | VOLUME 14 www.nature.com/reviews/immunol

© 2014 Macmillan Publishers Limited. All rights reserved


F O C U S O N H o m e os tat i c I m m un e R e sRpEons eS
V IE W s

29. Vaishnava, S. et al. The antibacterial lectin RegIIIγ 47. McDole, J. R. et al. Goblet cells deliver luminal antigen 65. Swanson, P. A. et al. Enteric commensal bacteria
promotes the spatial segregation of microbiota and to CD103+ dendritic cells in the small intestine. potentiate epithelial restitution via reactive oxygen
host in the intestine. Science 334, 255–258 (2011). Nature 483, 345–349 (2012). species-mediated inactivation of focal adhesion
In this study, the antimicrobial protein REGIIIγ This report shows that goblet cells can transport kinase phosphatases. Proc. Natl Acad. Sci. USA 108,
is identified as being necessary for the physical soluble luminal antigens to lamina propria DCs, 8803–8808 (2011).
separation of commensal bacteria from the surface implicating a pathway by which IECs directly 66. Leoni, G. et al. Annexin A1, formyl peptide receptor,
of the small intestinal epithelium, thus limiting the mediate antigen delivery to immune cells in and NOX1 orchestrate epithelial repair. J. Clin. Invest.
activation of the intestinal immune response. addition to M cell-mediated transport. 123, 443–454 (2012).
30. Meyer-Hoffert, U. et al. Secreted enteric antimicrobial 48. Rescigno, M. et al. Dendritic cells express tight 67. Lee, W.‑J. Bacterial-modulated host immunity and
activity localises to the mucus surface layer. Gut 57, junction proteins and penetrate gut epithelial stem cell activation for gut homeostasis. Genes Dev.
764–771 (2008). monolayers to sample bacteria. Nature Immunol. 2, 23, 2260–2265 (2009).
31. Hampe, J. et al. A genome-wide association scan of 361–367 (2001). 68. Hochmuth, C. E., Biteau, B., Bohmann, D. & Jasper, H.
nonsynonymous SNPs identifies a susceptibility 49. Chieppa, M., Rescigno, M., Huang, A. Y. C. & Redox regulation by Keap1 and Nrf2 controls
variant for Crohn disease in ATG16L1. Nature Genet. Germain, R. N. Dynamic imaging of dendritic cell intestinal stem cell proliferation in Drosophila. Cell
39, 207–211 (2007). extension into the small bowel lumen in response to Stem Cell 8, 188–199 (2011).
32. Rioux, J. D. et al. Genome-wide association study epithelial cell TLR engagement. J. Exp. Med. 203, 69. Jones, R. M. et al. Symbiotic lactobacilli stimulate gut
identifies new susceptibility loci for Crohn disease 2841–2852 (2006). epithelial proliferation via Nox-mediated generation of
and implicates autophagy in disease pathogenesis. This paper demonstrates a role for epithelial reactive oxygen species. EMBO J. 32, 3017–3028
Nature Genet. 39, 596–604 (2007). cell TLR signalling in promoting DC sampling of (2013).
References 31 and 32 report the association antigens from the intestinal lumen. 70. Rakoff-Nahoum, S. & Medzhitov, R. Regulation of
between genetic variants in the gene for the 50. Shan, M. et al. Mucus enhances gut homeostasis and spontaneous intestinal tumorigenesis through the
autophagy protein ATG16L1 and Crohn’s oral tolerance by delivering immunoregulatory signals. adaptor protein MyD88. Science 317, 124–127
disease susceptibility, establishing a Crohn’s Science 342, 447–453 (2013). (2007).
disease-specific genetic link between IBD and 51. Abreu, M. T. Toll-like receptor signalling in the 71. Lee, S. H. et al. ERK activation drives intestinal
the autophagy pathway. intestinal epithelium: how bacterial recognition tumorigenesis in Apcmin/+ mice. Nature Med. 16,
33. Cadwell, K. et al. A key role for autophagy and shapes intestinal function. Nature Rev. Immunol. 10, 665–670 (2010).
the autophagy gene Atg16l1 in mouse and human 131–144 (2010). 72. Vlantis, K. et al. Constitutive IKK2 activation in
intestinal Paneth cells. Nature 456, 259–263 52. Chen, G. Y. & Núñez, G. Inflammasomes in intestinal intestinal epithelial cells induces intestinal tumors
(2008). inflammation and cancer. Gastroenterology 141, in mice. J. Clin. Invest. 121, 2781–2793 (2011).
34. Kaser, A. et al. XBP1 links ER stress to intestinal 1986–1999 (2011). 73. Schwitalla, S. et al. Intestinal tumorigenesis initiated
inflammation and confers genetic risk for human 53. Elinav, E., Henao-Mejia, J. & Flavell, R. A. Integrative by dedifferentiation and acquisition of stem-cell-like
inflammatory bowel disease. Cell 134, 743–756 inflammasome activity in the regulation of intestinal properties. Cell 152, 25–38 (2013).
(2008). mucosal immune responses. Mucosal Immunol. 6, 74. Fukata, M. et al. Toll-Like receptor‑4 promotes the
35. Brandl, K. et al. Enhanced sensitivity to DSS colitis 4–13 (2013). development of colitis-associated colorectal tumors.
caused by a hypomorphic Mbtps1 mutation disrupting 54. Li, X.‑D. et al. Mitochondrial antiviral signaling protein Gastroenterology 133, 1869–1869 (2007).
the ATF6‑driven unfolded protein response. Proc. Natl (MAVS) monitors commensal bacteria and induces an 75. Xiao, H. et al. The Toll–interleukin‑1 receptor
Acad. Sci. USA 106, 3300–3305 (2009). immune response that prevents experimental colitis. member SIGIRR regulates colonic epithelial
36. Cadwell, K. et al. Virus-plus-susceptibility gene Proc. Natl Acad. Sci. USA 108, 17390–17395 homeostasis, inflammation, and tumorigenesis.
interaction determines Crohn’s disease gene Atg16L1 (2011). Immunity 26, 461–475 (2007).
phenotypes in intestine. Cell 141, 1135–1145 (2010). 55. Broquet, A. H., Hirata, Y., McAllister, C. S. & 76. Dupaul-Chicoine, J. et al. Control of intestinal
In this study, the interaction between Kagnoff, M. F. RIG‑I/MDA5/MAVS are required to homeostasis, colitis, and colitis-associated colorectal
environmental exposures and genetic susceptibility signal a protective IFN response in rotavirus-infected cancer by the inflammatory caspases. Immunity 32,
is shown to determine the penetrance of disease in intestinal epithelium. J. Immunol. 186, 1618–1626 367–378 (2010).
mouse models of intestinal inflammation. (2011). 77. Hu, B. et al. Inflammation-induced tumorigenesis
37. Khor, B., Gardet, A. & Xavier, R. J. Genetics and 56. Rakoff-Nahoum, S., Paglino, J., Eslami-Varzaneh, F., in the colon is regulated by caspase‑1 and NLRC4.
pathogenesis of inflammatory bowel disease. Nature Edberg, S. & Medzhitov, R. Recognition of commensal PNAS 107, 21635–21640 (2010).
474, 307–317 (2011). microflora by Toll-like receptors is required for 78. Zaki, M. H. et al. The NOD-like receptor NLRP12
38. Benjamin, J. L., Sumpter, R. Jr, Levine, B. & intestinal homeostasis. Cell 118, 229–241 (2004). attenuates colon inflammation and tumorigenesis.
Hooper, L. V. Intestinal epithelial autophagy is In this study, mice deficient in TLR signalling or Cancer Cell 20, 649–660 (2011).
essential for host defense against invasive bacteria. greatly depleted of commensal bacteria exhibited 79. Normand, S. et al. Nod-like receptor pyrin
Cell Host Microbe 13, 723–734 (2013). increased susceptibility to experimentally domain‑containing protein 6 (NLRP6) controls
39. Adolph, T. E. et al. Paneth cells as a site of origin for induced intestinal inflammation, implicating epithelial self-renewal and colorectal
intestinal inflammation. Nature 503, 272–276 (2013). commensal microorganism-dependent signals carcinogenesis upon injury. Proc. Natl Acad. Sci. USA
This study demonstrates interactions between the in the regulation of intestinal homeostasis and 108, 9601–9606 (2011).
regulation of the UPR and the autophagy pathway response to injury. 80. Allen, I. C. et al. NLRP12 suppresses colon
in Paneth cells and supports a model in which 57. Brandl, K. et al. MyD88 signaling in inflammation and tumorigenesis through the
alterations in these two responses regulate the nonhematopoietic cells protects mice against induced negative regulation of noncanonical NF‑κB signaling.
development of Crohn’s disease. colitis by regulating specific EGF receptor ligands. Immunity 36, 742–754 (2012).
40. Johansen, F.‑E. & Kaetzel, C. S. Regulation of the Proc. Natl Acad. Sci. USA 107, 19967–19972 81. Salcedo, R. et al. MyD88‑mediated signaling
polymeric immunoglobulin receptor and IgA (2010). prevents development of adenocarcinomas of the
transport: new advances in environmental factors 58. Podolsky, D. K., Gerken, G., Eyking, A. & Cario, E. colon: role of interleukin 18. J. Exp. Med. 207,
that stimulate pIgR expression and its role in Colitis-associated variant of TLR2 causes impaired 1625–1636 (2010).
mucosal immunity. Mucosal Immunol. 4, 598–602 mucosal repair because of TFF3 deficiency. 82. Huber, S. et al. IL‑22BP is regulated by the
(2011). Gastroenterology 137, 209–220 (2009). inflammasome and modulates tumorigenesis in
41. Johansen, F.‑E. et al. Absence of epithelial 59. Cario, E., Gerken, G. & Podolsky, D. K. Toll-like the intestine. Nature 491, 259–263 (2012).
immunoglobulin a transport, with increased mucosal receptor 2 enhances ZO‑1‑associated intestinal 83. Otte, J.‑M., Cario, E. & Podolsky, D. K. Mechanisms
leakiness, in polymeric immunoglobulin receptor/ epithelial barrier integrity via protein kinase C. of cross hyporesponsiveness to Toll-like receptor
secretory component–deficient mice. J. Exp. Med. Gastroenterology 127, 224–238 (2004). bacterial ligands in intestinal epithelial cells.
190, 915–922 (1999). 60. Greten, F. R. et al. IKKβ links inflammation and Gastroenterology 126, 1054–1070 (2004).
42. Shulzhenko, N. et al. Crosstalk between B tumorigenesis in a mouse model of colitis-associated 84. Lotz, M. et al. Postnatal acquisition of endotoxin
lymphocytes, microbiota and the intestinal epithelium cancer. Cell 118, 285–296 (2004). tolerance in intestinal epithelial cells. J. Exp. Med.
governs immunity versus metabolism in the gut. This study provides evidence that the NF‑κB 203, 973–984 (2006).
Nature Med. 17, 1585–1593 (2011). signalling pathway in IECs regulates the This study demonstrates the inhibition of TLR
In this study, a compensatory response of IECs development of inflammation-induced cancer signalling in IECs and the acquisition of tolerance
in the absence of adaptive IgA directed against through the regulation of apoptosis. to microbial colonization that occurs shortly after
commensal bacteria is characterized by the 61. Nenci, A. et al. Epithelial NEMO links innate birth to establish host–microbial homeostasis in
engagement of immune pathways and immunity to chronic intestinal inflammation. the intestine.
dysregulation of lipid metabolism. Nature 446, 557–561 (2007). 85. Vereecke, L. et al. Enterocyte-specific A20
43. Suzuki, K. et al. Aberrant expansion of segmented 62. Hugot, J.‑P. et al. Association of NOD2 leucine-rich deficiency sensitizes to tumor necrosis factor-induced
filamentous bacteria in IgA-deficient gut. Proc. Natl repeat variants with susceptibility to Crohn’s disease. toxicity and experimental colitis. J. Exp. Med. 207,
Acad. Sci. USA 101, 1981–1986 (2004). Nature 411, 599–603 (2001). 1513–1523 (2010).
44. Mabbott, N. A., Donaldson, D. S., Ohno, H., 63. Ogura, Y. et al. A frameshift mutation in NOD2 86. Chassin, C. et al. miR‑146a mediates protective
Williams, I. R. & Mahajan, A. Microfold (M) cells: associated with susceptibility to Crohn’s disease. innate immune tolerance in the neonate intestine.
important immunosurveillance posts in the intestinal Nature 411, 603–606 (2001). Cell Host Microbe 8, 358–368 (2010).
epithelium. Mucosal Immunol. 6, 666–677 (2013). References 62 and 63 provide the first examples 87. Guma, M. et al. Constitutive intestinal NF‑κB does not
45. Mowat, A. M. Anatomical basis of tolerance and of genetic variants linking bacterial recognition trigger destructive inflammation unless accompanied
immunity to intestinal antigens. Nature Rev. Immunol. with susceptibility to human IBD. by MAPK activation. J. Exp. Med. 208, 1889–1900
3, 331–341 (2003). 64. Zaki, M. H., Lamkanfi, M. & Kanneganti, T.‑D. (2011).
46. Hase, K. et al. Uptake through glycoprotein 2 The Nlrp3 inflammasome: contributions to 88. Neish, A. S. et al. Prokaryotic regulation of epithelial
of FimH+ bacteria by M cells initiates mucosal intestinal homeostasis. Trends Immunol. 32, responses by inhibition of IκB‑α ubiquitination.
immune response. Nature 462, 226–230 (2009). 171–179 (2011). Science 289, 1560–1563 (2000).

NATURE REVIEWS | IMMUNOLOGY VOLUME 14 | MARCH 2014 | 151

© 2014 Macmillan Publishers Limited. All rights reserved


R E V IE W S

89. Kumar, A. et al. The bacterial fermentation product 109. Xu, W. et al. Epithelial cells trigger frontline 133. Siracusa, M. C. et al. Thymic stromal lymphopoietin-
butyrate influences epithelial signaling via reactive immunoglobulin class switching through a pathway mediated extramedullary hematopoiesis promotes
oxygen species-mediated changes in cullin‑1 regulated by the inhibitor SLPI. Nature Immunol. 8, allergic inflammation. Immunity 39, 1158–1170
neddylation. J. Immunol. 182, 538–546 (2009). 294–303 (2007). (2013).
90. Kondo, T., Kawai, T. & Akira, S. Dissecting negative References 108 and 109 identify IEC production 134. Spits, H. & Cupedo, T. Innate lymphoid cells:
regulation of Toll-like receptor signaling. Trends of APRIL and BAFF in response to microbial emerging insights in development, lineage
Immunol. 33, 449–458 (2012). stimulation as important for the regulation of relationships, and function. Annu. Rev. Immunol. 30,
91. Blander, J. M. & Sander, L. E. Beyond pattern B cell CSR and the mucosal IgA response. 647–675 (2012).
recognition: five immune checkpoints for scaling the 110. Taylor, B. C. et al. TSLP regulates intestinal immunity 135. Monticelli, L. A., Sonnenberg, G. F. & Artis, D.
microbial threat. Nature Rev. Immunol. 12, 215–225 and inflammation in mouse models of helminth Innate lymphoid cells: critical regulators of allergic
(2012). infection and colitis. J. Exp. Med. 206, 655–667 inflammation and tissue repair in the lung.
92. Gewirtz, A. T., Navas, T. A., Lyons, S., Godowski, P. J. & (2009). Curr. Opin. Immunol. 24, 284–289 (2012).
Madara, J. L. Cutting edge: bacterial flagellin activates 111. Pabst, O. & Bernhardt, G. The puzzle of intestinal 136. Kim, B. S. et al. TSLP elicits IL‑33‑independent innate
basolaterally expressed TLR5 to induce epithelial lamina propria dendritic cells and macrophages. lymphoid cell responses to promote skin
proinflammatory gene expression. J. Immunol. 167, Eur. J. Immunol. 40, 2107–2111 (2010). inflammation. Sci. Transl. Med. 5, 170ra16 (2013).
1882–1885 (2001). 112. Bogunovic, M. et al. Origin of the lamina propria 137. Tait Wojno, E. D. & Artis, D. Innate lymphoid cells:
93. Rhee, S. H. et al. Pathophysiological role of Toll-like dendritic cell network. Immunity 31, 513–525 balancing immunity, inflammation, and tissue repair
receptor 5 engagement by bacterial flagellin in (2009). in the intestine. Cell Host Microbe 12, 445–457
colonic inflammation. Proc. Natl Acad. Sci. USA 102, 113. Varol, C. et al. Intestinal lamina propria dendritic cell (2012).
13610–13615 (2005). subsets have different origin and functions. Immunity 138. Walker, J. A., Barlow, J. L. & McKenzie, A. N. J.
94. Lee, J. et al. Maintenance of colonic homeostasis 31, 502–512 (2009). Innate lymphoid cells — how did we miss them?
by distinctive apical TLR9 signalling in intestinal 114. Rivollier, A., He, J., Kole, A., Valatas, V. & Kelsall, B. L. Nature Rev. Immunol. 13, 75–87 (2013).
epithelial cells. Nature Cell Biol. 8, 1327–1336 (2006). Inflammation switches the differentiation program 139. Cherrier, M., Ohnmacht, C., Cording, S. & Eberl, G.
95. Barnich, N., Aguirre, J. E., Reinecker, H.‑C., Xavier, R. of Ly6Chi monocytes from antiinflammatory Development and function of intestinal innate
& Podolsky, D. K. Membrane recruitment of NOD2 macrophages to inflammatory dendritic cells in the lymphoid cells. Curr. Opin. Immunol. 24, 277–283
in intestinal epithelial cells is essential for nuclear colon. J. Exp. Med. 209, 139–155 (2012). (2012).
factor-κB activation in muramyl dipeptide recognition. 115. Zigmond, E. & Jung, S. Intestinal macrophages: well 140. Spits, H. et al. Innate lymphoid cells — a proposal for
J. Cell Biol. 170, 21–26 (2005). educated exceptions from the rule. Trends Immunol. uniform nomenclature. Nature Rev. Immunol. 13,
96. Lipinski, S. et al. RNAi screening identifies mediators 34, 162–168 (2013). 145–149 (2013).
of NOD2 signaling: Implications for spatial specificity 116. Schulz, O. et al. Intestinal CD103+, but not CX3CR1+, 141. Bernink, J. H. et al. Human type 1 innate lymphoid
of MDP recognition. Proc. Natl Acad. Sci. USA 109, antigen sampling cells migrate in lymph and serve cells accumulate in inflamed mucosal tissues.
21426–21431 (2012). classical dendritic cell functions. J. Exp. Med. 206, Nature Immunol. 14, 221–229 (2013).
97. Matzinger, P. The danger model: A renewed sense of 3101–3114 (2009). 142. Fuchs, A. et al. Intraepithelial type 1 innate
self. Science 296, 301–305 (2002). 117. Macpherson, A. J. & Uhr, T. Induction of protective lymphoid cells are a unique subset of IL‑12- and
98. Sander, L. E. et al. Detection of prokaryotic mRNA IgA by intestinal dendritic cells carrying commensal IL‑15‑responsive IFN‑γ‑producing cells. Immunity 38,
signifies microbial viability and promotes immunity. bacteria. Science 303, 1662–1665 (2004). 769–781 (2013).
Nature 474, 385–389 (2011). 118. Coombes, J. L. et al. A functionally specialized 143. Moro, K. et al. Innate production of TH2 cytokines by
In this study, bacterial mRNA is identified as a population of mucosal CD103+ DCs induces Foxp3+ adipose tissue-associated c‑Kit+Sca‑1+ lymphoid cells.
signal that enables the gauging of infectious risk regulatory T cells via a TGF‑β– and retinoic acid– Nature 463, 540–544 (2009).
posed by live versus dead bacteria, exemplifying a dependent mechanism. J. Exp. Med. 204, 144. Neill, D. R. et al. Nuocytes represent a new innate
vita-PAMP. 1757–1764 (2007). effector leukocyte that mediates type‑2 immunity.
99. Hooper, L. V., Stappenbeck, T. S., Hong, C. V. & 119. Sun, C.‑M. et al. Small intestine lamina propria Nature 464, 1367–1370 (2010).
Gordon, J. I. Angiogenins: a new class of microbicidal dendritic cells promote de novo generation of Foxp3 References 143 and 144 identify a population of
proteins involved in innate immunity. Nature Immunol. T reg cells via retinoic acid. J. Exp. Med. 204, innate lymphoid cells that mediate early TH2
4, 269–273 (2003). 1775–1785 (2007). cytokine production in response to the cytokines
100. Kobayashi, K. S. et al. Nod2‑dependent regulation of 120. Mora, J. R. et al. Selective imprinting of gut-homing IL-25 and IL-33 and during intestinal helminth
innate and adaptive immunity in the intestinal tract. T cells by Peyer’s patch dendritic cells. Nature 424, infection, that promotes goblet cell hyperplasia
Science 307, 731–734 (2005). 88–93 (2003). and worm expulsion.
101. Vaishnava, S., Behrendt, C. L., Ismail, A. S., 121. Johansson-Lindbom, B. et al. Functional specialization 145. Price, A. E. et al. Systemically dispersed innate
Eckmann, L. & Hooper, L. V. Paneth cells directly of gut CD103+ dendritic cells in the regulation of IL‑13‑expressing cells in type 2 immunity. Proc. Natl
sense gut commensals and maintain homeostasis at tissue-selective T cell homing. J. Exp. Med. 202, Acad. Sci. USA 107, 11489–11494 (2010).
the intestinal host-microbial interface. Proc. Natl 1063–1073 (2005). 146. Monticelli, L. A. et al. Innate lymphoid cells promote
Acad. Sci. USA 105, 20858–20863 (2008). 122. Iwata, M. et al. Retinoic acid imprints gut-homing lung-tissue homeostasis after infection with influenza
102. Bruno, M. E. C., Frantz, A. L., Rogier, E. W., specificity on T cells. Immunity 21, 527–538 (2004). virus. Nature Immunol. 12, 1045–1054 (2011).
Johansen, F.‑E. & Kaetzel, C. S. Regulation of the 123. Jaensson, E. et al. Small intestinal CD103+ dendritic 147. Chang, Y.‑J. et al. Innate lymphoid cells mediate
polymeric immunoglobulin receptor by the classical cells display unique functional properties that are influenza-induced airway hyper-reactivity
and alternative NF‑κB pathways in intestinal conserved between mice and humans. J. Exp. Med. independently of adaptive immunity. Nature Immunol.
epithelial cells. Mucosal Immunol. 4, 468–478 205, 2139–2149 (2008). 12, 631–638 (2011).
(2011). 124. Mora, J. R. & von Andrian, U. H. Differentiation and 148. Wilhelm, C. et al. An IL‑9 fate reporter demonstrates
103. Rimoldi, M. et al. Intestinal immune homeostasis is homing of IgA-secreting cells. Mucosal Immunol. 1, the induction of an innate IL‑9 response in lung
regulated by the crosstalk between epithelial cells and 96–109 (2008). inflammation. Nature Immunol. 12, 1071–1077
dendritic cells. Nature Immunol. 6, 507–514 (2005). 125. Niess, J. H. et al. CX3CR1‑mediated dendritic cell (2011).
This study describes the conditioning of mucosal access to the intestinal lumen and bacterial clearance. 149. Halim, T. Y. F., Krauß, R. H., Sun, A. C. & Takei, F.
DCs towards a non-inflammatory phenotype by Science 307, 254–258 (2005). Lung natural helper cells are a critical source of Th2
interactions with IECs, representing a mechanism 126. Hadis, U. et al. Intestinal tolerance requires cell-type cytokines in protease allergen-induced
for the indirect IEC-mediated regulation of gut homing and expansion of FoxP3+ regulatory airway inflammation. Immunity 36, 451–463
adaptive immune cell priming. T cells in the lamina propria. Immunity 34, 237–246 (2012).
104. Zeuthen, L. H., Fink, L. N. & Frokiaer, H. Epithelial (2011). 150. Mjösberg, J. et al. The transcription factor GATA3 Is
cells prime the immune response to an array of gut- 127. Murai, M. et al. Interleukin 10 acts on regulatory essential for the function of human type 2 innate
derived commensals towards a tolerogenic phenotype T cells to maintain expression of the transcription lymphoid cells. Immunity 37, 649–659 (2012).
through distinct actions of thymic stromal factor Foxp3 and suppressive function in mice with 151. Mebius, R. E., Rennert, P. & Weissman, I. L.
lymphopoietin and transforming growth factor‑β. colitis. Nature Immunol. 10, 1178–1184 (2009). Developing lymph nodes collect, CD4+CD3− LTβ+ cells
Immunology 123, 197–208 (2008). 128. Kayama, H. et al. Intestinal CX3C chemokine receptor that can differentiate to APC, NK cells, and follicular
105. Zaph, C. et al. Epithelial-cell-intrinsic IKKβ expression 1high (CX3CR1high) myeloid cells prevent T‑cell- cells but not T or B cells. Immunity 7, 493–504
regulates intestinal immune homeostasis. Nature dependent colitis. Proc. Natl Acad. Sci. USA 109, (1997).
446, 552–556 (2007). 5010–5015 (2012). 152. Zenewicz, L. A. et al. Innate and adaptive
This study demonstrates a crucial role for 129. Kang, S. et al. Intestinal epithelial cell-derived interleukin‑22 protects mice from inflammatory
IEC-intrinsic NF-κB signalling and the production semaphorin 7A negatively regulates development of bowel disease. Immunity 29, 947–957 (2008).
of the cytokine TSLPthymic stromal lymphopoietin colitis via αvβ1 integrin. J. Immunol. 188, 1108–1116 153. Sonnenberg, G. F., Monticelli, L. A., Elloso, M. M.,
in regulating intestinal immune responses and (2012). Fouser, L. A. & Artis, D. CD4+ lymphoid tissue-inducer
coordinating anti-helminth immunity. 130. Saenz, S. A. et al. IL25 elicits a multipotent progenitor cells promote innate immunity in the gut. Immunity
106. Atarashi, K. et al. Induction of colonic regulatory T cell population that promotes TH2 cytokine responses. 34, 122–134 (2011).
cells by indigenous Clostridium species. Science 331, Nature 464, 1362–1366 (2010). 154. Sonnenberg, G. F. et al. Innate lymphoid cells
337–341 (2011). 131. Siracusa, M. C. et al. TSLP promotes promote anatomical containment of lymphoid-resident
107. Zaph, C. et al. Commensal-dependent expression of interleukin‑3‑independent basophil haematopoiesis commensal bacteria. Science 336, 1321–1325
IL‑25 regulates the IL‑23–IL‑17 axis in the intestine. and type 2 inflammation. Nature 477, 229–233 (2012).
J. Exp. Med. 205, 2191–2198 (2008). (2011). In this study, innate lymphoid cell-derived IL-22 is
108. He, B. et al. Intestinal bacteria trigger T cell- 132. Saenz, S. A. et al. IL‑25 simultaneously elicits distinct demonstrated to mediate the containment of
independent immunoglobulin A2 class switching by populations of innate lymphoid cells and multipotent specialized intestinal lymphoid tissue-resident
inducing epithelial-cell secretion of the cytokine progenitor type 2 (MPPtype2) cells. J. Exp. Med. 210, commensal bacteria to prevent systemic immune
APRIL. Immunity 26, 812–826 (2007). 1823–1837 (2013). activation.

152 | MARCH 2014 | VOLUME 14 www.nature.com/reviews/immunol

© 2014 Macmillan Publishers Limited. All rights reserved


F O C U S O N ho m e os tat i c i m m un e r e sRpEons eS
V IE W s

155. Hanash, A. M. et al. Interleukin‑22 protects intestinal 170. Sathaliyawala, T. et al. Distribution and 187. Nava, P. et al. Interferon‑γ regulates intestinal
stem cells from immune-mediated tissue damage and compartmentalization of human circulating and epithelial homeostasis through converging β‑catenin
regulates sensitivity to graft versus host disease. tissue-resident memory T cell subsets. Immunity 38, signaling pathways. Immunity 32, 392–402 (2010).
Immunity 37, 339–350 (2012). 187–197 (2013). 188. Alenghat, T. et al. Histone deacetylase 3 coordinates
156. Kirchberger, S. et al. Innate lymphoid cells sustain 171. Schön, M. P. et al. Mucosal T lymphocyte numbers are commensal-bacteria-dependent intestinal
colon cancer through production of interleukin‑22 in selectively reduced in integrin αE (CD103)-deficient homeostasis. Nature 504, 153–157 (2013).
a mouse model. J. Exp. Med. 210, 917–931 (2013). mice. J. Immunol. 162, 6641–6649 (1999). 189. Blair, S. A., Kane, S. V., Clayburgh, D. R. & Turner, J. R.
157. Muñoz, M. et al. Interleukin (IL)-23 mediates Toxoplasma 172. El‑Asady, R. et al. TGF‑β-dependent CD103 Epithelial myosin light chain kinase expression and
gondii–induced immunopathology in the gut via expression by CD8+ T cells promotes selective activity are upregulated in inflammatory bowel
matrixmetalloproteinase‑2 and IL‑22 but independent destruction of the host intestinal epithelium disease. Lab. Invest. 86, 191–201 (2006).
of IL‑17. J. Exp. Med. 206, 3047–3059 (2009). during graft-versus-host disease. J. Exp. Med. 201, 190. Marchiando, A. M. et al. Caveolin‑1–dependent
158. Buonocore, S. et al. Innate lymphoid cells drive 1647–1657 (2005). occludin endocytosis is required for TNF-induced tight
interleukin‑23‑dependent innate intestinal pathology. 173. Huang, Y. et al. Mucosal memory CD8+ T cells are junction regulation in vivo. J. Cell Biol. 189, 111–126
Nature 464, 1371–1375 (2010). selected in the periphery by an MHC class I molecule. (2010).
159. Geremia, A. et al. IL‑23–responsive innate lymphoid Nature Immunol. 12, 1086–1095 (2011). 191. Wang, F. et al. Active deformation of apoptotic
cells are increased in inflammatory bowel disease. 174. Mackay, L. K. et al. Long-lived epithelial immunity by intestinal epithelial cells with adhesion-restricted
J. Exp. Med. 208, 1127–1133 (2011). tissue-resident memory T (TRM) cells in the absence of polarity contributes to apoptotic clearance.
160. Coccia, M. et al. IL‑1β mediates chronic intestinal persisting local antigen presentation. Proc. Natl Acad. Lab. Invest. 91, 462–471 (2011).
inflammation by promoting the accumulation of Sci. USA 109, 7037–7042 (2012). 192. Eisenhoffer, G. T. et al. Crowding induces live cell
IL‑17A secreting innate lymphoid cells and CD4+ 175. Shin, H. & Iwasaki, A. A vaccine strategy that protects extrusion to maintain homeostatic cell numbers
Th17 cells. J. Exp. Med. 209, 1595–1609 (2012). against genital herpes by establishing local memory in epithelia. Nature 484, 546–549 (2012).
161. Sawa, S. et al. RORγt+ innate lymphoid cells regulate T cells. Nature 491, 463–467 (2012). 193. D’Incà, R. et al. Intestinal permeability test as a
intestinal homeostasis by integrating negative signals 176. Hansen, S. G. et al. Profound early control of highly predictor of clinical course in Crohn’s disease.
from the symbiotic microbiota. Nature Immunol. 12, pathogenic SIV by an effector memory T‑cell vaccine. Am. J. Gastroenterol. 94, 2956–2960 (1999).
320–326 (2011). Nature 473, 523–527 (2011). 194. Wyatt, J. & Vogelsang, H. Intestinal permeability
162. Vonarbourg, C. et al. Regulated expression of nuclear 177. Cerutti, A. The regulation of IgA class switching. and the prediction of relapse in Crohn’s disease.
receptor RORγt confers distinct functional fates to NK Nature Rev. Immunol. 8, 421–434 (2008). Lancet 341, 1437 (1993).
cell receptor-expressing RORγt+ innate lymphocytes. 178. Mora, J. R. et al. Generation of gut-homing IgA- 195. Khounlotham, M. et al. Compromised intestinal
Immunity 33, 736–751 (2010). secreting B cells by intestinal dendritic cells. epithelial barrier induces adaptive immune
163. Satoh-Takayama, N. et al. Microbial flora drives Science 314, 1157–1160 (2006). compensation that protects from colitis. Immunity 37,
interleukin 22 production in intestinal NKp46+ cells 179. Litinskiy, M. B. et al. DCs induce CD40‑independent 563–573 (2012).
that provide innate mucosal immune defense. immunoglobulin class switching through BLyS and 196. Su, L. et al. Targeted epithelial tight junction
Immunity 29, 958–970 (2008). APRIL. Nature Immunol. 3, 822–829 (2002). dysfunction causes immune activation and
164. Yu, Q. et al. MyD88‑dependent signaling for IL‑15 180. Castigli, E. et al. TACI and BAFF‑R mediate isotype contributes to development of experimental colitis.
production plays an important role in maintenance switching in B cells. J. Exp. Med. 201, 35–39 (2005). Gastroenterology 136, 551–563 (2009).
of CD8αα TCRαβ and TCRγδ intestinal intraepithelial 181. Salzer, U. et al. Mutations in TNFRSF13B encoding
lymphocytes. J. Immunol. 176, 6180–6185 (2006). TACI are associated with common variable Acknowledgements
165. Cheroutre, H., Lambolez, F. & Mucida, D. The light and immunodeficiency in humans. Nature Genet. 37, The authors thank all members of the Artis laboratory for
dark sides of intestinal intraepithelial lymphocytes. 820–828 (2005). discussions and critical reading of the manuscript. This work
Nature Rev. Immunol. 11, 445–456 (2011). 182. Castigli, E. et al. TACI is mutant in common variable is supported by US National Institutes of Health grants
166. Edelblum, K. L. et al. Dynamic migration of γδ immunodeficiency and IgA deficiency. Nature Genet. (AI061570, AI095608, AI087990, AI074878, AI095466,
intraepithelial lymphocytes requires occludin. 37, 829–834 (2005). AI106697, AI102942 and AI097333 to D.A.; T32AI00744
Proc. Natl Acad. Sci. USA 109, 7097–7102 (2012). 183. Barker, N. et al. Identification of stem cells in small to L.W.P.), the Burroughs Wellcome Fund Investigator in
167. Ismail, A. S. et al. γδ intraepithelial lymphocytes are intestine and colon by marker gene Lgr5. Nature 449, Pathogenesis of Infectious Disease Award (D.A.) and the
essential mediators of host-microbial homeostasis at 1003–1007 (2007). Crohn’s and Colitis Foundation of America (D.A.).
the intestinal mucosal surface. Proc. Natl Acad. Sci. 184. Takeda, N. et al. Interconversion between intestinal
USA 108, 8743–8748 (2011). stem cell populations in distinct niches. Science 334,
168. Mucida, D. et al. Transcriptional reprogramming 1420–1424 (2011). Competing interests statement
of mature CD4+ helper T cells generates distinct 185. Yan, K. S. et al. The intestinal stem cell markers Bmi1 The authors declare no competing interests.
MHC class II‑restricted cytotoxic T lymphocytes. and Lgr5 identify two functionally distinct populations.
Nature Immunol. 14, 281–289 (2013). Proc. Natl Acad. Sci. USA 109, 466–471 (2012).
169. Gebhardt, T., Mueller, S. N., Heath, W. R. & 186. Capaldo, C. T. et al. IFN‑γ and TNF‑α‑induced SUPPLEMENTARY INFORMATION
Carbone, F. R. Peripheral tissue surveillance and GBP‑1 inhibits epithelial cell proliferation through See online article: S1 (table)
residency by memory T cells. Trends Immunol. 34, suppression of β‑catenin/TCF signaling. ALL LINKS ARE ACTIVE IN THE ONLINE PDF
27–32 (2013). Mucosal Immunol. 5, 681–690 (2012).

NATURE REVIEWS | IMMUNOLOGY VOLUME 14 | MARCH 2014 | 153

© 2014 Macmillan Publishers Limited. All rights reserved

Das könnte Ihnen auch gefallen