Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Sep 30.
Published in final edited form as: Inflamm Bowel Dis. 2015 May;21(5):1166–1175. doi: 10.1097/MIB.0000000000000329

Very Early-onset Inflammatory Bowel Disease: Gaining Insight Through Focused Discovery

Christopher J Moran *,, Christoph Klein , Aleixo M Muise §,, Scott B Snapper ¶,**,††
PMCID: PMC6165626  NIHMSID: NIHMS989337  PMID: 25895007

Abstract

The pathogenesis of pediatric inflammatory bowel disease (IBD) is only partially understood. Strong evidence implicates a strong genetic component including high monozygotic twin concordance and familial disease phenotype concordance rates. Genome-wide association studies have identified associations between >160 genetic loci and the risk for developing IBD. The roles of implicated genes are largely immune-mediated, although other functions include cellular migration, oxidative stress, and carbohydrate metabolism. Additionally, growing literature describes monogenic causes of IBD that frequently present as infantile or very early-onset IBD. The interplay between IBD risk single nucleotide polymorphisms and rare genetic variants has yet to be determined. Studying patients with very early-onset IBD may elicit genetic factors that could be applied to broader populations of IBD. This review describes what is known about the genetic causes of very early-onset IBD and genetic strategies that may unravel more of the genetic causes of IBD.

Keywords: Crohn’s disease, ulcerative colitis, inflammatory bowel disease, genetics, very early onset IBD, immunodeficiency, pediatric, genome-wide association studies, whole exome sequencing


Inflammatory bowel disease (IBD) is a chronic inflammatory condition of the gastrointestinal tract that can present at any age.1 IBD encompasses 2 disease entities, ulcerative colitis (UC) in which the inflammation is limited to the large intestine and Crohn’s disease (CD) where the inflammatory pattern typically involves the small and/or large intestine but can affect any segment of the gastrointestinal tract. The initial evidence supporting a genetic cause for IBD is provided by the high concordance rates for CD in monozygotic twins and the increased risk of children with parents who suffer from IBD.27 However, despite decades of inquiry with advancing technology, the currently known genetic factors contribute only 20% to 25% of disease heritability; an understanding of the underlying genetic trigger(s) remains elusive.8,9

Children constitute a significant proportion of patients with IBD, with pediatric-onset IBD comprising 25% of IBD population in the United States.10,11 In contrast to adult-onset presentations, children with UC are likely to present with pancolitis, whereas children with CD more frequently present with ileocolonic disease and only rarely present as isolated ileal disease.1215 Those children diagnosed with CD in the first 8 years of life usually present with isolated colonic inflammation (rather than ileitis or ileocolitis).1518 Due to the frequent colonic presentation of very young patients with CD, a higher percentage of these patients may be inaccurately classified as UC during the initial evaluation.19 Children with IBD are more likely to receive corticosteroids, be initiated on immunomodulators, and require surgery in the first year after diagnosis than adults with IBD.13,20,21

Concomitant with the growing interest in studying pediatric-onset IBD, subclassifications of pediatric populations that reflect inherent distinctions in phenotypic and genotypic patterns have been developed. The Montreal Classification for IBD categorized all pediatric patients (diagnosed at <17 yr of age) in a single group.22 The subsequent Paris Modification divided pediatric-onset IBD into 2 groups: A1a (diagnosed <10 yr of age) and A1b (diagnosed between 10 and 17 yr of age), although these new pediatric categories may not reflect intrinsically different disease processes.23 As mentioned above, a notable phenotypic difference in very young patients with CD is the predominantly colonic involvement, contrasting with the increasing occurrence of ileal involvement starting at 8 years of age.1618,24 In addition, specific serologic patterns (CBir1 positivity in patients with otherwise negative serology) are associated with children diagnosed with CD <8 years of age.25 The fundamental differences between older A1a subjects and younger A1b may be quite small, and the disease present in many of these patients may behave similar to that of adult-onset IBD. Therefore, better classification schemes reflecting differences in pathophysiology and genetics are desirable.

Perhaps a more distinct population are patients with very early-onset IBD (VEO-IBD) who have disease onset in the first 6 years of life, which constitute 4% to 10% of pediatric IBD.16,26 By nature of their earlier diagnosis, patients with VEO-IBD will likely have longer exposure to immunosuppressive medications, more surgical interventions over their lifetime, and possibly an overall more complicated disease course. Limited series of infantile-onset IBD suggest that they often have severe presentations, and the courses are complicated by serious infections raising concerns of immunodeficiency.27 However, recent evidence shows that patients with VEO-IBD do not seem to have a more aggressive phenotype compared with adolescent onset IBD.28 Similarly, although patients with VEO-IBD are more likely to have a family history of IBD, and there exists a growing literature of individual cases with Mendelian IBD, it is unclear whether the genetic influences in this population as a whole are unique.29,30 In this regard, further studies are needed to determine whether the natural history and genetic susceptibility in patients with VEO-IBD is distinct from the more common presentation of IBD in older children. We believe that these patients may offer the unique opportunity to determine the role of genetic susceptibility in a context where chronic environmental influences may be less important, and at the very least, develop a better framework for studying this complex disease. This review will describe the overall pattern of genetics in VEO-IBD (in the context of pediatric IBD overall) and highlight some of the rapidly expanding areas of study.

COMMON POLYMORPHISMS ASSOCIATED WITH IBD RISK

Early work to identify genetic causes of IBD largely focused on genetic linkage analyses. Initially, genetic risk was found in major histocompatibility complex Class II molecules with both UC (DR2, DR9, and DRB1*0103) and CD (DR7, DRB3*0301, and DQ4).3134 Linkage studies requiring large pedigrees enriched for IBD led to the discovery of the CD risk loci of IBD1 and IBD5.35,36 NOD2, a key pattern recognition receptor in innate immune and epithelial cells, that recognizes microbial products was ultimately implicated at the IBD1 locus, illustrating a pivotal role of the crosstalk between intestinal microbiota and the innate immune system in IBD pathogenesis.3739 Three single nucleotide polymorphisms (SNPs) in NOD2 are linked to the majority of risk seen with the IBD1 locus (with a variety of other SNPs being less commonly reported), and increased risk has been demonstrated in both heterozygotes and homozygotes for these NOD2 SNPs.4042 Such linkage studies facilitated the discovery of high-impact CD variants whose significance has remained in the genome-wide association era.

Genome-wide association studies (GWAS) have rapidly expanded our knowledge of the role of common genetic variants in complex genetic disease pathogenesis. Currently, over 160 loci have been associated with IBD with most loci contributing to both CD and UC risk; however, some variants are unique to CD- or UC-specific risk (Tables 1 and 2).43,44,52 GWAS have identified SNPs associated with IBD risk that are in loci of genes involved in many immune-related pathways such as autophagy (ATG16L1, IRGM, LRRK2), adaptive immunity (IL2, IL12B, IL23R), and immunoregulation (STAT3, TYK2, IL10).43,44,53,5759 Additionally, these studies have highlighted risk associated with other relevant pathways including the maintenance of epithelial integrity (MUC19, CDH1), antigen presentation (ERAP2, DENND1B), and endoplasmic reticular stress (CPEB4, ORMDL3).4649,51,52,5456,60 Furthermore, deep resequencing of risk loci (e.g., NOD2, CARD9, IL23R) have identified additional and perhaps more functionally significant variants independent of the original SNPs found by GWAS.61,62

TABLE 1.

Single Nucleotide Polymorphisms Associated with Ulcerative Colitis in GWAS

Adult-Onset UC SNPs Pediatric-Onset and Adult-Onset UC SNPs
ADCY343 GPR65/GALC43 PTGER443,44 CAPN10/GPR354345
ARPC2/IL-8RA/IL-8RB44,46 GPR183/GPR1843 REL/PUS1043,44 HLA-DQ/HLA-DR4447
C1orf5343 HCK43 RELA43 ICOSLG4345
C1orf10644 HNF4A/ADA43,48 RFTN2/PLCL143 IFNG/IL-264345,49
CALM343 IFIH143 RORC43 IL-104345
CARD94446 IBD5 locus43 SERINC344 IL-12B4345
CARD11/GNA1243 IL-2/IL-21/EXOC343,44 SLC9A343 MST14345
CCDC88B43 IL-2RA/IL-15RA43 SMAD343 NKX2-343,45
CCL13/CCL243 IL-7R44 SMAD743 ORMDL34345
CCR6/RPS6KA243 IL-18RAP/IL-1R1/IL-1R243,44 SMURF143 OTUD3/PLA2G2E43,45,49
CD643 IL-23R43,44,47 SOCS1/LITAF43 PSMG143,50
CD4043 IL-2743 SPRED243 TNFRSF6B43,44,50
CD4843 IRF543,44 SPRY4/NDFIP143 ZMIZ1/ZPBP/IKZF143,45
CD22643 IRF843 STAT1/STAT443
CDH1/CDH348 IRGM43 STAT343
CEBPB43 ITGAL43 TAB143
CEBPG43 JAK243,44 TBD1/RPS6KB143
CISD1/IPMK43 JRKL/MAML243 TNFAIP343
CNTF/LPXN43 KIF21B43 TNFSF15/TNFSF843,44
CREM/CCNY43,44 LIF/OSM43 TNFRSF943
CRTC343 LOH12CR143,44 TNFRSF1443,44
CXCR543 MAP3K843 TNFRSF18/TNFRSF443
DAP43,44 MUC1943 TNNI2/LSP143,44
DLD/LAMB143,48,49,51 NFIL343 TRAF3IP243
DNMT3B43 NFKB1/MANBA43 TRIB143
DOK343 NXPE1/NXPE443 TSPAN14/C10orf5843
EPO43 PHACTR243 TUBD1/RPS6KB143
ERAP2/ERAP143,44 PRDM144 TYK243
FCGR2A/FCGR2B/FCGR3A43,51 PRKCB43 VDR43
FOSL2/BRE43 PRKCD/ITIH443 ZFP36L143
FOS/MLH343,44 ZFP9043,44
ZNF831/CTSZ43

TABLE 2.

Single Nucleotide Polymorphisms Associated with Crohn’s Disease in GWAS

Adult-Onset CD SNPs Adult-Onset and Pediatric-Onset CD SNPs
ADAM3043 GPR3543 RIPK243 ATG16L143,45,505254
ADCY343 GPR65/GALC4352 RORC43 C11orf3045,5254
BACH252 GPR183/GPR1843 SLC22A4/IRF1/IL-352 C13orf31455254
C1orf5343 GPX4/SBNO252 SMAD343,52 CCL2/CCL7/CCL1343,45,52
CARD943,52 HCK43 SMAD743 CCNY/CREM43,45,52,55
CCDC88B43 HMHA1/GPX443,52 SMURF143 FMO443,45
CCR643,525455 IFIH143 SOCS1/LITAF43 HLA-DQA50,52
CD643 IBD5 locus43 SP14043,52 ICOSLG43,45,52,54,55
CD4043 IFNG43 SPRED243 IL-1043,45,52
CD4843 IFNGR2/IFNAR143 SPRY4/NDFIP143 IL-12B43,45,52,54,55
CD22643 IL-2/IL-2143 STAT1/STAT443 IL-18RAP/IL-1R1/IL-18R1/IL12RL243,45,52
CD244/ITNL152,54 IL-2RA/IL-15RA43,52 STAT343,52,54,55 IL-23R45,50,52,55
CDKAL152,54 IL-6ST/IL-31RA43 TAB143 IL-27/CD1943,45,52
CEBPB43 IRF843 TAGAP43,52 JAK243,45,52,54,55
CEBPG43 IRGM43,52,54,55 TBD1/RPS6KB143 LRRK2/MUC1943,45,52,54
CISD1/IPMK43 KIF21B/C1orf10643,55 THADA52 MST1/PFKB443,45,5255
CNTF/LPXN43 LACC143 TNFAIP343 MTMR345,52
CPEB443,52 LGALS9/NOS243 TNFRSF18/TNFRSF4/FASLG43 NKX2-343,45,5255
CREB5/JAZF143 LIF/OSM43 TNFRSF6B43 NOD243,45,5254,56
CRTC343 LOH12CR143 TNFSF1152 ORMDL3/ZPBP/IKZF1/IKZF343,45,52,54,55
CXCR543 MAP3K7IP152 TNNI2/LSP143 PTGER443,45,5254
DAP43 MAP3K843 TRAF3IP243 PTPN2243,45,52,54
DENND1B52 MUC1/SCAMP343,52 TRIB143 TNFSF15/TNFSF8/SLC46A243,45,52,55
DNMT3A/DNMT3B43,52 NDFIP152 TSPAN14/C10orf5843 ZMIZ145,52
DOK343 NFIL343 TUBD1/RPS6KB143 ZNF30045,52
EPO43 PHACTR243 TYK2/ICAM1/ICAM343,52 ZNF36543,45,5254
ERAP2/ERAP143,52 PLCL152 UBE2D152 ZPBP45,52
FADS152 PRKCB43 UCN43
FCGR2A/FCGR2B/FCGR3A43 PRDM152 VAMP352
FOS/MLH343 PRDX5/ESRRA52 VDR43
FOSL2/BRE43 PTPN25254 YDJC52
FUT2/RASIP143,52 RASGRP1/SPRED143 ZFP36L143,52
GCKR52 RBX1-EP30056 ZNF831/CTSZ43
REL43
RELA43

Not only have GWAS of adult IBD reaffirmed the risk of NOD2 with CD, but they have identified specific clinical phenotypic presentations that are associated with NOD2 SNP carriers (ileal involvement and surgical resection).53,63,64 GWAS have identified additional genetic factors for colonic involvement of CD (ZPBP) and ileocolonic involvement of CD (JAK2, IL23R).65 Variation in genes involved in autophagy (IRGM, ITLN1, and ATG16L1) has been associated with ileal-predominant CD (as well as upper gastrointestinal tract involvement).66,67 Patients with CD who have a higher genetic burden (of the 140 SNPs associated with CD) have an earlier age of onset and are more likely to have ileal involvement.68

A complimentary approach to classic GWAS strategies are pathway/network analyses, which have led to a better understanding of the pathogenesis of type 2 diabetes, obesity, pancreatic and bladder cancer, psoriasis, lymphoma, and Behҫget’s disease.6974 These methods often rely on assignment of SNPs to a specific gene and specific tissue expression pattern information — tasks not always possible— and can benefit from some knowledge of gene function.75,76 Pathway and network analyses using data from adult-onset IBD GWAS have confirmed the involvement of known immune-mediated signaling pathways (IL-12, interferons) and antigen presentation but have also highlighted novel immune (activation of IL-9 and IL-2Rβ), and non-immune pathways (lipid metabolism).43,77,78 In addition, approaches agnostic to gene function have identified broader Nod2-focused IBD causal subnetworks that involve genes enriched in anti-inflammatory macrophages.43

GWAS and pathway/network analyses of adult-onset IBD have further emphasized that the pathophysiology of IBD likely results from either innate and/or adaptive immune defects although most SNPs detected in IBD GWAS have relatively modest effect sizes (with the exceptions of NOD2 and IL23R).52 However, identified variants may have stronger effects on specific IBD subsets (as discussed above for particular disease locations) and for pediatric-onset IBD. Regarding pediatric IBD pathogenesis, studies of adult cohorts may have overlooked some of the key genetic factors in pediatric IBD simply because younger patients were not included in initial GWAS. Furthermore, given that it is sometimes difficult to associate an SNP with a particular gene, the mechanistic explanation of the identified disease risk can be difficult to ascertain.

PEDIATRIC PERSPECTIVE OF IBD GENETICS

The SNPs identified initially by GWAS in adult IBD were based on cohorts with patients that have been drawn from both young and older adult patients with IBD. Multiple studies have demonstrated that many adult-onset CD SNPs (including IL23R, NOD2, and LRRK2) play a role in pediatric-onset CD, although these cohorts predominantly included teenage-onset (A1b) CD.45,79 Initial GWAS that included only pediatric-onset IBD replicated associations with 8 of 17 adult-onset UC SNPs (including IL-10).45,50 However, the results from GWAS of adult IBD may not be easily extrapolated to represent the risk in significantly younger subjects because they were largely missing from these studies.

In addition to replicating SNPs associated with adult-onset IBD, pediatric IBD GWAS identified novel SNPs not found in initial adult-onset IBD GWAS. Imielinski et al45 reported associations between pediatric CD and SNPs in the interleukin 27 (IL-27), ZMIZ1, and MTMR3 loci as well as an association of the CAPN10 locus with pediatric UC. IL-27 induces type 1 regulatory T cells and suppresses TH17 cells, and the IL27 risk SNP results in a 7-fold reduction of IL-27 production from lymphoblastoid cells.45,80 Kugathasan et al50 identified TNFRSF6B and PSMG1 loci as SNPs that confer risk for pediatric IBD. TNFRSF6B encodes DCR3, which is a soluble receptor that modulates the differentiation of dendritic cells and T cells and alters cellular sensitivity to FasL-induced death, whereas PSMG1 is critical in the formation of the 20S proteasome.8184 Subsequently, the IL-27, MTMR3, TNFRSF6B, PSMG1, and CAPN10 associations were detected in the large-scale metaanalyses of adult CD and UC.43,44,52 Interestingly, a ZMIZ1 SNP (rs1250550) had a protective effect in pediatric CD, whereas this SNP and others within the region were found to confer increased risk for adult-onset CD.45,52 These data indicate that many SNPs perceived to be pediatric-specific were ultimately found in the large-scale meta-analyses of adult IBD studies, although the effect sizes for some of these SNPs is different than that of their adult IBD counterparts.43,44,52 Similarly, GWAS-based pathway/network analyses that have included pediatric cohorts have identified pathways (e.g., IL-12/IL-23 signaling) in pediatric IBD that have also been identified in adult.43,85

GWAS assume that common yet complex diseases are the result of “common” polymorphisms. It is notable that 113 of the 163 IBD SNPs are also found to enhance risk for other autoimmune diseases (including lupus, rheumatoid arthritis, and ankylosing spondylitis), suggesting a partially shared pathogenesis.43 Early GWAS were powered to identify disease risk associated with SNPs with a minor allele frequency of >5%.86 Recent meta-analyses of GWAS data (benefiting from larger samples sizes) have lowered the detection threshold to ~1%. Patients with disease caused by highly penetrant, loss-of-function rare variants (with minor allele frequency <1%) are likely overlooked by current GWAS strategies. SNPs found by GWAS to be associated with IBD risk alone are unlikely to be independently causative given the high prevalence of these SNPs in nondiseased individuals.44,52 Additionally, although these SNPs are associated with risk for IBD, these specific SNPs may not be specifically responsible for the risk but rather they may be in strong linkage disequilibrium with the responsible genetic variant.

Moreover, the current collection of disease-associated SNPs can best explain 20% to 25% of the heritability of IBD.8,9 The applicability of GWAS is also a product of the study population. GWAS that involve largely adult populations of patients with IBD may overlook key regulators of pediatric IBD pathogenesis due to the absence of pediatric patients in their study cohorts. Because GWAS of pediatric IBD include predominantly adolescent-aged subjects, the applicability to patients with VEO-IBD may be limited. The study of these unique patients would be better served by a more focused approach that would be an important complement to GWAS.

LESSONS FROM PRIMARY IMMUNODEFICIENCY

Although many patients with IBD exhibit a proinflammatory state, the genetics underlying this immune dysregulation are highly variable which creates substantial obstacles to studying the immune physiology. Patients with primary immunodeficiency syndromes offer an alternative starting point by examining gastrointestinal pathophysiology in patients with well-defined and more uniform immune defects to observe the downstream defects. Chronic granulomatous disease (CGD), Wiskott—Aldrich syndrome (WAS), Immune dysregulation-Polyendocrinopathy-Enteropathy-X-linked syndrome, glycogen storage disease Type 1b, NEMO deficiency, and Hermansky—Pudlak syndrome all may show varying degrees of gastrointestinal symptoms.8791 Studying well-defined immune syndromes associated with gastrointestinal inflammation can act as an important complement for studying a complex genetically mediated disease such as IBD.30

WAS is an X-linked condition that is characterized by the classic triad of recurrent infections, thrombocytopenia, and eczema due to defective Wiskott-Aldrich Syndrome protein (WASp) expression.91 Mice deficient in WASp develop severe TH2-predominant colitis in part due to defective regulatory T-cell (Treg) function and aberrant antigen presenting cell function.9294 However, less than 10% of patients with WAS develop intestinal inflammation.95 Disease severity correlates with the degree of WASp dysfunction, as patients who possess mutations that result in full-length WASp with only altered amino acid sequences can develop a thrombocytopenia syndrome (X-linked thrombocytopenia) without eczema or recurrent infections, whereas mutations leading to truncated WASp often lead to classic WAS.96 Whether the incomplete penetrance of gastrointestinal symptoms in patients with WAS is due to the severity of the WASp defect, mutations in other immune-mediated genes in those patients, or a dysregulated intestinal microbiome has yet to be elucidated.

CGD is an immunodeficiency syndrome due to a defective oxidative burst from the NADPH oxidase complex.97 The NADPH oxidative complex consists of 5 proteins: one gene on the X chromosome (gp91phox) and 4 autosomal genes (p47phox, p67phox, p22phox, and p40phox).90,98 X-linked CGD is more common than all forms of autosomal CGD combined and presents at a significantly younger age than autosomal CGD.98100 These patients develop recurrent infections from Staphylococcus aureus, Serratia marcescens, Burkholderia cepacia, Nocardia, and Aspergillus due to defective intracellular killing by phagocytes.90,101 Nearly 50% of patients with CGD have gastrointestinal and hepatic complications, among these symptoms include colitis and gastrointestinal granulomas that resembles CD.102,103 Gastrointestinal symptoms occur more commonly in X-linked CGD and in autosomal recessive forms of CGD that have concomitant variants in myeloperoxidase and FcRγIIIb genes.100,104,105 Even in a well-defined immunodeficiency, genetic heterogeneity seems to modulate gastrointestinal presentations.

In a similar manner, genes involved in well-defined immunodeficiency syndromes may also contribute to states of immune dysregulation such as IBD. Patients with CD have long been described to display defects in oxidative burst, and GWAS have identified an SNP in the NCF4 locus (p40phox) that associates with the risk for ileal CD.106108 Taking a candidate gene approach, Muise et al109 recently showed that mutations in NCF2 (p67phox) are associated with VEO-IBD. More recently, additional genetic variants in NADPH oxidase genes have been shown to associate with VEO-IBD.110 These well-described syndromes described above are only but a few of the large list of immunodeficiency syndromes with both intestinal and extraintes-tinal manifestations.30 An equally important pool of knowledge can be gained from studying immunodeficiency syndromes with more gastrointestinal predominant symptoms.

IL-10 and genes downstream of IL-10 signaling have consistently been demonstrated by GWAS (e.g., Tyk2; Stat3) to be associated with IBD risk.43,44,52 Much stronger evidence for the involvement of defective IL-10 signaling comes from the analysis of infantile-onset IBD patients.111 Linkage studies (as well as targeted sequencing of IL-10R genes) in patients with VEO-IBD have identified homozygous mutations that lead to colitis and variable phenotypes of folliculitis and occasionally deepsited bacterial infections and polyarthritis.112117 Patients with IL-10 defects have a similar phenotype.118 These patients uniformly have aggressive IBD with significant perianal involvement (whose luminal disease is largely limited to the colon) that is recalcitrant to immunosuppressive agents and surgery, requiring allogeneic stem cell transplantation for control of the disease.112115,117,118 Interestingly, patients with mutations in IL10RA and IL10RB are also prone to develop intestinal diffuse large B-cell lymphomas, potentially caused by aberrant antitumor T-cell responses.119

IL-10R genes may not have been identified in pediatric IBD GWAS cohorts due to the primarily adolescent nature of pediatric GWAS cohorts or due to the limited coverage of SNPs in these genes. Extending from these initial studies in infantile-onset IBD, we recently completed a candidate gene analysis in children with VEO-IBD, identifying an association between IL-10RA SNPs and very early-onset UC.113 The expanding role of defective IL-10 signaling and the oxidative burst pathway in IBD pathogenesis illustrate the ongoing value that candidate gene analysis and linkage studies will have in studying IBD pathophysiology to pursue rare variants that elude GWAS detection.

NEW DIRECTIONS IN STUDYING EARLY-ONSET IBD

Significant progress in understanding the genetics of VEO-IBD has been made by studying families with high degrees of consanguinity. The evolving description of infantile-onset IBD patients with defective IL-10 signaling and the discovery of an IBD-like condition in patients with mutations in ADAM17 demonstrate the benefits of SNP-homozygosity mapping and linkage analysis.112,116,118,120 This allows for mutation discovery in consanguineous families given the high likelihood that a rare homozygous variant is present in these patients. However, these strategies may also overlook subjects who possess compound heterozygous mutations and dominant mutations with variable penetrance.

Whole exome sequencing (WES) has a developing role in the study of VEO-IBD and other rare disease processes.121 The exome includes all coding regions that comprise approximately 1% to 2% of the human genome.122 By using next generation sequencing, WES allows for the identification of coding variants across the exome and can play a vital role in gene discovery as most Mendelian disorders appear to result from genetic variation in the exome.123 This technology is becoming rapidly more accessible as the cost for sequencing an individual’s entire exome decreases. Limiting factors for WES are the ability to analyze (and store) the large amount of data generated by these sequencing efforts.124 Typically, individuals of European ancestry possess roughly 20,000 exomic variants (whereas the total in those of African ancestry averages closer to 24,000 variants), but only ~2% result in altered protein structure.125 Utilization of the enormity of these data is facilitated by limiting the analysis to those variants that are novel and alter amino acid sequence.124

Initially, WES was shown as an effective tool in the study of rare, Mendelian diseases.124 One only needs DNA samples from the probands and multiple first degree relatives to search for the causative genetic mutation. Studies using WES in Coffin–Siris syndrome, Miller syndrome, and mandibulofacial dysostosis with microcephaly have identified a number of genes not previously thought to be relevant to the underlying disease.126128 More recently, applications of WES in common diseases (such as Alzheimer’s disease, multiple sclerosis, and familial dyslipidemia) have identified genetic variants not found by GWAS.129132

The introduction of WES to the arena of IBD research led to the description of a hemizygous mutation in X-linked inhibitor of apoptosis protein gene (XIAP) in an infant with aggressive colitis.133 XIAP is an intracellular protein that interacts with caspases, NOD2, and NFκB and is expressed in all hematopoietic cells.134 Patients with XIAP deficiency classically develop an x-linked lymphoproli-ferative syndrome and hemophagocytic lymphohistiocytosis.135,136 Although XIAP-deficient patients had previously been described to have gastrointestinal manifestations, this report was the first of several to report a primary Crohn’s-like phenotype.133,137 Further analysis of the role that XIAP plays in pediatric IBD have shown that patients with XIAP deficiency can present with pediatric CD.138,139

The promise of WES includes the ability to identify causative mutations in genes that are not a priori judged to be involved in disease pathogenesis. WES has identified a homozygous PIK3R1 mutation in a patient with defective B-cell development and a homozygous mutation in LRBA in a patient with a common variable immunodeficiency, both of whom had colitis.140142 Most recently, patients presenting with recurrent intestinal atresia, combined immunodeficiency, and enterocolitis have been described due to mutations in TTC7A.143146 Much of the interest in WES has focused on its use in consanguineous families with extremely rare diseases, but WES has also been useful in identifying compound heterozygous mutations (e.g., IL10RA) in nonconsanguineous families.147

Although one obvious use of WES can be to ascertain the causative gene in extreme cases, it can also be applied to understand more common disease risk. More recently, WES was used in a cohort of patients with adult-onset CD to identify a novel risk SNP in NDP52 and to clarify that the risk for UC and CD seen at the 6q21 locus is due to genetic variation in PRDM1.148 Additionally, WES in 8 patients with pediatric-onset CD identified nonsynony-mous variants (mostly in the heterozygote state) in genes involved in IBD pathogenesis but did not identify a clear causative mutation in any of the patients.149 Although the potential yield of studying patients with VEO-IBD from consanguineous families may be the highest for gene discovery, there remains utility in studying patients apart from this very select population. The ability to study a multitude of genes simultaneously and the ability to identify compound heterozygotes in a particular disease make WES a rapidly developing field. However, WES does not address the ability of regulatory regions in the human genome to modulate disease.

An extension to WES, whole genome sequencing (WGS), offers the ability to obtain genetic information at every locus throughout the human genome. WGS has already displayed great benefit in identifying causative mutations in various cancers (multiple myeloma, melanoma, and hepatocellular carcinoma) and rare Mendelian disorders such as metachondromatosis, as well as a recent prenatal diagnosis of CHARGE syndrome.150154 The application of WGS technology to more complex genetic diseases has been largely limited to individual subject analysis due to a multitude of reasons.155 As WGS captures more data, it also carries a higher cost of sequencing and a greater burden of data analysis and storage. The ability of both WES and WGS to detect variants is dependent on the coverage of the sequence amplifiers used by each platform which also continues to improve. The current cost differential of WES and WGS may limit large-scale studies from using WGS, but accessibility to each of these technologies will increase as the cost of both continues to fall.

CONCLUSIONS

A better understanding of the genetics of IBD will undoubtedly occur from a combination of genetic strategies. Despite the significant progress made during the GWAS era, it is estimated that the currently known risk alleles only explain ~20% to 25% of the heritability of CD.52 A substantial proportion of the remaining risk may be due to rare variants. The advent of WES (and WGS) allows for the identification of genes not previously known to be involved in IBD and subsequently allows further analysis in larger cohorts by traditional genotyping and sequencing strategies. Although some genetic variants may contribute only modestly to overall pediatric IBD, they may have stronger effects on distinct subtypes of IBD (such as VEO-IBD) and allow stratification of patients into subsets to allow for better tailoring of therapy. Furthermore, the pathways identified by GWAS can allow for a direct pathway analysis in a broader IBD cohort. This will ultimately result in identification of patients with fundamental defects of their immune system that may benefit from more aggressive alternative therapy (such as allogeneic hematopoietic stem cell therapy) rather than profound immunosuppression and repeated surgical intervention. Although the underlying genetics of IBD remain elusive, recent advances in this field prepare us for the beginning of a quickly expanding research landscape.

ACKNOWLEDGMENTS

S. B. Snapper is supported by NIH Grants HL59561, DK034854, and AI50950, the Helmsley Charitable Trust, and the Wolpow Family Chair in IBD Treatment and Research.

Footnotes

The authors have no conflicts of interest to disclose.

REFERENCES

  • 1.Kaser A, Zeissig S, Blumberg RS. Inflammatory bowel disease. Annu Rev Immunol. 2010;28:573–621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Tysk C, Lindberg E, Jarnerot G, et al. Ulcerative colitis and Crohn’s disease in an unselected population of monozygotic and dizygotic twins. A study of heritability and the influence of smoking. Gut. 1988;29:990–996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Orholm M, Binder V, Sorensen TI, et al. Concordance of inflammatory bowel disease among Danish twins. Results of a nationwide study. Scand J Gastroenterol. 2000;35:1075–1081. [DOI] [PubMed] [Google Scholar]
  • 4.Probert CS, Jayanthi V, Hughes AO, et al. Prevalence and family risk of ulcerative colitis and Crohn’s disease: An epidemiological study among Europeans and south Asians in Leicestershire. Gut. 1993;34:1547–1551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Orholm M, Munkholm P, Langholz E, et al. Familial occurrence of inflammatory bowel disease. N Engl J Med. 1991;324:84–88. [DOI] [PubMed] [Google Scholar]
  • 6.Colombel JF, Grandbastien B, Gower-Rousseau C, et al. Clinical characteristics of Crohn’s disease in 72 families. Gastroenterology. 1996; 111:604–607. [DOI] [PubMed] [Google Scholar]
  • 7.Polito JM II, Childs B, Mellits ED, et al. Crohn’s disease: influence of age at diagnosis on site and clinical type of disease. Gastroenterology. 1996;111:580–586. [DOI] [PubMed] [Google Scholar]
  • 8.Manolio TA, Collins FS, Cox NJ, et al. Finding the missing heritability of complex diseases. Nature. 2009;461:747–753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Park JH, Wacholder S, Gail MH, et al. Estimation of effect size distribution from genome-wide association studies and implications for future discoveries. Nat Genet. 2010;42:570–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Loftus EV Jr. Clinical epidemiology of inflammatory bowel disease: incidence, prevalence, and environmental influences. Gastroenterology. 2004;126:1504–1517. [DOI] [PubMed] [Google Scholar]
  • 11.Kim SC, Ferry GD. Inflammatory bowel diseases in pediatric and adolescent patients: clinical, therapeutic, and psychosocial considerations. Gastroenterology. 2004;126:1550–1560. [DOI] [PubMed] [Google Scholar]
  • 12.Sawczenko A, Sandhu BK, Logan RF, et al. Prospective survey of childhood inflammatory bowel disease in the British Isles. Lancet. 2001;357: 1093–1094. [DOI] [PubMed] [Google Scholar]
  • 13.Hyams J, Markowitz J, Lerer T, et al. The natural history of corticosteroid therapy for ulcerative colitis in children. Clin Gastroenterol Hepatol. 2006;4:1118–1123. [DOI] [PubMed] [Google Scholar]
  • 14.Henriksen M, Jahnsen J, Lygren I, et al. Ulcerative colitis and clinical course: results of a 5-year population-based follow-up study (the IBSEN study). Inflamm Bowel Dis. 2006;12:543–550. [DOI] [PubMed] [Google Scholar]
  • 15.Van Limbergen J, Russell RK, Drummond HE, et al. Definition of phenotypic characteristics of childhood-onset inflammatory bowel disease. Gastroenterology. 2008;135:1114–1122. [DOI] [PubMed] [Google Scholar]
  • 16.Gupta N, Bostrom AG, Kirschner BS, et al. Presentation and disease course in early-compared to later-onset pediatric Crohn’s disease. Am J Gastroenterol. 2008;103:2092–2098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Mamula P, Telega GW, Markowitz JE, et al. Inflammatory bowel disease in children 5 years of age and younger. Am J Gastroenterol. 2002;97: 2005–2010. [DOI] [PubMed] [Google Scholar]
  • 18.Meinzer U, Idestrom M, Alberti C, et al. Ileal involvement is age dependent in pediatric Crohn’s disease. Inflamm Bowel Dis. 2005;11:639–644. [DOI] [PubMed] [Google Scholar]
  • 19.Abraham BP, Mehta S, El-Serag HB. Natural history of pediatric-onset inflammatory bowel disease: a systematic review. J Clin Gastroenterol. 2012;46:581–589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Jakobsen C, Bartek J Jr, Wewer V, et al. Differences in phenotype and disease course in adult and paediatric inflammatory bowel disease-a population-based study. Aliment Pharmacol Ther. 2011;34:1217–1224. [DOI] [PubMed] [Google Scholar]
  • 21.Pigneur B, Seksik P, Viola S, et al. Natural history of Crohn’s disease: comparison between childhood-and adult-onset disease. Inflamm Bowel Dis. 2010;16:953–961. [DOI] [PubMed] [Google Scholar]
  • 22.Silverberg MS, Satsangi J, Ahmad T, et al. Toward an integrated clinical, molecular and serological classification of inflammatory bowel disease: report of a Working Party of the 2005 Montreal World Congress of Gastroenterology. Can J Gastroenterol. 2005;19(suppl A):5–36. [DOI] [PubMed] [Google Scholar]
  • 23.Levine A, Griffiths A, Markowitz J, et al. Pediatric modification of the Montreal classification for inflammatory bowel disease: the Paris classification. Inflamm Bowel Dis. 2011;17:1314–1321. [DOI] [PubMed] [Google Scholar]
  • 24.Heyman MB, Kirschner BS, Gold BD, et al. Children with early-onset inflammatory bowel disease (IBD): analysis of a pediatric IBD consortium registry. J Pediatr. 2005;146:35–40. [DOI] [PubMed] [Google Scholar]
  • 25.Markowitz J, Kugathasan S, Dubinsky M, et al. Age of diagnosis influences serologic responses in children with Crohn’s disease: a possible clue to etiology? Inflamm Bowel Dis. 2009;15:714–719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Griffiths AM. Specificities of inflammatory bowel disease in childhood. Best Pract Res Clin Gastroenterol. 2004;18:509–523. [DOI] [PubMed] [Google Scholar]
  • 27.Ruemmele FM, El Khoury MG, Talbotec C, et al. Characteristics of inflammatory bowel disease with onset during the first year of life. J Pediatr Gastroenterol Nutr. 2006;43:603–609. [DOI] [PubMed] [Google Scholar]
  • 28.Benchimol EI, Mack DR, Nguyen GC, et al. Incidence, outcomes, and health services burden of very early onset inflammatory bowel disease. Gastroenterology. 2014;147:803–813. [DOI] [PubMed] [Google Scholar]
  • 29.Paul T, Birnbaum A, Pal DK, et al. Distinct phenotype ofearly childhood inflammatory bowel disease. J Clin Gastroenterol. 2006;40:583–586. [DOI] [PubMed] [Google Scholar]
  • 30.Uhlig HH. Monogenic diseases associated with intestinal inflammation: implications for the understanding of inflammatory bowel disease. Gut. 2013;62:1795–1805. [DOI] [PubMed] [Google Scholar]
  • 31.Stokkers PC, Reitsma PH, Tytgat GN, et al. HLA-DR and -DQ phenotypes in inflammatory bowel disease: a meta-analysis. Gut. 1999;45: 395–401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Forcione DG, Sands B, Isselbacher KJ, et al. An increased risk of Crohn’s disease in individuals who inherit the HLA class II DRB3*0301 allele. Proc Natl Acad Sci USA. 1996;93:5094–5098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Toyoda H, Wang SJ, Yang HY, et al. Distinct associations of HLA class II genes with inflammatory bowel disease. Gastroenterology. 1993;104: 741–748. [DOI] [PubMed] [Google Scholar]
  • 34.Asakura H, Tsuchiya M, Aiso S, et al. Association of the human lymphocyte-DR2 antigen with Japanese ulcerative colitis. Gastroenterology. 1982;82:413–18. [PubMed] [Google Scholar]
  • 35.Hugot JP, Laurent-Puig P, Gower-Rousseau C, et al. Mapping of a susceptibility locus for Crohn’s disease on chromosome 16. Nature. 1996; 379:821–823. [DOI] [PubMed] [Google Scholar]
  • 36.Rioux JD, Silverberg MS, Daly MJ, et al. Genomewide search in Canadian families with inflammatory bowel disease reveals two novel susceptibility loci. Am J Hum Genet. 2000;66:1863–1870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Hugot JP, Chamaillard M, Zouali H, et al. Association of NOD2 leucine-rich repeat variants with susceptibility to Crohn’s disease. Nature. 2001; 411:599–603. [DOI] [PubMed] [Google Scholar]
  • 38.Clarke TB, Weiser JN. Intracellular sensors of extracellular bacteria. Immunol Rev. 2011;243:9–25. [DOI] [PubMed] [Google Scholar]
  • 39.Ogura Y, Bonen DK, Inohara N, et al. A frameshift mutation in NOD2 associated with susceptibility to Crohn’s disease. Nature. 2001;411:603–606. [DOI] [PubMed] [Google Scholar]
  • 40.Economou M, Trikalinos TA, Loizou KT, et al. Differential effects of NOD2 variants on Crohn’s disease risk and phenotype in diverse populations: a metaanalysis. Am J Gastroenterol. 2004; 99:2393–2404. [DOI] [PubMed] [Google Scholar]
  • 41.Hampe J, Cuthbert A, Croucher PJ, et al. Association between insertion mutation in NOD2 gene and Crohn’s disease in German and British populations. Lancet. 2001;357:1925–1928. [DOI] [PubMed] [Google Scholar]
  • 42.Adler J, Rangwalla SC, Dwamena BA, et al. The prognostic power of the NOD2 genotype for complicated Crohn’s disease: a meta-analysis. Am J Gastroenterol. 2011;106:699–712. [DOI] [PubMed] [Google Scholar]
  • 43.Jostins L, Ripke S, Weersma RK, et al. Host-microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature. 2012;491:119–124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Anderson CA, Boucher G, Lees CW, et al. Meta-analysis identifies 29 additional ulcerative colitis risk loci, increasing the number of confirmed associations to 47. Nat Genet. 2011;43:246–252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Imielinski M, Baldassano RN, Griffiths A, et al. Common variants at five new loci associated with early-onset inflammatory bowel disease. Nat Genet. 2009;41:1335–1340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Franke A, Balschun T, Karlsen TH, et al. Sequence variants in IL10, ARPC2 and multiple other loci contribute to ulcerative colitis susceptibility. Nat Genet. 2008;40:1319–1323. [DOI] [PubMed] [Google Scholar]
  • 47.Silverberg MS, Cho JH, Rioux JD, et al. Ulcerative colitis-risk loci on chromosomes 1p36 and 12q15 found by genome-wide association study. Nat Genet. 2009;41:216–220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Consortium UIG, Barrett JC, Lee JC, et al. Genome-wide association study of ulcerative colitis identifies three new susceptibility loci, including the HNF4A region. Nat Genet. 2009;41:1330–1334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.McGovern DP, Gardet A, Torkvist L, et al. Genome-wide association identifies multiple ulcerative colitis susceptibility loci. Nat Genet. 2010; 42:332–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Kugathasan S, Baldassano RN, Bradfield JP, et al. Loci on 20q13 and 21q22 are associated with pediatric-onset inflammatory bowel disease. Nat Genet. 2008;40:1211–1215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Asano K, Matsushita T, Umeno J, et al. A genome-wide association study identifies three new susceptibility loci for ulcerative colitis in the Japanese population. Nat Genet. 2009;41:1325–1329. [DOI] [PubMed] [Google Scholar]
  • 52.Franke A, McGovern DP, Barrett JC, et al. Genome-wide meta-analysis increases to 71 the number of confirmed Crohn’s disease susceptibility loci. Nat Genet. 2010;42:1118–1125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Wellcome Trust Case Control C. Genome-wide association study of 14,000 cases of seven common diseases and 3,000 shared controls. Nature. 2007;447:661–678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Barrett JC, Hansoul S, Nicolae DL, et al. Genome-wide association defines more than 30 distinct susceptibility loci for Crohn’s disease. Nat Genet. 2008;40:955–962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Umeno J, Asano K, Matsushita T, et al. Meta-analysis of published studies identified eight additional common susceptibility loci for Crohn’s disease and ulcerative colitis. Inflamm Bowel Dis. 2011;17: 2407–2415. [DOI] [PubMed] [Google Scholar]
  • 56.Libioulle C, Louis E, Hansoul S, et al. Novel Crohn disease locus identified by genome-wide association maps to a gene desert on 5p13.1 and modulates expression of PTGER4. PLoS Genet. 2007;3:e58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Franke A, Balschun T, Sina C, et al. Genome-wide association study for ulcerative colitis identifies risk loci at 7q22 and 22q13 (IL17REL). Nat Genet. 2010;42:292–294. [DOI] [PubMed] [Google Scholar]
  • 58.Hampe J, Franke A, Rosenstiel P, 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. [DOI] [PubMed] [Google Scholar]
  • 59.Parkes M, Barrett JC, Prescott NJ, et al. Sequence variants in the autoph-agy gene IRGM and multiple other replicating loci contribute to Crohn’s disease susceptibility. Nat Genet. 2007;39:830–832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Barrett JC, Lee JC, Lees CW, et al. Genome-wide association study of ulcerative colitis identifies three new susceptibility loci, including the HNF4A region. Nat Genet. 2009;41:1330–1334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Momozawa Y, Mni M, Nakamura K, et al. Resequencing of positional candidates identifies low frequency IL23R coding variants protecting against inflammatory bowel disease. Nat Genet. 2011;43:43–7. [DOI] [PubMed] [Google Scholar]
  • 62.Rivas MA, Beaudoin M, Gardet A, et al. Deep resequencing of GWAS loci identifies independent rare variants associated with inflammatory bowel disease. Nat Genet. 2011;43:1066–1073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Cuthbert AP, Fisher SA, Mirza MM, et al. The contribution of NOD2 gene mutations to the risk and site of disease in inflammatory bowel disease. Gastroenterology. 2002;122:867–874. [DOI] [PubMed] [Google Scholar]
  • 64.Li E, Hamm CM, Gulati AS, et al. Inflammatory bowel diseases phenotype, C. difficile and NOD2 genotype are associated with shifts in human ileum associated microbial composition. PloS One. 2012;7:e26284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Cleynen I, Gonzalez JR, Figueroa C, et al. Genetic factors conferring an increased susceptibility to develop Crohn’s disease also influence disease phenotype: results from the IBDchip European Project. Gut. 2013;62:1556–1565. [DOI] [PubMed] [Google Scholar]
  • 66.Duraes C, Machado JC, Portela F, 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. [DOI] [PubMed] [Google Scholar]
  • 67.Prescott NJ, Fisher SA, Franke A, et al. A nonsynonymous SNP in ATG16L1 predisposes to ileal Crohn’s disease and is independent of CARD15 and IBD5. Gastroenterology. 2007;132:1665–1671. [DOI] [PubMed] [Google Scholar]
  • 68.Ananthakrishnan AN, Huang H, Nguyen DD, et al. Differential effect of genetic Burden on disease phenotypes in Crohn’s disease and ulcerative colitis: analysis of a North American cohort. Am J Gastroenterol. 2014;109:395–400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Perry JR, McCarthy MI, Hattersley AT, et al. Interrogating type 2 diabetes genome-wide association data using a biological pathway-based approach. Diabetes. 2009;58:1463–1467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Li D, Duell EJ, Yu K, et al. Pathway analysis of genome-wide association study data highlights pancreatic development genes as susceptibility factors for pancreatic Cancer. Carcinogenesis. 2012;33:1384–1390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Menashe I, Figueroa JD, Garcia-Closas M, et al. Large-scale pathway-based analysis of bladder cancer genome-wide association data from five studies of European background. PLoS One. 2012;7:e29396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Lee YH, Choi SJ, Ji JD, et al. Genome-wide pathway analysis of a genome-wide association study on psoriasis and Behcet’s disease. Mol Biol Rep. 2012;39:5953–5959. [DOI] [PubMed] [Google Scholar]
  • 73.Giallourakis CC, Benita Y, Molinie B, et al. Genome-wide analysis of immune system genes by expressed sequence Tag profiling. J Immunol. 2013;190:5578–5587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Yang X, Deignan JL, Qi H, et al. Validation of candidate causal genes for obesity that affect shared metabolic pathways and networks. Nat Genet. 2009;41:415–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Ramanan VK, Shen L, Moore JH, et al. Pathway analysis of genomic data: concepts, methods and prospects for future development. Trends Genet. 2012;28:323–332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Califano A, Butte AJ, Friend S, et al. Leveraging models of cell regulation and GWAS data in integrative network-based association studies. Nat Genet. 2012;44:841–847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Ballard D, Abraham C, Cho J, et al. Pathway analysis comparison using Crohn’s disease genome wide association studies. BMC Med Genomics. 2010;3:25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Torkamani A, Topol EJ, Schork NJ. Pathway analysis of seven common diseases assessedby genome-wide association. Genomics. 2008;92:265–272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Essers JB, Lee JJ, Kugathasan S, et al. Established genetic risk factors do not distinguish early and later onset Crohn’s disease. Inflamm Bowel Dis. 2009;15:1508–1514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Pot C, Apetoh L, Awasthi A, et al. Induction of regulatory Tr1 cells and inhibition of T(H)17 cells by IL-27. Semin Immunol. 2011;23:438–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Hirano Y, Hendil KB, Yashiroda H, et al. A heterodimeric complex that promotes the assembly of mammalian 20S proteasomes. Nature. 2005; 437:1381–1385. [DOI] [PubMed] [Google Scholar]
  • 82.Hsu TL, Chang YC, Chen SJ, et al. Modulation of dendritic cell differentiation and maturation by decoy receptor 3. J Immunol. 2002;168: 4846–4853. [DOI] [PubMed] [Google Scholar]
  • 83.Hsu TL, Wu YY, Chang YC, et al. Attenuation ofTh1 response in decoy receptor 3 transgenic mice. J Immunol. 2005;175:5135–5145. [DOI] [PubMed] [Google Scholar]
  • 84.Ina K, Itoh J, Fukushima K, et al. Resistance of Crohn’s disease T cells to multiple apoptotic signals is associated with a Bcl-2/Bax mucosal imbalance. J Immunol. 1999;163:1081–1090. [PubMed] [Google Scholar]
  • 85.Wang K, Zhang H, Kugathasan S, et al. Diverse genome-wide association studies associate the IL12/IL23 pathway with Crohn Disease. Am J Hum Genet. 2009;84:399–405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Visscher PM, Brown MA, McCarthy MI, et al. Five years of GWAS discovery. Am J Hum Genet. 2012;90:7–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Grucela AL, Patel P, Goldstein E, et al. Granulomatous enterocolitis associated with Hermansky-Pudlak syndrome. Am J Gastroenterol. 2006;101:2090–2095. [DOI] [PubMed] [Google Scholar]
  • 88.Yamaguchi T, Ihara K, Matsumoto T, et al. Inflammatory bowel disease like colitis in glycogen storage disease type 1b. Inflamm Bowel Dis. 2001;7:128–132. [DOI] [PubMed] [Google Scholar]
  • 89.Montalto M, D’Onofrio F, Santoro L, et al. Autoimmune enteropathy in children and adults. Scand J Gastroenterol. 2009;44:1029–1036. [DOI] [PubMed] [Google Scholar]
  • 90.Kang EM, Marciano BE, DeRavin S, et al. Chronic granulomatous disease: overview and hematopoietic stem cell transplantation. J Allergy Clin Immunol. 2011;127:1319–1326; quiz 27–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Ochs HD, Thrasher AJ. The Wiskott-Aldrich syndrome. J Allergy Clin Immunol. 2006;117:725–738; quiz 39. [DOI] [PubMed] [Google Scholar]
  • 92.Nguyen DD, Maillard MH, Cotta-de-Almeida V, et al. Lymphocyte-dependent and Th2 cytokine-associated colitis in mice deficient in Wiskott-Aldrich syndrome protein. Gastroenterology. 2007;133:1188–1197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Maillard MH, Cotta-de-Almeida V, Takeshima F, et al. The Wiskott-Aldrich syndrome protein is required for the function of CD4(+)CD25(+) Foxp3(+) regulatory T cells. J Exp Med. 2007;204:381–391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Nguyen DD, Wurbel MA, Goettel JA, et al. Wiskott-Aldrich syndrome protein deficiency in innate immune cells leads to mucosal immune dys-regulation and colitis in mice. Gastroenterology. 2012;143:719–729 e2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Dupuis-Girod S, Medioni J, Haddad E, et al. Autoimmunity in Wiskott-Aldrich syndrome: risk factors, clinical features, and outcome in a singlecenter cohort of 55 patients. Pediatrics. 2003;111:e622–e627. [DOI] [PubMed] [Google Scholar]
  • 96.Jin Y, Mazza C, Christie JR, et al. Mutations of the Wiskott-Aldrich Syndrome Protein (WASP): hotspots, effect on transcription, and translation and phenotype/genotype correlation. Blood. 2004;104:4010–4019. [DOI] [PubMed] [Google Scholar]
  • 97.Heyworth PG, Cross AR, Curnutte JT. Chronic granulomatous disease. Curr Opin Immunol. 2003;15:578–584. [DOI] [PubMed] [Google Scholar]
  • 98.Rahman FZ, Marks DJ, Hayee BH, et al. Phagocyte dysfunction and inflammatory bowel disease. Inflamm Bowel Dis. 2008;14:1443–1452. [DOI] [PubMed] [Google Scholar]
  • 99.Goldblatt D, Thrasher AJ. Chronic granulomatous disease. Clin Exp Immunol. 2000;122:1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Winkelstein JA, Marino MC, Johnston RB Jr, et al. Chronic granulomatous disease. Report on a national registry of 368 patients. Medicine. 2000;79:155–169. [DOI] [PubMed] [Google Scholar]
  • 101.Dinauer MC, Orkin SH. Chronic granulomatous disease. Annu Rev Med. 1992;43:117–124. [DOI] [PubMed] [Google Scholar]
  • 102.Agarwal S, Mayer L. Gastrointestinal manifestations in primary immune disorders. Inflamm Bowel Dis. 2010;16:703–711. [DOI] [PubMed] [Google Scholar]
  • 103.Marks DJ, Miyagi K, Rahman FZ, et al. Inflammatory bowel disease in CGD reproduces the clinicopathological features of Crohn’s disease. Am J Gastroenterol. 2009;104:117–124. [DOI] [PubMed] [Google Scholar]
  • 104.Marciano BE, Rosenzweig SD, Kleiner DE, et al. Gastrointestinal involvement in chronic granulomatous disease. Pediatrics. 2004;114:462–468. [DOI] [PubMed] [Google Scholar]
  • 105.Foster CB, Lehrnbecher T, Mol F, et al. Host defense molecule polymorphisms influence the risk for immune-mediated complications in chronic granulomatous disease. J Clin Invest. 1998;102:2146–2155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Segal AW, Loewi G. Neutrophil dysfunction in Crohn’s disease. Lancet. 1976;2:219–221. [DOI] [PubMed] [Google Scholar]
  • 107.Rioux JD, Xavier RJ, Taylor KD, et al. Genome-wide association study identifies new susceptibility loci for Crohn disease and implicates autophagy in disease pathogenesis. Nat Genet. 2007;39:596–604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Verspaget HW, Mieremet-Ooms MA, Weterman IT, et al. Partial defect of neutrophil oxidative metabolism in Crohn’s disease. Gut. 1984;25:849–853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Muise AM, Xu W, Guo CH, 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] [PMC free article] [PubMed] [Google Scholar]
  • 110.Dhillon SS, Fattouh R, Elkadri A, et al. Variants in NADPH oxidase complex components determine susceptibility to very early onset inflammatory bowel disease. Gastroenterology. 2014;147:680–689. [DOI] [PubMed] [Google Scholar]
  • 111.Shouval DS, Ouahed J, Biswas A, et al. Interleukin 10 receptor signaling: master regulator of intestinal mucosal homeostasis in mice and humans. Adv Immunol. 2014;122:177–210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Glocker EO, Kotlarz D, Boztug K, et al. Inflammatory bowel disease and mutations affecting the interleukin-10 receptor. N Engl J Med. 2009;361: 2033–2045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Moran CJ, Walters TD, Guo CH, et al. IL-10R polymorphisms are associated with very-early-onset ulcerative colitis. Inflamm Bowel Dis. 2013; 19:115–123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Engelhardt KR, Shah N, Faizura-Yeop I, et al. Clinical outcome in IL-10- and IL-10 receptor-deficient patients with or without hematopoietic stem cell transplantation. J Allergy Clin Immunol. 2013;131:825–830. [DOI] [PubMed] [Google Scholar]
  • 115.Pigneur B, Escher J, Elawad M, et al. Phenotypic Characterization of very early-onset IBD due to mutations in the IL10, IL10 receptor alpha or Beta gene: a survey of the GENIUS working group. Inflamm Bowel Dis. 2013;19:2820–2828. [DOI] [PubMed] [Google Scholar]
  • 116.Begue B, Verdier J, Rieux-Laucat F, et al. Defective IL10 signaling defining a subgroup of patients with inflammatory bowel disease. Am J Gastroenterol. 2011;106:1544–1555. [DOI] [PubMed] [Google Scholar]
  • 117.Murugan D, Albert MH, Langemeier J, et al. Very early onset inflammatory bowel disease associated with aberrant Trafficking of IL-10R1 and Cure by T Cell Replete Haploidentical Bone Marrow transplantation. J Clin Immunol. 2014;34:331–339. [DOI] [PubMed] [Google Scholar]
  • 118.Glocker EO, Frede N, Perro M, et al. Infant colitis-it’s in the genes. Lancet. 2010;376:1272. [DOI] [PubMed] [Google Scholar]
  • 119.Neven B, Mamessier E, Bruneau J, et al. A Mendelian predisposition to B-cell lymphoma caused by IL-10R deficiency. Blood. 2013;122:3713–3722. [DOI] [PubMed] [Google Scholar]
  • 120.Blaydon DC, Biancheri P, Di WL, et al. Inflammatory skin and bowel disease linked to ADAM17 deletion. NEngl J Med. 2011;365:1502–1508. [DOI] [PubMed] [Google Scholar]
  • 121.Muise AM, Snapper SB, Kugathasan S. The age of gene discovery in very early onset inflammatory bowel disease. Gastroenterology. 2012; 143:285–288. [DOI] [PubMed] [Google Scholar]
  • 122.Bick D, Dimmock D. Whole exome and whole genome sequencing. Curr Opin Pediatr. 2011;23:594–600. [DOI] [PubMed] [Google Scholar]
  • 123.Antonarakis SE, Beckmann JS. Mendelian disorders deserve more attention. Nat Rev Genet. 2006;7:277–282. [DOI] [PubMed] [Google Scholar]
  • 124.Bamshad MJ, Ng SB, Bigham AW, et al. Exome sequencing as a tool for Mendelian disease gene discovery. Nat Rev Genet. 2011;12:745–755. [DOI] [PubMed] [Google Scholar]
  • 125.Tennessen JA, Bigham AW, O’Connor TD, et al. Evolution and functional impact of rare coding variation from deep sequencing of human exomes. Science. 2012;337:64–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Tsurusaki Y, Okamoto N, Ohashi H, et al. Mutations affecting components of the SWI/SNF complex cause Coffin-Siris syndrome. Nat Genet. 2012;44:376–378. [DOI] [PubMed] [Google Scholar]
  • 127.Ng SB, Bigham AW, Buckingham KJ, et al. Exome sequencing identifies MLL2 mutations as a cause of Kabuki syndrome. Nat Genet. 2010; 42:790–793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Lines MA, Huang L, Schwartzentruber J, et al. Haploinsufficiency of a spliceosomal GTPase encoded by EFTUD2 causes mandibulofacial dysostosis with microcephaly. Am J Hum Genet. 2012;90:369–377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Ramagopalan SV, Dyment DA, Cader MZ, et al. Rare variants in the CYP27B1 gene are associated with multiple sclerosis. Ann Neurol. 2011; 70:881–886. [DOI] [PubMed] [Google Scholar]
  • 130.Guerreiro RJ, Lohmann E, Kinsella E, et al. Exome sequencing reveals an unexpected genetic cause of disease: NOTCH3 mutation in a Turkish family with Alzheimer’s disease. Neurobiol Aging. 2012;33:1008 e17–e23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Musunuru K, Pirruccello JP, Do R, et al. Exome sequencing, ANGPTL3 mutations, and familial combined hypolipidemia. N Engl J Med. 2010; 363:2220–2227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Dyment DA, Cader MZ, Chao MJ, et al. Exome sequencing identifies a novel multiple sclerosis susceptibility variant in the TYK2 gene. Neurology. 2012;79:406–411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Worthey EA, Mayer AN, Syverson GD, et al. Making a definitive diagnosis: successful clinical application of whole exome sequencing in a child with intractable inflammatory bowel disease. Genet Med. 2011; 13:255–262. [DOI] [PubMed] [Google Scholar]
  • 134.Filipovich AH, Zhang K, Snow AL, et al. X-linked lymphoproliferative syndromes: brothers or distant cousins? Blood. 2010;116:3398–3408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Rezaei N, Mahmoudi E, Aghamohammadi A, et al. X-linked lympho-proliferative syndrome: a genetic condition typified by the triad of infection, immunodeficiency and lymphoma. Br J Haematol. 2011; 152:13–30. [DOI] [PubMed] [Google Scholar]
  • 136.Rigaud S, Fondaneche MC, Lambert N, et al. XIAP deficiency in humans causes an X-linked lymphoproliferative syndrome. Nature. 2006; 444:110–114. [DOI] [PubMed] [Google Scholar]
  • 137.Pachlopnik Schmid J, Canioni D, Moshous D, et al. Clinical similarities and differences of patients with X-linked lymphoproliferative syndrome type 1 (XLP-1/SAP deficiency) versus type 2 (XLP-2/XIAP deficiency). Blood. 2011;117:1522–1529. [DOI] [PubMed] [Google Scholar]
  • 138.Speckmann C, Lehmberg K, Albert MH, et al. X-linked inhibitor of apoptosis (XIAP) deficiency: the spectrum of presenting manifestations beyond hemophagocytic lymphohistiocytosis. Clin Immunol. 2013;149:133–141. [DOI] [PubMed] [Google Scholar]
  • 139.Zeissig Y, Petersen BS, Milutinovic S, et al. XIAP variants in male Crohn’s disease. Gut. 2015;64:66–76. [DOI] [PubMed] [Google Scholar]
  • 140.Conley ME, Dobbs AK, Quintana AM, et al. Agammaglobulinemia and absent B lineage cells in a patient lacking the p85alpha subunit of PI3K. J Exp Med. 2012;209:463–470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.de la Morena M, Haire RN, Ohta Y, et al. Predominance of sterile immunoglobulin transcripts in a female phenotypically resembling Bruton’s agammaglobulinemia. Eur J Immunol. 1995;25:809–815. [DOI] [PubMed] [Google Scholar]
  • 142.Alangari A, Alsultan A, Adly N, et al. LPS-responsive beige-like anchor (LRBA) gene mutation in a family with inflammatory bowel disease and combined immunodeficiency. J Allergy Clin Immunol. 2012;130:481–488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Avitzur Y, Guo C, Mastropaolo LA, et al. Mutations in tetratricopeptide repeat domain 7A result in a severe form of very early onset inflammatory bowel disease. Gastroenterology. 2014;146:1028–1039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Chen R, Giliani S, Lanzi G, et al. Whole-exome sequencing identifies tetratricopeptide repeat domain 7A (TTC7A) mutations for combined immunodeficiency with intestinal atresias. J Allergy Clin Immunol. 2013;132:656–664. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Samuels ME, Majewski J, Alirezaie N, et al. Exome sequencing identifies mutations in the gene TTC7A in French-Canadian cases with hereditary multiple intestinal atresia. J Med Genet. 2013;50: 324–329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Bigorgne AE, Farin HF, Lemoine R, et al. TTC7A mutations disrupt intestinal epithelial apicobasal polarity. J Clin Invest. 2014; 124:328–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Mao H, Yang W, Lee PP, et al. Exome sequencing identifies novel compound heterozygous mutations of IL-10 receptor 1 in neonatal-onset Crohn’s disease. Genes Immun. 2012;13:437–442. [DOI] [PubMed] [Google Scholar]
  • 148.Ellinghaus D, Zhang H, Zeissig S, et al. Association between variants of PRDM1 and NDP52 and Crohns disease, based on exome sequencing and functional studies. Gastroenterology. 2013;145:339–347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Christodoulou K, Wiskin AE, Gibson J, et al. Next generation exome sequencing of paediatric inflammatory bowel disease patients identifies rare and novel variants in candidate genes. Gut. 2013;62:977–984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Berger MF, Hodis E, Heffernan TP, et al. Melanoma genome sequencing reveals frequent PREX2 mutations. Nature. 2012;485:502–506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Fujimoto A, Totoki Y, Abe T, et al. Whole-genome sequencing of liver cancers identifies etiological influences on mutation patterns and recurrent mutations in chromatin regulators. Nat Genet. 2012;44:760–764. [DOI] [PubMed] [Google Scholar]
  • 152.Egan JB, Shi CX, Tembe W, et al. Whole genome sequencing of multiple myeloma from diagnosis to plasma cell leukemia reveals genomic initiating events, evolution and clonal tides. Blood. 2012;120: 1060–1066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Sobreira NL, Cirulli ET, Avramopoulos D, et al. Whole-genome sequencing of a single proband together with linkage analysis identifies a Mendelian disease gene. PLoS Genet. 2010;6:e1000991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Talkowski ME, Ordulu Z, Pillalamarri V, et al. Clinical diagnosis by whole-genome sequencing of a prenatal sample. N Engl J Med. 2012; 367:2226–2232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Ashley EA, Butte AJ, Wheeler MT, et al. Clinical assessment incorporating a personal genome. Lancet. 2010;375:1525–1535. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES