Pathogenesis of

DETLEF SCHUPPAN Division of , Beth Israel Deaconess Medical Center, Harvard Medical School, Boston, MA 02215, USA

Introduction intestine. These first were described in 1954 [8] and refined further by Marsh [9]. With the Classical coeliac disease is an inflammatory advent of serum diagnostic tests such as im- disorder of the small intestine that is characte- munoglobulin A (IgA) antibodies to gliadin and rised by global malabsorption of macronu- especially to an autoantigen present in reticu- trients, minerals and vitamins. It is charac- lin and endomysium (see later), coeliac disease terised histologically by a predominant T-cell could be differentiated more easily from other infiltration in the epithelium and the lamina malabsorptive disorders [10]. More important- propria that may lead to complete destruction ly, population studies using the antibody tests of the small intestinal villi and to significant as screening tools and confirmatory small in- reduction of the gastrointestinal tract absorp- testinal biopsies revealed a hitherto unexpec- tive surface [1-4]. Coeliac disease was common- tedly high prevalence of coeliac disease in the ly fatal in the recent past as attested by the 12% Western world, Northern Africa, and the Near mortality among affected children reported and the Middle East, ranging from ~1:80-1:150 in a retrospective study done in 1939 [5]. The [11-13] (see C. Catassi, E. Fabiani and A. Fasano aetiology of this disorder remained unex- in this issue, pp. 83-94). In these studies, the plained until the Dutch paediatrician Willem vast majority (>80%) of affected individuals Dicke recognised an association between the identified by screening presented with clini- consumption of products and relapsing cally “silent” or atypical forms of the disease. If diarrhoea. He confirmed his suspicions du ring and to what extent asymptomatic or oligosymp- food shortages in the Second World War when tomatic affected individuals identified through the symptoms of his patients improved as screening subsequently develop overt coeliac bread was replaced by noncereal containing foods. After the war, Dicke, Weijers and van de Kamer performed controlled experiments Gluten proteins can be separated into by exposing affected children to defined diets the ethanol-insoluble glutenins and the and determining faecal weight and faecal fat as measures of malabsorption. These metabolic alcohol-soluble gliadins. Both gliadins studies demonstrated clearly that wheat, bar- and glutenins display a high content ley and rye triggered gastrointestinal signs and symptoms that defined coeliac disease and that of the amino acids glutamine (32-56%) these were reversed after their exclusion from and proline (15-30%), and due to their diets [6]. They also identified the toxic agents as present mainly in the alcohol-soluble gliadin cysteine content glutenins can fraction of wheat gluten [7]. form complex homopolymers and The characteristic coeliac microscopic le- sions are mucosal inflammation, crypt hyper- heteropolymers with gliadins. plasia and villous atrophy in the proximal small

Annales Nestlé 2004;62:95-106 95 Detlef Schuppan

disease, secondary autoimmune diseases or GLIADINPEPTIDE 4 CELLRESPONSE even malignancy remains to be shown. This is especially of concern because these indivi- (,! $1 duals are likely to continue consuming gluten- HETERODIMER containing diets. ^OFFIRSTDEGREE #$PATIENTS$1  RELATIVESAFFECTED $1 'ENERALPOPULATION^ Gluten: the causative agent

Gluten comprises the storage proteins of wheat endosperm that are responsible for many of B  A  $2A  $2 the desired baking properties. Gluten proteins can be separated into the ethanol-insoluble glu- B  $2 tenins and the alcohol-soluble gliadins. Each  wheat variant produces an estimated 40-50 glia- ENCODEDINCIS ENCODEDINTRANS GENEDOSE dins that are structurally similar proteins with ~250-500 amino acids, and a limited number of Figure 1: HLA-DQ2 (DQ8) and presentation of gluten high and low molecular weight glutenins, with peptides to T-cells. HLA-DQ2 (DQ8) is a heterodimer of an α1 and a β1 chain that is necessary to present ~650-800 and ~270-320 amino acids, respective- (modified) gluten peptides to gluten reactive T-cells. ly. Both gliadins and glutenins display a high In most populations, more than 90% of coeliac pa- content of the amino acids glutamine (32-56%) tients express DQ2, and the minority DQ8. The two and proline (15-30%), and due to their cysteine DQ2 chains can be encoded on a single allele (in cis) content glutenins can form complex homopoly- in association with DR3, or on the two alleles in as- mers and heteropolymers with gliadins [14, 15]. sociation with DR5 and DR7 (in trans). In case of the Storage proteins similar to gliadins (generally trans configuration, only 1 out of the 4 possible hete- termed prolamines) are found in rye (secalins) rodimers that can form contains the correct chains. and barley (hordeins). Avenins of oats and es- In case of homozygosity, the cis configuration can lead to 4 out of 4 heterodimers with the correct chain pecially the zeins of rice are more distantly re- composition. Homozygosity for DQ2 in cis underlies lated to these other cereal proteins (Table I). On the observed gene dose effect described in fig. 2.

Table I: Cereal proteins that cause intestinal inflam- mation in patients with coeliac disease. the basis of their electrophoretic properties and Cereal Prolamine Composition “Toxicity” primary structures the gliadins are subdivided Wheat α-gliadin 36% Q, 17-23% P +++ into the classes of α, γ and ω. These are further Barley hordeins 36% Q, 17-23% P ++ classified into several distinct proteins, such as Rye secalins 36% Q, 17-23% P ++ α1-11, γ1-6, and ω1-5. Although the proinflam- Oats avenins high Q, low P (+) matory effects of all gliadin fractions were Maize zeins low Q, high A, V – shown in vitro [16] and after intraduodenal or Millet ? low Q, high A, V – Rice ? low Q, high A, V – rectal challenge in vivo [17], most experimen- tal and human studies focused on α-gliadins, The “toxic” prolamines (gliadins, secalins and hor- and more specifically on proinflammatory pep- deins) that drive the destructive intestinal T-cell tides in its amino terminal region [18, 19]. response in coeliac disease are rich in glutamine (Q) and proline (P). The proteins from grains that do not cause intestinal inflammation contain less Q and P, and a higher proportion of alanine (A) and valine (V). The genetic component In addition to the prolamines, the structurally diffe- rent glutenins from wheat and likely related proteins The risk to coeliac disease is genetically de- from barley and rye harbour immunogenic epitopes termined. Moreover, this genetic component is relevant to this condition. more relevant than that of most other inherited

96 Annales Nestlé 2004;62:95-106 Pathogenesis of coeliac disease

diseases. Transmission is autosomal dominant In addition, the DR3 linked genes MICA-A4 and with incomplete penetrance, resulting in a di- -A5.1, MICB-CA24 and MIB-350 were found to sease prevalence of ~15% in first-degree rela- confer a higher risk for coeliac disease in two tives [20-22], and reaching a concordance rate European cohorts [30, 31]. of 75 % in monozygotic twins [23]. Coeliac di- Coeliac disease is associated with other, sease is associated uniquely with the HLA class classical autoimmune diseases, such as type 1 II antigens DQ2 and DQ8 located on chromo- diabetes, autoimmune thyroiditis, connective some 6p21 [21, 22]. This association is so tight tissue diseases, idiopathic cardiomyopathy that between 85 to 95% of coeliac patients carry and autoimmune hepatitis. The prevalence of HLA-DQ2 and virtually all remaining patients coexisting celiac disease and a single other au- express HLA-DQ8 [24]. DQ2 is a heterodimer toimmune disease ranges from ~3-15% [32-42]. encoded by the alleles DQA1*05 (α) and DQB1*02 A large cross-sectional study suggested that (β). These are associated with the DR3-DQ2 the duration of gluten exposure correlated with haplotype (when encoded on the same allele the co-occurrence of these autoimmune disea- in cis) or with the DR5/DR7-DQ2 haplotype ses. The reported prevalence of co-occurrence (when encoded on both alleles in trans) [21, 22] ranged from 5.1% to 23.6% with ages at diagno- (Fig. 1). DQ8 is a heterodimer encoded by sis <2 years and >10 years, respectively [32]. DQA1*0301 and DQB1*0302. DQ2 or DQ8 is ne- However, a subsequent study demonstrated cessary for the development of the disease, that susceptibility to secondary autoimmunity accounting for roughly 40% of the genetic pre- was related mainly to the ages of the coeliac pa- disposition. However, since these antigens are tients [33]. It is therefore most likely that older expressed in 25-30% and 5-15%, respectively, ages and shared genetic predispositions, such of healthy controls in susceptible populations, as DR4, HLA-DQ8 and genes on chromosomes large international studies have been initiated 2q33 and 15q26 associated with type 1 diabetes to search for additional genetic determinants. [36-38], or HLA-DQ2/DQ8 and CTLA-4 polymor- These studies suggested several genetic loci phisms associated with autoimmune thyroidi- linked to coeliac disease, with and without tis [36, 42] underlie the association, as opposed identification of target genes and with hetero- to causative roles for coeliac disease and the geneity among the investigated populations. duration and intensity of gluten ingestion. Importantly, their impact appears to be much lower than that of either DQ2 or DQ8. The most convincing data relate to chromosome 5q31-33, Tissue transglutaminase: encoding a cluster of cytokine genes [25-27]. the coeliac disease autoantigen Minor associations were found for chromosome 2q33 where immune regulatory genes such as Patients with untreated coeliac disease have CTLA-4, ICOS-1 and CD28 are encoded [28, 29]. circulating IgA class autoantibodies against endomysium (the connective tissue around smooth muscle) or reticulin (fibronectin and containing extracellular fibrils). Due to their high positive and negative predictive Coeliac disease is associated uniquely value, ranging between 90 and 100%, these with the HLA class II antigens DQ2 autoantibodies are valuable tools for disease diagnosis and population screening [10]. The and DQ8 located on chromosome 6p21. discovery that this autoantigen is an enzyme, This association is so tight that between namely tissue transglutaminase (tTG) was unexpected [43]. tTG or transglutaminase 2 85 to 95% of coeliac patients carry is expressed by almost all cell types and usu- HLA-DQ2 and virtually all remaining ally is retained intracellularly in an inactive form. Cells under mechanical or inflammatory patients express HLA-DQ8. stress release tTG into the extracellular space. This explains its association with the extracel-

Annales Nestlé 2004;62:95-106 97 Detlef Schuppan

lular matrix [44, 45]. tTG belongs to a family tTG expression [44-47]. On the other hand, tTG of at least eight calcium-dependent transami- tethers the inactive proform of the pluripotent dating enzymes that catalyze the covalent and cytokine TGF-β (latent TGF-β) to the cellular irreversible crosslinking of a protein with a surface which allows its proteolytic activation glutamine residue (glutamine donor) to a se- to mature TGF-β by plasmin [51]. This could be cond protein with a lysine residue (glutamine relevant to coeliac disease pathophysiology be- acceptor), resulting in the formation of an ε- cause TGF-β can induce the differentiation of (γ-glutamyl)-lysine isopeptide bond [44-46]. human intestinal epithelial cells that are less tTG displays a high specificity for only certain differentiated in active coeliac disease [52]. protein-bound glutamine residues as glutamine donor substrates, whereas the lysine-contai- ning glutamine acceptor substrates are nume- Tissue transglutaminase rous. The reaction is calcium-dependent and and coeliac disease: a missing link tTG is only active in the presence of high cal- cium concentrations, as are found in extracel- After it became clear that the disease-associa- lular spaces, where it contributes to the extra- ted HLA molecules are HLA-DQ2 and, to a lesser cellular matrix’s stabilisation [47]. Intracellular extent, HLA-DQ8, the existence of gluten spe- tTG activation and subsequent crosslinking also cific T-cells remained hypothetical until 1993 occurs when cellular integrity is destroyed and when such T-cells were isolated from patients intracellular calcium rises in apoptosis. The lat- and found to be absent in healthy controls [22]. ter function may prevent leakage of potentially HLA-DQ molecules harbour a peptide binding harmful molecules from virus-infected or dying groove that accommodates peptides of 9 amino cells. Notably, tTG and tTG-derived crosslin- acids length that must contain negatively king appears to play a role in a variety of neu- charged amino acids at so called anchor posi- rodegenerative disorders, such as Alzheimer’s tions [54-56]. However, such acidic peptides do and Huntington’s disease [48, 49]. Under cer- not occur in the gluten molecules. This mystery tain conditions, e.g., when primary lysines are was solved when it was found that tTG, the tar- unavailable as glutamine acceptors or at low get autoantigen, can crosslink gluten peptides pH, tTG merely deamidates target glutamine that are preferred glutamine donor substrates in the substrate protein, transforming neutral and deamidate certain glutamines in these pep- glutamine to a negatively charged glutamic acid tides, which introduces the negative charges residue [50]. required for binding to HLA-DQ molecules. tTG is upregulated in wound healing, angio- Thus, ingested gluten molecules, degraded to genesis and apoptosis. Transforming growth peptides by gastrointestinal enzymes and mo- factor-β (TGF-β), tumor necrosis factor-α (TNF- dified by tTG, bind to HLA-DQ2 or HLA-DQ8 and α), interleukin-6, retinoids and corticosteroids trigger an inflammatory T-cell response. stimulate, whereas bone morphogenetic pro- Further work revealed the identity of a series teins-2 and -4 and histamine downregulate of (deamidated) gluten peptides that can trigger such T-cell responses, both in vitro and ex vivo using cultures of biopsies from coeliac patients [57-64]. However, most T-cell clones derived Patients with untreated coeliac disease from paediatric patients and some T-cell clones from adult patients responded primarily to non- have circulating IgA class autoantibodies deamidated gliadins and glutenins, indicating against endomysium or reticulin. that the early immune response is directed to several unmodified gluten peptides, whereas The discovery that this autoantigen long-standing disease favours a few immunodo- is an enzyme, namely tissue minant and preferentially deamidated peptides [63]. transglutaminase (tTG) was unexpected. The peptide PQPQLPY (proline-glutamine- proline-glutamine-leucine-proline-tyrosine)

98 Annales Nestlé 2004;62:95-106 Pathogenesis of coeliac disease

from α-gliadin that can be deamidated to DQ2/8 gene dose and extent of gluten PQPELPY (E, exchange of a glutamine for a exposure determine disease manifestation glutamic acid residue) was identified as an immunodominant epitope [60]. Interestingly, Although the major determinants of coeliac a 33-mer peptide from α2-gliadin that con- disease, i.e. immunogenic gluten peptides, tains three partly overlapping stretches of this HLA-DQ2/DQ8 and tTG, and their interplay have epitope and is fairly resistant to digestion by been defined, their quantities, i.e. exposure, intestinal proteases [65, 68]. This stability is also contribute to the manifestation of coeliac attributable to a rigid, three-dimensional struc- disease in a given patient. Thus, an early and ture with a type II polyproline helix that makes massive gluten exposure favours coeliac di- the peptide inaccessible to endopeptidases. sease manifestation. This explains the 5-10 fold Neutralisation of the 33-mer peptide that can be higher prevalence of classical coeliac disease considered a gluten superantigen is therefore a in Swedish compared to Danish children, two key target for non-dietary approaches to treat otherwise genetically similar populations [74] coeliac disease. Clinical trials using bacterial (see A. Ivarsson and O. Hernell, in this issue, endopeptidase that degrades this and related pp. 107-118). Furthermore, patients can express peptides are underway. two copies or only a single copy of HLA-DQ2 Synthetic peptide libraries were used to (DQ8) that dramatically increases the likeli- define the primary sequence requirements for hood of effective antigen presentation and di- those glutamine residues that are substrates sease manifestation [75-78]. Lastly, tTG expres- for tTG. These studies showed that the rela- sion is upregulated by mechanical stress and by tive positions of proline to glutamine residues intestinal inflammation. [57, 79, 80] and it can are crucial. Thus, the sequences QXP and QXX be hypothesised that mechanical irritation and (X representing a hydrophobic amino acid) are proinflammatory non proliferating agents and targeted by tTG. On the otherhand, tTG is not microbes can trigger the disease in genetically active on the sequences QP or QXXP [61, 62, susceptible individuals whose coeliac disease 65, 68]. Based on these tTG recognition motifs, remained subliminal for years (Fig. 2). algorithms were designed and used to screen the available sequence data bases of cereal pro- teins for preferred tTG recognition sequences. The cytokine response in coeliac disease The authors detected ~50 T-cell epitopes in wheat gliadins, and similar epitopes in the Within the first two hours after a gliadin chal- phylogenetically related hordeins from bar- lenge, HLA molecules are upregulated on ley and secalins from rye [65, 68]. In addi- enterocytes and adjacent macrophages. This tion, some T-cell stimulatory sequences also response is followed by an increased expres- were found in the structurally unrelated wheat sion of intercellular adhesion molecule ICAM-1 glutenins, confirming prior observations with and T-cell activation [81]. Gluten ingestion in glutenin reactive T-cell clones [69]. The lack patients with untreated coeliac disease induces of optimal tTG recognition sequences and thus a non-proliferative activation of T-cell receptor ineffective deamidation by tTG in the avenins α/β positive CD4+ T-cells in the lamina propria of oats [65, 70], which have a high glutamine and a proliferative activation of T-cell receptor but a low proline content (Table I), lends scien- α/β- or γ/δ-positive CD8+ intraepithelial lym- tific support to clinical studies that showed phocytes (IEL) [82]. that pure oats are safe for coeliac patients [71, DQ2-restricted gliadin-specific T-cell clones 72]. A recent study from Norway, however, express mainly the Th1 type cytokine IFN-γ, described three patients who did not tolerate and an IFN-γ blocking antibody could prevent oats and presented with intestinal inflamma- histological damage to healthy mucosa in an tion following its consumption. Avenin reactive intestinal organ culture system exposed to T-cell clones were isolated from these patients supernatants of gliadin-specific T-cell clones [73], indicating that in rare occasions oats may from coeliac patients [83]. Immunohistochemi- result in coeliac disease. cal studies confirmed the increased expres-

Annales Nestlé 2004;62:95-106 99 Detlef Schuppan

α $1$1$1$1 li [88]. Therefore, IFN- treatment of patients with subliminal coeliac disease, e.g. for chronic hepatitis C, may be risky [89]. Contrary to its role in Crohn’s disease, IL-12, a prominent in- ducer of the Th1 reaction, does not appear to be rele vant in coeliac disease, since IL-12 is virtu- $1$1X

6ILLOUSATROPHY ally absent from biopsy specimens of coeliac -ICROBIALINFECTION patients [86, 90]. This role may be fulfilled by 4HRESHOLD IL-18 that is expressed in intestinal biopsies of OFGLUTEN INTOLERANCE patients with coeliac disease but not of controls [91]. IL-15 further adds to the complex cytokine profile of the disease. IL-15 shares many biologi- cal properties with IL-2, such as signalling via .ORMALVILLI the same β- and γ-chain of the IL-2 high affinity EARLY GLUTENDOSE receptor. It functions as a chemoattractant for Figure 2: DQ2 gene dose and gluten load. DQ2 gene peripheral blood lymphocytes and stimulates dose and the extent of (early) gluten exposure deter- the growth of activated T-cells and cytotoxic mine the manifestation and severity of the disease. Al- effector cells [92]. IL-15 can induce enterocyte though a low gene dose of DQ2 (see Fig. 1) and a care- proliferation and at the same time increase en- ful dietary introduction of gluten can keep intestinal terocyte apoptosis in biopsies of patients with inflammation subliminal, a high gene dose (and thus active coeliac disease [93]. A study using intes- surface expression) of DQ2, coupled with early and tinal biopsies suggested that the α-gliadin pep- massive gluten ingestion, can spark off overt coeliac disease. Additional triggers are probably mechani- tide p31-43, which does not bind to HLA-DQ2 or cal irritation and inflammation by chemicals or mi- -DQ8, can trigger IL-15 release from macropha- crobes that lead to enhanced intestinal permeability. ges and dendritic cells by activation of the innate (unspecific, pathogen recognising) immune system. This may prime the adaptive (HLA-me- diated, T-cell specific) immune system in pre- sion of IL-2, IFN-γ and TNF-α positive cells in disposed individuals for effective reaction with the lamina propria of patients with active coe- the immunodominant gluten peptides [94, 95] liac disease compared to patients in remission (Fig. 3). or controls [84]. Although a central role for Cytokines appear to be the major driving IFN-γ secreted by lamina propria lymphocytes force of tissue remodelling that results in vil- and IEL is undisputed, studies using quantita- lous atrophy and crypt hyperplasia, characte- tive polymerase chain reaction (PCR) in mate- ristic for coeliac disease. Using human foetal rial from intestinal biopsies did not show an intestinal explant cultures, Pender et al. increase of TNF-α and IL-2 in classical coeliac demonstrated that IFN-γ activates intestinal disease [85, 86]. Most studies demonstrated increased expression of IL-10 in active coeliac disease, with a shift from lamina propria lym- phocytes to IEL [85]. IL-10 suppresses Th1-cells Neutralisation of the 33-mer peptide and is likely acting as a counter-regulatory that can be considered a gluten cytokine. IFN-α that is found in the intestinal mucosa of coeliac patients but not of controls superantigen is a key target for non- [87] did not induce histological changes and dietary approaches to treat coeliac only minor IFN-γ and TNF-α production in foe- tal gut explants, but caused crypt hyperplasia disease. Clinical trials using bacterial and an an exaggerated Th1 response in com- endopeptidase that degrades this bination with a stimulatory anti-CD3 antibody, thus potentially facilitating coeliac disease and related peptides are underway. immunopathology in concert with other stimu-

100 Annales Nestlé 2004;62:95-106 Pathogenesis of coeliac disease

$IETARYGLUTENPEPTIDES "ACTERIALENDOPEPTIDASES 0ERMEABILITYINCREASE

)NFECTION CHEMICAL MECHANICALSTRESS 2ESORPTIVEVILLOUS EPITHELIUM

"ASEMENTMEMBRANE

2ELEASE ACTIVATION 'LUTENPEPTIDEDEAMIDATION ),  OFT4' CROSSLINKINGBYT4' ,AMINAPROPRIA &IBROBLASTS %PITHELIALREMODELLING %NDOTHELIAL INFLAMMATORYCELLS ,ATEEXPOSURETO,03 ),  )& A !NTIGENPRESENTATION $1 BLOCKINGPEPTIDES T4'INHIBITORS VIA(,! $1 !0# 4 CELLRECEPTOR STIMULATION -UCOSALDESTRUCTION KINES )MMUNEMODULATORS 4HCYTO 6ILLOUSATROPHY -ATURE %ARLYEXPOSURETO,03 ),  DENDRITICCELL )& G ),  4'&B ),  4H 4OLEROGENIC #$ CYTOKINES DENTRICCELL 4 CELL 'LUTENPEPTIDES ),  ),  0RODUCTIONOFANTIBODIES $EAMIDATED TOGLUTEN T4'NEOEPITOPES GLUTENPEPTIDES 4 CELLANERGY #$ ACTIVESUPRESSION 4R4HCELL T4' " CELL

Figure 3: Mechanisms of mucosal injury in coeliac disease. Dietary gluten peptides that reach lamina propria lead to enhanced intestinal expression of tTG. Cross-linking and particularly deamidation of gluten peptides by tTG create potent immunostimulatory epitopes that are presented via HLA-DQ2 or -DQ8 on antigen presen- ting cells, i.e. mature dendritic cells, macrophages or B-lymphocytes. Subsequently, CD4+ T-cells are activated, secreting mainly Th1-cytokines that via activation of macrophages cause the release and activation of matrix metalloproteinases (MMPs) by mesenchymal cells, that finally result in mucosal destruction. macro phages that in turn secrete TNF-α. TNF-α autoantibodies to tTG and since these autoan- potently stimulates the expression of matrix tibodies slowly disappear after institution of metalloproteases (MMPs) by intestinal myofi- a gluten free diet, coeliac disease is not con- broblasts, with MMP-1 and especially MMP-3 sidered a classical autoimmune disease. A mediating extracellular matrix degradation likely explanation for these autoantibodies is and remodelling [96]. In active coeliac disease, the incorporation of gliadins into complexes myofibroblast MMP-1 and MMP-3 mRNA are in- with tTG. In these complexes, tTG serves as creased in the subepithelial region, returning glutamine acceptor protein in a process of au- to normal after adherence to a gluten-free diet tocatalysis [43, 98]. Due to the close vicinity [97]. of gliadin and tTG in these complexes, gliadin- specific T-cells could help to stimulate the au- toantibody production by residual tTG-specific Autoantibodies to tTG B-cells that have escaped thymic selection [99]. Anti-tTG autoantibodies may not be innocent Since gluten ingestion drives the production bystanders but cause pathology themselves. of IgA- (and in the case of IgA deficiency IgG-) Indeed, in vitro gliadin challenge of small in-

Annales Nestlé 2004;62:95-106 101 Detlef Schuppan

testinal biopsies from treated coeliac patients structure resolution of the complex between induce production of IgA anti-tTG [100]. These DQ2 and the immunodominant α-gliadin pep- autoantibodies can block tTG activity and thus tide [104]; and 4) modulation of the intestinal inhibit the activation of latent TGF-β that is cytokine milieu, e.g. by immunomodulatory dependent on tTG [51]. The inhibition of TGF- cytokines or by cytokine antagonists [101] β activation by the anti-tTG antibodies could (Fig. 3). Apart from benefiting patients with contribute to observed intestinal epithelium coeliac disease, such novel therapies may final- dedifferentiation and dysfunction in coeliac ly provide a template for the treatment of other disease. Nonetheless, the inhibitory capacity of autoimmune diseases. the autoantibodies that are directed mainly to the amino- and carboxyterminal regions of tTG remains controversial. Acknowledgements

The author was supported by Deutsche For- Novel therapies and conclusions schungsgemeinschaft (Grant Schu 646/11-3), by the project QLK1-CT-1999-00037 of the European Due to tremendous recent advances, coeliac Union and the German Coeliac Association. dise ase could be considered the best under- stood HLA-linked disorder. The disease: i) is driven by a defined external (nutritional) trig- References ger, namely gluten proteins from wheat, rye and barley; ii) develops on the background of a 1. Trier JS. Celiac sprue. N Engl J Med 1991;325:1709- strong genetic association with HLA-DQ2/DQ8; 19. and iii) involves tTG as an autoantigen that 2. Schuppan D. Current concepts of celiac disease plays a key role in its pathogenesis, namely an- pathogenesis. Gastroenterology 2000;119:234-42. tigenic potentiation of gluten peptides. In ad- 3. Ciclitira PJ, King AL, Fraser JS. AGA technical re- dition, the amount of gluten ingested, the gene view on celiac sprue. American Gastroenterologi- cal Association. Gastroenterology 2001;120:1526- dose of HLA-DQ2 (DQ8) and the local expression 40. of tTG appear to be important determinants 4. Farrell RJ, Kelly CP. Celiac sprue. N Engl J Med of coeliac disease manifestation and severity. 2002;346:180-8. A sizable number of immunodominant gluten 5. Hardwick, C. Prognosis in coeliac disease. Arch peptides have been identified, part of them re- Dis Child 1939;14:279-89. sisting degradation by intestinal peptidases 6. Dicke WK, Weijers HA, van de Kamer JH. Coeliac and thus reaching the intestinal mucosa where disease. II. The presence in wheat of a factor hav- T-cells are recruited. Th1 cytokines released ing a deleterious effect in cases of coeliac disease. from these activated gluten-specific T-cells or Acta Paediatr 1953;42:34-42. from IEL cause the typical mucosal damage. 7. Van de Kamer JH, Weyers HA, Dicke WK. Coeliac An effective therapy of coeliac disease is ad- disease IV. An investigation into the injurious con- herence to a strictly gluten-free diet, a burden stituents of wheat in connection with their action for most patients. Therefore, alternative treat- on patients with coeliac disease. Acta Paediatr ments are being explored [101]. These include: 1953;42: 223-31. 1) degradation of immunodominant gliadin pep- 8. Paulley JW. Observation on the aetiology of idio- tides that resist intestinal proteases by use of pathic steatorrhoea; jejunal and lymph-node biop- exogenous bacterial prolyl endopeptidases [64, sies. Br Med J 1954;4900:1318-21. 9. Marsh MN. Gluten, major histocompatibility com- TG 67]; 2) inhibition of intestinal t activity by plex, and the small intestine: a molecular and im- specific inhibitors [102]; 3) inhibition of specif- munobiologic approach to the spectrum of gluten ic T-cell stimulation by peptides that only bind sensitivity (“celiac sprue”). Gastroenterology to HLA-DQ2 or -DQ8 DQ but not to the gluten pep- 1992; 102:330-54. tide-specific T-cell receptors [103], the targeted 10. Wong RC, Steele RH, Reeves GE, et al. Antibody design of such “inhibitors” being now feasible and genetic testing in coeliac disease. Pathology because of the successful crystallographic fine 2003;35:285-304.

102 Annales Nestlé 2004;62:95-106 Pathogenesis of coeliac disease

11. Catassi C, Ratsch IM, Fabiani E, et al. Coeliac dis- 27. Percopo S, Babron MC, Whalen M, et al. Saturation ease in the year 2000: exploring the iceberg. Lan- of the 5q31-q33 candidate region for coeliac dis- cet 1994;343:200-3. ease. Ann Hum Genet 2003;67:265-8. 12. Fasano A, Catassi C. Current approaches to diag- 28. Popat S, Hearle N, Hogberg L, et al. Variation in the nosis and treatment of celiac disease: an evolving CTLA4/CD28 gene region confers an increased risk spectrum. Gastroenterology 2001;120:636-51. of celiac disease. Ann Hum Genet 2002;66:125-37. 13. Mäki M, Mustalahti K, Kokkonen J, et al. Preva- 29. Holopainen P, Naluai AT, Moodie S, et al. Members lence of celiac disease among children in Finland. of the European Genetics Cluster on Coeliac Dis- N Engl J Med 2003;348:2517-24. ease. Candidate gene region 2q33 in European fam- 14. Wieser H. The precipitating factor in coeliac ilies with coeliac disease. Tissue Antigens 2004; disease. Baillieres Clin Gastroenterol 1995;9:191- 63:212-22. 207. 30. Bilbao JR, Martin-Pagola A, Vitoria JC, et al. HLA- 15. Shewry PR, Halford NG, Belton PS, Tatham AS. DRB1 and MHC class 1 chain-related A haplotypes The structure and properties of gluten: an elas- in Basque families with celiac disease. Tissue An- tic protein from wheat grain. Philos Trans R Soc tigens 2002;60:71-6. Lond B Biol Sci 2002;357:133-42. 31. Bolognesi E, Karell K, Percopo S, et al. Additional 16. Howdle PD, Ciclitira PJ, Simpson FG, Losowsky factor in some HLA DR3/DQ2 haplotypes confers MS. Are all gliadins toxic in coeliac disease? An in a fourfold increased genetic risk of celiac disease. vitro study of alpha, beta, gamma, and omega glia- Tissue Antigens 2003;61:308-16. dins. Scand J Gastroenterol 1984;19:41-7. 32. Ventura A, Magazzu G, Greco L. Duration of expo- 17. Ciclitira PJ, Evans DJ, Fagg NL, et al. Clinical test- sure to gluten and risk for autoimmune disorders ing of gliadin fractions in coeliac patients. Clin Sci in patients with celiac disease. SIGEP Study Group 1984;66:357-64. for Autoimmune Disorders in Celiac Disease. Gas- 18. De Ritis G, Auricchio S, Jones HW, et al. In vitro troenterology 1999;117:297-303. (organ culture) studies of the toxicity of specific 33. Sategna Guidetti C, Solerio E, et al. Duration of A-gliadin peptides in celiac disease. Gastroente- gluten exposure in adult coeliac disease does not rology 1988;94:41-9. correlate with the risk for autoimmune disorders. 19. Sturgess R, Day P, Ellis HJ, et al. Wheat peptide Gut 2001;49:502-5. challenge in coeliac disease. Lancet 1994;343:758- 34. Schuppan D, Ciccocioppo R. Coeliac disease and 61. secondary autoimmunity. Dig Liver Dis 2002; 20. Howell MD, Austin RK, Kelleher D, et al. An HLA-D 34:13-5. region restriction length polymorphism associated with celiac disease. J Exp Med 1986;164:333-9. 35. Luft LM, Barr SG, Martin LO, et al. Autoantibodies 21. Sollid LM, Markussen G, Ek J, et al. Evidence to tissue transglutaminase in Sjogren’s syndrome for a primary association of celiac disease to a and related rheumatic diseases. J Rheumatol 2003; particular HLA-DQ α/β heterodimer. J Exp Med 30:2613-9. 1989;169:345-50. 36. King AL, Moodie SJ, Fraser JS, et al. Coeliac dis- 22. Sollid LM, Thorsby E. HLA susceptibility genes in ease: investigation of proposed causal variants in celiac disease: genetic mapping and role in patho- the CTLA4 gene region. Eur J Immunogenet 2003; genesis. Gastroenterology 1993;105:910-22. 30:427-32. 23. Greco L, Romino R, Coto I, et al. The first large 37. Naluai AT, Nilsson S, Samuelsson L, et al. The population based twin study of coeliac disease. CTLA4/CD28 gene region on chromosome 2q33 Gut 2002;50:624-8. confers susceptibility to celiac disease in a way 24. Karell K, Louka AS, Moodie SJ, et al. HLA types in possibly distinct from that of type 1 diabetes and celiac disease patients not carrying the DQA1*05- other chronic inflammatory disorders. Tissue An- DQB1*02 (DQ2) heterodimer: results from the Eu- tigens 2000;56:350-5. ropean Genetics Cluster on Celiac Disease. Hum 38. Susi M, Holopainen P, Mustalahti K, et al. Candi- Immunol 2003;64:469-77. date gene region 15q26 and genetic susceptibil- 25. Greco L, Babron MC, Corazza GR, et al. Exist- ity to coeliac disease in Finnish families. Scand J ence of a genetic risk factor on chromosome 5q in Gastroenterol 2001;36:372-4. Italian coeliac disease families. Ann Hum Genet 39. Collin P, Salmi J, Hallstrom O, et al. Autoimmune 2001;65:35-41. thyroid disorders and coeliac disease. Eur J Endo- 26. Babron MC, Nilsson S, Adamovic S, et al. Europe- crinol 1994;130:137-40. an Genetics Cluster on Coeliac Disease. Meta and 40. Valentino R, Savastano S, Tommaselli AP, et al. pooled analysis of European coeliac disease data. Prevalence of coeliac disease in patients with thy- Eur J Hum Genet 2003;11:828-34. roid autoimmunity. Horm Res 1999;51:124-7.

Annales Nestlé 2004;62:95-106 103 Detlef Schuppan

41. Volta U, Ravaglia G, Granito A, et al. Coeliac dis- 56. Godkin A, Friede T, Davenport M, et al. Use of elu- ease in patients with autoimmune thyroiditis. Di- ted peptide sequence data to identify the binding gestion 2001;64:61-5. characteristics of peptides to the insulin-depen- 42. Badenhoop K, Dieterich W, Segni M, et al. HLA dent diabetes susceptibility allele HLA-DQ8 (DQ DQ2 and/or DQ8 is associated with celiac disease- 3.2). Int Immunol 1997;9:905-11. specific autoantibodies to tissue transglutaminase 57. Molberg Ø, McAdam SN, Koerner R, et al. Tissue in families with thyroid autoimmunity. Am J Gas- transglutaminase selectively modifies gliadin pep- troenterol 2001;96:1648-9. tides that are recognized by gut-derived T cells in 43. Dieterich W, Ehnis T, Bauer M, et al. Identification celiac disease. Nat Med 1998;4:713-7. of tissue transglutaminase as the autoantigen of 58. Van de Wal Y, Kooy Y, van Veelen P, et al. Selective celiac disease. Nat Med 1997;3:797-801. deamidation by tissue transglutaminase strongly 44. Piacentini M, Rodolfo C, Farrace MG, Autuori F. enhances gliading-specific T cell reactivity. J Im- Tissue transglutaminase in animal development. munol 1998;161:1585-8. Int J Dev Biol 2000;44:655-62. 59. Arentz-Hansen H, Körner R, Molberg Ø, et al. The α 45. Aeschlimann D, Thomazy V. Protein crosslinking intestinal T cell response to -gliadin in adult ce- in assembly and remodelling of extracellular ma- liac disease is focused on a single deamidated trices: the role of transglutaminases. Connect Tis- glutamine targeted by tissue transglutaminase. J sue Res 2000;41:1-27. Exp Med 2000;191: 603-12. 46. Lorand L, Graham RM. Transglutaminases: 60. Anderson RP, Degano P, Godkin AJ, et al. In vivo crosslinking enzymes with pleiotropic functions. antigen challenge in celiac disease identifies a sin- Nat Rev Mol Cell Biol 2003;4:140-56. gle transglutaminase-modified peptide as the dom- inant A-gliadin T-cell epitope. Nat Med 2000;6:337- 47. Haroon ZA, Hettasch JM, Lai TS, et al. Tissue 42. transglutaminase is expressed, active, and direct- 61. Arentz-Hansen H, McAdam SN, Molberg Ø, et al. ly involved in rat dermal wound healing and angio- Celiac lesion T cells recognize epitopes that clus- genesis. FASEB J 1999;13:1787-95. ter in regions of gliadins rich in proline residues. 48. Lesort M, Tucholski J, Miller ML, Johnson GV. Tis- Gastroenterology 2002;123:803-9. sue transglutaminase: a possible role in neurode- 62. Fleckenstein B, Molberg Ø, Qiao SW, et al. Gliadin generative diseases. Prog Neurobiol 2000;61:439- T cell epitope selection by tissue transglutaminase 63. in celiac disease. J Biol Chem 2002;277:34109-16. 49. Gentile V, Cooper AJ. Transglutaminases (possible 63. Vader W, Kooy Y, Van Veelen P, et al. The gluten re- drug targets in human diseases. Curr Drug Tar- sponse in children with celiac disease is directed gets CNS Neurol Disord 2004;3:99-104. toward multiple gliadin and glutenin peptides. 50. Folk JE, Chung SI. Transglutaminases. Meth En- Gastroenterology 2002;122:1729-37. zymol 1985;113:358-75. 64. Shan L, Molberg Ø, Parrot I, et al. Structural basis 51. Nunes I, Gleizes PE, Metz CN, Rifkin DB. Latent for gluten intolerance in celiac sprue. Science transforming growth factor-β binding protein 2002;297:2275-9. domains involved in activation and transglutami- 65. Vader WL, de Ru A, van der Wal Y, et al. Specificity nase-dependent cross-linking of latent transform- of tissue transglutaminase explains cereal toxicity ing growth factor-β. J Cell Biol 1997;136:1151-63. in celiac disease. J Exp Med 2002;195:643-9. 52. Halttunen T, Marttinen A, Rantala I, et al. Fibrob- 66. Skovbjerg H, Koch C, Anthonsen D, Sjostrom H. lasts and transforming growth factor ß induce Deamidation and cross-linking of gliadin peptides organization and differentiation of T84 human epi- by transglutaminases and the relation to celiac thelial cells. Gastroenterology 1996;111:1252-62. disease. Biochim Biophys Acta 2004;1690:220-30. 53. Nakaoka H, Perez DM, Baek KJ, et al. Gh: A GTP- 67. Qiao SW, Bergseng E, Molberg O, et al. Antigen binding protein with transglutaminase activity presentation to celiac lesion-derived T cells of a and receptor signaling function. Science 1994;264: 33-mer gliadin peptide naturally formed by gastro- 1593-6. intestinal digestion. J Immunol 2004;173:1757-62. 54. Vartdal F, Johansen BH, Friede T, et al. The pep- 68. Vader LW, Stepniak DT, Bunnik EM, et al. Charac- tide binding motif of the disease associated HLA- terization of cereal toxicity for celiac disease pa- DQ (α1*0501, β1*0201) molecule. Eur J Immunol tients based on protein homology in grains. Gas- 1996;26:2764-72. troenterology 2003;125:1105-13. 55. Van der Wal Y, Kooy YMC, Drijfhout JW, et al. Pep- 69. Molberg O, Solheim Flaete N, Jensen T, et al. In- tide binding characteristics of the coeliac disease- testinal T-cell responses to high-molecular-weight associated DQ (α1*0501, β1*0201) molecule. Im- glutenins in celiac disease. Gastroenterology 2003; munogenetics 1996;44:246-53. 125:337-44.

104 Annales Nestlé 2004;62:95-106 Pathogenesis of coeliac disease

70. Janatuinen EK, Pikkarainen PH, Kemppainen 83. Przemioslo RT, Lundin KE, Sollid LM, et al. Histo- TA, et al. A comparison of diets with and without logical changes in small bowel mucosa induced by oats in adults with celiac disease. N Engl J Med gliadin sensitive T lymphocytes can be blocked by 1995;333:1033-7. anti-interferon gamma antibody. Gut 1995;36:874- 71. Hogberg L, Laurin P, Falth-Magnusson K, et al. Oats 9. to children with newly diagnosed coeliac disease: 84. Westerholm-Ormio M, Garioch J, Ketola I, Savi- a randomised double blind study. Gut 2004;53:649- lahti E. Inflammatory cytokines in small intestinal 54. mucosa of patients with potential coeliac disease. 72. Kilmartin C, Lynch S, Abuzakouk M, et al. Avenin Clin Exp Immunol 2002;128:94-101. fails to induce a Th1 response in coeliac tissue fol- 85. Forsberg G, Hernell O, Melgar S, et al. Paradoxical lowing in vitro culture. Gut 2003;52:47-52. coexpression of proinflammatory and down-regu- 73. Arentz-Hansen H, Fleckenstein B, Molberg O, et al. latory cytokines in intestinal T cells in childhood The molecular basis for oat intolerance in patients celiac disease. Gastroenterology 2002;123:667-78. with celiac disease. Plos Med 2004;1:in press. 86. Ciccocioppo R, Di Sabatino A, Bauer M, et al. Cy- 74. Weile B, Cavell B, Nivenius K, Krasilnikoff PR. tokine and matrix metalloproteinase pattern in Striking difference in the incidence of childhood coeliac duodenal mucosa. Lab Invest 2005, in celiac disease between Denmark and Sweden: a press. plausible explanation. J Pediatr Gastroenterol 87. Monteleone G, Pender SL, Alstead E, et al. Role of Nutr 1995;21:64-8. interferon alpha in promoting T helper cell type 1 responses in the small intestine in coeliac disease. 75. Petronzelli F, Multari G, Ferrante P, et al. Different Gut 2001;48:425-9. dose effect of HLA-DQ alpha beta heterodimers in 88. Monteleone G, Pender SL, Wathen NC, MacDonald insulin-dependent diabetes mellitus and celiac dis- TT. Interferon-alpha drives T cell-mediated im- ease susceptibility. Hum Immunol 1993;36:156-62. munopathology in the intestine. Eur J Immunol 76. Congia M, Cucca F, Frau F, et al. A gene dosage ef- 2001;31:2247-55. fect of the DQA1*0501/DQB1*0201 allelic combina- 89. Cammarota G, Cuoco L, Cianci R, et al. Onset of tion influences the clinical heterogeneity of celiac coeliac disease during treatment with interferon disease. Hum Immunol 1994;40:138-42. for chronic hepatitis C. Lancet 2000;356:1494-5. 77. Zubillaga P, Vidales MC, Zubillaga I, et al. HLA- 90. Nilsen EM, Jahnsen FL, Lundin KE, et al. Gluten DQA1 and HLA-DQB1 genetic markers and clinical induces an intestinal cytokine response strongly presentation in celiac disease. J Pediatr Gastroen- dominated by interferon gamma in patients with terol Nutr 2002;34:548-54. celiac disease. Gastroenterology 1998;115:551-63. 78. Vader W, Stepniak D, Kooy Y, et al. The HLA-DQ2 91. Salvati VM, MacDonald TT, Bajaj-Elliott M, et gene dose effect in celiac disease is directly related al. Interleukin 18 and associated markers of T to the magnitude and breadth of gluten-specific T helper cell type 1 activity in coeliac disease. Gut cell responses. Proc Natl Acad Sci USA 2003;100: 2002;50:186-90. 12390-5. 92. Waldmann TA. The IL-2/IL-15 receptor systems: 79. Hansson T, Ulfgren AK, Lindroos E, et al. Trans- targets for immunotherapy. J Clin Immunol 2002; forming growth factor-β (TGF-β) and tissue trans- 22:51-6. glutaminase expression in the small intestine in 93. Maiuri L, Ciacci C, Auricchio S, et al. Interleukin 15 children with coeliac disease. Scand J Immunol mediates epithelial changes in celiac disease. Gas- 2002;56:530-7. troenterology 2000;119:996-1006. 80. Brusco G, Muzi P, Ciccocioppo R, et al. Trans- 94. Maiuri L, Ciacci C, Ricciardelli I, et al. Association glutaminase and coeliac disease: endomysial re- between innate response to gliadin and activation activity and small bowel expression. Clin Exp Im- of pathogenic T cells in coeliac disease. Lancet munol 1999;118:371-5. 2003:362:30-37. 81. Maiuri L, Picarelli A, Boirivant M, et al. Definition 95. Schuppan D, Esslinger B, Dieterich W. Innate im- of the initial immunologic modifications upon in munity and coeliac disease. Lancet 2003:362;3-4. vitro gliadin challenge in the small intestine of ce- 96. Pender SLF, Tickle SP, Docherty AJP, et al. A major liac patients. Gastroenterology 1996;110:1368-78. role for matrix metalloproteinases in T cell injury 82. Halstensen TS, Brandtzaeg P. Activated T lym- in the gut. J Immunol 1997;158:1582-10. phocytes in the celiac lesion: non-proliferative 97. Daum S, Bauer U, Foss HD, et al. Increased expres- activation (CD25) of CD4+ alpha/beta cells in the sion of mRNA from matrix metalloproteinases-1 lamina propria but proliferation (Ki-67) of alpha/ and -3 and tissue inhibitor of metalloproteinases-1 beta and gamma/delta cells in the epithelium. Eur in intestinal biopsy specimens from patients with J Immunol 1993;23:505-10. coeliac disease. Gut 1999;44:17-25.

Annales Nestlé 2004;62:95-106 105 Detlef Schuppan

98. Fleckenstein B, Qiao SW, Larsen MR, et al. Mo- 102. Hausch F, Halttunen T, Mäki M, Khosla C. Design, lecular characterization of covalent complexes synthesis and evaluation of gluten peptide ana- between tissue transglutaminase and gliadin pep- logs as selective inhibitors of human tissue trans- tides. J Biol Chem 2004;279:17607-16. glutaminase. Chem Biol 2003;10:225-31. 99. Sollid LM, Molberg O, McAdam S, Lundin KEA. Au- 103. Biagi F, Ellis HJ, Parnell NDJ, et al. A non-toxic toantibodies in coeliac disese: tissue transglutam- analogue of a coeliac-activating gliadin peptide: a inase - guilt by association? Gut 1997;41:851-2. basis for immunomodulation? Aliment Pharmacol 100. Picarelli A, Maiuri L, Frate A, et al. Produc- Ther 1999;13:945-50. tion of antiendomysial antibodies after in vitro 104. Kim CY, Quarsten H, Bergseng E, et al. Structural gliadin challenge of small intestine biopsy sam- basis for HLA-DQ2-mediated presentation of glu- ples from patients with coeliac disease. Lancet ten epitopes in celiac disease. Proc Natl Acad Sci 1996;348:1065-7. USA 2004;101:4175-9. 101. Schuppan D, Hahn EG. Gluten and the gut-lessons for immune regulation. Science 2002;297:2218-20.

106 Annales Nestlé 2004;62:95-106