Next Article in Journal
A Cautionary Note on the Use of Split-YFP/BiFC in Plant Protein-Protein Interaction Studies
Next Article in Special Issue
Erythropoietin Action in Stress Response, Tissue Maintenance and Metabolism
Previous Article in Journal
Differential Proteomic Analysis of the Pancreas of Diabetic db/db Mice Reveals the Proteins Involved in the Development of Complications of Diabetes Mellitus
Previous Article in Special Issue
Adipose-Derived Stem Cells: A Review of Signaling Networks Governing Cell Fate and Regenerative Potential in the Context of Craniofacial and Long Bone Skeletal Repair
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

NOD-Like Receptors in Intestinal Homeostasis and Epithelial Tissue Repair

by
Marianna Parlato
1 and
Garabet Yeretssian
1,2,*
1
Department of Medicine, Immunology Institute, Icahn School of Medicine at Mount Sinai, New York, NY 10029, USA
2
Tisch Cancer Institute, Icahn School of Medicine at Mount Sinai, New York, NY 10029, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2014, 15(6), 9594-9627; https://doi.org/10.3390/ijms15069594
Submission received: 20 March 2014 / Revised: 16 May 2014 / Accepted: 20 May 2014 / Published: 30 May 2014
(This article belongs to the Special Issue Signal Transduction of Tissue Repair)

Abstract

:
The intestinal epithelium constitutes a dynamic physical barrier segregating the luminal content from the underlying mucosal tissue. Following injury, the epithelial integrity is restored by rapid migration of intestinal epithelial cells (IECs) across the denuded area in a process known as wound healing. Hence, through a sequence of events involving restitution, proliferation and differentiation of IECs the gap is resealed and homeostasis reestablished. Relapsing damage followed by healing of the inflamed mucosa is a hallmark of several intestinal disorders including inflammatory bowel diseases (IBD). While several regulatory peptides, growth factors and cytokines stimulate restitution of the epithelial layer after injury, recent evidence in the field underscores the contribution of innate immunity in controlling this process. In particular, nucleotide-binding and oligomerization domain-like receptors (NLRs) play critical roles in sensing the commensal microbiota, maintaining homeostasis, and regulating intestinal inflammation. Here, we review the process of intestinal epithelial tissue repair and we specifically focus on the impact of NLR-mediated signaling mechanisms involved in governing epithelial wound healing during disease.

Graphical Abstract

1. Introduction

Recurrent damage following injury of the intestinal epithelial barrier is at the basis of several intestinal disorders including inflammatory bowel disease (IBD), ischemic colitis and enteroinvasive bacterial infections. IBD, comprising the two clinically distinct forms, Crohn’s disease (CD) and ulcerative colitis (UC), is characterized by inappropriate and intermittent episodes of intestinal inflammation. IBD affects the health of more than 1.4 million people in North America and over 2.2 million individuals in Europe [1,2]. The incidence and prevalence of IBD continue to increase with time particularly in low frequency areas such as Asia, North Africa, Australia and New Zealand [3]. The pathophysiology and etiology of IBD are still unclear, however, it is thought to result from an exuberant inflammatory response to commensals and an impaired mucosal healing in genetically predisposed individuals [4,5]. The intestinal mucosa is colonized by tens of trillions of microbes that are either harmless or provide beneficial functions to the host [6]. This complex microbiota shapes the intestinal physiology predominantly by maintaining immune homeostasis [7], protecting against enteric pathogen colonization [8], and procuring nutrient metabolism [9]. It appears that tolerance mechanisms towards the commensal microbiota are dysregulated in IBD patients [10,11,12]. These patients show profound alterations in their gut microbiota composition, suggestive of an imbalance or “dysbiosis” between certain microbes that exacerbate inflammation and others that mitigate inflammation [13,14,15,16]. Sequencing of the 16S ribosomal DNA of the “normal” gut microbiota highlights the predominance of species from two bacterial divisions represented by the Bacteroidetes and the Firmicutes phyla [17,18]. Indeed, clinical dysbiosis was linked to CD with greater preponderance of certain Bacteroidetes and reduced complexity of the Firmicutes phylum [19,20]. Likewise, increased Proteobacteria and Actinobacteria and a concomitant depletion in Bacteroidetes was found in mucosal biopsies from UC patients compared to non-IBD controls [21,22]. Evidence obtained from animal models emphasizes the key role of certain commensals in dictating mucosal immunity by directly suppressing the inflammatory response or promoting the expansion of regulatory T (Treg) cells [23]. For instance, depletion of the mouse commensal segmented filamentous bacteria (SFB) leads to an impaired capacity of CD4+ T cells to differentiate into IL-17 producing T helper (TH) 17 subset, essential for host resistance to enteric infections [24,25,26]. Moreover, certain Clostridium strains promote Foxp3+ Treg accumulation in a transforming-growth factor-β (TGF-β) rich milieu, indicating that indigenous species contribute to directly control systemic immune responses by inducing polarization of protective T cells [27].
While microbiota and environmental factors are essential contributors to the pathogenesis of IBD, genetic predispositions play a central role in IBD by disturbing the balanced and symbiotic/mutualistic host-microbiota interactions [28]. Recently, genetic studies have identified 163 IBD risk loci with 110 loci associated with both disease phenotypes, 30 specifically linked with CD and 23 with UC [29]. Interestingly, classified loci converge into common biological pathways essential for intestinal homeostasis- or inflammation-related processes, such as epithelial wound repair (e.g., REL [30], STAT3 [31], HNF4A [32]) and barrier function (e.g., ITLN1 [33], MUC19 [31]), innate immune responses (e.g., NOD2 [34], CARD9 [35], RIPK2 [29], NLRP3 [36]), autophagy (e.g., ATG16L1 [37], IRGM [38], LRRK2 [39]), T-cell differentiation (IL23R [40,41]; IL10R [21,42], JAK2 [43]), etc.
The intestinal mucosal surface is covered with epithelial cells constituting a physical barrier that prevents the entry of luminal antigens into the system [44,45]. Upon injury, the epithelial barrier undergoes a tightly regulated wound healing process involving various cytokines, growth factors, regulatory peptides, short chain fatty acids, bile acids, etc. [46]. Moreover, signaling cascades downstream of the innate immune Toll-like receptors (TLRs) and Nucleotide binding and oligomerization domain (NOD)-like receptors (NLRs) are also implicated in modulating epithelial barrier repair and integrity [47,48,49]. In this review, we will describe the mechanisms by which the intestinal epithelium is repaired following injury or during IBD. We will pay particular attention to the emerging role of NLRs in intestinal homeostasis and inflammation and discuss their potential in controlling signal transduction of intestinal tissue repair.

2. Intestinal Epithelial Barrier

The epithelium is a single layer of cells lining the inner surface of the gastrointestinal tract presenting different secretory and absorptive properties along the proximo-distal axis [50]. The small intestine consists of invaginations called crypts of Lieberkühn lined with finger-like protrusions or villi, whereas epithelial cells in the colon are regularly arranged in crypts. Four major cell lineages, namely columnar cells or enterocytes, goblet cells, enteroendocrine cells and Paneth cells, reside in the intestinal epithelium and display all the features of differentiated mature epithelial cells. Absorptive enterocytes are the most abundant population of polarized cells that not only regulate paracellular antigen trafficking via the tight junctions but also, in an undefined mechanism, process and present antigens directly to T cells [51,52,53]. Goblet cells and hormone-producing enteroendocrine cells are present in the entire intestinal epithelium of both crypts and villi [54]. Goblet cells secrete highly glycosylated gel-forming mucins (e.g., MUC2) that constitute the mucus, a paramount defense line of the gut epithelial barrier [54,55,56]. MUC2 is the major secreted intestinal mucin, however it acts differently throughout the proximo-distal axis of the intestine probably due to differential processing mechanisms [57]. The small intestine has a single layer of loose mucus, whereas the colon has a double-layered mucus, an outer loose and an inner attached mucus [54]. Increasing evidence underscores the importance of the concept that a defective mucus layer enhances the epithelial cell-microbiota interactions and therefore host susceptibility to intestinal injury. In fact, lack of MUC2 in mice results in severe colitis propagated by increased bacterial adhesion to the epithelium and intestinal permeability [58]. Paneth cells, the major source of antimicrobial peptides in the small intestine, are confined to the bottom of the crypts where they constitute a niche for multipotent leucine-rich repeat-containing G-protein coupled receptor 5 (Lgr5)-expressing intestinal stem cells (ISCs) [59].
During homeostatic turnover, epithelial cells are constantly expelled from the epithelial monolayer and die by anoikis, however barrier integrity is maintained [60]. Indeed, the intestinal epithelium undergoes rapid self-renewal every 4–5 days, and in order to retain optimal barrier functions and homeostatic cell numbers the epithelium employs a highly dynamic process involving crowding-induced extrusion of cells and their subsequent death [60,61,62]. The epithelial barrier is organized into three compartments along the crypt-to-villi axis: the base corresponds to the stem cell niche, the middle compartment comprises the transit-amplifying (TA) cells, and the upper part contains the differentiation zone, which expands until the surface epithelium of the crypts in the colon and the tip of the villi in the small intestine [63]. ISCs at the base of the crypts regularly divide to produce highly proliferative daughter TA cells that migrate upwards and become progressively committed to the absorptive or secretory cell lineages through activation of complex signal transduction cascades [64]. The Lgr5-expressing ISCs are virtually able to raise all cell types present in the epithelium [65,66,67], while another subset of ISCs known as +4 stem cells are quiescent and might serve as “reserve” stem cells [68]. Paneth cells have a slower turnover compared to other intestinal cell lineages; they arise from a small subset of dedicated Lgr5-expressing precursors and differentiate following a downward migratory path to the crypt bottom [69]. Animal models emphasize the long-term regenerative abilities of the intestinal epithelium during radiation- or chemically-induced injury providing key mechanistic insights into ISC-mediated tissue repair processes [70]. Several ex vivo culture systems and in particular the Matrigel-based three-dimensional culture system are built around the expansion of purified LGR5-expressing cells in a milieu enriched with growth factors, surrogate of the environment in the stem cell niche [59,71]. These in vitro expanded cells or epithelial organoids were shown to successfully repair and regenerate damaged colonic epithelia in mice [72]. Likewise, human and mouse immature fetal intestinal progenitors can be exploited to regenerate a differentiated region-specific epithelium, suggesting that stem cell-based therapies might be considered as potential treatment for IBD patients [73]. The lamina propria underneath the epithelial compartment consists of sub-epithelial connective tissue, non-hematopoietic mesenchymal or stromal cells and, among others, immune cells such as macrophages, dendritic cells, T and B lymphocytes [74].

3. Process of Intestinal Epithelial Tissue Repair

Impaired intestinal epithelium is a hallmark for various diseases and may result in the dissemination of toxic and immunogenic luminal substances into the system causing dysregulated gut homeostasis and extensive inflammation [75,76]. Hence, rapid sealing of the epithelium or epithelial wound healing after injury is necessary to reinstate this balance [77,78,79]. Complex and highly coordinated processes restore the continuity of the epithelial barrier in three distinct but overlapping steps: epithelial restitution, proliferation and differentiation [48,80]. Restitution begins within minutes to hours when epithelial cells lining the injured surface undergo de-differentiation, remodel the actin cytoskeleton and migrate over the wound to cover the damaged area [46]. Proliferation starts hours or days after the injury to increase the pool of enterocytes and is followed by the re-differentiation step mainly to reestablish barrier integrity and function. During epithelial restitution, actin reorganization provides a structural framework that defines the shape and polarity of the epithelial cells and that creates traction forces necessary to push the cell forward [81]. This dynamic mechanism is tightly controlled by the RHO family of small guanine triphosphatases (GTPases; e.g., RHO, RAC and CDC42) [82]. Principally, CDC42 and RAC are respectively required for the development of filopodia and lamellipodia protrusions that facilitate the transient adherence of the epithelial cells to the underlying matrix [83], whereas RHO regulates the polymerization of actin to produce stress fibers [84]. Crawling of intestinal epithelial cells (IECs) during restitution can be enhanced via chemokine activation. In fact, RHO-GTPase activation enhances CXCL12-CXCR4-mediated actin rearrangement promoting epithelial cell migration and barrier restitution [85]. Similarly, CCL20 and the human β-defensin 2 (HBD2) through engagement with CCR6 direct actin accumulation and IEC migration [86]. Certain intestinal epithelial molecules such as Villin, an actin-binding protein expressed predominantly in the villi, also participate in tissue repair by apical pole remodeling of migrating cells [87,88]. Villin localizes at the lamellipodia favoring cellular motility in vitro and in vivo through calcium-dependent wrapping, capping, and severing of actin filaments [89,90,91,92]. Low levels of Villin expression are observed in IECs derived from IBD patients, and Villin-deficient mice are more susceptible to dextran sulfate sodium (DSS) [93,94,95], a sulfated polysaccharide highly toxic to crypt enterocytes and commonly used for experimental acute injury and intestinal tissue repair [96,97].

4. Mechanisms of Intestinal Epithelial Tissue Repair

Multiple regulatory growth factors, immunoregulatory cytokines and peptides control intestinal wound healing by either impacting the restitution or the proliferation processes. Many of these factors promote tissue repair by enhanced production of bioactive TGF-β [98], however some other factors such as trefoil factors (TFF) and galectins promote wound healing in a TGF-β-independent mechanism [99,100]. Galectins are a family of β-galactoside binding lectins of which Galectins-2, -3 and -4 are expressed in IECs supporting cell migration and wound re-epithelialization [100,101]. Recent discoveries show that Galectin-3 levels are impaired in IBD patients and propose that Matrix metalloproteinase-7 (MMP7) cleaves Galectin-3 during intestinal inflammation hampering its tissue repair functions [102]. TFF, small protease-resistant peptides secreted by Goblet cells, play critical roles in intestinal homeostasis through potentiating epithelial restitution [99,103,104]. For instance, TFF3 deficiency renders mice highly susceptible to DSS-induced colitis due to an impaired mucosal protection [105]. Moreover, administration of human hTFF2 and transgenic overexpression of hTFF1 accelerate mucosal healing and decrease inflammation during experimental models of colitis [106,107,108,109]. So far no specific TFF receptor was described, however in vitro evidence put forward the idea that TFF3 induces the transactivation of ErbB family members of receptor tyrosine kinases, including epidermal growth factor receptor (EGFR/ErbB1) and ErbB2, and enhances IEC survival and proliferation in a p38 mitogen activated protein kinase (MAPK)- as well as phosphoinositol 3-kinase (PI3K)-dependent manner [110,111,112] (Figure 1). Additionally, tumor necrosis factor (TNF)-α and nuclear factor-kappa B (NF-κB) repress the transcription of intestinal TFF3 in vitro; suggesting that perpetual activation of NF-κB in the intestinal mucosa of IBD patients may promote ulcerations and hence decrease TFF3-dependent tissue repair [113]. It appears that NF-κB, master regulator of inflammation [114], has tissue-protective functions in the intestine as IEC-specific inhibition of NF-κB in mice through deletion of NEMO (also called IKKγ; IκB kinase-γ) [115], or both IKKα and IKKβ subunits spontaneously cause severe chronic intestinal inflammation [116,117,118].
Apart from TFF and Galectins, tissue repair factors expressed within the intestinal mucosa such as EGF, TGF-α, TGF-β, hepatocyte growth factor (HGF), granulocyte macrophage colony stimulating factor (GM-CSF), vascular endothelial cell growth factor (VEGF), keratinocyte growth factor (KGF), platelet-derived growth factor (PDGF), as well as interleukin (IL)-1β, IL-2 and interferon (IFN)-γ cytokines activate specific signaling pathways that converge to induce bioactive TGF-β-mediated epithelial cell restitution [46,98,119,120,121,122,123,124,125,126]. Most of these regulatory molecules reduce mucosal damage and intestinal inflammation in in vivo models of colitis [127,128,129,130,131], however EGF family peptides (EGF, TGF-α and TGF-β) occupy a central role in the healing mechanisms and therefore will be discussed more extensively. In the intestine, EGF and TGF-α control IEC proliferation and regeneration. Both factors are necessary for healing of the injured intestinal mucosa where the majority of cells express EGFR, the common receptor for EGF and TGF-α. Interestingly, mucosal inflammation enhances EGFR expression, yet IBD patients have impaired TGF-α expression and low serum EGF levels [132,133,134]. Transactivation of EGFR triggers MAPK p38 and ERK (extracellular signal regulated-kinases) signaling, which is critical for mucosal defense and for protecting IECs from TNF-α-mediated apoptosis [135,136,137] (Figure 1). Mice lacking TGF-α are highly susceptible to DSS or trinitrobenzene sulphonic acid (TNBS) induced colitis, a phenotype that is reversible by systemic administration of TGF-α [138]. Similarly, EGF largely contributes to restore the epithelial barrier integrity after chemical injury [139]. Alongside, pharmacological inhibition of MAPK signaling aggravates DSS- and TNBS-induced colitis in mice [140].
Figure 1. Signal transduction pathways involved in intestinal epithelial tissue repair. TLR2 engagement induces TFF3 release from mucin (MUC2) secreting Goblet cells. Released TFF3 is detected by adjacent enterocytes via EGFR (ErbB) transactivation leading to PI3K and MAPK (ERK, p38) signaling. Similarly, TLR4 activation at the basal side of the IECs permits COX1/2-derived PGE2 secretion essential for efficient cell proliferation. Both EGF and TGF-α induce epithelial restitution and proliferation by stimulating ErbB receptors and activating downstream p38, ERK and NF-κB signaling cascades. On the other hand, TGF-β via its type I and II serine/threonine receptors induces the translocation of SMAD2/3/4 complex to the nucleus and regulates a gene transcription program fundamental for epithelial cell differentiation and restitution.
Figure 1. Signal transduction pathways involved in intestinal epithelial tissue repair. TLR2 engagement induces TFF3 release from mucin (MUC2) secreting Goblet cells. Released TFF3 is detected by adjacent enterocytes via EGFR (ErbB) transactivation leading to PI3K and MAPK (ERK, p38) signaling. Similarly, TLR4 activation at the basal side of the IECs permits COX1/2-derived PGE2 secretion essential for efficient cell proliferation. Both EGF and TGF-α induce epithelial restitution and proliferation by stimulating ErbB receptors and activating downstream p38, ERK and NF-κB signaling cascades. On the other hand, TGF-β via its type I and II serine/threonine receptors induces the translocation of SMAD2/3/4 complex to the nucleus and regulates a gene transcription program fundamental for epithelial cell differentiation and restitution.
Ijms 15 09594 g001
Although TGF-β plays a crucial role in reestablishing homeostasis and tissue repair after injury, it inhibits IEC proliferation [141]. TGF-β deficiency in mice mediates spontaneous intestinal inflammation [142], extending earlier discoveries that compromised TGF-β signaling contributes to the development of IBD through upregulation of the negative regulator Smad-7 [143,144]. TGF-β signals through type I and type II serine/threonine receptors and sequentially triggers the activation of Smad-2, -3, and -4 transcription cofactors [145]. Lack of TGF-β signaling in transgenic mice expressing a dominant negative functionally inactive form of TGFβRII on epithelial cells increases susceptibility to intestinal injury and delays wound healing in vivo [146]. Various studies indicate the presence of Smad-independent TGF-β signal transduction pathways critical for intestinal tissue repair [147,148,149]. In fact, TGF-β rapidly activates PI3K/AKT, MAPK p38 and ERK and TGF-β activated kinase 1 (TAK1) pathways to induce wound healing. For example, IEC-specific TAK1 signaling prevents TNF-dependent apoptosis and its absence disrupts intestinal homeostasis favoring inflammation [150,151,152] (Figure 1).
While most intestinal tissue repair mechanisms involve the epithelial factors described above, deeper lesions require non-epithelial cell populations and other inflammatory processes. Interestingly, stromal cells underlying the epithelial barrier are in perpetual crosstalk with IECs during homeostasis and repair after injury [153]. Recent studies show that Wnt5a-expressing stromal cells are key for intestinal tissue regeneration [154,155]. Secretion of the non-canonical ligand Wnt5a amplifies TGF-β signaling, which in turn represses proliferation of adjacent IECs allowing these later to arrange into new crypt units [155,156]. Stromal cells as well as IECs are a major source of Prostaglandin-E2 (PGE2) that prevents diminished epithelial proliferation accompanying radiation- or DSS-mediated injury in mice [157,158]. PGE2 is synthesized through activation of cyclooxygenase-1 (COX-1) or COX-2 reflected by the fact that mice deficient of either of these enzymes show decreased epithelial proliferation following DSS administration [159,160,161,162] (Figure 1). Interestingly, PGE2 producing-stromal cells relocate from the lamina propria to the bottom of the crypts during tissue injury and trigger mechanisms to actively support the intestinal epithelial restitution and tissue repair [74,161].

5. Innate Immunity Controlling Intestinal Tissue Repair

Commensal microbiota extensively contributes to intestinal homeostasis through a wide range of mechanisms specifically involving epithelial wound healing. Early studies put forward the idea that mice bred in germ-free conditions or wild-type mice treated with broad-spectrum antibiotics are highly susceptible to DSS-induced injury, suggesting that commensal microbiota sensing is required for optimal epithelial cell restitution and proliferation [163,164]. Interestingly however, seminal studies support the concept that microbiota is required for the initiation of intestinal inflammation with an important implication of innate immune recognition receptors in the development of commensal-dependent colitis [165,166,167].

5.1. Toll-Like Receptors TLRs in Intestinal Tissue Repair

Toll-like receptor (TLR) expression and activation is specifically regulated in the intestine to contrive continuous reaction to commensal derived molecular patterns. In the meantime, intestinal epithelial TLRs retain the ability to sense and combat invading pathogens [168]. Surface expression of TLRs is decreased in IECs and is relocated to intracellular compartments or to the basolateral surface of the cell [169,170,171,172,173,174,175]. Numerous in vivo and in vitro studies highlight the contribution of TLR signaling to intestinal tissue repair after injury [168,176]. In fact, mice deficient in the shared TLR signaling adaptor protein Myeloid Differentiation Primary Response 88 (MyD88) show severe intestinal inflammation and increased susceptibility to experimental or infectious colitis, due to impaired epithelial restitution and intestinal tissue repair [163,177,178,179,180,181]. Similarly, mice lacking TLR2, TLR4, TLR5 or TLR9 present unbalanced microbial sensing that provokes epithelial barrier injury and exacerbated intestinal inflammation during colitis [182,183,184,185,186]. Engaging the TLR signaling preserves intestinal epithelial barrier integrity by enhancing transepithelial resistance of IECs [187], by protecting intercellular tight junctions [188] or by promoting the expression of various proliferative and anti-apoptotic factors [151,189,190,191]. Indeed, TLR4 signaling is required to induce COX-2 activation, which leads to PGE2 production by IECs and subsequent induction of an EGF family growth factor, amphiregulin [192] (Figure 1). Moreover, TLR signaling within IECs confers protection in acute DSS-induced colitis by increasing the expression of amphiregulin in a MyD88 and TIR-domain-containing adapter-inducing interferon-β (TRIF) dependent manner [193]. Following intestinal injury, TLR9 upregulates VEGF and an intestinal progenitor cell differentiation factor hairy enhancer of split 1 (HES1) to potently drive epithelial tissue repair [194]. Also, TLR2 activation induces secretion of TFF3 by Goblet cells, which once sensed by the IECs fosters their restitution and proliferation [191,195]. Besides, during intestinal inflammation stromal cells and myeloid cells are recruited to the intestinal stem cell niche, where they promote proliferation and survival of epithelial progenitors through secretion of a myriad of regulatory factors that render the milieu highly restitutive and proliferative [161,196].

5.2. NOD-Like Receptors (NLRs) in Intestinal Homeostasis, Inflammation and Tissue Repair

The human NOD-like receptor (NLR) family of intracellular sensors comprises 22 members, which share a common modular tripartite organization: C-terminal leucine-rich repeats (LRR) implicated in ligand sensing, central nucleotide-binding and oligomerization domain (NOD also called NACHT) and N-terminal protein-protein–interaction domain. This later prompts NLR activity and their classification into four subfamilies: NLRA containing an acidic domain, NLRB with a baculovirus inhibitor of apoptosis repeat (BIR) domain, NLRC consisting of NLRs with a caspase activation and recruitment domain (CARD), and NLRP with a pyrin domain (PYD) [197]. NLRX1 is an exception, as it comprises a non-identified CARD-related X effector domain [198,199,200]. Upon sensing a wide array of microbial- or danger-associated molecular patterns, MAMPs and DAMPs respectively, the NLRs undergo conformational changes with the LRR domain, exposing the N-terminal domain and allowing protein-protein interactions with downstream signaling adaptor molecules. The culmination of these interactions drives the expression of inflammatory and antimicrobial genes that govern both innate and adaptive immune responses.

5.2.1. Non-Inflammasome NLRs

NOD1 (CARD4) and NOD2 (CARD15) are the first NLRs identified and play a major role in triggering innate immune signaling following bacterial sensing. NOD1 and NOD2 discriminate between Gram-negative and Gram-positive bacteria by sensing specific peptidoglycan motifs. For instance, NOD2 detects the common bacterial muramyl-dipeptide (MDP), while NOD1 recognizes a subset of bacteria that contain the d-glutamyl-meso-diaminopimelic acid (DAP) found in Gram-negative bacteria but also in some Gram-positive bacteria like Listeria monocytogenes and Bacillus species [201,202,203]. Upon ligand sensing, NOD1 and NOD2 via homotypic CARD–CARD interactions recruit the adaptor kinase receptor-interacting protein 2 (RIP2, also known as RIPK2) and form large macromolecular complexes [204,205,206]. Lack of RIP2 in mice abrogates ligand-mediated NOD1 and NOD2 signaling, emphasizing the central role that RIP2 plays in this cascade [205,207]. RIP2 activation occurs via Lysine (K)63-polyubiquitination mediated by several E3 ubiquitin ligases, including X-linked inhibitor of apoptosis (XIAP) [208], cellular inhibitor of apoptosis protein 1 (cIAP1) and cIAP2 [209], TNFR-associated factor 2 (TRAF2), TRAF5 and TRAF6 [210,211]. Polyubiquitination of RIP2 is accompanied with its autophosphorylation on tyrosine (Tyr474), which promotes NF-κB signaling [212]. RIP2 then directly binds to the TAK-1/TAB-2/TAB-3 (TAK-1-binding protein 2 and 3) complex and recruits IKK complex, which in turn leads to downstream activation of both NF-κB and MAPK pathways [204,213,214] (Figure 2). Recent genome-wide RNAi screens put forward novel regulators of the NOD1/2-NF-kB signaling axis [215,216,217]. For instance, NOD1 and NOD2 engage in a crosstalk with apoptosis proteins through an interaction with the pro-apoptotic protein BH3-interacting domain death agonist (BID) [215,218]. Notably, loss of BID impairs NOD1- and NOD2-mediated innate immunity by differentially regulating NF-κB and ERK signaling [215]. Activation of NOD1 and NOD2 results in the production of an array of antimicrobial peptides (e.g., defensins [45]) as well as pro-inflammatory cytokines (e.g., TNF-α, IL-6, etc.) and chemokines (e.g., CC-chemokine ligand 2 (CCL2), CXC-chemokine ligand 8 (CXCL8 or IL-8), CXCL2, etc.) [49,219,220].
While structurally similar, NOD1 and NOD2 diverge in their expression profiles as NOD1 is found in a wide variety of cell types whereas the expression of NOD2 is confined to myeloid and lymphoid cells [221]. However, IECs as well as small intestinal Paneth cells express NOD2 [201,222]. Consistently, increased levels of NOD2 are observed in Paneth cells from IBD patients [223]. In fact, treatment of colonic IECs with TNF-α, IFNγ or butyrate regulate the synthesis of NOD2 at the mRNA level, likely indicating that inflammation during IBD drives the increase in NOD2 signaling [224,225,226]. NOD2 is the first identified gene associated with CD susceptibility and until now the strongest genetic risk factor for the development of CD [29,34,227]. In particular, three coding variants within NOD2 (p.G908R, p.R702W and p.L1007fs) possess the greatest risk associated with CD so far [34,227,228]. All three variants fall within the LRR region, and therefore result in decreased MDP responsiveness [229]. Unlike NOD2, genetic studies linking polymorphisms in NOD1 with IBD susceptibility are still conflicting and necessitate further investigation [230,231]. Given the implication of NOD1 and NOD2 in bacterial sensing, their expression in IECs and the strong association of NOD2 with CD, it is speculated that altered detection of commensals might impair intestinal homeostasis and contribute to intestinal inflammation. Evidence from in vitro and in vivo experiments support this concept and underscore the importance of the cellular localization of NOD proteins and their antimicrobial functions (reviewed in [45]) in the intestine. During infections with enteric pathogens, cytosolic NOD1 and NOD2 interact with the autophagy-related protein 16-like 1 (ATG16L1) and are recruited to the plasma membrane. Also, FERM and PDZ domain protein-containing 2 (FRMPD2) cooperate with NOD2 at the basolateral membrane of IECs providing a spatial specificity to MDP-mediated NOD2 signaling [216]. Concomitantly, FRMPD2 links NOD2 to the membrane-associated ERBB2-interacting protein ERBIN [216,232,233]. Small RHO GTPases are well-described targets of bacterial virulence factors but also critical for epithelial cell migration and barrier restitution after injury [85,234]. Recent data shows that RHO GTPases form a complex with NOD1 and NOD2, further promoting their assembly at the plasma membrane and suggesting that cytoskeleton remodeling regulates NOD signaling [235,236,237]. Notably, activation of RHOA, RAC1 and CDC42 prompt NOD1 signaling in both ligand dependent- and independent-manner [236].
Murine models are invaluable tools to unravel the mechanisms governing NOD1- and NOD2-mediated intestinal homeostasis and inflammation. For instance, administration of peptidoglycan or MDP negatively regulates TLR signaling and protects wild type mice from colitis, suggesting that the microbial sensing by NOD proteins at the gut immune barrier is key for host protection [238,239]. Furthermore, mice deficient of NOD1, NOD2 or double knockout mice for NOD1 and NOD2 are extremely susceptible to chemically-induced colitis due to increased barrier permeability and lack of efficient epithelial tissue repair [240,241,242,243,244]. Given its expression in Paneth cells, NOD2 controls the production of antimicrobial peptides and thus preserves the balance between the gut commensal communities and confers protection during intestinal injury [245]. Similar to CD patients, NOD2-deficient mice present dysbiosis with reduced ability to prevent intestinal colonization of pathogenic bacteria and enhanced sensitivity to colitis [244,246,247]. Interestingly, when wild type mice are cohoused with NOD2-deficient mice and subjected to DSS-induced injury, they develop severe intestinal inflammation and epithelial barrier injury, a phenotype reversed by systemic neutralization of the pro-inflammatory cytokine IL-6 [244]. Akin to NOD2, NOD1 controls the composition of the intestinal commensal flora and is required for optimal generation of intestinal lymphoid tissue by directly detecting microbial components [248]. Along the same line of investigation, mice deficient in NOD2 or both NOD1 and NOD2 have a delayed response to Gram-negative Citrobacter rodentium, a rodent surrogate of human enteropathogenic Escherichia coli infection [219,249]. Indeed, NOD1 and NOD2 double deficient mice present an impaired capacity to mount appropriate TH17-type immune responses against C. rodentium infection [249]. On the other hand, NOD2 deficiency renders mice more susceptible to C. rodentium due to reduced CCL2 chemokine production, decreased chemotaxis of inflammatory monocytes to the intestine and subsequent compromised TH1 immune responses [219].

5.2.2. Inflammasome NLRs

Apart from NOD1 and NOD2, other NLRs mediate the activation of the inflammatory caspases-1 and -11 in macromolecular platforms called inflammasomes subsequently promoting the secretion of the pro-inflammatory cytokines IL-1β and IL-18 and pyroptosis, a form of cell death induced by bacterial pathogens [250,251,252] (Figure 2). Inflammasome activation requires two signals: signal 1 is provided by TLR stimulation and NF-κB-dependent synthesis of the pro-forms of IL-1β and IL18; whereas signal 2 specifically induced by MAMPs or DAMPs promotes caspase-1 activation and processing of the pro-forms of IL-1β and IL-18 into their active mature forms. The inflammasome NLRs share the same tripartite architecture as all NLRs with either an N-terminal PYD or CARD domain, which allows homotypic interactions with the PYD domain of the adaptor protein apoptosis-associated speck-like protein (ASC; also called PYCARD) or the CARD domain of caspase-1, respectively [253,254,255]. ASC also contains a CARD domain that upon inflammasome activation with microbial- or danger-associated molecular patterns (MAMPs or DAMPs) binds to the CARD of caspase-1 [256]. Most inflammasomes play a central role in infectious, autoimmune and inflammatory diseases, however only few regulate intestinal homeostasis and inflammation including NLRP3, NLRC4, NLRP6 and NLRP12 [49,257].
Figure 2. Simplified NLR pathways engaged in intestinal homeostasis. Upon microbial sensing, NOD1 and NOD2 recruit the adaptor protein RIP2 via CARD-CARD interactions. Activated RIP2 binds to the E3 ubiquitin ligases TRAFs and IAPs triggering NF-κB and MAPK signaling. Similarly, the adaptor protein ASC bridges the activated inflammasome NLRs (NLRP3, 6 and 12) to pro-caspase-1 through both PYD and CARD domains, respectively. Formation of the inflammasome complex triggers auto-proteolytic cleavage and activation of caspase-1 subsequently leading to maturation and release of IL-1β and IL-18. Activation of both NOD signaling and inflammasome pathways contributes to the maintenance of intestinal homeostasis against resident commensal microbes.
Figure 2. Simplified NLR pathways engaged in intestinal homeostasis. Upon microbial sensing, NOD1 and NOD2 recruit the adaptor protein RIP2 via CARD-CARD interactions. Activated RIP2 binds to the E3 ubiquitin ligases TRAFs and IAPs triggering NF-κB and MAPK signaling. Similarly, the adaptor protein ASC bridges the activated inflammasome NLRs (NLRP3, 6 and 12) to pro-caspase-1 through both PYD and CARD domains, respectively. Formation of the inflammasome complex triggers auto-proteolytic cleavage and activation of caspase-1 subsequently leading to maturation and release of IL-1β and IL-18. Activation of both NOD signaling and inflammasome pathways contributes to the maintenance of intestinal homeostasis against resident commensal microbes.
Ijms 15 09594 g002
Observations made using genetic studies emphasize the role of inflammasomes in the pathogenesis of IBD. Interestingly, polymorphisms in genes encoding IL-18 and IL-18 receptor accessory protein (IL-18RAP) are associated with increased CD pathology [258,259]. Consistently, polymorphisms in the regulatory region downstream of the human NLRP3 gene are also linked to CD susceptibility, due for instance to decreased NLRP3 mRNA expression and impaired IL-1β production by lipopolysaccharide-challenged monocytes [36]. Recent evidence from murine models of experimental colitis support the genetic linkage analysis and accentuate the central role played by inflammasomes in the intestine [29,36]. Initial studies show that pharmacological inhibition of caspase-1 with Pralnacasan or neutralization of IL-18 confers resistance to experimental colitis [260,261]. Similarly, caspase-1-deficient mice on the Balb/c background have reduced disease after DSS-injury [262]. Interestingly however, recent discoveries refuted these earlier findings by purporting the importance of inflammasome signaling in maintaining tissue homeostasis and protection during injury [263,264,265,266]. Mice deficient in the inflammasome components ASC and caspase-1 are susceptible to DSS colitis and display heightened intestinal permeability and commensal bacterial translocation probably due to an impaired epithelial restitution and proliferation [263]. These mice exhibit a hampered IL-18 production by IECs upon acute intestinal injury, a phenotype rescued by exogenous administration of recombinant IL-18, further underscoring the requirement of IL-18 to efficiently repair and regenerate the epithelial barrier upon injury [263]. Equally, lack of epithelial IL-18 expression in NLRP3 deficient mice enhanced susceptibility of these mice to DSS-induced colitis when compared to control wild-type mice [265,267,268]. Notably, increased intestinal inflammation in NLRP3 knockout mice is associated with reduced levels of IL-1β, anti-inflammatory IL-10, and the IEC restitution factor TGF-β [268]. While most studies outlined above underscore a protective role for NLRP3 during intestinal inflammation, recent studies contradict these observations by showing that NLRP3-deficient mice are resistant to colitis [269,270]. In an infectious model of colitis with C. rodentium, early activation of the NLRP3 or NLRC4 inflammasomes within the IECs limit pathogen colonization and prevent pathology in mice [271,272]. These findings were confirmed in mice deficient of various inflammasome components including NLRP3, NLRC4, ASC, Caspase-1, IL-1β and IL-18 [271,272,273]. IL-18 receptor (IL-18R) signals through the common adaptor MyD88, which suggests that the IL-18/IL-18R/MyD88 signaling might be implicated in COX-1- and COX-2-mediated secretion of PGE2 essential for intestinal epithelial restitution and proliferation [160,161,189]. Consistently, recent reports further emphasize that IL1R, IL18R or IL18 deficiency renders mice highly susceptible to DSS-induced colitis indicating that signaling through MyD88 is required to preserve the intestinal barrier and to promote repair following injury [266,274,275]. Hence, IL-18 exerts both protective (in IECs) and harmful (in the lamina propria) effects in intestinal inflammation depending on the site of activation and the level of IL-18 produced [263,264,276]. Unlike other inflammasomes, NLRC4 activates caspase-1 and mediates IL-1β and IL-18 production in an ASC-dependent or independent manner [255,277]. The NLRC4 inflammasome is important in host defense against Gram-negative bacteria and its activation involves sensing of flagellin and components of the bacterial Type III secretion system [255,278,279]. Although loss of NLRC4 does not alter basal intestinal homeostasis, DSS-induced epithelial injury in mice lacking NLRC4 results in low expression of IL-1β and IL-18 and consequently severe disease [280]. However, another study using the same chemically induced colitis model shows opposite results, demonstrating an ameliorated severity of colitis in NLRC4-deficient mice [270]. These discrepancies may stem from the length and concentration of DSS used as well as differences in the composition of the intestinal microflora between the two facilities.
In addition, two recently described NLRP6 and NLRP12 inflammasomes also play a key role in the maintenance of gut homeostasis [281,282,283,284,285]. Both NLRP6 and NLRP12 cooperate with ASC to promote caspase-1 activation and production of IL-1β and IL-18 [286,287,288]. Moreover, in vitro studies suggest that both receptors potentiate NF-κB and MAPK signaling and require ASC to promote these responses [286,287]. In marked contrast, later findings show that NLRP6 is a negative regulator of canonical NF-κB signaling and that mice deficient in NLRP6 survive systemic bacterial infections through upregulation of NF-κB-mediated cytokines and enhanced bacterial clearance [289]. Moreover, NLRP12 negatively regulates the noncanonical NF-κB pathway by directly associating with the NF-κB inducing kinase (NIK) or IRAK-1 and interfering with TLR-mediated signaling [284,290,291]. NLRP6 deficient mice develop sever colitis in response to DSS when compared to wild type mice, manifested with reduced IL-18 expression by IECs and increased pathology [281,282,283]. The phenotype of NLRP6 mice is transmissible when cohoused together with wild type mice for 4 weeks, highlighting the colitogenic origin of the microbiota of NLRP6 mice. Notably, NLRP6 mice have an altered microbiota with expansion of pathobiont bacteria belonging to the genus Prevotellaceae as well as the phylum TM7 and present increased CCL5 production [281]. Broad-spectrum antibiotic treatment of these mice depletes both Prevotellaceae and TM7 and protects the mice from DSS-induced injury. In addition, another possible mechanism for NLRP6 mediated protection might be that NLRP6 limits epithelial cell depolarization and tissue disintegration upon injury by regulating downstream target genes of both Wnt and Notch signaling cascades [282]. Akin to NLRP6, loss of NLRP12 in mice causes more severe colitis, in line with the role of NLRP12 as negative regulator of NF-κB and MAPK pathways [284,285]. Interestingly however, both studies describing the colitic phenotype of the NLRP12-deficient mice propose opposing mechanisms using bone marrow chimera experiments without further characterization whether NLRP12 exerts its protective functions from the hematopoietic or non-hematopoietic compartments [257,284,285]. While lack of NLRs affects the composition of the commensal flora, little is know whether the protective functions of NLRs in the context of intestinal injury are direct. A recent report by Wlodarska et al. emphasizes that the NLRP6 inflammasome is a critical orchestrator of goblet cell mucin exocytosis and lack of NLRP6 leads to defective autophagy and mucin secretion, two key regulators implicated in maintaining intestinal homeostasis and protection during colitis [292]. This suggests that NLRs in general and NLRP6 in particular might indirectly protect the host from intestinal commensals during injury by employing the host’s physiological defense mechanisms such as the autophagy machinery.

6. Conclusions

Exciting progress has been made in understanding the signal transduction mechanisms governing mucosal wound healing after physical, chemical or bacterial injury. Based on the above discussions, intestinal tissue repair arises as a highly orchestrated and finely tuned process involving, not only regulatory peptides, growth factors and cytokines, but also innate immune recognition. A large body of recent research emphasizes the significant role of host innate immune NLRs in promoting intestinal health. NLRs maintain intestinal mucosal integrity and homeostasis by controlling tolerance to the commensal microbiota, by regulating inflammatory signaling events, and by amplifying the crosstalk between IECs and underlying hematopoietic and non-hematopoietic cells. Genetic studies correlating loss of function mutations in the NOD2 gene or in the regulatory region of the NLRP3 gene with CD susceptibility further accentuate the prominence of NLRs in tissue repair processes at the gut immune barrier. Emerging experimental and clinical evidence suggests that NLR-mediated signals protect the host from intestinal inflammation endorsing the concept that these innate immune sensors prompt IEC restitution and elicit reestablishment of homeostasis. However, various mechanisms may be involved in driving mucosal healing by NLRs. Hence, additional studies are required to definitively clarify the relative contributions of the different NLR-mediated signal transduction mechanisms in the regulation of intestinal homeostasis and disease. The full understanding of these cascades of events may set the stage for novel therapeutic strategies targeting epithelial tissue repair in IBD patients by specifically modulating NLR functions within the IECs.

Abbreviations

IECs
intestinal epithelial cells
IBD
inflammatory bowel diseases
NLRs
nucleotide-binding and oligomerization domain-like receptors
TLRs
toll-like receptors
CD
Crohn’s disease
UC
ulcerative colitis
TH
T helper
TGF-β
transforming-growth factor-β
Foxp3
forkhead box P3
STAT3
signal transducer and activator of transcription 3
HNF4A
hepatocyte nuclear factor 4 alpha
ITLN1
intelectin 1
MUC19
mucin 19
NOD2
nucleotide-binding oligomerization domain containing 2
CARD9
caspase recruitment domain family, member 9
RIPK2
receptor interacting protein kinase 2
NLRP3
NLR family, pyrin domain containing 3
ATG16L1
autophagy related 16-like 1
IRGM
immunity-related GTPase family, M
LRRK2
leucine-rich repeat kinase 2
IL23R
interleukin 23 receptor
IL10R
interleukin 10 receptor
JAK2
janus kinase 2
MUC2
mucin 2
Lgr5
leucine-rich repeat-containing G-protein coupled receptor 5
TA
transit-amplifying
ISCs
intestinal stem cells
GTPases
guanine triphosphatases
RHO
rhodopsin
CDC42
cell division cycle 42
CXCL
chemokine (C-X-C motif) ligand
HBD2
human β-defensin 2
CCL2
chemokine (C-C motif) ligand 2
CCR6
chemokine (C-C motif) receptor 6
DSS
dextran sulfate sodium
MMP7
Matrix metalloproteinase-7
TFF
trefoil factor
EGFR
epidermal growth factor receptor
EGF
epidermal growth factor
ErbB
v-erb-b2 avian erythroblastic leukemia viral oncogene homolog
MAPK
mitogen activated protein kinase
PI3K
phosphoinositol 3-kinase
TNF-α
tumor necrosis factor alpha
NF-κB
nuclear factor-kappa B
ΙΚΚα
IκB kinase alpha
ΙΚΚβ
IκB kinase beta
ΙΚΚγ
IκB kinase gamma
NEMO
NF-κB Essential Modulator
HGF
hepatocyte growth factor
VEGF
vascular endothelial cell growth factor
GM-CSF
granulocyte macrophage colony stimulating factor
KGF
keratinocyte growth factor
PDGF
platelet-derived growth factor
IFN
interferon
ERK
extracellular signal regulated-kinases
TNBS
trinitrobenzene sulphonic acid
COX1/2
cyclooxygenase 1/2
SMAD
Sma- And Mad-Related Protein
TGFβRII
TGFβ receptor II
TAK1
TGF-β activated kinase 1
Wnt5a
wingless-type MMTV integration site family, member 5A
PGE2
Prostaglandin-E2
MyD88
Myeloid Differentiation Primary Response 88
TRIF
TIR-domain-containing adapter-inducing interferon-β
HES1
hairy enhancer of split 1
LRR
leucine-rich repeats
NACHT
nucleotide-binding and oligomerization domain
BIR
baculovirus inhibitor of apoptosis repeat
CARD
caspase activation and recruitment domain
PYD
pyrin domain
MAMPs
microbial-associated molecular patterns
DAMPs
danger-associated molecular patterns
DAP
d-glutamyl-meso-diaminopimelic acid
MDP
muramyl dipeptide
XIAP
X-linked inhibitor of apoptosis
cIAP
cellular inhibitor of apoptosis protein
TRAF
TNFR-associated factor
TAB2-TAB3
TAK-1-binding protein 2 and 3
RNAi
ribonucleic acid interference
FRMPD2
FERM and PDZ domain protein-containing 2
ERBIN
ERBB2-interacting protein
ASC
apoptosis-associated speck-like protein
IL18RAP
Interleukin 18 Receptor Accessory Protein

Acknowledgments

We thank all the Yeretssian lab members for the insightful and critical reading of the manuscript. This work was supported by the Helmsley Foundation.

Author Contributions

Marianna Parlato and Garabet Yeretssian co-conceived and wrote the review. All authors approved the manuscript for submission.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Loftus, E.V., Jr. Clinical epidemiology of inflammatory bowel disease: Incidence, prevalence, and environmental influences. Gastroenterology 2004, 126, 1504–1517. [Google Scholar] [CrossRef]
  2. Cho, J.H. The genetics and immunopathogenesis of inflammatory bowel disease. Nat. Rev. Immunol. 2008, 8, 458–466. [Google Scholar] [CrossRef]
  3. Long, M.D.; Hutfless, S.; Kappelman, M.D.; Khalili, H.; Kaplan, G.G.; Bernstein, C.N.; Colombel, J.F.; Gower-Rousseau, C.; Herrinton, L.; Velayos, F.; et al. Challenges in designing a national surveillance program for inflammatory bowel disease in the united states. Inflamm. Bowel Dis. 2014, 20, 398–415. [Google Scholar] [CrossRef]
  4. Abraham, C.; Cho, J.H. Inflammatory bowel disease. N. Engl. J. Med. 2009, 361, 2066–2078. [Google Scholar] [CrossRef]
  5. Garrett, W.S.; Gordon, J.I.; Glimcher, L.H. Homeostasis and inflammation in the intestine. Cell 2010, 140, 859–870. [Google Scholar] [CrossRef]
  6. Lochner, M.; Ohnmacht, C.; Presley, L.; Bruhns, P.; Si-Tahar, M.; Sawa, S.; Eberl, G. Microbiota-induced tertiary lymphoid tissues aggravate inflammatory disease in the absence of rorgamma T and LTI cells. J. Exp. Med. 2011, 208, 125–134. [Google Scholar] [CrossRef]
  7. Olszak, T.; An, D.; Zeissig, S.; Vera, M.P.; Richter, J.; Franke, A.; Glickman, J.N.; Siebert, R.; Baron, R.M.; Kasper, D.L.; et al. Microbial exposure during early life has persistent effects on natural killer t cell function. Science 2012, 336, 489–493. [Google Scholar] [CrossRef]
  8. Fukuda, S.; Toh, H.; Hase, K.; Oshima, K.; Nakanishi, Y.; Yoshimura, K.; Tobe, T.; Clarke, J.M.; Topping, D.L.; Suzuki, T.; et al. Bifidobacteria can protect from enteropathogenic infection through production of acetate. Nature 2011, 469, 543–547. [Google Scholar] [CrossRef]
  9. Yatsunenko, T.; Rey, F.E.; Manary, M.J.; Trehan, I.; Dominguez-Bello, M.G.; Contreras, M.; Magris, M.; Hidalgo, G.; Baldassano, R.N.; Anokhin, A.P.; et al. Human gut microbiome viewed across age and geography. Nature 2012, 486, 222–227. [Google Scholar]
  10. Casellas, F.; Borruel, N.; Papo, M.; Guarner, F.; Antolin, M.; Videla, S.; Malagelada, J.R. Antiinflammatory effects of enterically coated amoxicillin-clavulanic acid in active ulcerative colitis. Inflamm. Bowel Dis. 1998, 4, 1–5. [Google Scholar] [CrossRef]
  11. Macpherson, A.; Khoo, U.Y.; Forgacs, I.; Philpott-Howard, J.; Bjarnason, I. Mucosal antibodies in inflammatory bowel disease are directed against intestinal bacteria. Gut 1996, 38, 365–375. [Google Scholar] [CrossRef]
  12. Pirzer, U.; Schonhaar, A.; Fleischer, B.; Hermann, E.; Meyer zum Buschenfelde, K.H. Reactivity of infiltrating T lymphocytes with microbial antigens in crohn’s disease. Lancet 1991, 338, 1238–1239. [Google Scholar] [CrossRef]
  13. Frank, D.N.; St. Amand, A.L.; Feldman, R.A.; Boedeker, E.C.; Harpaz, N.; Pace, N.R. Molecular-phylogenetic characterization of microbial community imbalances in human inflammatory bowel diseases. Proc. Natl. Acad. Sci. USA 2007, 104, 13780–13785. [Google Scholar]
  14. Dicksved, J.; Halfvarson, J.; Rosenquist, M.; Jarnerot, G.; Tysk, C.; Apajalahti, J.; Engstrand, L.; Jansson, J.K. Molecular analysis of the gut microbiota of identical twins with crohn’s disease. ISME J. 2008, 2, 716–727. [Google Scholar] [CrossRef]
  15. Sartor, R.B. Microbial influences in inflammatory bowel diseases. Gastroenterology 2008, 134, 577–594. [Google Scholar] [CrossRef]
  16. Guarner, F. What is the role of the enteric commensal flora in IBD? Inflamm. Bowel Dis. 2008, 14, S83–S84. [Google Scholar] [CrossRef]
  17. Eckburg, P.B.; Bik, E.M.; Bernstein, C.N.; Purdom, E.; Dethlefsen, L.; Sargent, M.; Gill, S.R.; Nelson, K.E.; Relman, D.A. Diversity of the human intestinal microbial flora. Science 2005, 308, 1635–1638. [Google Scholar] [CrossRef]
  18. Salzman, N.H.; Hung, K.; Haribhai, D.; Chu, H.; Karlsson-Sjoberg, J.; Amir, E.; Teggatz, P.; Barman, M.; Hayward, M.; Eastwood, D.; et al. Enteric defensins are essential regulators of intestinal microbial ecology. Nat. Immun. 2010, 11, 76–83. [Google Scholar] [CrossRef]
  19. Scanlan, P.D.; Shanahan, F.; O’Mahony, C.; Marchesi, J.R. Culture-independent analyses of temporal variation of the dominant fecal microbiota and targeted bacterial subgroups in crohn’s disease. J. Clin. Microbiol. 2006, 44, 3980–3988. [Google Scholar]
  20. Manichanh, C.; Rigottier-Gois, L.; Bonnaud, E.; Gloux, K.; Pelletier, E.; Frangeul, L.; Nalin, R.; Jarrin, C.; Chardon, P.; Marteau, P.; et al. Reduced diversity of faecal microbiota in crohn’s disease revealed by a metagenomic approach. Gut 2006, 55, 205–211. [Google Scholar] [CrossRef]
  21. Franke, A.; Balschun, T.; Karlsen, T.H.; Hedderich, J.; May, S.; Lu, T.; Schuldt, D.; Nikolaus, S.; Rosenstiel, P.; Krawczak, M.; et al. Replication of signals from recent studies of crohn’s disease identifies previously unknown disease loci for ulcerative colitis. Nat. Genet. 2008, 40, 713–715. [Google Scholar] [CrossRef]
  22. Lepage, P.; Hasler, R.; Spehlmann, M.E.; Rehman, A.; Zvirbliene, A.; Begun, A.; Ott, S.; Kupcinskas, L.; Dore, J.; Raedler, A.; et al. Twin study indicates loss of interaction between microbiota and mucosa of patients with ulcerative colitis. Gastroenterology 2011, 141, 227–236. [Google Scholar] [CrossRef]
  23. Round, J.L.; Mazmanian, S.K. The gut microbiota shapes intestinal immune responses during health and disease. Nat. Rev. Immunol. 2009, 9, 313–323. [Google Scholar] [CrossRef]
  24. Gaboriau-Routhiau, V.; Rakotobe, S.; Lecuyer, E.; Mulder, I.; Lan, A.; Bridonneau, C.; Rochet, V.; Pisi, A.; de Paepe, M.; Brandi, G.; et al. The key role of segmented filamentous bacteria in the coordinated maturation of gut helper t cell responses. Immunity 2009, 31, 677–689. [Google Scholar]
  25. Wu, H.J.; Ivanov, II; Darce, J.; Hattori, K.; Shima, T.; Umesaki, Y.; Littman, D.R.; Benoist, C.; Mathis, D. Gut-residing segmented filamentous bacteria drive autoimmune arthritis via T helper 17 cells. Immunity 2010, 32, 815–827. [Google Scholar] [CrossRef]
  26. Ivanov, II; Atarashi, K.; Manel, N.; Brodie, E.L.; Shima, T.; Karaoz, U.; Wei, D.; Goldfarb, K.C.; Santee, C.A.; Lynch, S.V.; et al. Induction of intestinal Th17 cells by segmented filamentous bacteria. Cell 2009, 139, 485–498. [Google Scholar] [CrossRef]
  27. Atarashi, K.; Tanoue, T.; Shima, T.; Imaoka, A.; Kuwahara, T.; Momose, Y.; Cheng, G.; Yamasaki, S.; Saito, T.; Ohba, Y.; et al. Induction of colonic regulatory T cells by indigenous clostridium species. Science 2011, 331, 337–341. [Google Scholar] [CrossRef]
  28. Lozupone, C.A.; Stombaugh, J.I.; Gordon, J.I.; Jansson, J.K.; Knight, R. Diversity, stability and resilience of the human gut microbiota. Nature 2012, 489, 220–230. [Google Scholar] [CrossRef]
  29. Jostins, L.; Ripke, S.; Weersma, R.K.; Duerr, R.H.; McGovern, D.P.; Hui, K.Y.; Lee, J.C.; Schumm, L.P.; Sharma, Y.; Anderson, C.A.; et al. Host-microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature 2012, 491, 119–124. [Google Scholar] [CrossRef]
  30. Janse, M.; Lamberts, L.E.; Franke, L.; Raychaudhuri, S.; Ellinghaus, E.; Muri Boberg, K.; Melum, E.; Folseraas, T.; Schrumpf, E.; Bergquist, A.; et al. Three ulcerative colitis susceptibility loci are associated with primary sclerosing cholangitis and indicate a role for IL2, REL, and CARD9. Hepatology 2011, 53, 1977–1985. [Google Scholar] [CrossRef]
  31. Danoy, P.; Pryce, K.; Hadler, J.; Bradbury, L.A.; Farrar, C.; Pointon, J.; Australo-Anglo-American Spondyloarthritis Consortium; Ward, M.; Weisman, M.; Reveille, J.D.; et al. Association of variants at 1q32 and STAT3 with ankylosing spondylitis suggests genetic overlap with crohn’s disease. PLoS Genet. 2010, 6. [Google Scholar] [CrossRef] [Green Version]
  32. Consortium, U.I.G.; Barrett, J.C.; Lee, J.C.; Lees, C.W.; Prescott, N.J.; Anderson, C.A.; Phillips, A.; Wesley, E.; Parnell, K.; Zhang, H.; et al. Genome-wide association study of ulcerative colitis identifies three new susceptibility loci, including the HNF4a region. Nat. Genet. 2009, 41, 1330–1334. [Google Scholar] [CrossRef]
  33. Barrett, J.C.; Hansoul, S.; Nicolae, D.L.; Cho, J.H.; Duerr, R.H.; Rioux, J.D.; Brant, S.R.; Silverberg, M.S.; Taylor, K.D.; Barmada, M.M.; et al. Genome-wide association defines more than 30 distinct susceptibility loci for crohn’s disease. Nat. Genet. 2008, 40, 955–962. [Google Scholar] [CrossRef]
  34. Hugot, J.P.; Chamaillard, M.; Zouali, H.; Lesage, S.; Cezard, J.P.; Belaiche, J.; Almer, S.; Tysk, C.; O’Morain, C.A.; Gassull, M.; et al. Association of NOD2 leucine-rich repeat variants with susceptibility to crohn’s disease. Nature 2001, 411, 599–603. [Google Scholar] [CrossRef]
  35. Beaudoin, M.; Goyette, P.; Boucher, G.; Lo, K.S.; Rivas, M.A.; Stevens, C.; Alikashani, A.; Ladouceur, M.; Ellinghaus, D.; Torkvist, L.; et al. Deep resequencing of gwas loci identifies rare variants in CARD9, IL23R and RNF186 that are associated with ulcerative colitis. PLoS Genet. 2013, 9. [Google Scholar] [CrossRef] [Green Version]
  36. Villani, A.C.; Lemire, M.; Fortin, G.; Louis, E.; Silverberg, M.S.; Collette, C.; Baba, N.; Libioulle, C.; Belaiche, J.; Bitton, A.; et al. Common variants in the nlrp3 region contribute to crohn’s disease susceptibility. Nat. Genet. 2009, 41, 71–76. [Google Scholar] [CrossRef]
  37. Hampe, J.; Franke, A.; Rosenstiel, P.; Till, A.; Teuber, M.; Huse, K.; Albrecht, M.; Mayr, G.; de La Vega, F.M.; Briggs, J.; et al. A genome-wide association scan of nonsynonymous SNPs identifies a susceptibility variant for crohn disease in ATG16L1. Nat. Genet. 2007, 39, 207–211. [Google Scholar] [CrossRef]
  38. Parkes, M.; Barrett, J.C.; Prescott, N.J.; Tremelling, M.; Anderson, C.A.; Fisher, S.A.; Roberts, R.G.; Nimmo, E.R.; Cummings, F.R.; Soars, D.; et al. Sequence variants in the autophagy gene IRGM and multiple other replicating loci contribute to crohn’s disease susceptibility. Nat. Genet. 2007, 39, 830–832. [Google Scholar] [CrossRef]
  39. Duraes, C.; Machado, J.C.; Portela, F.; Rodrigues, S.; Lago, P.; Cravo, M.; Ministro, P.; Marques, M.; Cremers, I.; Freitas, J.; et al. Phenotype-genotype profiles in crohn’s disease predicted by genetic markers in autophagy-related genes (GOIA study II). Inflamm. Bowel Dis. 2013, 19, 230–239. [Google Scholar] [CrossRef]
  40. Duerr, R.H.; Taylor, K.D.; Brant, S.R.; Rioux, J.D.; Silverberg, M.S.; Daly, M.J.; Steinhart, A.H.; Abraham, C.; Regueiro, M.; Griffiths, A.; et al. A genome-wide association study identifies IL23R as an inflammatory bowel disease gene. Science 2006, 314, 1461–1463. [Google Scholar]
  41. Tremelling, M.; Cummings, F.; Fisher, S.A.; Mansfield, J.; Gwilliam, R.; Keniry, A.; Nimmo, E.R.; Drummond, H.; Onnie, C.M.; Prescott, N.J.; et al. IL23R variation determines susceptibility but not disease phenotype in inflammatory bowel disease. Gastroenterology 2007, 132, 1657–1664. [Google Scholar] [CrossRef]
  42. Glocker, E.O.; Kotlarz, D.; Boztug, K.; Gertz, E.M.; Schaffer, A.A.; Noyan, F.; Perro, M.; Diestelhorst, J.; Allroth, A.; Murugan, D.; et al. Inflammatory bowel disease and mutations affecting the interleukin-10 receptor. N. Engl. J. Med. 2009, 361, 2033–2045. [Google Scholar] [CrossRef]
  43. Anderson, C.A.; Massey, D.C.; Barrett, J.C.; Prescott, N.J.; Tremelling, M.; Fisher, S.A.; Gwilliam, R.; Jacob, J.; Nimmo, E.R.; Drummond, H.; et al. Investigation of crohn’s disease risk loci in ulcerative colitis further defines their molecular relationship. Gastroenterology 2009, 136, 523–529. [Google Scholar] [CrossRef]
  44. Laukoetter, M.G.; Bruewer, M.; Nusrat, A. Regulation of the intestinal epithelial barrier by the apical junctional complex. Curr. Opin. Gastroenterol. 2006, 22, 85–89. [Google Scholar] [CrossRef]
  45. Muniz, L.R.; Knosp, C.; Yeretssian, G. Intestinal antimicrobial peptides during homeostasis, infection, and disease. Front. Immun. 2012, 3, 310. [Google Scholar]
  46. Sturm, A.; Dignass, A.U. Epithelial restitution and wound healing in inflammatory bowel disease. World J. Gastroenterol. 2008, 14, 348–353. [Google Scholar] [CrossRef]
  47. Harris, G.; KuoLee, R.; Chen, W. Role of toll-like receptors in health and diseases of gastrointestinal tract. World J. Gastroenterol. 2006, 12, 2149–2160. [Google Scholar]
  48. Iizuka, M.; Konno, S. Wound healing of intestinal epithelial cells. World J. Gastroenterol. 2011, 17, 2161–2171. [Google Scholar] [CrossRef]
  49. Yeretssian, G. Effector functions of NLRs in the intestine: Innate sensing, cell death, and disease. Immun. Res. 2012, 54, 25–36. [Google Scholar] [CrossRef]
  50. Turner, J.R. Intestinal mucosal barrier function in health and disease. Nat. Rev. Immunol. 2009, 9, 799–809. [Google Scholar] [CrossRef]
  51. Allez, M.; Brimnes, J.; Dotan, I.; Mayer, L. Expansion of CD8+ T cells with regulatory function after interaction with intestinal epithelial cells. Gastroenterology 2002, 123, 1516–1526. [Google Scholar] [CrossRef]
  52. Rabinowitz, K.; Mayer, L. Working out mechanisms of controlled/physiologic inflammation in the GI tract. Immun. Res. 2012, 54, 14–24. [Google Scholar] [CrossRef]
  53. Roda, G.; Jianyu, X.; Park, M.S.; Demarte, L.; Hovhannisyan, Z.; Couri, R.; Stanners, C.P.; Yeretssian, G.; Mayer, L. Characterizing CEACAM5 interaction with CD8α and CD1D in intestinal homeostasis. Mucosal Immunol. 2013, 7, 615–624. [Google Scholar]
  54. Johansson, M.E.; Sjovall, H.; Hansson, G.C. The gastrointestinal mucus system in health and disease. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 352–361. [Google Scholar] [CrossRef]
  55. Thornton, D.J.; Rousseau, K.; McGuckin, M.A. Structure and function of the polymeric mucins in airways mucus. Ann. Rev. Physiol. 2008, 70, 459–486. [Google Scholar] [CrossRef]
  56. Kim, Y.S.; Ho, S.B. Intestinal goblet cells and mucins in health and disease: Recent insights and progress. Curr. Gastroenterol. Rep. 2010, 12, 319–330. [Google Scholar] [CrossRef]
  57. Johansson, M.E.; Phillipson, M.; Petersson, J.; Velcich, A.; Holm, L.; Hansson, G.C. The inner of the two MUC2 mucin-dependent mucus layers in colon is devoid of bacteria. Proc. Natl. Acad. Sci. USA 2008, 105, 15064–15069. [Google Scholar]
  58. Shan, M.; Gentile, M.; Yeiser, J.R.; Walland, A.C.; Bornstein, V.U.; Chen, K.; He, B.; Cassis, L.; Bigas, A.; Cols, M.; et al. Mucus enhances gut homeostasis and oral tolerance by delivering immunoregulatory signals. Science 2013, 342, 447–453. [Google Scholar] [CrossRef]
  59. Sato, T.; van Es, J.H.; Snippert, H.J.; Stange, D.E.; Vries, R.G.; van den Born, M.; Barker, N.; Shroyer, N.F.; van de Wetering, M.; Clevers, H. Paneth cells constitute the niche for LGR5 stem cells in intestinal crypts. Nature 2011, 469, 415–418. [Google Scholar] [CrossRef]
  60. Watson, A.J.; Hughes, K.R. Tnf-α-induced intestinal epithelial cell shedding: Implications for intestinal barrier function. Ann. N. Y. Acad. Sci. 2012, 1258, 1–8. [Google Scholar] [CrossRef]
  61. Eisenhoffer, G.T.; Loftus, P.D.; Yoshigi, M.; Otsuna, H.; Chien, C.B.; Morcos, P.A.; Rosenblatt, J. Crowding induces live cell extrusion to maintain homeostatic cell numbers in epithelia. Nature 2012, 484, 546–549. [Google Scholar] [CrossRef]
  62. Marinari, E.; Mehonic, A.; Curran, S.; Gale, J.; Duke, T.; Baum, B. Live-cell delamination counterbalances epithelial growth to limit tissue overcrowding. Nature 2012, 484, 542–545. [Google Scholar] [CrossRef]
  63. Clevers, H. The intestinal crypt, a prototype stem cell compartment. Cell 2013, 154, 274–284. [Google Scholar] [CrossRef]
  64. Barker, N.; Bartfeld, S.; Clevers, H. Tissue-resident adult stem cell populations of rapidly self-renewing organs. Cell Stem Cell 2010, 7, 656–670. [Google Scholar] [CrossRef]
  65. Barker, N.; van Es, J.H.; Kuipers, J.; Kujala, P.; van den Born, M.; Cozijnsen, M.; Haegebarth, A.; Korving, J.; Begthel, H.; Peters, P.J.; et al. Identification of stem cells in small intestine and colon by marker gene LGR5. Nature 2007, 449, 1003–1007. [Google Scholar] [CrossRef]
  66. Sato, T.; Vries, R.G.; Snippert, H.J.; van de Wetering, M.; Barker, N.; Stange, D.E.; van Es, J.H.; Abo, A.; Kujala, P.; Peters, P.J.; et al. Single LGR5 stem cells build crypt-villus structures in vitro without a mesenchymal niche. Nature 2009, 459, 262–265. [Google Scholar] [CrossRef]
  67. Clevers, H. Stem cells: A unifying theory for the crypt. Nature 2013, 495, 53–54. [Google Scholar] [CrossRef]
  68. Barker, N.; van Oudenaarden, A.; Clevers, H. Identifying the stem cell of the intestinal crypt: Strategies and pitfalls. Cell Stem Cell 2012, 11, 452–460. [Google Scholar] [CrossRef]
  69. Buczacki, S.J.; Zecchini, H.I.; Nicholson, A.M.; Russell, R.; Vermeulen, L.; Kemp, R.; Winton, D.J. Intestinal label-retaining cells are secretory precursors expressing LGR5. Nature 2013, 495, 65–69. [Google Scholar] [CrossRef]
  70. Sato, T.; Clevers, H. Growing self-organizing mini-guts from a single intestinal stem cell: Mechanism and applications. Science 2013, 340, 1190–1194. [Google Scholar] [CrossRef]
  71. Huch, M.; Dorrell, C.; Boj, S.F.; van es, J.H.; Li, V.S.; van de Wetering, M.; Sato, T.; Hamer, K.; Sasaki, N.; Finegold, M.J.; et al. In vitro expansion of single LGR5+ liver stem cells induced by wnt-driven regeneration. Nature 2013, 494, 247–250. [Google Scholar] [CrossRef]
  72. Yui, S.; Nakamura, T.; Sato, T.; Nemoto, Y.; Mizutani, T.; Zheng, X.; Ichinose, S.; Nagaishi, T.; Okamoto, R.; Tsuchiya, K.; et al. Functional engraftment of colon epithelium expanded in vitro from a single adult LGR5+ stem cell. Nat. Med. 2012, 18, 618–623. [Google Scholar]
  73. Fordham, R.P.; Yui, S.; Hannan, N.R.; Soendergaard, C.; Madgwick, A.; Schweiger, P.J.; Nielsen, O.H.; Vallier, L.; Pedersen, R.A.; Nakamura, T.; et al. Transplantation of expanded fetal intestinal progenitors contributes to colon regeneration after injury. Cell Stem Cell 2013, 13, 734–744. [Google Scholar] [CrossRef]
  74. Owens, B.M.; Simmons, A. Intestinal stromal cells in mucosal immunity and homeostasis. Mucosal Immunol. 2013, 6, 224–234. [Google Scholar] [CrossRef]
  75. Banan, A.; Choudhary, S.; Zhang, Y.; Fields, J.Z.; Keshavarzian, A. Oxidant-induced intestinal barrier disruption and its prevention by growth factors in a human colonic cell line: Role of the microtubule cytoskeleton. Free Radic. Biol. Med. 2000, 28, 727–738. [Google Scholar] [CrossRef]
  76. Dean, P.; Kenny, B. Intestinal barrier dysfunction by enteropathogenic Escherichia coli is mediated by two effector molecules and a bacterial surface protein. Mol. Microbiol. 2004, 54, 665–675. [Google Scholar] [CrossRef]
  77. Moore, R.; Carlson, S.; Madara, J.L. Rapid barrier restitution in an in vitro model of intestinal epithelial injury. Lab. Investig. 1989, 60, 237–244. [Google Scholar]
  78. McCormack, S.A.; Viar, M.J.; Johnson, L.R. Migration of IEC-6 cells: A model for mucosal healing. Am. J. Physiol. 1992, 263, G426–G435. [Google Scholar]
  79. Nusrat, A.; Delp, C.; Madara, J.L. Intestinal epithelial restitution. Characterization of a cell culture model and mapping of cytoskeletal elements in migrating cells. J. Clin. Investig. 1992, 89, 1501–1511. [Google Scholar] [CrossRef]
  80. Karrasch, T.; Jobin, C. Wound healing responses at the gastrointestinal epithelium: A close look at novel regulatory factors and investigative approaches. Z. Gastroenterol. 2009, 47, 1221–1229. [Google Scholar] [CrossRef]
  81. Hall, A. Rho GTPases and the control of cell behaviour. Biochem. Soc. Trans. 2005, 33, 891–895. [Google Scholar]
  82. Jaffe, A.B.; Hall, A. Rho GTPases: Biochemistry and biology. Ann. Rev. Cell Dev. Biol. 2005, 21, 247–269. [Google Scholar] [CrossRef]
  83. Cau, J.; Hall, A. CDC42 controls the polarity of the actin and microtubule cytoskeletons through two distinct signal transduction pathways. J. Cell Sci. 2005, 118, 2579–2587. [Google Scholar] [CrossRef]
  84. Raftopoulou, M.; Hall, A. Cell migration: Rho GTPases lead the way. Dev. Biol. 2004, 265, 23–32. [Google Scholar] [CrossRef]
  85. Moyer, R.A.; Wendt, M.K.; Johanesen, P.A.; Turner, J.R.; Dwinell, M.B. Rho activation regulates CXCL12 chemokine stimulated actin rearrangement and restitution in model intestinal epithelia. Lab. Investig. 2007, 87, 807–817. [Google Scholar] [CrossRef]
  86. Vongsa, R.A.; Zimmerman, N.P.; Dwinell, M.B. CCR6 regulation of the actin cytoskeleton orchestrates human β defensin-2 and CCL20-mediated restitution of colonic epithelial cells. J. Biol. Chem. 2009, 284, 10034–10045. [Google Scholar]
  87. Glenney, J.R., Jr.; Geisler, N.; Kaulfus, P.; Weber, K. Demonstration of at least two different actin-binding sites in villin, a calcium-regulated modulator of F-actin organization. J. Biol. Chem. 1981, 256, 8156–8161. [Google Scholar]
  88. Athman, R.; Louvard, D.; Robine, S. Villin enhances hepatocyte growth factor-induced actin cytoskeleton remodeling in epithelial cells. Mol. Biol. Cell 2003, 14, 4641–4653. [Google Scholar] [CrossRef]
  89. Bretscher, A.; Weber, K. Villin is a major protein of the microvillus cytoskeleton which binds both G and F actin in a calcium-dependent manner. Cell 1980, 20, 839–847. [Google Scholar] [CrossRef]
  90. Glenney, J.R., Jr.; Kaulfus, P.; Matsudaira, P.; Weber, K. F-actin binding and bundling properties of fimbrin, a major cytoskeletal protein of microvillus core filaments. J. Biol. Chem. 1981, 256, 9283–9288. [Google Scholar]
  91. Glenney, J.R., Jr.; Kaulfus, P.; Weber, K. F actin assembly modulated by villin: Ca+-dependent nucleation and capping of the barbed end. Cell 1981, 24, 471–480. [Google Scholar] [CrossRef]
  92. Ubelmann, F.; Chamaillard, M.; El-Marjou, F.; Simon, A.; Netter, J.; Vignjevic, D.; Nichols, B.L.; Quezada-Calvillo, R.; Grandjean, T.; Louvard, D.; et al. Enterocyte loss of polarity and gut wound healing rely upon the F-actin-severing function of villin. Proc. Natl. Acad. Sci. USA 2013, 110, E1380–E1389. [Google Scholar] [CrossRef]
  93. Ferrary, E.; Cohen-Tannoudji, M.; Pehau-Arnaudet, G.; Lapillonne, A.; Athman, R.; Ruiz, T.; Boulouha, L.; El Marjou, F.; Doye, A.; Fontaine, J.J.; et al. In vivo, villin is required for ca2+-dependent F-actin disruption in intestinal brush borders. J. Cell Biol. 1999, 146, 819–830. [Google Scholar] [CrossRef]
  94. Wang, Y.; Srinivasan, K.; Siddiqui, M.R.; George, S.P.; Tomar, A.; Khurana, S. A novel role for villin in intestinal epithelial cell survival and homeostasis. J. Biol. Chem. 2008, 283, 9454–9464. [Google Scholar] [CrossRef]
  95. Kersting, S.; Bruewer, M.; Schuermann, G.; Klotz, A.; Utech, M.; Hansmerten, M.; Krieglstein, C.F.; Senninger, N.; Schulzke, J.D.; Naim, H.Y.; et al. Antigen transport and cytoskeletal characteristics of a distinct enterocyte population in inflammatory bowel diseases. Am. J. Pathol. 2004, 165, 425–437. [Google Scholar] [CrossRef]
  96. Dieleman, L.A.; Elson, C.O.; Tennyson, G.S.; Beagley, K.W. Kinetics of cytokine expression during healing of acute colitis in mice. Am. J. Pathol. 1996, 271, G130–G136. [Google Scholar]
  97. Perse, M.; Cerar, A. Dextran sodium sulphate colitis mouse model: Traps and tricks. J. Biomed. Biotechnol. 2012, 2012. [Google Scholar] [CrossRef]
  98. Dignass, A.U.; Podolsky, D.K. Cytokine modulation of intestinal epithelial cell restitution: Central role of transforming growth factor β. Gastroenterology 1993, 105, 1323–1332. [Google Scholar] [CrossRef]
  99. Playford, R.J.; Marchbank, T.; Chinery, R.; Evison, R.; Pignatelli, M.; Boulton, R.A.; Thim, L.; Hanby, A.M. Human spasmolytic polypeptide is a cytoprotective agent that stimulates cell migration. Gastroenterology 1995, 108, 108–116. [Google Scholar] [CrossRef]
  100. Paclik, D.; Lohse, K.; Wiedenmann, B.; Dignass, A.U.; Sturm, A. Galectin-2 and -4, but not galectin-1, promote intestinal epithelial wound healing in vitro through a TGF-β-independent mechanism. Inflamm. Bowel Dis. 2008, 14, 1366–1372. [Google Scholar] [CrossRef]
  101. Cao, Z.; Said, N.; Amin, S.; Wu, H.K.; Bruce, A.; Garate, M.; Hsu, D.K.; Kuwabara, I.; Liu, F.T.; Panjwani, N. Galectins-3 and -7, but not galectin-1, play a role in re-epithelialization of wounds. J. Biol. Chem. 2002, 277, 42299–42305. [Google Scholar]
  102. Puthenedam, M.; Wu, F.; Shetye, A.; Michaels, A.; Rhee, K.J.; Kwon, J.H. Matrilysin-1 (MMP7) cleaves galectin-3 and inhibits wound healing in intestinal epithelial cells. Inflamm. Bowel Dis. 2011, 17, 260–267. [Google Scholar] [CrossRef]
  103. Wong, W.M.; Poulsom, R.; Wright, N.A. Trefoil peptides. Gut 1999, 44, 890–895. [Google Scholar] [CrossRef]
  104. Taupin, D.; Podolsky, D.K. Trefoil factors: Initiators of mucosal healing. Nat. Rev. Mol. Cell Biol. 2003, 4, 721–732. [Google Scholar] [CrossRef]
  105. Mashimo, H.; Wu, D.C.; Podolsky, D.K.; Fishman, M.C. Impaired defense of intestinal mucosa in mice lacking intestinal trefoil factor. Science 1996, 274, 262–265. [Google Scholar] [CrossRef]
  106. Playford, R.J.; Marchbank, T.; Goodlad, R.A.; Chinery, R.A.; Poulsom, R.; Hanby, A.M. Transgenic mice that overexpress the human trefoil peptide PS2 have an increased resistance to intestinal damage. Proc. Natl. Acad. Sci. USA 1996, 93, 2137–2142. [Google Scholar] [CrossRef]
  107. Tran, C.P.; Cook, G.A.; Yeomans, N.D.; Thim, L.; Giraud, A.S. Trefoil peptide TFF2 (spasmolytic polypeptide) potently accelerates healing and reduces inflammation in a rat model of colitis. Gut 1999, 44, 636–642. [Google Scholar] [CrossRef]
  108. Poulsen, S.S.; Kissow, H.; Hare, K.; Hartmann, B.; Thim, L. Luminal and parenteral TFF2 and TFF3 dimer and monomer in two models of experimental colitis in the rat. Regul. Pept. 2005, 126, 163–171. [Google Scholar] [CrossRef]
  109. Soriano-Izquierdo, A.; Gironella, M.; Massaguer, A.; May, F.E.; Salas, A.; Sans, M.; Poulsom, R.; Thim, L.; Pique, J.M.; Panes, J. Trefoil peptide TFF2 treatment reduces VCAM-1 expression and leukocyte recruitment in experimental intestinal inflammation. J. Leukoc. Biol. 2004, 75, 214–223. [Google Scholar]
  110. Taupin, D.; Wu, D.C.; Jeon, W.K.; Devaney, K.; Wang, T.C.; Podolsky, D.K. The trefoil gene family are coordinately expressed immediate-early genes: EGF receptor- and map kinase-dependent interregulation. J. Clin. Investig. 1999, 103, R31–R38. [Google Scholar] [CrossRef]
  111. Taupin, D.R.; Kinoshita, K.; Podolsky, D.K. Intestinal trefoil factor confers colonic epithelial resistance to apoptosis. Proc. Natl. Acad. Sci. USA 2000, 97, 799–804. [Google Scholar] [CrossRef]
  112. Kinoshita, K.; Taupin, D.R.; Itoh, H.; Podolsky, D.K. Distinct pathways of cell migration and antiapoptotic response to epithelial injury: Structure-function analysis of human intestinal trefoil factor. Mol. Cell. Biol. 2000, 20, 4680–4690. [Google Scholar] [CrossRef]
  113. Loncar, M.B.; Al-azzeh, E.D.; Sommer, P.S.; Marinovic, M.; Schmehl, K.; Kruschewski, M.; Blin, N.; Stohwasser, R.; Gott, P.; Kayademir, T. Tumour necrosis factor α and nuclear factor κb inhibit transcription of human TFF3 encoding a gastrointestinal healing peptide. Gut 2003, 52, 1297–1303. [Google Scholar] [CrossRef]
  114. Karin, M.; Greten, F.R. NF-κb: Linking inflammation and immunity to cancer development and progression. Nat. Rev. Immunol. 2005, 5, 749–759. [Google Scholar] [CrossRef]
  115. Nenci, A.; Becker, C.; Wullaert, A.; Gareus, R.; van Loo, G.; Danese, S.; Huth, M.; Nikolaev, A.; Neufert, C.; Madison, B.; et al. Epithelial nemo links innate immunity to chronic intestinal inflammation. Nature 2007, 446, 557–561. [Google Scholar] [CrossRef]
  116. Egan, L.J.; Eckmann, L.; Greten, F.R.; Chae, S.; Li, Z.W.; Myhre, G.M.; Robine, S.; Karin, M.; Kagnoff, M.F. Iκb-kinaseβ-dependent NF-κb activation provides radioprotection to the intestinal epithelium. Proc. Natl. Acad. Sci. USA 2004, 101, 2452–2457. [Google Scholar] [CrossRef]
  117. Li, Q.; Verma, I.M. NF-κb regulation in the immune system. Nat. Rev. Immunol. 2002, 2, 725–734. [Google Scholar] [CrossRef]
  118. Spehlmann, M.E.; Eckmann, L. Nuclear factor-κ b in intestinal protection and destruction. Curr. Opin. Gastroenterol. 2009, 25, 92–99. [Google Scholar] [CrossRef]
  119. Dignass, A.U.; Lynch-Devaney, K.; Podolsky, D.K. Hepatocyte growth factor/scatter factor modulates intestinal epithelial cell proliferation and migration. Biochem. Biophys. Res. Commun. 1994, 202, 701–709. [Google Scholar] [CrossRef]
  120. Dignass, A.U.; Tsunekawa, S.; Podolsky, D.K. Fibroblast growth factors modulate intestinal epithelial cell growth and migration. Gastroenterology 1994, 106, 1254–1262. [Google Scholar]
  121. Housley, R.M.; Morris, C.F.; Boyle, W.; Ring, B.; Biltz, R.; Tarpley, J.E.; Aukerman, S.L.; Devine, P.L.; Whitehead, R.H.; Pierce, G.F. Keratinocyte growth factor induces proliferation of hepatocytes and epithelial cells throughout the rat gastrointestinal tract. J. Clin. Investig. 1994, 94, 1764–1777. [Google Scholar] [CrossRef]
  122. Ohneda, K.; Ulshen, M.H.; Fuller, C.R.; D’Ercole, A.J.; Lund, P.K. Enhanced growth of small bowel in transgenic mice expressing human insulin-like growth factor I. Gastroenterology 1997, 112, 444–454. [Google Scholar] [CrossRef]
  123. Dignass, A.U.; Podolsky, D.K. Interleukin 2 modulates intestinal epithelial cell function in vitro. Exp. Cell Res. 1996, 225, 422–429. [Google Scholar] [CrossRef]
  124. Ciacci, C.; Lind, S.E.; Podolsky, D.K. Transforming growth factor β regulation of migration in wounded rat intestinal epithelial monolayers. Gastroenterology 1993, 105, 93–101. [Google Scholar]
  125. Liu, F.; Wu, H.Y.; Wesselschmidt, R.; Kornaga, T.; Link, D.C. Impaired production and increased apoptosis of neutrophils in granulocyte colony-stimulating factor receptor-deficient mice. Immunity 1996, 5, 491–501. [Google Scholar] [CrossRef]
  126. Ramsay, R.G.; Micallef, S.J.; Williams, B.; Lightowler, S.; Vincan, E.; Heath, J.K.; Mantamadiotis, T.; Bertoncello, I. Colony-stimulating factor-1 promotes clonogenic growth of normal murine colonic crypt epithelial cells in vitro. J. Interferon Cytokine Res. 2004, 24, 416–427. [Google Scholar] [CrossRef]
  127. Egger, B.; Procaccino, F.; Sarosi, I.; Tolmos, J.; Buchler, M.W.; Eysselein, V.E. Keratinocyte growth factor ameliorates dextran sodium sulfate colitis in mice. Dig. Dis. Sci. 1999, 44, 836–844. [Google Scholar] [CrossRef]
  128. Greenwood-Van Meerveld, B.; Venkova, K.; Connolly, K. Efficacy of repifermin (keratinocyte growth factor-2) against abnormalities in gastrointestinal mucosal transport in a murine model of colitis. J. Pharm. Pharmacol. 2003, 55, 67–75. [Google Scholar] [CrossRef]
  129. Scaldaferri, F.; Vetrano, S.; Sans, M.; Arena, V.; Straface, G.; Stigliano, E.; Repici, A.; Sturm, A.; Malesci, A.; Panes, J.; et al. VEGF-A links angiogenesis and inflammation in inflammatory bowel disease pathogenesis. Gastroenterology 2009, 136, 585–595. [Google Scholar] [CrossRef]
  130. Nair, D.G.; Miller, K.G.; Lourenssen, S.R.; Blennerhassett, M.G. Inflammatory cytokines promote growth of intestinal smooth muscle cells by induced expression of PDGF-Rβ. J. Cell. Mol. Med. 2014, 18, 444–454. [Google Scholar] [CrossRef]
  131. Sainathan, S.K.; Hanna, E.M.; Gong, Q.; Bishnupuri, K.S.; Luo, Q.; Colonna, M.; White, F.V.; Croze, E.; Houchen, C.; Anant, S.; et al. Granulocyte macrophage colony-stimulating factor ameliorates dss-induced experimental colitis. Inflamm. Bowel Dis. 2008, 14, 88–99. [Google Scholar] [CrossRef]
  132. Hoffmann, P.; Reinshagen, M.; Zeeh, J.M.; Lakshmanan, J.; Wu, V.S.; Goebell, H.; Gerken, G.; Eysselein, V.E. Increased expression of epidermal growth factor-receptor in an experimental model of colitis in rats. Scand. J. Gastroenterol. 2000, 35, 1174–1180. [Google Scholar] [CrossRef]
  133. Oikonomou, K.A.; Kapsoritakis, A.N.; Kapsoritaki, A.I.; Manolakis, A.C.; Tsiopoulos, F.D.; Germenis, A.E.; Potamianos, S.P. Downregulation of serum epidermal growth factor in patients with inflammatory bowel disease. Is there a link with mucosal damage? Growth F. 2010, 28, 461–466. [Google Scholar] [CrossRef]
  134. Hormi, K.; Cadiot, G.; Kermorgant, S.; Dessirier, V.; le Romancer, M.; Lewin, M.J.; Mignon, M.; Lehy, T. Transforming growth factor-α and epidermal growth factor receptor in colonic mucosa in active and inactive inflammatory bowel disease. Growth Factors 2000, 18, 79–91. [Google Scholar] [CrossRef]
  135. Frey, M.R.; Golovin, A.; Polk, D.B. Epidermal growth factor-stimulated intestinal epithelial cell migration requires SRC family kinase-dependent p38 MAPK signaling. J. Biol. Chem. 2004, 279, 44513–44521. [Google Scholar] [CrossRef]
  136. Frey, M.R.; Dise, R.S.; Edelblum, K.L.; Polk, D.B. P38 kinase regulates epidermal growth factor receptor downregulation and cellular migration. EMBO J. 2006, 25, 5683–5692. [Google Scholar] [CrossRef]
  137. Yamaoka, T.; Yan, F.; Cao, H.; Hobbs, S.S.; Dise, R.S.; Tong, W.; Polk, D.B. Transactivation of EGF receptor and ERBB2 protects intestinal epithelial cells from TNF-induced apoptosis. Proc. Natl. Acad. Sci. USA 2008, 105, 11772–11777. [Google Scholar]
  138. Hoffmann, P.; Zeeh, J.M.; Lakshmanan, J.; Wu, V.S.; Procaccino, F.; Reinshagen, M.; McRoberts, J.A.; Eysselein, V.E. Increased expression of transforming growth factor α precursors in acute experimental colitis in rats. Gut 1997, 41, 195–202. [Google Scholar] [CrossRef]
  139. Procaccino, F.; Reinshagen, M.; Hoffmann, P.; Zeeh, J.M.; Lakshmanan, J.; McRoberts, J.A.; Patel, A.; French, S.; Eysselein, V.E. Protective effect of epidermal growth factor in an experimental model of colitis in rats. Gastroenterology 1994, 107, 12–17. [Google Scholar]
  140. Nishimura, T.; Andoh, A.; Nishida, A.; Shioya, M.; Koizumi, Y.; Tsujikawa, T.; Fujiyama, Y. Fr167653, a p38 mitogen-activated protein kinase inhibitor, aggravates experimental colitis in mice. World J. Gastroenterol. 2008, 14, 5851–5856. [Google Scholar] [CrossRef]
  141. Margadant, C.; Sonnenberg, A. Integrin-TGF-β crosstalk in fibrosis, cancer and wound healing. EMBO Rep. 2010, 11, 97–105. [Google Scholar] [CrossRef]
  142. Kulkarni, A.B.; Karlsson, S. Transforming growth factor-β 1 knockout mice. A mutation in one cytokine gene causes a dramatic inflammatory disease. Am. J. Pathol. 1993, 143, 3–9. [Google Scholar]
  143. Fiocchi, C. TGF-β/smad signaling defects in inflammatory bowel disease: Mechanisms and possible novel therapies for chronic inflammation. J. Clin. Investig. 2001, 108, 523–526. [Google Scholar] [CrossRef]
  144. Monteleone, G.; Pallone, F.; MacDonald, T.T. SMAD7 in TGF-β-mediated negative regulation of gut inflammation. Trends Immun. 2004, 25, 513–517. [Google Scholar] [CrossRef]
  145. Feng, X.H.; Derynck, R. Specificity and versatility in TGF-β signaling through SMADs. Ann. Rev. Cell Dev. Biol. 2005, 21, 659–693. [Google Scholar] [CrossRef]
  146. Beck, P.L.; Rosenberg, I.M.; Xavier, R.J.; Koh, T.; Wong, J.F.; Podolsky, D.K. Transforming growth factor-β mediates intestinal healing and susceptibility to injury in vitro and in vivo through epithelial cells. Am. J. Pathol. 2003, 162, 597–608. [Google Scholar] [CrossRef]
  147. Bhowmick, N.A.; Ghiassi, M.; Bakin, A.; Aakre, M.; Lundquist, C.A.; Engel, M.E.; Arteaga, C.L.; Moses, H.L. Transforming growth factor-β1 mediates epithelial to mesenchymal transdifferentiation through a RHOA-dependent mechanism. Mol. Biol. Cell 2001, 12, 27–36. [Google Scholar] [CrossRef]
  148. Derynck, R.; Zhang, Y.E. Smad-dependent and smad-independent pathways in TGF-β family signalling. Nature 2003, 425, 577–584. [Google Scholar] [CrossRef]
  149. Moustakas, A.; Heldin, C.H. Non-smad TGF-β signals. J. Cell Sci. 2005, 118, 3573–3584. [Google Scholar] [CrossRef]
  150. Kim, S.I.; Kwak, J.H.; Na, H.J.; Kim, J.K.; Ding, Y.; Choi, M.E. Transforming growth factor-β (TGF-β1) activates TAK1 via TAB1-mediated autophosphorylation, independent of TGF-β receptor kinase activity in mesangial cells. J. Biol. Chem. 2009, 284, 22285–22296. [Google Scholar] [CrossRef]
  151. Kim, J.Y.; Kajino-Sakamoto, R.; Omori, E.; Jobin, C.; Ninomiya-Tsuji, J. Intestinal epithelial-derived TAK1 signaling is essential for cytoprotection against chemical-induced colitis. PLoS One 2009, 4. [Google Scholar] [CrossRef]
  152. Kajino-Sakamoto, R.; Inagaki, M.; Lippert, E.; Akira, S.; Robine, S.; Matsumoto, K.; Jobin, C.; Ninomiya-Tsuji, J. Enterocyte-derived TAK1 signaling prevents epithelium apoptosis and the development of ileitis and colitis. J. Immunol. 2008, 181, 1143–1152. [Google Scholar]
  153. Stappenbeck, T.S.; Miyoshi, H. The role of stromal stem cells in tissue regeneration and wound repair. Science 2009, 324, 1666–1669. [Google Scholar] [CrossRef]
  154. Barrett, T.A. Developmental biology. Intestinal wound healing requires a WNT balancing act. Science 2012, 338, 51–52. [Google Scholar] [CrossRef]
  155. Miyoshi, H.; Ajima, R.; Luo, C.T.; Yamaguchi, T.P.; Stappenbeck, T.S. WNT5A potentiates TGF-β signaling to promote colonic crypt regeneration after tissue injury. Science 2012, 338, 108–113. [Google Scholar] [CrossRef]
  156. Moses, H.L.; Yang, E.Y.; Pietenpol, J.A. TGF-β stimulation and inhibition of cell proliferation: New mechanistic insights. Cell 1990, 63, 245–247. [Google Scholar] [CrossRef]
  157. Tessner, T.G.; Cohn, S.M.; Schloemann, S.; Stenson, W.F. Prostaglandins prevent decreased epithelial cell proliferation associated with dextran sodium sulfate injury in mice. Gastroenterology 1998, 115, 874–882. [Google Scholar] [CrossRef]
  158. Powell, D.W.; Saada, J.I. Mesenchymal stem cells and prostaglandins may be critical for intestinal wound repair. Gastroenterology 2012, 143, 19–22. [Google Scholar] [CrossRef]
  159. Newberry, R.D.; Stenson, W.F.; Lorenz, R.G. Cyclooxygenase-2-dependent arachidonic acid metabolites are essential modulators of the intestinal immune response to dietary antigen. Nat. Med. 1999, 5, 900–906. [Google Scholar]
  160. Cohn, S.M.; Schloemann, S.; Tessner, T.; Seibert, K.; Stenson, W.F. Crypt stem cell survival in the mouse intestinal epithelium is regulated by prostaglandins synthesized through cyclooxygenase-1. J. Clin. Investig. 1997, 99, 1367–1379. [Google Scholar] [CrossRef]
  161. Brown, S.L.; Riehl, T.E.; Walker, M.R.; Geske, M.J.; Doherty, J.M.; Stenson, W.F.; Stappenbeck, T.S. Myd88-dependent positioning of PTGS2-expressing stromal cells maintains colonic epithelial proliferation during injury. J. Clin. Investig. 2007, 117, 258–269. [Google Scholar] [CrossRef]
  162. Morteau, O.; Morham, S.G.; Sellon, R.; Dieleman, L.A.; Langenbach, R.; Smithies, O.; Sartor, R.B. Impaired mucosal defense to acute colonic injury in mice lacking cyclooxygenase-1 or cyclooxygenase-2. J. Clin. Investig. 2000, 105, 469–478. [Google Scholar] [CrossRef]
  163. Rakoff-Nahoum, S.; Paglino, J.; Eslami-Varzaneh, F.; Edberg, S.; Medzhitov, R. Recognition of commensal microflora by toll-like receptors is required for intestinal homeostasis. Cell 2004, 118, 229–241. [Google Scholar] [CrossRef]
  164. Maslowski, K.M.; Vieira, A.T.; Ng, A.; Kranich, J.; Sierro, F.; Yu, D.; Schilter, H.C.; Rolph, M.S.; Mackay, F.; Artis, D.; et al. Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 2009, 461, 1282–1286. [Google Scholar] [CrossRef]
  165. Rakoff-Nahoum, S.; Hao, L.; Medzhitov, R. Role of toll-like receptors in spontaneous commensal-dependent colitis. Immunity 2006, 25, 319–329. [Google Scholar] [CrossRef]
  166. Sellon, R.K.; Tonkonogy, S.; Schultz, M.; Dieleman, L.A.; Grenther, W.; Balish, E.; Rennick, D.M.; Sartor, R.B. Resident enteric bacteria are necessary for development of spontaneous colitis and immune system activation in interleukin-10-deficient mice. Infect. Immun. 1998, 66, 5224–5231. [Google Scholar]
  167. Nell, S.; Suerbaum, S.; Josenhans, C. The impact of the microbiota on the pathogenesis of IBD: Lessons from mouse infection models. Nat. Rev. Microbiol. 2010, 8, 564–577. [Google Scholar] [CrossRef]
  168. Abreu, M.T. Toll-like receptor signalling in the intestinal epithelium: How bacterial recognition shapes intestinal function. Nat. Rev. Immunol. 2010, 10, 131–144. [Google Scholar] [CrossRef]
  169. Abreu, M.T.; Vora, P.; Faure, E.; Thomas, L.S.; Arnold, E.T.; Arditi, M. Decreased expression of toll-like receptor-4 and MD-2 correlates with intestinal epithelial cell protection against dysregulated proinflammatory gene expression in response to bacterial lipopolysaccharide. J. Immunol. 2001, 167, 1609–1616. [Google Scholar] [CrossRef]
  170. Furrie, E.; Macfarlane, S.; Thomson, G.; Macfarlane, G.T.; Microbiology; Gut Biology, G.; Tayside, T.; Tumour, B. Toll-like receptors-2, -3 and -4 expression patterns on human colon and their regulation by mucosal-associated bacteria. Immunology 2005, 115, 565–574. [Google Scholar] [CrossRef]
  171. Melmed, G.; Thomas, L.S.; Lee, N.; Tesfay, S.Y.; Lukasek, K.; Michelsen, K.S.; Zhou, Y.; Hu, B.; Arditi, M.; Abreu, M.T. Human intestinal epithelial cells are broadly unresponsive to toll-like receptor 2-dependent bacterial ligands: Implications for host-microbial interactions in the gut. J. Immunol. 2003, 170, 1406–1415. [Google Scholar] [CrossRef]
  172. Cario, E.; Brown, D.; McKee, M.; Lynch-Devaney, K.; Gerken, G.; Podolsky, D.K. Commensal-associated molecular patterns induce selective toll-like receptor-trafficking from apical membrane to cytoplasmic compartments in polarized intestinal epithelium. Am. J. Pathol. 2002, 160, 165–173. [Google Scholar]
  173. Ortega-Cava, C.F.; Ishihara, S.; Rumi, M.A.; Kawashima, K.; Ishimura, N.; Kazumori, H.; Udagawa, J.; Kadowaki, Y.; Kinoshita, Y. Strategic compartmentalization of toll-like receptor 4 in the mouse gut. J. Immunol. 2003, 170, 3977–3985. [Google Scholar] [CrossRef]
  174. Hornef, M.W.; Frisan, T.; Vandewalle, A.; Normark, S.; Richter-Dahlfors, A. Toll-like receptor 4 resides in the golgi apparatus and colocalizes with internalized lipopolysaccharide in intestinal epithelial cells. J. Exp. Med. 2002, 195, 559–570. [Google Scholar] [CrossRef]
  175. Hornef, M.W.; Normark, B.H.; Vandewalle, A.; Normark, S. Intracellular recognition of lipopolysaccharide by toll-like receptor 4 in intestinal epithelial cells. J. Exp. Med. 2003, 198, 1225–1235. [Google Scholar] [CrossRef]
  176. Cario, E. Toll-like receptors in inflammatory bowel diseases: A decade later. Inflamm. Bowel Dis. 2010, 16, 1583–1597. [Google Scholar] [CrossRef]
  177. Gibson, D.L.; Ma, C.; Bergstrom, K.S.; Huang, J.T.; Man, C.; Vallance, B.A. MYD88 signalling plays a critical role in host defence by controlling pathogen burden and promoting epithelial cell homeostasis during citrobacter rodentium-induced colitis. Cell. Microbiol. 2008, 10, 618–631. [Google Scholar] [CrossRef]
  178. Fukata, M.; Breglio, K.; Chen, A.; Vamadevan, A.S.; Goo, T.; Hsu, D.; Conduah, D.; Xu, R.; Abreu, M.T. The myeloid differentiation factor 88 (MYD88) is required for CD4+ T cell effector function in a murine model of inflammatory bowel disease. J. Immunol. 2008, 180, 1886–1894. [Google Scholar] [CrossRef]
  179. Tomita, T.; Kanai, T.; Fujii, T.; Nemoto, Y.; Okamoto, R.; Tsuchiya, K.; Totsuka, T.; Sakamoto, N.; Akira, S.; Watanabe, M. MYD88-dependent pathway in T cells directly modulates the expansion of colitogenic CD4+ T cells in chronic colitis. J. Immunol. 2008, 180, 5291–5299. [Google Scholar] [CrossRef]
  180. Asquith, M.J.; Boulard, O.; Powrie, F.; Maloy, K.J. Pathogenic and protective roles of MYD88 in leukocytes and epithelial cells in mouse models of inflammatory bowel disease. Gastroenterology 2010, 139, 519–529. [Google Scholar] [CrossRef]
  181. Hapfelmeier, S.; Stecher, B.; Barthel, M.; Kremer, M.; Muller, A.J.; Heikenwalder, M.; Stallmach, T.; Hensel, M.; Pfeffer, K.; Akira, S.; et al. The salmonella pathogenicity island (SPI)-2 and SPI-1 type III secretion systems allow salmonella serovar typhimurium to trigger colitis via MYD88-dependent and MYD88-independent mechanisms. J. Immunol. 2005, 174, 1675–1685. [Google Scholar] [CrossRef]
  182. Fukata, M.; Michelsen, K.S.; Eri, R.; Thomas, L.S.; Hu, B.; Lukasek, K.; Nast, C.C.; Lechago, J.; Xu, R.; Naiki, Y.; et al. Toll-like receptor-4 is required for intestinal response to epithelial injury and limiting bacterial translocation in a murine model of acute colitis. Am. J. Physiol. 2005, 288, G1055–G1065. [Google Scholar]
  183. Vijay-Kumar, M.; Sanders, C.J.; Taylor, R.T.; Kumar, A.; Aitken, J.D.; Sitaraman, S.V.; Neish, A.S.; Uematsu, S.; Akira, S.; Williams, I.R.; et al. Deletion of TLR5 results in spontaneous colitis in mice. J. Clin. Investig. 2007, 117, 3909–3921. [Google Scholar]
  184. Lee, J.; Mo, J.H.; Katakura, K.; Alkalay, I.; Rucker, A.N.; Liu, Y.T.; Lee, H.K.; Shen, C.; Cojocaru, G.; Shenouda, S.; et al. Maintenance of colonic homeostasis by distinctive apical TLR9 signalling in intestinal epithelial cells. Nat. Cell Biol. 2006, 8, 1327–1336. [Google Scholar] [CrossRef]
  185. Ey, B.; Eyking, A.; Klepak, M.; Salzman, N.H.; Gothert, J.R.; Runzi, M.; Schmid, K.W.; Gerken, G.; Podolsky, D.K.; Cario, E. Loss of TLR2 worsens spontaneous colitis in MDR1A deficiency through commensally induced pyroptosis. J. Immunol. 2013, 190, 5676–5688. [Google Scholar]
  186. Rachmilewitz, D.; Katakura, K.; Karmeli, F.; Hayashi, T.; Reinus, C.; Rudensky, B.; Akira, S.; Takeda, K.; Lee, J.; Takabayashi, K.; et al. Toll-like receptor 9 signaling mediates the anti-inflammatory effects of probiotics in murine experimental colitis. Gastroenterology 2004, 126, 520–528. [Google Scholar] [CrossRef]
  187. Cario, E.; Gerken, G.; Podolsky, D.K. Toll-like receptor 2 enhances ZO-1-associated intestinal epithelial barrier integrity via protein kinase C. Gastroenterology 2004, 127, 224–238. [Google Scholar] [CrossRef]
  188. Cario, E.; Gerken, G.; Podolsky, D.K. Toll-like receptor 2 controls mucosal inflammation by regulating epithelial barrier function. Gastroenterology 2007, 132, 1359–1374. [Google Scholar] [CrossRef]
  189. Fukata, M.; Chen, A.; Klepper, A.; Krishnareddy, S.; Vamadevan, A.S.; Thomas, L.S.; Xu, R.; Inoue, H.; Arditi, M.; Dannenberg, A.J.; et al. Cox-2 is regulated by toll-like receptor-4 (TLR4) signaling: Role in proliferation and apoptosis in the intestine. Gastroenterology 2006, 131, 862–877. [Google Scholar] [CrossRef]
  190. Zheng, L.; Riehl, T.E.; Stenson, W.F. Regulation of colonic epithelial repair in mice by toll-like receptors and hyaluronic acid. Gastroenterology 2009, 137, 2041–2051. [Google Scholar] [CrossRef]
  191. Podolsky, D.K.; Gerken, G.; Eyking, A.; Cario, E. Colitis-associated variant of TLR2 causes impaired mucosal repair because of TFF3 deficiency. Gastroenterology 2009, 137, 209–220. [Google Scholar] [CrossRef]
  192. Fukata, M.; Abreu, M.T. TLR4 signalling in the intestine in health and disease. Biochem. Soc. Transact. 2007, 35, 1473–1478. [Google Scholar] [CrossRef]
  193. Brandl, K.; Sun, L.; Neppl, C.; Siggs, O.M.; le Gall, S.M.; Tomisato, W.; Li, X.; Du, X.; Maennel, D.N.; Blobel, C.P.; et al. MYD88 signaling in nonhematopoietic cells protects mice against induced colitis by regulating specific EGF receptor ligands. Proc. Natl. Acad. Sci. USA 2010, 107, 19967–19972. [Google Scholar] [CrossRef]
  194. Rose, W.A., 2nd; Sakamoto, K.; Leifer, C.A. TLR9 is important for protection against intestinal damage and for intestinal repair. Sci. Rep. 2012, 2, 574. [Google Scholar]
  195. Lin, N.; Xu, L.F.; Sun, M. The protective effect of trefoil factor 3 on the intestinal tight junction barrier is mediated by toll-like receptor 2 via a PI3K/AKT dependent mechanism. Biochem. Biophys. Res. Commun. 2013, 440, 143–149. [Google Scholar] [CrossRef]
  196. Pull, S.L.; Doherty, J.M.; Mills, J.C.; Gordon, J.I.; Stappenbeck, T.S. Activated macrophages are an adaptive element of the colonic epithelial progenitor niche necessary for regenerative responses to injury. Proc. Natl. Acad. Sci. USA 2005, 102, 99–104. [Google Scholar]
  197. Ting, J.P.; Lovering, R.C.; Alnemri, E.S.; Bertin, J.; Boss, J.M.; Davis, B.K.; Flavell, R.A.; Girardin, S.E.; Godzik, A.; Harton, J.A.; et al. The NLR gene family: A standard nomenclature. Immunity 2008, 28, 285–287. [Google Scholar] [CrossRef]
  198. Moore, C.B.; Bergstralh, D.T.; Duncan, J.A.; Lei, Y.; Morrison, T.E.; Zimmermann, A.G.; Accavitti-Loper, M.A.; Madden, V.J.; Sun, L.; Ye, Z.; et al. NLRX1 is a regulator of mitochondrial antiviral immunity. Nature 2008, 451, 573–577. [Google Scholar] [CrossRef]
  199. Tattoli, I.; Carneiro, L.A.; Jehanno, M.; Magalhaes, J.G.; Shu, Y.; Philpott, D.J.; Arnoult, D.; Girardin, S.E. NLRX1 is a mitochondrial NOD-like receptor that amplifies NF-κb and JNK pathways by inducing reactive oxygen species production. EMBO Rep. 2008, 9, 293–300. [Google Scholar] [CrossRef]
  200. Meylan, E.; Tschopp, J. NLRX1: Friend or foe? EMBO Rep. 2008, 9, 243–245. [Google Scholar] [CrossRef]
  201. Chamaillard, M.; Hashimoto, M.; Horie, Y.; Masumoto, J.; Qiu, S.; Saab, L.; Ogura, Y.; Kawasaki, A.; Fukase, K.; Kusumoto, S.; et al. An essential role for NOD2 in host recognition of bacterial peptidoglycan containing diaminopimelic acid. Nat. Immunol. 2003, 4, 702–707. [Google Scholar] [CrossRef]
  202. Girardin, S.E.; Boneca, I.G.; Carneiro, L.A.; Antignac, A.; Jehanno, M.; Viala, J.; Tedin, K.; Taha, M.K.; Labigne, A.; Zahringer, U.; et al. NOD2 detects a unique muropeptide from gram-negative bacterial peptidoglycan. Science 2003, 300, 1584–1587. [Google Scholar] [CrossRef]
  203. McDonald, C.; Inohara, N.; Nunez, G. Peptidoglycan signaling in innate immunity and inflammatory disease. J. Biol. Chem. 2005, 280, 20177–20180. [Google Scholar] [CrossRef]
  204. Inohara, N.; Koseki, T.; Lin, J.; del Peso, L.; Lucas, P.C.; Chen, F.F.; Ogura, Y.; Nunez, G. An induced proximity model for NF-κb activation in the NOD2/rick and RIP signaling pathways. J. Biol. Chem. 2000, 275, 27823–27831. [Google Scholar]
  205. Park, J.H.; Kim, Y.G.; McDonald, C.; Kanneganti, T.D.; Hasegawa, M.; Body-Malapel, M.; Inohara, N.; Nunez, G. RICK/RIP2 mediates innate immune responses induced through NOD1 and NOD2 but not TLRS. J. Immunol. 2007, 178, 2380–2386. [Google Scholar] [CrossRef]
  206. Sorbara, M.T.; Ellison, L.K.; Ramjeet, M.; Travassos, L.H.; Jones, N.L.; Girardin, S.E.; Philpott, D.J. The protein ATG16L1 suppresses inflammatory cytokines induced by the intracellular sensors NOD1 and NOD2 in an autophagy-independent manner. Immunity 2013, 39, 858–873. [Google Scholar] [CrossRef]
  207. Magalhaes, J.G.; Lee, J.; Geddes, K.; Rubino, S.; Philpott, D.J.; Girardin, S.E. Essential role of RIP2 in the modulation of innate and adaptive immunity triggered by NOD2 and NOD2 ligands. Eur. J. Immunol. 2011, 41, 1445–1455. [Google Scholar] [CrossRef]
  208. Krieg, A.; Correa, R.G.; Garrison, J.B.; le Negrate, G.; Welsh, K.; Huang, Z.; Knoefel, W.T.; Reed, J.C. XIAP mediates NOD signaling via interaction with RIP2. Proc. Natl. Acad. Sci. USA 2009, 106, 14524–14529. [Google Scholar] [CrossRef]
  209. Bertrand, M.J.; Doiron, K.; Labbe, K.; Korneluk, R.G.; Barker, P.A.; Saleh, M. Cellular inhibitors of apoptosis CIAP1 and CIAP2 are required for innate immunity signaling by the pattern recognition receptors NOD1 and NOD2. Immunity 2009, 30, 789–801. [Google Scholar] [CrossRef]
  210. Hasegawa, M.; Fujimoto, Y.; Lucas, P.C.; Nakano, H.; Fukase, K.; Nunez, G.; Inohara, N. A critical role of RICK/RIP2 polyubiquitination in NOD-induced NF-κb activation. EMBO J. 2008, 27, 373–383. [Google Scholar] [CrossRef]
  211. Yang, Y.; Yin, C.; Pandey, A.; Abbott, D.; Sassetti, C.; Kelliher, M.A. NOD2 pathway activation by MDP or mycobacterium tuberculosis infection involves the stable polyubiquitination of RIP2. J. Biol. Chem. 2007, 282, 36223–36229. [Google Scholar]
  212. Tigno-Aranjuez, J.T.; Asara, J.M.; Abbott, D.W. Inhibition of RIP2’s tyrosine kinase activity limits NOD2-driven cytokine responses. Genes Dev. 2010, 24, 2666–2677. [Google Scholar] [CrossRef]
  213. Lecine, P.; Esmiol, S.; Metais, J.Y.; Nicoletti, C.; Nourry, C.; McDonald, C.; Nunez, G.; Hugot, J.P.; Borg, J.P.; Ollendorff, V. The NOD2-RICK complex signals from the plasma membrane. J. Biol. Chem. 2007, 282, 15197–15207. [Google Scholar] [CrossRef]
  214. da Silva Correia, J.; Miranda, Y.; Leonard, N.; Hsu, J.; Ulevitch, R.J. Regulation of NOD2-mediated signaling pathways. Cell Death Differ. 2007, 14, 830–839. [Google Scholar] [CrossRef]
  215. Yeretssian, G.; Correa, R.G.; Doiron, K.; Fitzgerald, P.; Dillon, C.P.; Green, D.R.; Reed, J.C.; Saleh, M. Non-apoptotic role of BID in inflammation and innate immunity. Nature 2011, 474, 96–99. [Google Scholar] [CrossRef]
  216. Lipinski, S.; Grabe, N.; Jacobs, G.; Billmann-Born, S.; Till, A.; Hasler, R.; Aden, K.; Paulsen, M.; Arlt, A.; Kraemer, L.; et al. Rnai screening identifies mediators of NOD2 signaling: Implications for spatial specificity of mdp recognition. Proc. Natl. Acad. Sci. USA 2012, 109, 21426–21431. [Google Scholar] [CrossRef]
  217. Warner, N.; Burberry, A.; Franchi, L.; Kim, Y.G.; McDonald, C.; Sartor, M.A.; Nunez, G. A genome-wide sirna screen reveals positive and negative regulators of the NOD2 and NF-κb signaling pathways. Sci. Signal. 2013, 6. [Google Scholar] [CrossRef]
  218. Askari, N.; Correa, R.G.; Zhai, D.; Reed, J.C. Expression, purification, and characterization of recombinant NOD2 (NLRC1): A NLR family member. J. Biotechnol. 2012, 157, 75–81. [Google Scholar] [CrossRef]
  219. Kim, Y.G.; Kamada, N.; Shaw, M.H.; Warner, N.; Chen, G.Y.; Franchi, L.; Nunez, G. The NOD2 sensor promotes intestinal pathogen eradication via the chemokine CCL2-dependent recruitment of inflammatory monocytes. Immunity 2011, 34, 769–780. [Google Scholar] [CrossRef]
  220. Correa, R.G.; Milutinovic, S.; Reed, J.C. Roles of NOD1 (NLRC1) and NOD2 (NLRC2) in innate immunity and inflammatory diseases. Biosci. Rep. 2012, 32, 597–608. [Google Scholar] [CrossRef]
  221. Ogura, Y.; Inohara, N.; Benito, A.; Chen, F.F.; Yamaoka, S.; Nunez, G. NOD2, a NOD2/APAF-1 family member that is restricted to monocytes and activates NF-κb. J. Biol. Chem. 2001, 276, 4812–4818. [Google Scholar]
  222. Barnich, N.; Aguirre, J.E.; Reinecker, H.C.; Xavier, R.; Podolsky, D.K. Membrane recruitment of NOD2 in intestinal epithelial cells is essential for nuclear factor-κb activation in muramyl dipeptide recognition. J. Cell Biol. 2005, 170, 21–26. [Google Scholar] [CrossRef]
  223. Ogura, Y.; Lala, S.; Xin, W.; Smith, E.; Dowds, T.A.; Chen, F.F.; Zimmermann, E.; Tretiakova, M.; Cho, J.H.; Hart, J.; et al. Expression of NOD2 in paneth cells: A possible link to crohn’s ileitis. Gut 2003, 52, 1591–1597. [Google Scholar] [CrossRef]
  224. Hisamatsu, T.; Suzuki, M.; Reinecker, H.C.; Nadeau, W.J.; McCormick, B.A.; Podolsky, D.K. CARD15/NOD2 functions as an antibacterial factor in human intestinal epithelial cells. Gastroenterology 2003, 124, 993–1000. [Google Scholar]
  225. Leung, C.H.; Lam, W.; Ma, D.L.; Gullen, E.A.; Cheng, Y.C. Butyrate mediates nucleotide-binding and oligomerisation domain NOD2-dependent mucosal immune responses against peptidoglycan. Eur. J. Immunol. 2009, 39, 3529–3537. [Google Scholar] [CrossRef]
  226. Rosenstiel, P.; Fantini, M.; Brautigam, K.; Kuhbacher, T.; Waetzig, G.H.; Seegert, D.; Schreiber, S. TNF-α and IFN-γ regulate the expression of the NOD2 (CARD15) gene in human intestinal epithelial cells. Gastroenterology 2003, 124, 1001–1009. [Google Scholar]
  227. Ogura, Y.; Bonen, D.K.; Inohara, N.; Nicolae, D.L.; Chen, F.F.; Ramos, R.; Britton, H.; Moran, T.; Karaliuskas, R.; Duerr, R.H.; et al. A frameshift mutation in NOD2 associated with susceptibility to crohn’s disease. Nature 2001, 411, 603–606. [Google Scholar] [CrossRef]
  228. Hampe, J.; Cuthbert, A.; Croucher, P.J.; Mirza, M.M.; Mascheretti, S.; Fisher, S.; Frenzel, H.; King, K.; Hasselmeyer, A.; MacPherson, A.J.; et al. Association between insertion mutation in NOD2 gene and crohn’s disease in german and british populations. Lancet 2001, 357, 1925–1928. [Google Scholar] [CrossRef]
  229. Billmann-Born, S.; Till, A.; Arlt, A.; Lipinski, S.; Sina, C.; Latiano, A.; Annese, V.; Hasler, R.; Kerick, M.; Manke, T.; et al. Genome-wide expression profiling identifies an impairment of negative feedback signals in the crohn’s disease-associated NOD2 variant L1007fsinsC. J. Immunol. 2011, 186, 4027–4038. [Google Scholar] [CrossRef]
  230. McGovern, D.P.; Hysi, P.; Ahmad, T.; van Heel, D.A.; Moffatt, M.F.; Carey, A.; Cookson, W.O.; Jewell, D.P. Association between a complex insertion/deletion polymorphism in NOD2 (CARD4) and susceptibility to inflammatory bowel disease. Human Mol. Genet. 2005, 14, 1245–1250. [Google Scholar] [CrossRef]
  231. Tremelling, M.; Hancock, L.; Bredin, F.; Sharpstone, D.; Bingham, S.A.; Parkes, M. Complex insertion/deletion polymorphism in NOD2 (CARD4) is not associated with inflammatory bowel disease susceptibility in east anglia panel. Inflamm. Bowel Dis. 2006, 12, 967–971. [Google Scholar] [CrossRef]
  232. McDonald, C.; Chen, F.F.; Ollendorff, V.; Ogura, Y.; Marchetto, S.; Lecine, P.; Borg, J.P.; Nunez, G. A role for erbin in the regulation of NOD2-dependent NF-κb signaling. J. Biol. Chem. 2005, 280, 40301–40309. [Google Scholar] [CrossRef]
  233. Kufer, T.A.; Kremmer, E.; Banks, D.J.; Philpott, D.J. Role for erbin in bacterial activation of NOD2. Infect. Immun. 2006, 74, 3115–3124. [Google Scholar] [CrossRef]
  234. Aktories, K.; Schmidt, G.; Just, I. Rho GTPases as targets of bacterial protein toxins. Biol. Chem. 2000, 381, 421–426. [Google Scholar]
  235. Fukazawa, A.; Alonso, C.; Kurachi, K.; Gupta, S.; Lesser, C.F.; McCormick, B.A.; Reinecker, H.C. GEF-H1 mediated control of NOD2 dependent NF-κb activation by shigella effectors. PLoS Pathog. 2008, 4. [Google Scholar] [CrossRef]
  236. Keestra, A.M.; Winter, M.G.; Auburger, J.J.; Frassle, S.P.; Xavier, M.N.; Winter, S.E.; Kim, A.; Poon, V.; Ravesloot, M.M.; Waldenmaier, J.F.; et al. Manipulation of small Rho GTPases is a pathogen-induced process detected by NOD2. Nature 2013, 496, 233–237. [Google Scholar] [CrossRef]
  237. Boyer, L.; Magoc, L.; Dejardin, S.; Cappillino, M.; Paquette, N.; Hinault, C.; Charriere, G.M.; Ip, W.K.; Fracchia, S.; Hennessy, E.; et al. Pathogen-derived effectors trigger protective immunity via activation of the RAC2 enzyme and the IMD or RIP kinase signaling pathway. Immunity 2011, 35, 536–549. [Google Scholar] [CrossRef]
  238. Watanabe, T.; Asano, N.; Murray, P.J.; Ozato, K.; Tailor, P.; Fuss, I.J.; Kitani, A.; Strober, W. Muramyl dipeptide activation of nucleotide-binding oligomerization domain 2 protects mice from experimental colitis. J. Clin. Investig. 2008, 118, 545–559. [Google Scholar]
  239. Macho Fernandez, E.; Valenti, V.; Rockel, C.; Hermann, C.; Pot, B.; Boneca, I.G.; Grangette, C. Anti-inflammatory capacity of selected lactobacilli in experimental colitis is driven by NOD2-mediated recognition of a specific peptidoglycan-derived muropeptide. Gut 2011, 60, 1050–1059. [Google Scholar] [CrossRef]
  240. Chen, G.Y.; Shaw, M.H.; Redondo, G.; Nunez, G. The innate immune receptor NOD2 protects the intestine from inflammation-induced tumorigenesis. Cancer Res. 2008, 68, 10060–10067. [Google Scholar] [CrossRef]
  241. Watanabe, T.; Kitani, A.; Murray, P.J.; Wakatsuki, Y.; Fuss, I.J.; Strober, W. Nucleotide binding oligomerization domain 2 deficiency leads to dysregulated TLR2 signaling and induction of antigen-specific colitis. Immunity 2006, 25, 473–485. [Google Scholar] [CrossRef]
  242. Barreau, F.; Meinzer, U.; Chareyre, F.; Berrebi, D.; Niwa-Kawakita, M.; Dussaillant, M.; Foligne, B.; Ollendorff, V.; Heyman, M.; Bonacorsi, S.; et al. CARD15/NOD2 is required for peyer’s patches homeostasis in mice. PLoS One 2007, 2. [Google Scholar] [CrossRef]
  243. Natividad, J.M.; Petit, V.; Huang, X.; de Palma, G.; Jury, J.; Sanz, Y.; Philpott, D.; Garcia Rodenas, C.L.; McCoy, K.D.; Verdu, E.F. Commensal and probiotic bacteria influence intestinal barrier function and susceptibility to colitis in Nod1−/−; Nod2−/− mice. Inflamm. Bowel Dis. 2012, 18, 1434–1446. [Google Scholar] [CrossRef]
  244. Couturier-Maillard, A.; Secher, T.; Rehman, A.; Normand, S.; de Arcangelis, A.; Haesler, R.; Huot, L.; Grandjean, T.; Bressenot, A.; Delanoye-Crespin, A.; et al. Nod2-mediated dysbiosis predisposes mice to transmissible colitis and colorectal cancer. J. Clin. Investig. 2013, 123, 700–711. [Google Scholar]
  245. Kobayashi, K.S.; Chamaillard, M.; Ogura, Y.; Henegariu, O.; Inohara, N.; Nunez, G.; Flavell, R.A. NOD2-dependent regulation of innate and adaptive immunity in the intestinal tract. Science 2005, 307, 731–734. [Google Scholar] [CrossRef]
  246. Manichanh, C.; Borruel, N.; Casellas, F.; Guarner, F. The gut microbiota in IBD. Nat. Rev. Gastroenterol. Hepatol. 2012, 9, 599–608. [Google Scholar] [CrossRef]
  247. Petnicki-Ocwieja, T.; Hrncir, T.; Liu, Y.J.; Biswas, A.; Hudcovic, T.; Tlaskalova-Hogenova, H.; Kobayashi, K.S. NOD2 is required for the regulation of commensal microbiota in the intestine. Proc. Natl. Acad. Sci. USA 2009, 106, 15813–15818. [Google Scholar] [CrossRef]
  248. Bouskra, D.; Brezillon, C.; Berard, M.; Werts, C.; Varona, R.; Boneca, I.G.; Eberl, G. Lymphoid tissue genesis induced by commensals through NOD2 regulates intestinal homeostasis. Nature 2008, 456, 507–510. [Google Scholar] [CrossRef]
  249. Geddes, K.; Rubino, S.J.; Magalhaes, J.G.; Streutker, C.; le Bourhis, L.; Cho, J.H.; Robertson, S.J.; Kim, C.J.; Kaul, R.; Philpott, D.J.; et al. Identification of an innate T helper type 17 response to intestinal bacterial pathogens. Nat. Med. 2011, 17, 837–844. [Google Scholar] [CrossRef]
  250. Martinon, F.; Burns, K.; Tschopp, J. The inflammasome: A molecular platform triggering activation of inflammatory caspases and processing of proil-β. Mol. Cell 2002, 10, 417–426. [Google Scholar] [CrossRef]
  251. Kayagaki, N.; Warming, S.; Lamkanfi, M.; Vande Walle, L.; Louie, S.; Dong, J.; Newton, K.; Qu, Y.; Liu, J.; Heldens, S.; et al. Non-canonical inflammasome activation targets caspase-11. Nature 2011, 479, 117–121. [Google Scholar] [CrossRef]
  252. Latz, E.; Xiao, T.S.; Stutz, A. Activation and regulation of the inflammasomes. Nat. Rev. Immunol. 2013, 13, 397–411. [Google Scholar] [CrossRef]
  253. Srinivasula, S.M.; Poyet, J.L.; Razmara, M.; Datta, P.; Zhang, Z.; Alnemri, E.S. The pyrin-card protein asc is an activating adaptor for caspase-1. J. Biol. Chem. 2002, 277, 21119–21122. [Google Scholar]
  254. Agostini, L.; Martinon, F.; Burns, K.; McDermott, M.F.; Hawkins, P.N.; Tschopp, J. NALP3 forms an IL-1β-processing inflammasome with increased activity in muckle-wells autoinflammatory disorder. Immunity 2004, 20, 319–325. [Google Scholar] [CrossRef]
  255. Mariathasan, S.; Newton, K.; Monack, D.M.; Vucic, D.; French, D.M.; Lee, W.P.; Roose-Girma, M.; Erickson, S.; Dixit, V.M. Differential activation of the inflammasome by caspase-1 adaptors ASC and IPAF. Nature 2004, 430, 213–218. [Google Scholar] [CrossRef]
  256. Schroder, K.; Tschopp, J. The inflammasomes. Cell 2010, 140, 821–832. [Google Scholar] [CrossRef]
  257. Chen, G.Y. Role of NLRP6 and NLRP12 in the maintenance of intestinal homeostasis. Eur. J. Immunol. 2014, 44, 321–327. [Google Scholar] [CrossRef]
  258. Tamura, K.; Fukuda, Y.; Sashio, H.; Takeda, N.; Bamba, H.; Kosaka, T.; Fukui, S.; Sawada, K.; Tamura, K.; Satomi, M.; et al. IL18 polymorphism is associated with an increased risk of crohn’s disease. J. Gastroenterol. 2002, 37, 111–116. [Google Scholar] [CrossRef]
  259. Zhernakova, A.; Festen, E.M.; Franke, L.; Trynka, G.; van Diemen, C.C.; Monsuur, A.J.; Bevova, M.; Nijmeijer, R.M.; van Slot, R.; Heijmans, R.; et al. Genetic analysis of innate immunity in crohn’s disease and ulcerative colitis identifies two susceptibility loci harboring CARD9 and IL18RAP. Am. J. Hum. Genet. 2008, 82, 1202–1210. [Google Scholar] [CrossRef]
  260. Bauer, C.; Loher, F.; Dauer, M.; Mayer, C.; Lehr, H.A.; Schonharting, M.; Hallwachs, R.; Endres, S.; Eigler, A. The ice inhibitor pralnacasan prevents DSS-induced colitis in C57BL/6 mice and suppresses IP-10 mRNA but not TNF-α mRNA expression. Dig. Dis. Sci. 2007, 52, 1642–1652. [Google Scholar] [CrossRef]
  261. Siegmund, B.; Fantuzzi, G.; Rieder, F.; Gamboni-Robertson, F.; Lehr, H.A.; Hartmann, G.; Dinarello, C.A.; Endres, S.; Eigler, A. Neutralization of interleukin-18 reduces severity in murine colitis and intestinal IFN-γ and TNF-α production. Am. J. Physiol. Regul. 2001, 281, R1264–R1273. [Google Scholar]
  262. Ishikura, T.; Kanai, T.; Uraushihara, K.; Iiyama, R.; Makita, S.; Totsuka, T.; Yamazaki, M.; Sawada, T.; Nakamura, T.; Miyata, T.; et al. Interleukin-18 overproduction exacerbates the development of colitis with markedly infiltrated macrophages in interleukin-18 transgenic mice. J. Gastroenterol. Hepatol. 2003, 18, 960–969. [Google Scholar] [CrossRef]
  263. Dupaul-Chicoine, J.; Yeretssian, G.; Doiron, K.; Bergstrom, K.S.; McIntire, C.R.; LeBlanc, P.M.; Meunier, C.; Turbide, C.; Gros, P.; Beauchemin, N.; et al. Control of intestinal homeostasis, colitis, and colitis-associated colorectal cancer by the inflammatory caspases. Immunity 2010, 32, 367–378. [Google Scholar] [CrossRef]
  264. Zaki, M.H.; Boyd, K.L.; Vogel, P.; Kastan, M.B.; Lamkanfi, M.; Kanneganti, T.D. The NLRP3 inflammasome protects against loss of epithelial integrity and mortality during experimental colitis. Immunity 2010, 32, 379–391. [Google Scholar] [CrossRef]
  265. Allen, I.C.; TeKippe, E.M.; Woodford, R.M.; Uronis, J.M.; Holl, E.K.; Rogers, A.B.; Herfarth, H.H.; Jobin, C.; Ting, J.P. The NLRP3 inflammasome functions as a negative regulator of tumorigenesis during colitis-associated cancer. J. Exp. Med. 2010, 207, 1045–1056. [Google Scholar] [CrossRef]
  266. Salcedo, R.; Worschech, A.; Cardone, M.; Jones, Y.; Gyulai, Z.; Dai, R.M.; Wang, E.; Ma, W.; Haines, D.; O’HUigin, C.; et al. MYD88-mediated signaling prevents development of adenocarcinomas of the colon: Role of interleukin 18. J. Exp. Med. 2010, 207, 1625–1636. [Google Scholar] [CrossRef]
  267. Zaki, M.H.; Vogel, P.; Body-Malapel, M.; Lamkanfi, M.; Kanneganti, T.D. IL-18 production downstream of the NLRP3 inflammasome confers protection against colorectal tumor formation. J. Immunol. 2010, 185, 4912–4920. [Google Scholar] [CrossRef]
  268. Hirota, S.A.; Ng, J.; Lueng, A.; Khajah, M.; Parhar, K.; Li, Y.; Lam, V.; Potentier, M.S.; Ng, K.; Bawa, M.; et al. NLRP3 inflammasome plays a key role in the regulation of intestinal homeostasis. Inflamm. Bowel Dis. 2011, 17, 1359–1372. [Google Scholar] [CrossRef]
  269. Bauer, C.; Duewell, P.; Mayer, C.; Lehr, H.A.; Fitzgerald, K.A.; Dauer, M.; Tschopp, J.; Endres, S.; Latz, E.; Schnurr, M. Colitis induced in mice with dextran sulfate sodium (DSS) is mediated by the NLRP3 inflammasome. Gut 2010, 59, 1192–1199. [Google Scholar] [CrossRef] [Green Version]
  270. Hu, B.; Elinav, E.; Huber, S.; Booth, C.J.; Strowig, T.; Jin, C.; Eisenbarth, S.C.; Flavell, R.A. Inflammation-induced tumorigenesis in the colon is regulated by caspase-1 and NLRC4. Proc. Natl. Acad. Sci. USA 2010, 107, 21635–21640. [Google Scholar]
  271. Liu, Z.; Zaki, M.H.; Vogel, P.; Gurung, P.; Finlay, B.B.; Deng, W.; Lamkanfi, M.; Kanneganti, T.D. Role of inflammasomes in host defense against citrobacter rodentium infection. J. Biol. Chem. 2012, 287, 16955–16964. [Google Scholar]
  272. Song-Zhao, G.X.; Srinivasan, N.; Pott, J.; Baban, D.; Frankel, G.; Maloy, K.J. NLRP3 activation in the intestinal epithelium protects against a mucosal pathogen. Mucosal Immunol. 2013, 2013. [Google Scholar] [CrossRef]
  273. Nordlander, S.; Pott, J.; Maloy, K.J. NLRC4 expression in intestinal epithelial cells mediates protection against an enteric pathogen. Mucosal Immunol. 2013, 2013. [Google Scholar] [CrossRef]
  274. Lebeis, S.L.; Powell, K.R.; Merlin, D.; Sherman, M.A.; Kalman, D. Interleukin-1 receptor signaling protects mice from lethal intestinal damage caused by the attaching and effacing pathogen citrobacter rodentium. Infect. Immun. 2009, 77, 604–614. [Google Scholar] [CrossRef]
  275. Takagi, H.; Kanai, T.; Okazawa, A.; Kishi, Y.; Sato, T.; Takaishi, H.; Inoue, N.; Ogata, H.; Iwao, Y.; Hoshino, K.; et al. Contrasting action of Il-12 and Il-18 in the development of dextran sodium sulphate colitis in mice. Scand. J. Gastroenterol. 2003, 38, 837–844. [Google Scholar] [CrossRef]
  276. Siegmund, B. Interleukin-18 in intestinal inflammation: Friend and foe? Immunity 2010, 32, 300–302. [Google Scholar] [CrossRef]
  277. Molofsky, A.B.; Byrne, B.G.; Whitfield, N.N.; Madigan, C.A.; Fuse, E.T.; Tateda, K.; Swanson, M.S. Cytosolic recognition of flagellin by mouse macrophages restricts legionella pneumophila infection. J. Exp. Med. 2006, 203, 1093–1104. [Google Scholar] [CrossRef]
  278. Miao, E.A.; Mao, D.P.; Yudkovsky, N.; Bonneau, R.; Lorang, C.G.; Warren, S.E.; Leaf, I.A.; Aderem, A. Innate immune detection of the type III secretion apparatus through the NLRC4 inflammasome. Proc. Natl. Acad. Sci. USA 2010, 107, 3076–3080. [Google Scholar] [CrossRef]
  279. Miao, E.A.; Alpuche-Aranda, C.M.; Dors, M.; Clark, A.E.; Bader, M.W.; Miller, S.I.; Aderem, A. Cytoplasmic flagellin activates caspase-1 and secretion of interleukin 1β via IPAF. Nat. Immunol. 2006, 7, 569–575. [Google Scholar] [CrossRef]
  280. Carvalho, F.A.; Nalbantoglu, I.; Aitken, J.D.; Uchiyama, R.; Su, Y.; Doho, G.H.; Vijay-Kumar, M.; Gewirtz, A.T. Cytosolic flagellin receptor NLRC4 protects mice against mucosal and systemic challenges. Mucosal Immunol. 2012, 5, 288–298. [Google Scholar]
  281. Elinav, E.; Strowig, T.; Kau, A.L.; Henao-Mejia, J.; Thaiss, C.A.; Booth, C.J.; Peaper, D.R.; Bertin, J.; Eisenbarth, S.C.; Gordon, J.I.; et al. NLRP6 inflammasome regulates colonic microbial ecology and risk for colitis. Cell 2011, 145, 745–757. [Google Scholar] [CrossRef]
  282. Normand, S.; Delanoye-Crespin, A.; Bressenot, A.; Huot, L.; Grandjean, T.; Peyrin-Biroulet, L.; Lemoine, Y.; Hot, D.; Chamaillard, M. NOD-like receptor pyrin domain-containing protein 6 (NLRP6) controls epithelial self-renewal and colorectal carcinogenesis upon injury. Proc. Natl. Acad. Sci. USA 2011, 108, 9601–9606. [Google Scholar] [CrossRef]
  283. Chen, G.Y.; Liu, M.; Wang, F.; Bertin, J.; Nunez, G. A functional role for nlrp6 in intestinal inflammation and tumorigenesis. J. Immunol. 2011, 186, 7187–7194. [Google Scholar]
  284. Zaki, M.H.; Vogel, P.; Malireddi, R.K.; Body-Malapel, M.; Anand, P.K.; Bertin, J.; Green, D.R.; Lamkanfi, M.; Kanneganti, T.D. The NOD-like receptor NLRP12 attenuates colon inflammation and tumorigenesis. Cancer Cell 2011, 20, 649–660. [Google Scholar] [CrossRef]
  285. Allen, I.C.; Wilson, J.E.; Schneider, M.; Lich, J.D.; Roberts, R.A.; Arthur, J.C.; Woodford, R.M.; Davis, B.K.; Uronis, J.M.; Herfarth, H.H.; et al. NLRP12 suppresses colon inflammation and tumorigenesis through the negative regulation of noncanonical NF-κb signaling. Immunity 2012, 36, 742–754. [Google Scholar] [CrossRef]
  286. Wang, L.; Manji, G.A.; Grenier, J.M.; Al-Garawi, A.; Merriam, S.; Lora, J.M.; Geddes, B.J.; Briskin, M.; DiStefano, P.S.; Bertin, J. PYPAF7, a novel pyrin-containing APAF1-like protein that regulates activation of NF-κb and caspase-1-dependent cytokine processing. J. Biol. Chem. 2002, 277, 29874–29880. [Google Scholar] [CrossRef]
  287. Grenier, J.M.; Wang, L.; Manji, G.A.; Huang, W.J.; Al-Garawi, A.; Kelly, R.; Carlson, A.; Merriam, S.; Lora, J.M.; Briskin, M.; et al. Functional screening of five pypaf family members identifies PYPAF5 as a novel regulator of NF-κb and caspase-1. FEBS Lett. 2002, 530, 73–78. [Google Scholar] [CrossRef]
  288. Fernandes-Alnemri, T.; Alnemri, E.S. Assembly, purification, and assay of the activity of the ASC pyroptosome. Method. Enzym. 2008, 442, 251–270. [Google Scholar] [CrossRef]
  289. Anand, P.K.; Malireddi, R.K.; Lukens, J.R.; Vogel, P.; Bertin, J.; Lamkanfi, M.; Kanneganti, T.D. NLRP6 negatively regulates innate immunity and host defence against bacterial pathogens. Nature 2012, 488, 389–393. [Google Scholar] [CrossRef]
  290. Lich, J.D.; Williams, K.L.; Moore, C.B.; Arthur, J.C.; Davis, B.K.; Taxman, D.J.; Ting, J.P. Monarch-1 suppresses non-canonical NF-κb activation and p52-dependent chemokine expression in monocytes. J. Immunol. 2007, 178, 1256–1260. [Google Scholar] [CrossRef]
  291. Williams, K.L.; Lich, J.D.; Duncan, J.A.; Reed, W.; Rallabhandi, P.; Moore, C.; Kurtz, S.; Coffield, V.M.; Accavitti-Loper, M.A.; Su, L.; et al. The caterpiller protein monarch-1 is an antagonist of toll-like receptor-, tumor necrosis factor α-, and mycobacterium tuberculosis-induced pro-inflammatory signals. J. Biol. Chem. 2005, 280, 39914–39924. [Google Scholar] [CrossRef]
  292. Wlodarska, M.; Thaiss, C.A.; Nowarski, R.; Henao-Mejia, J.; Zhang, J.P.; Brown, E.M.; Frankel, G.; Levy, M.; Katz, M.N.; Philbrick, W.M.; et al. Nlrp6 inflammasome orchestrates the colonic host-microbial interface by regulating goblet cell mucus secretion. Cell 2014, 156, 1045–1059. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Parlato, M.; Yeretssian, G. NOD-Like Receptors in Intestinal Homeostasis and Epithelial Tissue Repair. Int. J. Mol. Sci. 2014, 15, 9594-9627. https://doi.org/10.3390/ijms15069594

AMA Style

Parlato M, Yeretssian G. NOD-Like Receptors in Intestinal Homeostasis and Epithelial Tissue Repair. International Journal of Molecular Sciences. 2014; 15(6):9594-9627. https://doi.org/10.3390/ijms15069594

Chicago/Turabian Style

Parlato, Marianna, and Garabet Yeretssian. 2014. "NOD-Like Receptors in Intestinal Homeostasis and Epithelial Tissue Repair" International Journal of Molecular Sciences 15, no. 6: 9594-9627. https://doi.org/10.3390/ijms15069594

Article Metrics

Back to TopTop