Summary
Advancements in human genetics now poise the field to illuminate the pathophysiology of complex genetic disease. In particular, genome-wide association studies have generated insights into the mechanisms driving inflammatory bowel disease (IBD) and implicated genes shared by multiple autoimmune and autoinflammatory diseases. Thus, emerging evidence suggests a central role for the mucosal immune system in mediating immune homeostasis and highlights the complexity of genetic and environmental interactions that collectively modulate risk of disease. Nevertheless, the challenge remains to determine how genetic variation can precipitate and sustain the inappropriate inflammatory response to commensals that is observed in IBD. Here, we highlight recent advancements in immunogenetics and provide a forward-looking view of the innovations that will deliver mechanistic insights from human genetics.
Keywords: Inflammatory Bowel Disease (IBD), Crohn’s disease, ulcerative colitis, genome wide association study (GWAS), genomics, mucosal immunity, host defense
Genetics of IBD
Genetic predisposition to autoimmune and autoinflammatory diseases is well established, yet the interrelationship between multiple genetic components and environmental triggers remains to be elucidated. The observation that some individuals develop IBD at childhood and others during adulthood, suggests distinct environmental contributions to disease initiation versus disease progression. While the genetic elements predisposing to IBD are present from birth, disease onset occurs later in life, and there is further need to understand how host genetic and epigenetic factors interact with environmental triggers at disease onset and how genetics influence immune regulatory networks that sustain disease. The most recent advancement in the field came from a meta analysis of 15 prior GWA studies to aid in the design of the Immunochip [1]. This custom designed genotyping array contains high-density genomic coverage of SNPs implicated by GWAS to fine map disease loci and implicate new genes. In combination with imputed GWAS data, Immunochip validation studies on over 25,000 IBD cases identified 71 new loci for a total of 163 loci associated with IBD [1]. As a result, these studies substantially increase the estimated disease heritability for both Crohn’s Disease (CD) and Ulcerative Colitis (UC). Furthermore, they highlight 110 shared loci for both disease subtypes, while 30 are classified as specific for CD and 23 for UC [1]. IBD shares the largest number of loci with type I diabetes and shows substantial enrichment of overlap with ankylosing spondylitis, psoriasis, and susceptibility to mycobacterial infection [1]. Such overlap in the genetics of autoimmune/autoinflammatory diseases points to several common immune processes, such as regulation of mucosal immunity, in mediating inflammatory pathology. Together, the immediate impact of implicating 163 loci in IBD resulted in identification of new candidate genes and pathways that may drive disease. However, taking human genetic studies a step further requires building new experimental systems to define the biological functions of these genes. For example, genetic studies in psoriasis patients implicated Act1 (Traf3ip2), an adaptor functioning downstream of the IL-17R [2–4]. Generation of a knockout mouse that modeled TRAF3IP2 loss of function variants in humans revealed hyperactive TH17 responses driving IL-22-dependent inflammation [5]. Thus, GWAS led to the development of a novel mouse model that provided key insight into disease mechanisms. Here, we discuss how advancements in genomics and functional genetics have contributed to our current understanding of IBD, and how these approaches can be applied to provide new mechanistic insights and therapeutic opportunities (Figure 1).
Figure 1. From Genetics to Disease Mechanisms.
Advancements in genomics technology have facilitated the discovery of 163 loci associated with IBD. Identifying causal mutations within coding and noncoding regions of these loci has begun to reveal gene function and shed new light on disease mechanisms. Progressing from genetics to mechanistic insight is accelerating as a result of new technology in the field of functional genetics.
Genetic Variation and Functional Repercussions
Many single nucleotide polymorphisms (SNPs) implicated by GWAS are not directly causal with respect to phenotype, rather they exist in linkage disequilibrium with yet to be discovered variants that are functional. This point highlights the fact that GWAS cover relatively common genetic variants including SNPs with >1% frequency within the population and fail to capture rare or undiscovered variants. Such germline encoded DNA variation can lead to nonsynonymous coding variants that change protein function and/or posttranslational regulation. However, the majority of the SNPs implicated by GWAS represent variation in noncoding regions of the genome. Thus, alterations in gene expression are likely to be important factors in immune dysregulation, and highlight the need to comprehensively characterize DNA and RNA regulatory elements.
Genetic variation impacts gene expression at the level of transcription, RNA stability, splicing, and epigenetic modification. Accordingly, new genomics tools have been developed to capture genetic regulation at multiple levels. Studies merging proteomics with genomics identified differential binding of transcription factors to SNPs in the IL2RA promoter that regulate gene expression [6]. As an additional mechanism of transcriptional regulation, long noncoding (lnc) RNAs within the genome have been shown to play key roles in modulating gene expression [7, 8]. Recent studies have identified a critical requirement for lncRNA in maintaining pluripotency and enforcing lineage specific gene expression profiles [9]. It is now clear that expression of lncRNAs is regulated in a cell type-specific manner. Thus, analysis of the human beta cell transcriptome identified tissue-specific lncRNAs that were transcriptionally dysregulated in type 2 diabetes (T2D) or that mapped to loci previously associated with T2D [10]. In the context of host defense, the lncRNA NeST has been identified as an epigenetic regulator of IFN-γ expression in CD8 T cells and determines susceptibility to viral and bacterial infection in mice [11]. However, the potential role for lncRNAs in regulating human inflammatory diseases remains largely unexplored.
Following transcription, mRNA splicing and stability are tightly regulated. In this context, inflammatory stimuli and microbial components induce a coordinated program of miRNA expression in monocytes that regulates inflammatory responses [12]. Consequently, variants in the gene encoding IL23R that associate with IBD are resistant to downmodulation by miRNA, which results in upregulation of IL23R expression [13]. Similarly, variants in the 3’UTR of genes encoding CTLA-4 and IL-10 alter recognition by miRNAs, resulting in dysregulated expression [14]. In addition to regulation at the level of mRNA stability, RNA splicing impacts gene function. SNPs in the 3’ untranslated region of Ctla4 associate with T1D and correlate with reduced expression of a soluble CTLA-4 splice variant [15]. Accordingly, transgenic mice with germline integration of a cassette encoding shRNA specifically targeting soluble CTLA-4 impaired Treg function and resulted in accelerated onset of diabetes in the NOD system and severe colitis in the CD45RBhi transfer model [16].
Determining the impact of genetic variation on expression profiles has been a major focus in the field, as highlighted in studies of expression quantitative trait loci (eQTL) that aim to correlate SNPs with transcriptome data [17]. Recent studies have begun to identify SNPs that alter transcription of neighboring genes (cis eQTLs) and distant genes (trans eQTLs). Analysis of 92 strains of inbred mice identified several thousand eQTLs associated with macrophage responses to LPS [18]. Similarly, eQTL studies in human dendritic cells (DCs) identified 198 eQTLs associated with Mycobacterium tuberculosis infection [19]. These findings are notable given the genetic overlap between susceptibility to Mycobacterium tuberculosis infection and IBD and given the enrichment of IBD candidate genes in DCs [1]. In a direct example of eQTLs in IBD, SNPs associated with CD correlate with increased expression of Card9, which has been suggested to potentiate inflammatory signaling cascades [1]. Similarly, SNPs associated with T1D have been shown to act in trans on an antiviral inflammatory network driven by IRF7 in monocytes [20]. While many of the eQTLs highlighted above are cell type-specific and the mechanistic basis of this specificity is not entirely clear, it is likely to be regulated by differential pathway activity and epigenetic effects. In this context, the Encyclopedia of DNA Elements (ENCODE) project continues to deposit rich datasets of genome-wide epigenetic modifications across multiple cell types. Recent chromatin mapping studies highlight SNPs associated with autoimmunity that are enriched within enhancer elements containing epigenetic modifications in specific cell types [21, 22].
As the number of noncoding variants and SNPs associated with autoimmune disease has expanded, so has identification of coding variants. Exome sequencing at IBD loci has identified novel coding variants and helped pinpoint specific genes within these loci that are likely to impact disease [23]. For example, many IBD loci are gene-dense, and SNP signals are not able to precisely pinpoint the causative gene. A locus on chromosome 12 implicated by GWAS is comprised of the LRRK2 gene and MUC19. While a great deal of attention has focused on study of the kinase LRRK2, the discovery of new coding variants in MUC19 indicates the need for more detailed analysis of this gene in IBD pathogenesis [23]. MUC19 is a gel-forming apomucin, thus genetic variants of MUC19 may contribute to mucosal barrier dysfunction by causing quantitative changes in its production or structural changes in the glycoprotein core. The barrier function of the mucous layer acts in concert with its ability to retain antimicrobial effector molecules such as defensins and secreted IgA. Thus, B cell IgA production is a central pathway in mucosal immunity, which is consistent with a newly discovered role for the IBD candidate gene BACH2 in class switch recombination [24]. In fact, exome sequencing recently defined coding variants in BACH2 and identified a distinct profile in severe UC characterized by mutations in both BACH2 and IL10 [25]. Although B cell function was not directly tested in this study, BACH2 and IL-10 can induce immunoglobulin class switching [24, 26], thus highlighting the potential of gene-gene interactions to impact cell type-specific phenotypes. Collectively, these studies demonstrate the diverse mechanisms by which genetic variation impacts gene function and highlight the importance of genetic interactions.
Addressing Genetic Epistasis
The diversity of genetic backgrounds in human subjects complicates phenotyping endeavors, but highlights the notion that phenotypes derive from the aggregate effects of multiple genetic factors. Furthermore, IBD genes interact, as highlighted by the discovery of a MAST3-regulated transcriptional program that broadly controls expression of inflammatory genes associated with NFkB activity, such as those induced by NOD2 and TLRs [27, 28]. It is now possible to discern the impact of genetic perturbation on transcriptional programs by employing RNAi-mediated knockdown of candidate genes followed by RNAseq. Accordingly, this approach may provide a deeper understanding of how IBD genes interact with one another at the genetic level. Although genetic epistasis is thought to be an important component of disease, it is not captured well by GWAS [29]. Nevertheless, studies have shown that T1D risk alleles for HLA class II and Ctla4 statistically interact and support a role for gene-gene interactions [30]. Epistasis is difficult to study systematically, because of the sheer number of possible genetic contributors. Furthermore, perturbing multiple genes simultaneously has practical limits, and addressing the issue of genetic epistasis in complex disease will ultimately require an unbiased discovery-based approach invoking GWAS on a much larger scale.
Exploring Gene-Environment Interactions
Following the logic that host genetic epistasis contributes to disease, emerging evidence indicates that host genetic interactions with microbes also shapes immunologic phenotypes. Twin studies suggest that infant and early childhood infections may be associated with IBD [31]. In addition, stable interactions of commensal communities with the host shape immune development, as demonstrated by the observation that segmented filamentous bacteria (SFB) promote development of Th17 cells in mice [32]. Similarly, innate lymphoid cell (ILC) development is shaped by the microbiome [33], and ILCs accumulate in inflamed mucosal tissues [34]. Dietary factors also promote ILC development, as aryl hydrocarbon receptor (AHR) ligands induce lymphoid follicle development in the intestine by acting on ILCs [35]. Conversely, ILCs shape the composition of the microbiome by limiting dissemination of lymphoid resident commensals [36]. These key observations begin to provide insight into the interactions between host immune system, microbiome, and dietary metabolites [37].
In addition to stable host-commensal interactions, transient interactions between host and pathogens irrevocably change the immune system. Prior pathogen exposure elicits more robust responses to subsequent infection. While memory within the adaptive immune system provides antigen-specific protection upon rechallenge, the innate immune system can be “trained” by prior infection and poised to respond more robustly to a variety of pathogens [38–41]. Thus, previous infection may condition the immune system to mount a pathologic response, which is not to suggest that IBD is caused by a particular pathogen. Rather, emerging evidence suggests a “multiple hit” model for conferring susceptibility to IBD [42]. Similar to IBD patients bearing ATG16L1 risk alleles, Atg16l1 hypomorphic mice exhibit paneth cell defects associated with abnormalities in granule packaging. Importantly, pathology associated with paneth cell function depended on prior exposure to virus [42]. Given the interconnection between host and microbes, the future challenge is to catalogue the metagenome and correlate host genotype to alterations in the microbiome. Still more challenging is the prospect of determining past history of pathogen infection in patients and identifying the immunologic consequences of prior infection. Notably, clues to prior pathogen exposure may be encoded in the specificity of gut IgA [43].
Extrapolating Pathways from Human Genetic Data
To date, GWAS studies have identified 163 loci associated with IBD [1]. Using pathway analysis tools to examine genes within these loci immediately reveals patterns related to immunity. For example, cytokines and their respective receptors are abundantly represented. Upon closer examination, additional immunoregulatory pathways can be distilled and implicate genetic programs associated with Treg function, Th17 development, negative regulation of TCR signaling, innate pathogen sensing pathways, antigen presentation, and apoptosis (Figure 2). While the function of caspases in eliciting apoptosis of T cells during activation-induced cell death is well characterized, caspases have additional functions in inflammasome signaling and processing of IL-1 family cytokines. Given that dysregulated caspase activity and inflammasome function have been associated with IBD, study of additional caspase-dependent pathways and caspase substrates may reveal novel insights into IBD pathophysiology [44].
Figure 2. IBD Pathways and Key Cell Types.
Genetic studies have helped to identify critical biological processes and cell types that regulate intestinal inflammation. Genetic variation that perturbs any stage of immune homeostasis, inflammation, or resolution can result in the inappropriate inflammatory response to commensals that is observed in IBD. M cell (M), antibody secreting cells (ASC), dendritic cell (DC), retinoic acid (RA), T cell (T), T regulatory cell (Treg), innate lymphoid cell (ILC), goblet cell (GC), paneth cell (PC), antimicrobial proteins (AMPs), neutrophil (Nφ), inflammatory monocyte (Mo), macrophage (Mφ), reactive oxygen species (ROS), nitric oxide (NO), matrix metalloproteinase (MMP).
Gene annotation and functional associations have also highlighted key pathways that surpass the traditional boundaries of immunology and would not have been easily predicted by previous studies (Box 1). Paramount among these are cellular processes related to autophagy, redox signaling and ER stress. Emerging evidence points to multiple pathway interactions; for example, ER stress and autophagy mutually regulate one another to maintain metabolic homeostasis during inflammation and to promote antimicrobial effector responses [45, 46]. In addition, macrophage exposure to bacterial lipopolysaccharide induces a metabolic switch to glycolysis resulting in accumulation of succinate that promotes IL-1p production by stabilizing HIF1-a (Tannahill et al. Nature, in press). Conversely, metabolism impinges upon the immune system by regulating tolerance. Tryptophan catabolism mediated by indoleamine 2,3-dioxygenase(IDO) generates kynurenine, which promotes IL-10 production by DCs and facilitates Treg differentiation, thus inhibiting Th17 responses [47]. Th17 differentiation is also regulated by environmental factors such as NaCl derived from dietary sources. In this context, elevated levels of salt stimulate expression of SGK1, which in turn induces expression of IL-23R to enhance Th17 differentiation [48]. Further work elucidating Th17 cytokine networks has uncovered previously unrecognized connections between IL-17 and maintenance of epithelial barrier function [49], thus indicating an important role for the immune system in maintaining epithelial homeostasis and restitution. In particular, neutrophil influx into inflamed tissues initially amplifies inflammation, and later promotes an anti-inflammatory healing response through macrophage-dependent clearance of apoptotic neutrophils [50]. In this context, mice deficient in the NADPH oxidase subunit P40phox exhibit an impairment in the healing response following acute colonic inflammation [51], thus revealing an additional role for reactive oxygen species (ROS) in epithelial restitution in addition to its known role as an antimicrobial agent. The importance of ROS in mucosal immunity is further supported by the discovery of rare variants of NCF4 and NCF2 that associate with IBD [52, 53]. These genetic studies implicating IBD genes in the ROS pathway suggest that neutrophils are key cell types in pathology, and further identify new functions for these cells that influence clinical manifestations of disease (Figure 3).
Box 1. Key Genes Implicated in IBD.
Advancements in genomics have identified specific loci and rare coding variants associated with risk or protection from IBD. The genes identified by these studies highlight specific cellular pathways that may contribute to disease onset and/or progression (reviewed in [76]). Here, we highlight candidate IBD genes that implicate additional pathways that collectively suggest connections between cellular metabolism, inflammation, and mucosal microbial communities.
| Select IBD Genes Identified by ImmunoChip [1] | ||
|---|---|---|
| Gene | Locus | Putative Function |
| RNF186 | 1p36.13 | Highly expressed in intestine and contains a RING-type zinc finger that may function as a ubiquitin ligase. Association with IBD has been validated in several populations [77, 78]. Evidence suggests genetic interaction with another IBD gene, HNF4A [79]. |
| SP110 | 2q37.1 | Associated with primary immunodeficiency. Expressed in hematopoietic cells and contains a bromodomain with potential involvement in epigenetic regulation. Loss of function mutations can decrease IL-10 production by B cells [80]. |
| SP140 | 2q37.1 | Expressed in hematopoietic cells and contains a bromodomain with potential involvement in epigenetic regulation. |
| MST1 | 3p21 | Hepatocyte growth factor-like protein produced in the liver. Activates the receptor tyrosine kinase MST1R on epithelial cells (and some subsets of macrophages). Gain of function variants enhance macrophage motility [81]. |
| FUT2 | 19q13.3 | Golgi protein expressed in gastrointestinal tract. Enzymatic activity generates a secreted oligosaccharide that functions as a substrate for synthesis of A and B blood group antigens. Loss of function mutations (nonsecretor phenotype) lack expression of blood group antigens in mucosal surfaces. Secretor status correlates with alterations in the microbiome [82] and risk of IBD [83] and T1D [84]. |
| SLC22A4 | 5q31.1 | Ergothioneine transporter expressed in intestine and subsets of myeloid cells. May regulate cellular redox state, potentially linking metabolism with inflammatory responses [85]. |
| GSDMB | 17q12 | May be involved in regulation of epithelial cell apoptosis [86]. It is also highly expressed in CD8 T cells. |
| ORMDL3 | 17q12 | Regulates ER stress response associated with inflammation [87]. |
| TNFSF15 | 9q32 | Expressed on endothelial cells and activated APCs. One of its receptors (TNFRSF25) promotes Treg expansion in a ligand-dependent manner [88]. |
| TNFAIP3 | 6q23 | Ubiquitin modifying enzyme expressed in myeloid cells. Negatively regulates NFkB signaling and inflammatory cytokines [72]. |
| SLC6A7 | 5q32 | Proline transporter that may regulate cellular metabolic state and inflammation. |
| IL10RA | 11q23 | Receptor for IL-10 broadly expressed on hematopoietic cells. Transduces immunosuppressive signal through STAT3 and TYK2. Associated with early onset IBD [89]. |
| Select IBD Genes with Coding Variants Identified by Exome Sequencing [23] | ||
| Gene | Locus | Putative function |
| IL23R | 1p31.3 | Receptor for IL-23 expressed predominantly in T cells. Promotes differentiation of pathogenic Th17 cells [90]. |
| CARD9 | 9q34.3 | Expressed in myeloid cells where it promotes activation of NFkB and inflammatory cytokines downstream of pattern recognition receptors (PRRs) that are associated with immunoreceptor tyrosine-based activation motifs (ITAMs) or hemi-ITAMs [91]. Promotes cytokine environment conducive to Th17 differentiation. |
| NOD2 | 16q21 | Intracellular PRR specific for bacterial peptidoglycans and is expressed in myeloid cells. Activates NFkB and promotes inflammatory cytokines. Can induce bacterial killing in an autophagy-dependent manner [92]. |
| IL18RAP | 2q12 | Accessory protein for IL-18 receptor expressed on NK and T cells. Promotes stimulatory effect of IL-18 on T cell IFN-γ production [93]. |
| MUC19 | 12q12 | Gel-forming mucin expressed in epithelial tissues. Potential role in barrier function and interaction with microbial communities. |
| CUL2 | 10p11.21 | Component of E3 ubiquitin-protein ligase complex potentially linking proteosomal system with autophagy. |
| PTPN22 | 1p13.2 | Protein tyrosine phosphatase that regulates T and B cell responses at the level of antigen receptor signaling [94]. |
| C1orf106 | 1q32.1 | Expressed in epithelial cells of the gastrointestinal tract. May promote epithelial integrity and barrier function. |
Figure 3. Molecular Pathogenesis of IBD: Assembling the Evidence.
Identification of genes in the ROS pathway point to neutrophils as key cellular mediators of IBD, while functional studies implicated ROS in antibacterial defense, autophagy, and epithelial healing. Thus, genetics can be used to define a mechanism-based subset of IBD patients and guide treatment.
Pathway interactions identified by GWAS implicate several cell types in inflammatory pathology. By cross referencing candidate disease genes and pathways with gene expression patterns in immune cell subsets, DCs and memory T cells feature prominently as central cell types contributing to IBD [1]. Consistent with this notion, Ly6Chi monocytes in the inflamed colon generate inflammatory DCs and antigen presenting DCs that drive autoinflammatory T cell responses [54]. Additional evidence implicates ILCs as key mediators of host defense and inflammatory pathology [33, 55]. Furthermore, expression of IBD genes identified by GWAS highlight the gut epithelium and innate immune mechanisms including barrier function, goblet cell secretion of mucins, and paneth cell secretion of antimicrobial mediators [56]. With several functional pathways implicated in IBD, the next challenge will be to place unannotated genes in pathways that drive disease and to elucidate regulatory networks.
New Approaches in Assigning Function to Genes
Although many of the candidate IBD genes implicated by GWAS can be assembled into known pathways, a substantial fraction of these genes (> 40%) are poorly characterized at the functional level. A significant challenge is to identify functions for candidate IBD genes and to determine how signaling pathways work together to regulate mucosal immune effector mechanisms. One can posit that candidate IBD genes with unknown function may control key immune functions that drive disease, and that assigning function to these genes will expand our understanding of immunoregulatory mechanisms. To meet the challenge of assigning functions to candidate IBD genes in relevant immune responses, high throughput RNAi screening approaches have been developed. Implementing RNAi screens has shown initial signs of success, but has yet to be systematically applied to assign IBD candidate genes with immunologic function. As a recent example, RNAi screening was used to identify mediators of innate pathogen recognition through the NOD2 complex. Here, screening approaches identified NOD2-dependent regulators of NFkB and further demonstrated a novel mechanism for the spatial assembly of the NOD2 signaling complex [57]. In another recent study, genome wide RNAi screening approaches identified 190 cofactors required for mediating endosomal pathogen sensing pathways mediated by TLR7 and TLR9 [58].
Elucidating Mechanisms
Most functional genetic strategies aimed at assigning function to genes involve knockdown or overexpression of candidates. While this approach can establish the requirement, or define a regulatory role for a specific gene in a defined biological process, it does not effectively capture how genetic variation in human populations impacts immune regulation. For example, coding variants may result in gain of function, loss of function, or exist on a functional continuum somewhere in between. These are important distinctions when attempting to infer disease mechanisms or design new therapeutics. By the same token, determining how genetic variation in risk alleles versus protective alleles impacts a cellular response requires mechanistic insight. The next challenge will be to develop reliable approaches to demonstrate the causal effect of a genetic variant on disease progression and to determine the underlying mechanism of action.
With accessibility of exome sequencing technology on the rise, genetic diversity can be quantified and increasing numbers of coding variants have been identified [59, 60]. It has been estimated that human genomes typically harbor up to 100 loss of function variants and approximately 20 genes that are nonfunctional [60]. Due to limitations in the ability to predict which variants cause a given phenotype and which are benign, rigorous mechanistic studies must necessarily follow. IBD GWAS led to the identification of the autophagy gene ATG16L1 and a putative loss of function coding variant thereof (T300A) [61, 62]. Knockdown of endogenous ATG16L1 in epithelial cells and overexpression of RNAi-resistant ATG16L1 T300A resulted in impaired antibacterial autophagy and formally demonstrated a role for this coding variant in a relevant biological readout [63]. While overexpression of coding variant cDNAs is ideally suited for high throughput analyses of many candidate mutations, overexpression can also mask subtle effects or exaggerate phenotypes. In gene replacement, or exon replacement approaches, a coding variant is introduced into the endogenous locus so as to retain regulation at the level of chromatin remodeling, transcription, and splicing. Knockin mice have proven successful in this regard and can be crossed-bred to introduce mutations at multiple loci. More recent innovations in genome engineering now enable efficient generation of isogenic human cells. Approaches that employ transcription activator like effectors (TALEs) allow for targeting endogenous genes to introduce coding variants [64–68], and the CRISPR/Cas system has been recently adapted to target multiple genes simultaneously [69–71].
Assessing Disease Relevance
In progressing from identification of genetic loci by GWAS to defining functions for disease genes and variants, the next step towards determining the role for a candidate gene in disease pathology requires study in vivo. Here, multiple cell types and signaling pathways coordinately interact to mediate loss of immune tolerance and persistent inflammation. In this context, murine models have earned a prominent position as a mainstay for determining if a candidate gene is necessary or sufficient for disease. Mouse models have been used to demonstrate that the ubiquitin modifying enzyme A20 limits inflammatory responses, as conditional deletion in DCs resulted in spontaneous colitis and spondyloarthritis [72]. Furthermore, the role of DCs in maintaining tolerance is highlighted by the observation that regulatory DCs pulsed with bacterial antigen can mitigate experimental colitis in mice [73].
An emerging concept from mouse models and human genetic studies is that IBD cases are likely comprised of distinct subsets of mechanism-driven disorders [74]. In this context, systematic immunophenotyping of genotyped human cohorts will be required to define mechanism-based disease subsets. In disease free individuals bearing a T1D risk allele at the IL2R locus, IL-2R expression was reduced and Treg function was impaired [75]. Here, a rigorous approach with large sample sizes combined with rigorous quality control was implemented to address effects of different genetic backgrounds. In addition, inclusion of healthy controls with disease risk genotypes can help to exclude complicating effects of chronic disease that may alter immune phenotypes.
Concluding Remarks
Human genetics has provided key insights into complex disease. Significant overlap in genes implicated across several autoimmune/autoinflammatory diseases indicates common immunologic mechanisms as well as unique disease-specific pathways that must be tightly regulated to balance host defense against the risk for pathological inflammation. Only with a deeper understanding of the mechanisms driving disease and their underlying genetic components, will the goal of interpreting patient genotypes become feasible. Towards this end, the diagnostic power of genetics bears potential to stratify patient subsets based on disease mechanisms and treat them accordingly. Moreover, progress towards deciphering the genetic components of IBD pathophysiology will identify new points of entry for mechanism-based therapeutics. While it remains a challenge to mitigate pathological inflammation without compromising host defense, advancements in genetics offer the opportunity to treat the underlying mechanisms that incite IBD rather than broadly suppressing immune function.
Highlights.
163 loci have been identified that associate with IBD
Genetic variation within IBD loci may dysregulate immune homeostasis in the gut
Innovations in genomics pave the way to decipher genotype-phenotype relationships
Acknowledgments
The authors are funded by the Center for the Study of Inflammatory Bowel Diseases (Massachusetts General Hospital), the Helmsley Trust, and NIH R01 DK092405.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Jostins L, et al. Host-microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature. 2012;491:119–124. doi: 10.1038/nature11582. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Strange A, et al. A genome-wide association study identifies new psoriasis susceptibility loci and an interaction between HLA-C and ERAP1. Nature genetics. 2010;42:985–990. doi: 10.1038/ng.694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Ellinghaus E, et al. Genome-wide association study identifies a psoriasis susceptibility locus at TRAF3IP2. Nature genetics. 2010;42:991–995. doi: 10.1038/ng.689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Huffmeier U, et al. Common variants at TRAF3IP2 are associated with susceptibility to psoriatic arthritis and psoriasis. Nature genetics. 2010;42:996–999. doi: 10.1038/ng.688. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Wang C, et al. The psoriasis-associated D10N variant of the adaptor Act1 with impaired regulation by the molecular chaperone hsp90. Nature immunology. 2013;14:72–81. doi: 10.1038/ni.2479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Butter F, et al. Proteome-Wide Analysis of Disease-Associated SNPs That Show Allele-Specific Transcription Factor Binding. PLoS genetics. 2012;8:e1002982. doi: 10.1371/journal.pgen.1002982. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Guttman M, et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature. 2009;458:223–227. doi: 10.1038/nature07672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Derrien T, et al. The GENCODE v7 catalog of human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome research. 2012;22:1775–1789. doi: 10.1101/gr.132159.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Guttman M, et al. lincRNAs act in the circuitry controlling pluripotency and differentiation. Nature. 2011;477:295–300. doi: 10.1038/nature10398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Moran I, et al. Human beta cell transcriptome analysis uncovers lncRNAs that are tissue-specific, dynamically regulated, and abnormally expressed in type 2 diabetes. Cell metabolism. 2012;16:435–448. doi: 10.1016/j.cmet.2012.08.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Gomez JA, et al. The NeST Long ncRNA Controls Microbial Susceptibility and Epigenetic Activation of the Interferon-gamma Locus. Cell. 2013;152:743–754. doi: 10.1016/j.cell.2013.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Hasler R, et al. Microbial pattern recognition causes distinct functional micro-RNA signatures in primary human monocytes. PloS one. 2012;7:e31151. doi: 10.1371/journal.pone.0031151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Zwiers A, et al. Cutting edge: a variant of the IL-23R gene associated with inflammatory bowel disease induces loss of microRNA regulation and enhanced protein production. J Immunol. 2012;188:1573–1577. doi: 10.4049/jimmunol.1101494. [DOI] [PubMed] [Google Scholar]
- 14.de Jong VM, et al. Post-transcriptional control of candidate risk genes for type 1 diabetes by rare genetic variants. Genes and immunity. 2012 doi: 10.1038/gene.2012.38. [DOI] [PubMed] [Google Scholar]
- 15.Ueda H, et al. Association of the T-cell regulatory gene CTLA4 with susceptibility to autoimmune disease. Nature. 2003;423:506–511. doi: 10.1038/nature01621. [DOI] [PubMed] [Google Scholar]
- 16.Gerold KD, et al. The soluble CTLA-4 splice variant protects from type 1 diabetes and potentiates regulatory T-cell function. Diabetes. 2011;60:1955–1963. doi: 10.2337/db11-0130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Knight JC. Resolving the variable genome and epigenome in human disease. Journal of internal medicine. 2012;271:379–391. doi: 10.1111/j.1365-2796.2011.02508.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Orozco LD, et al. Unraveling inflammatory responses using systems genetics and gene-environment interactions in macrophages. Cell. 2012;151:658–670. doi: 10.1016/j.cell.2012.08.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Barreiro LB, et al. Deciphering the genetic architecture of variation in the immune response to Mycobacterium tuberculosis infection. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:1204–1209. doi: 10.1073/pnas.1115761109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Heinig M, et al. A trans-acting locus regulates an anti-viral expression network and type 1 diabetes risk. Nature. 2010;467:460–464. doi: 10.1038/nature09386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Ernst J, et al. Mapping and analysis of chromatin state dynamics in nine human cell types. Nature. 2011;473:43–49. doi: 10.1038/nature09906. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Maurano MT, et al. Systematic localization of common disease-associated variation in regulatory DNA. Science. 2012;337:1190–1195. doi: 10.1126/science.1222794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Rivas MA, et al. Deep resequencing of GWAS loci identifies independent rare variants associated with inflammatory bowel disease. Nature genetics. 2011;43:1066–1073. doi: 10.1038/ng.952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Muto A, et al. Bach2 represses plasma cell gene regulatory network in B cells to promote antibody class switch. The EMBO journal. 2010;29:4048–4061. doi: 10.1038/emboj.2010.257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Christodoulou K, et al. Next generation exome sequencing of paediatric inflammatory bowel disease patients identifies rare and novel variants in candidate genes. Gut. 2012 doi: 10.1136/gutjnl-2011-301833. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Briere F, et al. Interleukin 10 induces B lymphocytes from IgA-deficient patients to secrete IgA. The Journal of clinical investigation. 1994;94:97–104. doi: 10.1172/JCI117354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Labbe C, et al. MAST3: a novel IBD risk factor that modulates TLR4 signaling. Genes and immunity. 2008;9:602–612. doi: 10.1038/gene.2008.57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Labbe C, et al. Genome-wide expression profiling implicates a MAST3-regulated gene set in colonic mucosal inflammation of ulcerative colitis patients. Inflammatory bowel diseases. 2012;18:1072–1080. doi: 10.1002/ibd.21887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Zuk O, et al. The mystery of missing heritability: Genetic interactions create phantom heritability. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:1193–1198. doi: 10.1073/pnas.1119675109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Howson JM, et al. Evidence of Gene-Gene Interaction and Age-at-Diagnosis Effects in Type 1 Diabetes. Diabetes. 2012;61:3012–3017. doi: 10.2337/db11-1694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Ng SC, et al. Role of genetic and environmental factors in British twins with inflammatory bowel disease. Inflammatory bowel diseases. 2012;18:725–736. doi: 10.1002/ibd.21747. [DOI] [PubMed] [Google Scholar]
- 32.Ivanov II, et al. Induction of intestinal Th17 cells by segmented filamentous bacteria. Cell. 2009;139:485–498. doi: 10.1016/j.cell.2009.09.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Sonnenberg GF, Artis D. Innate lymphoid cell interactions with microbiota: implications for intestinal health and disease. Immunity. 2012;37:601–610. doi: 10.1016/j.immuni.2012.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Bernink JH, et al. Human type 1 innate lymphoid cells accumulate in inflamed mucosal tissues. Nature immunology. 2013;14:221–229. doi: 10.1038/ni.2534. [DOI] [PubMed] [Google Scholar]
- 35.Kiss EA, et al. Natural aryl hydrocarbon receptor ligands control organogenesis of intestinal lymphoid follicles. Science. 2011;334:1561–1565. doi: 10.1126/science.1214914. [DOI] [PubMed] [Google Scholar]
- 36.Sonnenberg GF, et al. Innate lymphoid cells promote anatomical containment of lymphoid-resident commensal bacteria. Science. 2012;336:1321–1325. doi: 10.1126/science.1222551. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Fukuda S, et al. Bifidobacteria can protect from enteropathogenic infection through production of acetate. Nature. 2011;469:543–547. doi: 10.1038/nature09646. [DOI] [PubMed] [Google Scholar]
- 38.Sun JC, et al. Adaptive immune features of natural killer cells. Nature. 2009;457:557–561. doi: 10.1038/nature07665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Cooper MA, et al. Cytokine-induced memory-like natural killer cells. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:1915–1919. doi: 10.1073/pnas.0813192106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.O'Leary JG, et al. T cell- and B cell-independent adaptive immunity mediated by natural killer cells. Nature immunology. 2006;7:507–516. doi: 10.1038/ni1332. [DOI] [PubMed] [Google Scholar]
- 41.Quintin J, et al. Candida albicans infection affords protection against reinfection via functional reprogramming of monocytes. Cell host & microbe. 2012;12:223–232. doi: 10.1016/j.chom.2012.06.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Cadwell K, et al. Virus-plus-susceptibility gene interaction determines Crohn's disease gene Atg16L1 phenotypes in intestine. Cell. 2010;141:1135–1145. doi: 10.1016/j.cell.2010.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Lindner C, et al. Age, microbiota, and T cells shape diverse individual IgA repertoires in the intestine. The Journal of experimental medicine. 2012;209:365–377. doi: 10.1084/jem.20111980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Becker C, et al. Complex Roles of Caspases in the Pathogenesis of Inflammatory Bowel Disease. Gastroenterology. 2012 doi: 10.1053/j.gastro.2012.11.035. [DOI] [PubMed] [Google Scholar]
- 45.Rath E, Haller D. Unfolded protein responses in the intestinal epithelium: sensors for the microbial and metabolic environment. Journal of clinical gastroenterology. 2012;46 Suppl:S3–S5. doi: 10.1097/MCG.0b013e318264e632. [DOI] [PubMed] [Google Scholar]
- 46.Chang JS, et al. Endoplasmic reticulum stress response promotes cytotoxic phenotype of CD8alphabeta+ intraepithelial lymphocytes in a mouse model for Crohn's disease-like ileitis. J Immunol. 2012;189:1510–1520. doi: 10.4049/jimmunol.1200166. [DOI] [PubMed] [Google Scholar]
- 47.Nguyen NT, et al. Aryl hydrocarbon receptor negatively regulates dendritic cell immunogenicity via a kynurenine-dependent mechanism. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:19961–19966. doi: 10.1073/pnas.1014465107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Wu C, et al. Induction of pathogenic T17 cells by inducible salt-sensing kinase SGK1. Nature. 2013 doi: 10.1038/nature11984. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Reynolds JM, et al. Cutting edge: regulation of intestinal inflammation and barrier function by IL-17C. J Immunol. 2012;189:4226–4230. doi: 10.4049/jimmunol.1103014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Stark MA, et al. Phagocytosis of apoptotic neutrophils regulates granulopoiesis via IL-23 and IL-17. Immunity. 2005;22:285–294. doi: 10.1016/j.immuni.2005.01.011. [DOI] [PubMed] [Google Scholar]
- 51.Conway KL, et al. p40phox expression regulates neutrophil recruitment and function during the resolution phase of intestinal inflammation. J Immunol. 2012;189:3631–3640. doi: 10.4049/jimmunol.1103746. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Matute JD, et al. A new genetic subgroup of chronic granulomatous disease with autosomal recessive mutations in p40 phox and selective defects in neutrophil NADPH oxidase activity. Blood. 2009;114:3309–3315. doi: 10.1182/blood-2009-07-231498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Muise AM, et al. NADPH oxidase complex and IBD candidate gene studies: identification of a rare variant in NCF2 that results in reduced binding to RAC2. Gut. 2012;61:1028–1035. doi: 10.1136/gutjnl-2011-300078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Zigmond E, et al. Ly6C(hi) Monocytes in the Inflamed Colon Give Rise to Proinflammatory Effector Cells and Migratory Antigen-Presenting Cells. Immunity. 2012;37:1076–1090. doi: 10.1016/j.immuni.2012.08.026. [DOI] [PubMed] [Google Scholar]
- 55.Tait Wojno ED, Artis D. Innate lymphoid cells: balancing immunity, inflammation, and tissue repair in the intestine. Cell host & microbe. 2012;12:445–457. doi: 10.1016/j.chom.2012.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Bevins CL, Salzman NH. Paneth cells, antimicrobial peptides and maintenance of intestinal homeostasis. Nature reviews. Microbiology. 2011;9:356–368. doi: 10.1038/nrmicro2546. [DOI] [PubMed] [Google Scholar]
- 57.Lipinski S, et al. RNAi screening identifies mediators of NOD2 signaling: Implications for spatial specificity of MDP recognition. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:21426–21431. doi: 10.1073/pnas.1209673109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Chiang CY, et al. Cofactors required for TLR7- and TLR9-dependent innate immune responses. Cell host & microbe. 2012;11:306–318. doi: 10.1016/j.chom.2012.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Abecasis GR, et al. An integrated map of genetic variation from 1,092 human genomes. Nature. 2012;491:56–65. doi: 10.1038/nature11632. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.MacArthur DG, et al. A systematic survey of loss-of-function variants in human protein-coding genes. Science. 2012;335:823–828. doi: 10.1126/science.1215040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Hampe J, et al. A genome-wide association scan of nonsynonymous SNPs identifies a susceptibility variant for Crohn disease in ATG16L1. Nature genetics. 2007;39:207–211. doi: 10.1038/ng1954. [DOI] [PubMed] [Google Scholar]
- 62.Rioux JD, et al. Genome-wide association study identifies new susceptibility loci for Crohn disease and implicates autophagy in disease pathogenesis. Nature genetics. 2007;39:596–604. doi: 10.1038/ng2032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Kuballa P, et al. Impaired autophagy of an intracellular pathogen induced by a Crohn's disease associated ATG16L1 variant. PloS one. 2008;3:e3391. doi: 10.1371/journal.pone.0003391. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Moscou MJ, Bogdanove AJ. A simple cipher governs DNA recognition by TAL effectors. Science. 2009;326:1501. doi: 10.1126/science.1178817. [DOI] [PubMed] [Google Scholar]
- 65.Sanjana NE, et al. A transcription activator-like effector toolbox for genome engineering. Nature protocols. 2012;7:171–192. doi: 10.1038/nprot.2011.431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Hockemeyer D, et al. Genetic engineering of human pluripotent cells using TALE nucleases. Nature biotechnology. 2011;29:731–734. doi: 10.1038/nbt.1927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Reyon D, et al. FLASH assembly of TALENs for high-throughput genome editing. Nature biotechnology. 2012;30:460–465. doi: 10.1038/nbt.2170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Bedell VM, et al. In vivo genome editing using a high-efficiency TALEN system. Nature. 2012 doi: 10.1038/nature11537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Garneau JE, et al. The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid DNA. Nature. 2010;468:67–71. doi: 10.1038/nature09523. [DOI] [PubMed] [Google Scholar]
- 70.Cong L, et al. Multiplex Genome Engineering Using CRISPR/Cas Systems. Science. 2013 [Google Scholar]
- 71.Mali P, et al. RNA-Guided Human Genome Engineering via Cas9. Science. 2013 doi: 10.1126/science.1232033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Hammer GE, et al. Expression of A20 by dendritic cells preserves immune homeostasis and prevents colitis and spondyloarthritis. Nature immunology. 2011;12:1184–1193. doi: 10.1038/ni.2135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Yamanishi H, et al. Regulatory dendritic cells pulsed with carbonic anhydrase I protect mice from colitis induced by CD4+CD25- T cells. J Immunol. 2012;188:2164–2172. doi: 10.4049/jimmunol.1100559. [DOI] [PubMed] [Google Scholar]
- 74.Virgin HW, Todd JA. Metagenomics and personalized medicine. Cell. 2011;147:44–56. doi: 10.1016/j.cell.2011.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Garg G, et al. Type 1 diabetes-associated IL2RA variation lowers IL-2 signaling and contributes to diminished CD4+CD25+ regulatory T cell function. J Immunol. 2012;188:4644–4653. doi: 10.4049/jimmunol.1100272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Khor B, et al. Genetics and pathogenesis of inflammatory bowel disease. Nature. 2011;474:307–317. doi: 10.1038/nature10209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Yang SK, et al. Genome-Wide Association Study of Ulcerative Colitis in Koreans Suggests Extensive Overlapping of Genetic Susceptibility With Caucasians. Inflammatory bowel diseases. 2013 doi: 10.1097/MIB.0b013e3182802ab6. [DOI] [PubMed] [Google Scholar]
- 78.Juyal G, et al. An investigation of genome-wide studies reported susceptibility loci for ulcerative colitis shows limited replication in north Indians. PloS one. 2011;6:e16565. doi: 10.1371/journal.pone.0016565. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Garrison WD, et al. Hepatocyte nuclear factor 4alpha is essential for embryonic development of the mouse colon. Gastroenterology. 2006;130:1207–1220. doi: 10.1053/j.gastro.2006.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Bloch DB, et al. Decreased IL-10 production by EBV-transformed B cells from patients with VODI: implications for the pathogenesis of Crohn disease. The Journal of allergy and clinical immunology. 2012;129:1678–1680. doi: 10.1016/j.jaci.2012.01.046. [DOI] [PubMed] [Google Scholar]
- 81.Hauser F, et al. Macrophage-stimulating protein polymorphism rs3197999 is associated with a gain of function: implications for inflammatory bowel disease. Genes and immunity. 2012;13:321–327. doi: 10.1038/gene.2011.88. [DOI] [PubMed] [Google Scholar]
- 82.Rausch P, et al. Colonic mucosa-associated microbiota is influenced by an interaction of Crohn disease and FUT2 (Secretor) genotype. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:19030–19035. doi: 10.1073/pnas.1106408108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Miyoshi J, et al. Ectopic expression of blood type antigens in inflamed mucosa with higher incidence of FUT2 secretor status in colonic Crohn's disease. Journal of gastroenterology. 2011;46:1056–1063. doi: 10.1007/s00535-011-0425-7. [DOI] [PubMed] [Google Scholar]
- 84.Smyth DJ, et al. FUT2 nonsecretor status links type 1 diabetes susceptibility and resistance to infection. Diabetes. 2011;60:3081–3084. doi: 10.2337/db11-0638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Kato Y, et al. Gene knockout and metabolome analysis of carnitine/organic cation transporter OCTN1. Pharmaceutical research. 2010;27:832–840. doi: 10.1007/s11095-010-0076-z. [DOI] [PubMed] [Google Scholar]
- 86.Saeki N, et al. Distinctive expression and function of four GSDM family genes (GSDMA-D) in normal and malignant upper gastrointestinal epithelium. Genes, chromosomes & cancer. 2009;48:261–271. doi: 10.1002/gcc.20636. [DOI] [PubMed] [Google Scholar]
- 87.McGovern DP, et al. Genome-wide association identifies multiple ulcerative colitis susceptibility loci. Nature genetics. 2010;42:332–337. doi: 10.1038/ng.549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Khan SQ, et al. Cloning, expression, and functional characterization of TL1A–Ig. J Immunol. 2013;190:1540–1550. doi: 10.4049/jimmunol.1201908. [DOI] [PubMed] [Google Scholar]
- 89.Moran CJ, et al. IL-10R Polymorphisms Are Associated with Very-early-onset Ulcerative Colitis. Inflammatory bowel diseases. 2013;19:115–123. doi: 10.1002/ibd.22974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Ghoreschi K, et al. Generation of pathogenic T(H)17 cells in the absence of TGF-beta signalling. Nature. 2010;467:967–971. doi: 10.1038/nature09447. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Hara H, et al. The adaptor protein CARD9 is essential for the activation of myeloid cells through ITAM-associated and Toll-like receptors. Nature immunology. 2007;8:619–629. doi: 10.1038/ni1466. [DOI] [PubMed] [Google Scholar]
- 92.Homer CR, et al. ATG16L1 and NOD2 interact in an autophagy-dependent antibacterial pathway implicated in Crohn's disease pathogenesis. Gastroenterology. 2010;139:1630–1641. doi: 10.1053/j.gastro.2010.07.006. 1641 e1631–1632. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Cheung H, et al. Accessory protein-like is essential for IL-18-mediated signaling. J Immunol. 2005;174:5351–5357. doi: 10.4049/jimmunol.174.9.5351. [DOI] [PubMed] [Google Scholar]
- 94.Rhee I, Veillette A. Protein tyrosine phosphatases in lymphocyte activation and autoimmunity. Nature immunology. 2012;13:439–447. doi: 10.1038/ni.2246. [DOI] [PubMed] [Google Scholar]



