Open access peer-reviewed chapter

Immune Checkpoints as a Novel Source for Diagnostic and Therapeutic Target in Celiac Disease

Written By

Isabel Torres, Miguel Ángel López Casado, Teresa Palomeque and Pedro Lorite

Submitted: 30 September 2020 Reviewed: 14 January 2021 Published: 16 March 2021

DOI: 10.5772/intechopen.96022

From the Edited Volume

Celiac Disease

Edited by Jianyuan Chai

Chapter metrics overview

481 Chapter Downloads

View Full Metrics

Abstract

Celiac disease, as an autoimmune disorder, is a disease which appears in sensing and immune reaction responses to gluten. It has been confirmed that both genetic and environmental factors are involved. CD is strongly associated with the HLA alleles DQB1*02 (serological DQ2) or DQB1*0302 (serological DQ8). These HLA alleles are necessary but not sufficient for the development of CD and non-HLA risk genes also contribute to disease susceptibility. Several studies have identified linkage or association of CD with the 2q33 locus, a region harboring the candidate genes CD28, CTLA4 and ICOS, important immune checkpoints regulators of T-cell activity. Immune checkpoints are crucial to maintain self-tolerance and protect self-tissue from damage during an ongoing immune response.

Keywords

  • immune checkpoints
  • celiac disease
  • PD1
  • PDL
  • HLA-G
  • CTLA4
  • IDO
  • tryptophan

1. Introduction

Celiac disease is a unique autoimmune disorder in that the key genetic components (HLA class II genes DQ2 and/or DQ8) are present in almost all patients, the autoantigen is known (tTG), and, most importantly, the environmental trigger is known (gluten) [1, 2, 3, 4, 5]. The HLA-DQ molecules predispose to disease by preferential presentation of gluten antigens to CD4+ T cells [6, 7, 8]. These genotypes are necessary for the development of the disease, but they are not the only ones responsible, since these genes are present in the population, and only 1% develop CD [9]. Furthermore, in recent years, other areas of the genome outside the HLA region have also been identified that could influence susceptibility to CD, many are related to immunity, especially with T-cell and B-cell function [10, 11].

Gluten ingestion by patients with CD leads to a cascade of inflammatory reactions and eventually to the hallmark small-intestinal lesion. The most important consequence is reduced nutrient uptake characterized by CD4+ T-cell activation, increasing numbers of intraepithelial lymphocytes with partial to total intestinal villus atrophy [12, 13, 14, 15]. A common feature of gluten-derived epitopes is the presence of multiple proline and glutamine residues that are selectively deamidated by tTG. The passage of immunogenic peptides to the lamina propria stimulates specific CD4 + T lymphocytes when they occur together with HLA-DQ2/DQ8 molecules, after having a modification by tissue transglutaminase (TGt). Proinflammatory cytokine responses are activated and mechanisms causing mucosal alteration. Activation of these gluten-reactive CD4 + T cells lead to a pro-inflammatory response dominated by IFN-γ production [16, 17, 18].

The response of CD4 + T cells to post-translationally modified gluten and highly disease specific B cells to deamidated gluten and transglutaminase 2 (TG2) autoprotein are present in the pathogenesis in CD [12]. When immunogenic gluten peptides cross the intestinal lumen, they can trigger an innate and adaptive immune response, leading to the development of clinical and histological manifestations of CD [19].

The immune homeostasis has to be precisely maintained in a physiological state, through a balance of costimulatory (e.g. CD28) and coinhibitory (e.g. CTLA-4 or PD-1) immune signals known as “immune checkpoints”. Immune checkpoints are essential for maintaining self-tolerance, protecting tissues from damage caused by the immune system, and providing protective immunity [20]. An imbalance in immune homeostasis can lead to costimulation and the upregulation of T-cell activation in autoimmune diseases [21].

During the normal activation state, CD4+ and CD8+ T cells express multiple immune checkpoint molecules, and some of them also serve a costimulatory function of T cells activation. T cells obtained from individuals with autoimmune conditions have enhanced expression of these molecules that represent an activated T cell state. T lymphocytes play a central role in the induction of an effective adaptive immune response and responsible for maintaining immune homeostasis. Signaling through two well-known negative regulators or checkpoints of T cells, CTLA4 and PD1 leads to direct inhibition of T cell responses [20].

The present chapter discuss the role of the immune checkpoints in intestinal tissue homeostasis and tolerance, and speculate how genetic and environmental factors can regulate them in celiac disease.

Advertisement

2. Immune checkpoints

Immune checkpoints play an essential role in the function and regulation of effector T cells (Teff) and regulator T cells (Treg). Immune checkpoint molecules limits excessive T cell-mediated inflammatory responses and in their signaling processes include a series of ligands that are expressed on the membrane of antigen-presenting cells (APCs) transmitting inhibitory signals (Figure 1). These molecules employ specific receptor partners expressed by T lymphocytes and drive their activation and differentiation or promote immunoregulatory effects. Dysregulation of these signaling processes has been associated with autoimmunity and chronic inflammation.

Figure 1.

Schematic representation of immune checkpoint molecules including a series of ligands expressed on the membrane of antigen-presenting cells (APCs), that engage specific receptor partners expressed by T lymphocytes and either drive their activation and differentiation (positive immune checkpoint molecules) or promote immunoregulatory effects (negative immune checkpoint molecules).

Autoimmune diseases are heterogeneous conditions involving breakdown of tolerance and consequent activation of autoreactive immune cells [22]. The failure of immune checkpoints has been described in inflammatory myopathies with the involvement of autoimmune features [23], as well as in diabetes, multiple sclerosis, systemic lupus erythematosus and celiac disease [24]. Inhibitory checkpoint molecules have been considered as new targets in personalized cancer immunotherapy for their potential in multiple cancer types [21, 25]. An active area of research is the analysis of the functions of these checkpoint molecules and its ligands in tolerance and autoimmunity.

The immune system has the difficult dual function of discerning and defending against a variety of pathogens and avoiding self-reactivity. To further control the development of autoimmunity, multiple mechanisms of peripheral tolerance have evolved, including T-cell anergy, deletion and suppression by Tregs cells. Treg and Teff cells help maintain immune homeostasis through mutual regulation. Loss of homeostatic balance between Teff/Treg cells is often associated with autoimmunity [24, 26].

In this chapter, we will discuss the biology of immune checkpoints, highlight research strategies that may help reduce the incidence of immune related adverse events associated with celiac disease, and also suggest investigational approaches to manipulate immune checkpoints to treat its autoimmune disorder.

2.1 IDO/kynurenine pathway

An important inhibitory checkpoint is now considered to be the tolerogenic mechanism of the enzyme indoleamine-2, 3-dioxygenase (IDO), an intracellular protein involved in the oxidative catabolism of tryptophan (Trp). It catalyzes the conversion of Trp to N-formyl kynurenine via the kynurenine pathway. Depletion of Trp reduces T cell proliferation, whereas the production of kynurenine induces apoptosis of type 1 T helper (Th1) cells and naïve T cell differentiation into Tregs cells [27, 28].

IDO is expressed intracellularly in a constitutive manner in the placenta, epididymis, prostate, esophagus, intestine, colon, cecum, spleen, thymus, lung, brain, and skin [29, 30]. Notably, the morphological features of many IDO-expressing cells closely resemble those of antigen-presenting cells and epithelial cells [29].

IDO expression is inducible by inflammatory stimuli including cytokines and toll like receptor agonists. IDO is expressed in antigen presenting cells, macrophages and dendritic cells, and activation of IDO during the inflammatory response leads to a decrease in local trp levels [27]. These decreased levels have an inhibitory effect on the proliferation of T lymphocytes, directly or indirectly via activation of regulatory T cells [27, 31].

In most cell types, IDO is induced at the transcriptional level in response to specific inflammatory stimuli. IFN-γ is the principal IDO inducer in vitro and in vivo. Exposure to IFN-γ increases IDO transcription in monocyte/macrophages [32] and dendritic cells [33], fibroblasts [34], epithelial cells [35], smooth muscle cells [36]. Other inflammatory stimuli, such as IFN-α, IFN-β, LPS and cytotoxic T lymphocyte-associated antigen (CTLA)-4, also induce IDO to a lesser degree than that of IFN-γ [37, 38].

A dysfunctional IDO has recently been associated with a specific single nucleotide polymorphism (SNP) and with the occurrence of autoimmune diabetes and multiple sclerosis. The elevated levels of kynurenines that are present contain the proliferation of Teff cells and favor the differentiation of Treg cells [39]. Several genetic variations of the IDO gene have been associated with the occurrence and severity of autoimmune/chronic inflammatory diseases; however, the functional relevance of these variations, which mainly affect the intron regions and the promoter portion of IDO, has not been well characterized yet [40, 41].

In celiac disease, mechanisms dependent on tryptophan catabolism may be involved in the regulation of the immune response. Thus, in intestinal biopsies from celiac patients, a high expression of the anti-inflammatory enzyme IDO appears [42]. This increase in IDO levels results in an increase in serum levels of kynurenine in patients with celiac disease, which potentially contributes to intensify inflammation. Likewise, higher levels of kynurenine were found in celiac patients with other associated diseases, such as Down syndrome or autoimmune thyroiditis, contributing to the pathology [43].

Once an autoimmune disorder is established, the presence of chronic inflammation might provoke sustained IDO production and IDO fails to limit immune deregulation under these pathological conditions. Several inflammatory mediators including the most potent, IFN-γ, induce IDO production. Wolf et al. [44], have described overexpression of IDO in other Th1-associated chronic inflammatory disease of the gastrointestinal tract, such as Crohn’s disease, with increased kynurenine levels and a higher kynurenine/tryptophan ratio. In this pathology has been demonstrated that T helper 1 (Th1)-like cytokines such as IFN-γ and TNF-α are potent inducers of IDO expression.

2.2 HLA-G/ILT interaction is an immune checkpoint

The HLA-G gene is a non-classical class I HLA composed of eight exons and seven introns located on chromosome 6 at region 6p21.3, [45]. As a result of alternative RNA splicing, seven isoforms can be formed, comprising four membrane-bound isoforms (HLA-G1, G2, G3, and G4), and three secreted soluble isoforms (HLA-G5, G6, and G7) [46, 47]. Most studies focus on the full-length molecule (HLA-G1) and its soluble isoform (HLA-G5). These isoforms are identical, except HLA-G5 is missing the transmembrane domain.

HLA-G is considered to be an immune checkpoint molecule, a function that is closely linked to the structure and dynamics of the different HLA-G isoforms. The expression of HLA-G can be induced in several conditions, including cancer, transplantation, viral infections, and autoimmune and inflammatory diseases [48, 49, 50, 51].

HLA-G mediates its function by binding to receptors on immune cells. The known receptors are leukocyte Ig-like receptor subfamily B member 1 (LILRB1) and member 2 (LILRB2), also known as ILT2 and ILT4, and the killer immunoglobulin-like receptor 2DL4 (KIR2DL4) [52]. ILT2 is expressed by B cells, some subtypes of T cells and NK cells, and all monocytes/dendritic cells. On the other hand, ILT4 is myeloid-specific and only expressed by monocytes/dendritic cells [53].

HLA-G expression and gene polymorphisms have been associated with several disorders [54, 55, 56]. HLA-G has an important role in regulating the immune system; indeed, the molecule is able to inhibit the cytotoxic activity of Natural Killer cells (NK) and T cell-mediated cytolysis (CTL) [57]. HLAG can inhibit the response of alloproliferative CD4+ T cells, proliferation of T and NK cells, and the maturation and function of antigen presenting cells (APC) [58, 59]. In addition, HLA-G has a tolerogenic effect due to its capacity of generating suppressor cells by binding to specific receptors and it can induce apoptosis in endothelial cells [60].

The soluble form of HLA-G is of special interest in celiac disease because its molecule plays an important role in the induction of immune tolerance [61]. In this sense, soluble HLA-G has the function to inhibit the proliferation of activated T cells, and to induce apoptosis of T cells dose dependently, reinforcing the immune inhibitory role of soluble HLA-G capable to be secreted during CD as part of a mechanism to restore the tolerance process towards oral antigens [61, 62]. A potent anti-inflammatory response to gliadin may occur during disease development as a result of the adaptive response in CD. In celiac patients, gluten intake appears to cause an overreaction in intraepithelial T lymphocytes, with uncontrolled production of the HLA-G molecule [61]. This can cause the recruitment of intraepithelial lymphocytes, leading to amplified immune activity and maintenance of intestinal lesions. The increased expression of the soluble form of HLA-G in patients with CD could be part of a mechanism to restore gluten tolerance [61, 62].

Moreover, an association between HLA-G polymorphism and CD has already been described by Fabris et al. [63]. The 14 bp inserted (I) allele and the homozygous I/I genotype were significantly more frequent in CD patients than in healthy controls. The effect of the HLA-G D/I polymorphism is restricted for HLA-DQ2, and not simply due to the presence of linkage disequilibrium with the major known risk factor. In this sense, the risk conferred by HLA-DQ2 alone and that subjects that carry both DQ2 and HLA-G I alleles have an increased risk of CD than subjects that carry DQ2 but not the 14 bp inserted (I) allele [63]. The modulation of the HLA-G transcript stability is especially known for the 14 bp D/I polymorphism, which has been associated with autoimmunity [64, 65, 66, 67]. Based on the findings of Torres et al. [61] and considering that the presence of HLA-G SNPs affect the mRNA stability in CD patients, lower basal levels of HLA-G molecule, possibly due to the presence of genetic variations, can increase the risk of celiac disease development. Once that the disease has occurred the organism produces higher levels of soluble HLA-G trying to restore the immune tolerance [61, 62]. Similarly to 5’URR, also 3’UTR presents numerous polymorphic sites that could affect HLA-G transcription and/or translation [68]. By sequencing this region, there are 4 polymorphisms showing some significant associations with CD [64].

In summary, it has been shown that both HLA-G and IDO suppressor molecules are expressed in CD. The expression of these molecules, IDO and HLA-G, would be an essential mechanism to try to restore tolerance towards antigens in the diet [61, 62]. Therefore, the increase in IDO activity reflects an attempt to control chronic antigenic stimulation by downregulating the T cell-mediated autoimmune reaction. IDO and HLA-G could cooperate to suppress the immune response in CD in their active form [61].

López et al. [69] showed that suppressive molecules IDO and HLA-G are both expressed in dendritic cells, and these molecules can produce immunosuppression. Besides, IDO was shown to induce HLA-G expression during monocyte differentiation into DCs [69]. IDO and HLA-G share some properties: both have tolerogenic capacity, are highly expressed in human placenta and tumors [70, 71] and their expression can be regulated by the same cytokines (IFN-γ, IL-10) [72, 73]. The effect of IDO on HLA-G cell-surface expression seemed to be dependent on the type of cell studied and is likely to involve posttranslational mechanisms [74]. The inhibition of IDO function with 1-methyl tryptophan in antigen-presenting cells (APCs), which are originally HLA-G cell-surface negatives, increases the levels of HLA-G1 cell-surface expression, whereas high concentrations of tryptophan caused a loss of HLA-G1 expression in HLA-G1-positive cells [75].

2.3 CD28/CTLA4-B7 pathway

The immune regulatory proteins cytotoxic T lymphocyte antigen (CTLA-4) is important immune regulatory protein collectively referred to as immune checkpoint receptor. CTLA-4 is known to be crucial for tolerance induction in the early stages of the immune response being an important negative regulator of T-cell activation and proliferation, interacting with B7 molecules on antigen-presenting cells [76, 77]. The CTLA4 gene encodes a receptor involved in the control of T-cell proliferation and mediates T-cell apoptosis. CTLA-4 is therefore a plausible candidate for a susceptibility gene in diseases with T-cell mediated pathogenesis [77].

The chromosomal region 2q33 contains immunologically important genes, CD28 and ICOS, that has been associated with autoimmune disease, but the exact causal genetic sequence variation has yet to be established in CD [78]. There is good evidence that the CTLA-4 region on chromosome 2q33 contains a non-HLA susceptibility locus for celiac disease, although its participation may vary according to the geographic origin of the patients [79]. The association of several SNPs in the CTLA4 gene with CD, among them, the functional SNP, CT60, suggested that the CT60 polymorphism influences alternate RNA splicing of CTLA4, resulting in differing ratios of a full-length form, flCTLA4, and a soluble form, sCTLA4 of the protein [78].

High levels of serum soluble CTLA-4 in active celiac patients were found and are related to gluten intake. A positive correlation exists between autoantibodies to tissue transglutaminase, the grade of gut mucosa damage and soluble CTLA-4 concentration [80]. This correlation between the amount of serum sCTLA-4 and the grade of gut mucosa damage strongly suggests a possible immunomodulatory effect of this soluble molecule on cytotoxic T lymphocyte functions. Thus, soluble CTLA4 appears to be related to autoantibody production per se, independently from dietary gluten [80]. Soluble CTLA-4 could play a critical role in modulating the immune response, especially in the early stage. The immunomodulatory effect of soluble CTLA-4 could be involved in the regulation of B cell activation directly or via T helper function modulation [80].

The detection of the spliced/soluble variant from CD patients suggests that the soluble CTLA-4 does not result from a cleavage of the full-length form [80]. The potential genetic associations of several CTLA-4 polymorphisms to susceptibility to autoimmune diseases have been described, although the relationship between CTLA-4 polymorphisms and the ability to produce the soluble form is not fully clarified. CTLA-4 is a strong actor in the adaptive response.

2.4 PD1/PDL pathway

Programmed cell death-1 (PD-1) is a well-established immune checkpoint and co-inhibitory regulator critical to the maintenance of immune tolerance. PD-1 through binding to its PD-L1 and PD-L2 ligands, generate inhibitory signals that regulate the balance between immune system activation, tolerance and immunopathology [81]. The PD1 expression has been noted in activated CD4+/CD8+ T cells, a subset of Tregs, B-cells, myeloid DCs, monocytes, exhausted T cells and basal mesenchymal stem cells [82]. Basal levels of expression of ligand PDL1 was observed in mesenchymal stem cells and vascular endothelium. In addition, activated B-cells, DCs and monocytes also express both PD-L1 and PD-L2 [82]. Lower levels of PDL1 expression was reported in unstimulated CD4+/CD8+ T-cells which was increased upon activation.

Cytokines are potent stimuli for PD-L1 and PD-L2 expression. Type 1 and type 2 interferons and TNF-α induce PD-L1 expression in T cells, B cells, endothelial cells, and epithelial cells [83]. IL-2, IL-7, and IL-15 cytokines increase PD-L1 on human T cells. IL-21 can stimulate PD-L1 expression on B cells and IL-10 induces PD-L1 on monocytes. Expression of PD-L2 is stimulated by interferons, IL-4, and GMCSF on dendritic cells in vitro, and the common γ chain cytokines can induce PD-L1 and to a lesser extent PD-L2 on human monocytes/macrophages [83].

PD-1 plays a role in differentiating naive T cells into Treg cells and can inhibit T-cell responses by developing Treg cells [84]. The PD-1 upregulation is a consequence of the activation of T cells, which is essential to the immune responses. PD-1 expression in Tregs is indispensable for their suppressive functions, and loss of PD1 expression accelerates the generation of Tregs which lose Foxp3 expression and produce pro-inflammatory cytokines and thereby flare autoimmunity [84]. Tregs and the PD-1/ PD-L axis are both critical to immune responses, elimination of either can result in the breakdown of tolerance and the development of autoimmunity. The PD-1/PD-L pathway can prevent autoreactive T cells and protect against autoimmunity. Treg cells induced by the PD-1 pathway can also help maintain immune homeostasis, maintaining the activation threshold for T cells to protect against autoimmunity [81, 82].

Although the PD1 pathway has received considerable attention for its roles in T cell exhaustion and tumor immunosuppression, PD1 cannot be considered a specific molecule for cell exhaustion [85]. In fact, T cells express PD1 during activation, thus being a marker for effector T cells. PD1 is expressed by subsets of tolerant T cells, regulatory T cells, follicular helper T cells, follicular regulatory T cells, and memory T cells. In addition, it is expressed on B cells, NK cells, and some myeloid cells. Expression of PD1 can be found in CD8 + T cells of healthy humans, and these cells do not resemble exhausted T cell populations [85, 86].

Polymorphisms have been described in the gene Pdcd1 that confer susceptibility to development of autoimmune diseases in humans. Many single nucleotide polymorphisms have been reported and approximately most of them are located in the intron regions of the structural gene [87].

Some of the most studied are the PD-1.1 located in the promoter region, PD-1.2 located in intron 2, PD-1.3 and PD-1.4 located at intron 4, PD-1.9 and PD-1.5 in exon 5 and PD-1.6 at position 32 of the untranslated region. Exist different haplotypes of these SNPs in families Caucasian and it is known that PD-1.1, PD-1.2 and PD-1.9 are in linkage imbalance, while the PD-1.4 and PD-1.5 positions they form a different block [88, 89].

PD1 polymorphisms are associated with susceptibility to a variety of autoimmune conditions including systemic lupus erythematosus, rheumatoid arthritis, and progression in multiple sclerosis [89], but it is not yet clear if these SNPs are causative or simply correlative. Furthermore, autoantibodies against PDL1 have been found in patients with rheumatoid arthritis and correlate with disease activity [89].

Ponce de León et al. [90] have focused the alteration of PD-1/PD-L1 pathway in celiac disease. Levels of sPD-1 was considerably higher in the serum of patients with celiac disease compared with health controls. A negative expression of PD1 in intestinal epithelial cells and lamina propria cells of active CD patients. PD-1 protein expression in CD4 + and CD8 + T cells decreases significantly in patients with CD. In this way, PD-1 function would be compromised in CD4 + and CD8 + T cells, indicating an inappropriate activation state [90]. In CD, a deregulation of immune suppression mechanisms appears, which can lead to abnormal and persistent activation of T cells and the production of cytokines. Without PD1, excessive immune-mediated tissue damage can lead to devastating consequences, because PD1 plays crucial roles in central and peripheral T cell tolerance, aiding in the protection of self-tissues from autoimmune responses. The co-delivery of soluble PD-1 could increase the maturation of DCs, which could accompanied by upregulation of DC maturation markers such as major histocompatibility complex II (MHC II) [90]. DC maturation is mediated by activated T lymphocytes, therefore sPD-1-regulated DC maturation may be influenced by increased T cell responses [90].

The soluble isoform is likely to have antagonistic effects on PD-1 by interfering with its signaling pathway, particularly considering that PD-1Δ3 still retains the ability to bind to PD-L1/PD-L2 receptors [90]. In patients with CD, excessive soluble PD-1 could serve as an “antibody” to block the PD-1/PD-Ls pathway and lead to aberrant T-cell proliferation. If, for example, CD8 + T-cell responses are not adequately controlled, severe immunopathology can result from the production of pro-inflammatory cytokines, such as IFN-γ and TNF-α [91].

Soluble PD-1 can promote T-cell responses through blocking the PD-1/PD-Ls pathway. IFN-γ is crucial for this process but also contributes to the upregulation of PD-L1 which indicates that sPD-1 plays a crucial role not only during the phase of T-cell exhaustion but also during primary T-cell activation and that sPD-1 can be used as an adjuvant to increase T-cell immunity [91, 92]. These findings suggest that at the time of clinical diagnosis of CD, T cells can exhibit features of immune exhaustion. It is not yet known what factor(s) contribute to the dysregulated PD-1 expression and may have increased susceptibility to the autoimmune complications of CD. The PD-1 and PD-L1 levels in the serum and intestinal biopsies of CD patients may be relevant to the determination of a possible correlation between markers of the autoimmune response, inflammation, and disease activity.

Advertisement

3. Future directions

The immune cell activation in the setting of immune checkpoint inhibitors results in unmasking of gluten sensitivity in genetically susceptible people, leading to expansion of previously self-reactive CD4+ T cells and subsequent CD8+ T cell-induced tissue destruction. Soluble immune checkpoint molecules constitute the emerging novel mediators in immune regulation. The relationship between celiac disease and the level of soluble immune checkpoints as sCTLA4, sHLA-G, sPD-1, and sPD-L1 has been shown.

3.1 Immune checkpoints cooperation

The immune checkpoint molecules may be implicated in biological mechanisms underlying celiac disease. Some immune checkpoint molecules serve as inhibitory signaling mediators to maintain immune tolerance, especially in the adaptive immune compartment. There are two forms of these molecules: the surface receptor or membrane-bound and cell-free soluble molecules. The membrane-bound CTLA4, HLA-G, PD-1, PD-L1 regulate T cell homeostasis, inhibit autoreactive T cells, and drive peripheral tolerance in cancer, pregnancy, and sepsis. 12,13 They also promote T regulatory cell development and inhibit the effector T cell differentiation and cytokine production leading to immunosuppression [93]. On the other hand, the soluble forms of these immune checkpoint molecules were discovered later, and their biological functions have gradually been elucidated. The immune regulatory effect of soluble PD-L1, sCTLA-4, and sHLA-G can trigger Treg differentiation and T cells apoptosis due to retention of their receptor [94].

Circulating soluble PD-1, CTLA4 and HLA-G could take part in modulating immune tolerance causing disturbances in the molecular mechanisms responsible for maintenance of balance between effector and regulatory components of the immune system in celiac disease. HLA-G and IDO acting independently, both molecules would be complementary in inducing efficient tolerance status. HLA-G1/HLA-G5 and IDO molecules act on alloreactive T-cell proliferation through two distinct inhibitory pathways. However, as IDO expression is tightly regulated and responsive to inflammatory mediators, HLA-G may indirectly modulate IDO by up-regulating the production of such mediators. For instance, by up-regulating the expression of IL-10, HLA-G may boost the IDO pathway.

The HLA-G/ILT2/ILT4 interactions actually target a broader array of immune effectors than the B7/CTLA4 and PD-1/PD-L1 pathways, since CTLA4 and PD-1 are expressed only on T cells, whereas ILT2 and ILT4 are differentially expressed on NK, T, and B cells as well as monocytes, dendritic cells (DCs), and neutrophils and thus may inhibit the early phases of an immune response (PD-1/PD-L1), or the later phases (B7/CTLA4) [93].

3.2 Gene dysregulation in celiac disease

The detection of the spliced/soluble variant of these immune checkpoints from CD patients suggests that the soluble form of HLA-G, CTLA-4 and PD1 molecules does not result from a cleavage of the full-length form. The potential genetic associations of several polymorphisms to susceptibility to autoimmune diseases have been described, Splicing machinery would act as a biosensor to adapt gene expression to pathophysiological conditions.

Gene dysregulation of these genes could lead to an imbalance in the splice variants present in the cells at any given time. The existence of specific factors in the serum of celiac patients, such as peptides derived from gliadin, would be able to modulate the expression of relevant components of the splice and the function of the splicing machinery. Dietary intervention of gluten peptides can clearly alter the expression pattern of the splicing machinery in humans at risk for CD.

The alternative splicing process may represent a physiological mechanism for maintaining cellular homeostasis, as suggested by different studies that demonstrate that the nutrients can modulate gene expression and, in particular, the splicing of pre-mRNAs that encode regulatory proteins. Minimal disturbances in the alternative splicing process can lead to the generation of deficient proteins that contribute to several human diseases. So, the splicing process may represent an adaptive mechanism in response to different nutritional conditions, and that this mechanism could be in place not only in circulating PBMCs but may also operate in cell types from other tissues and organs tightly coupled to nutrient-dependent metabolic homeostasis, e.g., intestine. So, specifically gluten can modulate processes required for cell homeostasis through the alteration of gene expression and, particularly, the splicing of pre-mRNAs encoding key regulatory proteins.

Advertisement

4. Conclusions

Further investigation on the determination of immunological interactions and biological functions by immune checkpoints in celiac disease is needed to deepen our understanding of the underlying disease mechanism in ourquest for diagnostic and therapeutic target in celiac disease.

Advertisement

Acknowledgments

This work was supported by Research group BIO220, Junta de Andalucía.

References

  1. 1. Lundin KE, Scott H, Hansen T, Paulsen G, Halstensen TS, Fausa O, Thorsby E, Sollid LM. Gliadin-specific, HLADQ (1*0501 1*0201) restricted T cells isolated from thesmall intestinal mucosa of celiac disease patients. J Exp Med 1993; 178:187-196. DOI: 10.1084/jem.178.1.187
  2. 2. Sollid LM, Jabri B. Triggers and drivers of autoimmunity: lessons from coeliac Disease. Nat Rev Immunol 2013; 13(4): 294-302. DOI: 10.1038/nri3407
  3. 3. Viljamaa M, Kaukinen K, Huhtala H Kyrönpalo S, Rasmussen M, Collin P. Coeliac disease, autoimmune diseases and gluten exposure. Scand J Gastroenterol 2005; 40: 437-443. DOI: 10.1080/00365520510012181
  4. 4. López Casado MA, Lorite P, Ponce de León C, Palomeque T, Torres MI. Celiac disease autoimmunity. Arch Immunol Ther Exp. 2018; 66(6): 423-430. DOI: 10.1007/s00005-018-0520-z
  5. 5. Torres MI, Palomeque T, Lorite P. Celiac disease and other autoimmune disorders. In: Chatzidionysiou K (ed) Autoimmunity pathogenesis, clinical aspects and therapy of specific autoinmune diseases. Intech Croatia. 2015; 131-151. DOI: DOI: 10.5772/60695
  6. 6. Bodd M, Kim CY, Lundin KE, Sollid LM. T-cell response to gluten inpatients with HLA-DQ2.2 reveals requirement of peptide-MHC stability in celiac disease. Gastroenterology. 2012; 142:552-561. DOI: 10.1053/j.gastro.2011.11.021
  7. 7. Hunt KA, Zhernakova A, Turner G Heap GAR, Franke L, Bruinenberg M, et al. Newly identified genetic risk variants for celiac disease related to the immune response. Nat Genet 2008; 40:395-402. DOI: 10.1038/ng.102
  8. 8. Trynka G, Hunt KA, Bockett NA, Romanos J, Mistry V, Szperl A et al. Dense genotying identifies and localizes multiple common and rare variant association signals in celiac disease. Nat Genet 2011; 43:1193-1201. DOI: 10.1038/ng.998
  9. 9. Alaedini A, Green PH. Narrative review: celiac disease: understanding a complex autoimmune disorder. Ann Intern Med 2005; 142: 289-298. DOI: 10.7326/0003-4819-142-4-200502150-00011
  10. 10. Fasano A. Clinical presentation of celiac disease in the pediatric population. Gastroenterology 2005; 128: S68–S73. DOI: 10.1053/j. gastro. 2005 .02.015
  11. 11. Kim CY, Quarsten H, Bergseng E, Khosla C, Sollid LM. Structural basis for HLA-DQ2-mediated presentation of gluten epitopes in celiac disease. Proc Natl Acad Sci USA 2004; 101: 4175-4179. DOI: 10.1073/pnas.0306885101
  12. 12. Torres MI, López Casado MA, Ríos A. New aspects in celiac disease. World J Gastroenterol 2007; 13: 1156-1161. DOI: 10.3748/wjg.v13.i8.1156
  13. 13. Di Sabatino A, Vanoli A, Giuffrida P, Luinetti O, Solcia E, Corazza GR. The function of tissue transglutaminase in celiac disease. Autoimmun Rev 11: 746-753. DOI: 10.1016/j.autrev. 2012.01.007
  14. 14. Sollid LM, Markussen G, Ek J, Gjerde H, Vartdal F, Thorsby E. Evidence for a primary association of celiac disease to a particular HLA-DQ alpha/beta heterodimer. J Exp Med 1989; 169: 345-350. DOI: 10.1084/jem.169.1.345
  15. 15. Anderson RP, Degano P, Godkin AJ. In vivo antigen challenge in celiac disease identifies a single transglutaminase-modified peptide as the dominant A-gliadin T-cell epitope. Nat Med 2000; 6: 337-342. DOI: 10.1038/73200
  16. 16. Jabri B, Kasarda DD, Green PHR. Innate and adaptive immunity: the yin and yang of celiac disease. Immunol Rev 2005; 206: 219-231. DOI: 10.1111/j.0105-2896.2005.00294.x
  17. 17. Sollid LM. Coeliac disease: dissecting a complex inflammatory disorder. Nat Rev Immunol. 2002; 2(9), 647-655. DOI: 10.1038/nri885
  18. 18. Alaedini A, Green PHR. Narrative review: celiac disease: understanding a complex autoimmune disorder. Ann Intern Med. 2005; 142(4), 289-298. DOI: 10.7326 /0003-4819-142-4-200502150-00011
  19. 19. Fasano A, Catassi C. Clinical practice. Celiac disease. N Engl J Med 2012; 367: 2419-26. DOI: 10.1056/NEJMcp1113994
  20. 20. Orabona C, Mondanelli G, Puccetti P, Grohmann G. Immune Checkpoint Molecules, Personalized Immunotherapy, and Autoimmune Diabetes. Trends Mol Med. 2018, 24 (11): 931-41. Doi: 10.1016/j.molmed.2018.08.005
  21. 21. Calabrese L, Velcheti V. Checkpoint immunotherapy: good for cancer therapy, bad for rheumatic diseases. Ann Rheum Dis 2017; 76: 1-3. DOI: 10.1136/ annrheumdis-2016-209782
  22. 22. Abbas AK, Lohr J, Knoechel B, Nagabhushanam V. T cell tolerance and autoimmunity. Autoimmunity Reviews 2004; 3: 471-475. DOI:10.1016/j.autrev. 2004.07.004
  23. 23. Orabona C, Mondanelli G, Puccetti P, Grohmann G. Immune Checkpoint Molecules, Personalized Immunotherapy, and Autoimmune Diabetes. Trends Mol Med. 2018, 24 (11): 931-41. DOi: 10.1016/j.molmed.2018.08.005
  24. 24. Torres MI, Palomeque T, Lorite P. Celiac Disease and Other Autoimmune Disorders. In: Chatzidionysiou K (ed) Autoimmunity-Pathogenesis, clinical aspects and therapy of specific autoimmune diseases. Intech, Croatia. 2015; 131-151. DOI: 10.5772/60695
  25. 25. Pardoll DM. The blockade of immune checkpoints in cancer immunotherapy, Nat Rev Canc. 2012; 12: 252-264. DOI: 10.1038/nrc3239
  26. 26. Kumar P, Saini S, Khan S, Lele SS, Prabhakar BS. Restoring self-tolerance in autoimmune diseases by enhancing regulatory T-cells. Cell Immunol 2019; 339: 41-49. doi: 10.1016/ j.cellimm.2018.09.008
  27. 27. Grohmann, U. FallarinoF, Puccetti P. Tolerance, DCs and tryptophan: much ado about IDO. Trends Immunol. 2003; 24: 242-248
  28. 28. Puccetti P, Grohmann U. IDO and regulatory T cells: a role for reverse signalling and non-canonical NF-kappaB acti-vation. Nat Rev Immunol. 2007; 7: 817-823. DOI: 10.1038/nri2163
  29. 29. Dai X, Zhu BT. Indoleamine 2, 3-dioxygenase tissue distribution and cellular localization in mice: implications for its biological functions. J Histochem Cytochem. 2010; 58(1): 17-28. doi: 10.1369/jhc.2009.953604
  30. 30. Sedlmayr P, Blaschitz A, Wintersteiger R, Semlitsch M, Hammer A, MacKenzie et al. Localization of indoleamine 2, 3-dioxygenase in human female reproductive organs and the placenta. Mol Hum Reprod. 2002; 8(4): 385-391. DOI: 10.1093 /molehr/8.4.385
  31. 31. Mellor AL, Munn DH. Tryptophan catabolism and regulation of adaptive immunity. J Immunol. 2003; 170: 5809-5813
  32. 32. Mellor AL, Munn DH. Tryptophan catabolism and T-cell tolerance: immunosuppression by starvation? Immunol Today 1999; 10: 469-473. DOI: 10.1016/s0167-5699(99)01520-0
  33. 33. Mellor AL, Munn DH. IDO expression by dendritic cells: tolerance and tryptophan catabolism. Nat Rev Immunol. 2004; 4: 762-774. DOI: 10.1038 /nri1457
  34. 34. Scheler M, Wenzel J, Tüting T, Takikawa O, Bieber T, Von Bubnoff D. Indoleamine 2,3-Dioxygenase (IDO) : The Antagonist of Type I Interferon-Driven Skin Inflammation?. Am J Pathol. 2007; 171(6): 1936-1943. DOI: 10.2353 /ajpath .2007.070281
  35. 35. Mailankot M, Staniszewska MM, Butler H, Caprara MH, Howell S, Wang B, et al. Indoleamine 2, 3-dioxygenase overexpression causes kynurenine-modification of proteins, fiber cell apoptosis and cataract formation in the mouse lens. Lab Invest. 2009; 89(5):498-512. DOI: 10.1038/ labinvest.2009.22
  36. 36. Mailankot M, Nagaraj RH. Induction of indoleamine 2, 3-dioxygenase by interferon-gamma in human lens epithelial cells: apoptosis through the formation of 3-hydroxykynurenine. Int J Biochem Cell Biol. 2010; 42(9):1446-54. DOI: 10.1016/j.biocel.2010.04.014
  37. 37. Cuffy MC, Silverio AM, Qin L, Wang Y, Eid R, Brandacher G, et al. Induction of indoleamine 2, 3-dioxygenase in vascular smooth muscle cells by interferon-gamma contributes to medial immunoprivilege. J Immunol. 2007; 179(8): 5246-5254. DOI: 10.4049/jimmunol.179.8.5246
  38. 38. Grohmann U, Orabona C, Fallarino F, Vacca C, Calcinaro F, Falorni A, et al. CTLA-4-Ig regulates tryptophan catabolism in vivo. Nat Immunol. 2002; 3(11): 1097-1101. DOI: 10.1038/ni846
  39. 39. Jung ID, Lee CM, Jeong YI, Lee JS, Park WS, Han J, Park YM. Differential regulation of indoleamine 2, 3-dioxygenase by lipopolysaccharide and interferon gamma in murine bone marrow derived dendritic cells. FEBS Lett. 2007; 581(7): 1449-1456. DOI: 10.1016/ j.febslet. 2007.02.073
  40. 40. Puccetti P, Grohmann U. IDO and regulatory T cells: a role for reverse signalling and non-canonical NF-kappaB activation, Nat Rev Immunol 2007; 7: 817-823 DOI: 10.1038/nri2163
  41. 41. Mondanelli G, Iacono A, Carvalho A, Orabona C, Volpi C, Pallotta MT, et al. Amino acid metabolism as drug target in autoimmune diseases, Autoimmun Rev 2019; 18: 334-348, DOI: 10. 1016/j.autrev.2019.02.004
  42. 42. Boros FA, Bohár Z, Vécsei L. Genetic alterations affecting the genes encoding the enzymes of the kynurenine pathway and their association with human diseases, Mutat Res. 2018; 776: 32-45. DOI: 10.1016/j.mrrev.2018.03.001
  43. 43. Torres MI, López-Casado MA, Lorite P, Ríos A. Tryptophan metabolism and indoleamine 2,3-dioxygenase expression in coeliac disease. Clin Exp Immunol. 2007; 148(3): 419-424. DOI: 10.1111/j.1365-2249.2007.03365.x
  44. 44. Wolf AM, Wolf D, Rumpold H, Moschen AR, Kaser A, Obrist P et al. Overexpression of indoleamine 2,3-dioxygenase in human inflammatory bowel disease. Clin Immunol. 2004; 113: 47-55. DOI: 10.1016/j.clim.2004.05.004
  45. 45. Klein J, Sato A. The HLA system. First of two parts. N Engl J Med 2000; 343:702-9. DOI: 10.1056/NEJM200009073431006
  46. 46. Lin A, Yan W-H. Heterogeneity of HLA-G Expression in Cancers: Facing the Challenges. Front Immunol. 2018; 9: 2164. DOI: 10.3389/fimmu.2018.02164
  47. 47. Paul P, Cabestre F, Ibrahim E, Lefebvre S, Khalil-Daher I, Vazeux G, et al. Identification of HLA-G7 as a New Splice Variant of the HLA-G mRNA and Expression of Soluble HLA-G5, −G6, and -G7 Transcripts in Human Transfected Cells. Hum Immunol 2000; 61: 1138-1149 (2000). DOI: 10.1016/S0198-8859(00)00197-X
  48. 48. Rouas-Freiss N, Moreau P, Ferrone S, Carosella ED. HLA-G Proteins in Cancer: Do They Provide Tumor Cells with an Escape Mechanism? Cancer Res. 2005; 65(22): 10139-10144. DOI: 10.1158/ 0008-5472.CAN-05-0097
  49. 49. Lila N, Carpentier A, Amrein C, Khalil-Daher I, Dausset J, Carosella ED. Implication of HLA-G Molecule in Heart-Graft Acceptance. Lancet (London England) 2000; 355(9221): 2138. DOI: 10.1016/S0140-6736(00)02386-2
  50. 50. Lozano JM, González R, Kindelán JM, Rouas-Freiss N, Caballos R, Dausset J, et al. Monocytes and T Lymphocytes in HIV-1-Positive Patients Express HLA-G Molecule. AIDS (London England) 2002; 16(3): 347-351. DOI: 10.1097/ 00002030-200202150-00005
  51. 51. Torres MI, Le Discorde M, Lorite P, Ríos A, Gassull MA, Gil A, et al. Expression of HLA-G in inflammatory bowel disease provides a potential way to distinguish between ulcerative colitis and Crohn’s disease. Int Immunol. 2004; 16(4):579-83. DOI: 10.1093/intimm/dxh061
  52. 52. Shiroishi M, Tsumoto K, Amano K, Shirakihara Y, Colonna M, Braud V, et al. Human inhibitory receptors Ig-like transcript 2 (ILT2) and ILT4 compete with CD8 for MHC class I binding and bind preferentially to HLA-G. Proc Natl Acad Sci U S A. 2003; 100(15): 8856-8861. DOI: 10.1073/pnas.1431057100
  53. 53. Anderson KJ, Allen R. Regulation of T-cell immunity by leucocyte immunoglobulin-like receptors: innate immune receptors for self on antigen-presenting cells. Immunology. 2009; 127(1): 8-17. DOI: 10.1111/j.1365-2567. 2009. 03097.x
  54. 54. Veit TD, Cordero EA, Mucenic T, Monticielo OA, Brenol JC, Xavier RM, et al. Association of the HLA-G 14bp polymorphism with systemic lupus erythematosus. Lupus 2009; 18: 424-30. DOI: 10.1177/0961203308098187
  55. 55. Cordero EA, Veit TD, da Silva MA, Jacques SM, Silla LM, Chies JA. HLA-G polymorphism influences the susceptibility to HCV infection in sickle cell disease patients. Tissue Antigens 2009; 74: 308-13. DOI: 10.1111/j.1399-0039.2009 .01331.x
  56. 56. Glas J, Torok HP, Tonenchi L, Wetzke M, Beynon V, Teshome MY, et al. The 14bp deletion polymorphism in the HLA-G gene display significant difference between ulcerative colitis and Crohn’s disease and is associated with ileocecal resection in Crohn’disease. Int Immunol 2007; 16: 621-626. DOI: 10.1093/intimm /dxm027
  57. 57. Rouas-Freiss N, Goncalves RM, Menier C, Dausset J, Carosella ED. Direct evidence to support the role of HLA-G in protecting the fetus from maternal uterine natural killer cytolysis. Proc Natl Acad Sci U S A 1997; 94: 11520-11525. DOI: 10.1073/pnas.94.21.11520
  58. 58. LeMaoult J, Caumartin J, Daouya M, Favier B, Le Rond S, Gonzalez A, et al. Immune regulation by pretenders: cell-to-cell transfers of HLA-G make effector T cells act as regulatory cells. Blood 2007; 109: 2040-2048. DOI: 10.1182/blood-2006-05-024547
  59. 59. Horuzsko A, Lenfant F, Munn DH, Mellor AL. Maturation of antigen presenting cells is compromised in HLA-G transgenic mice. Int Immunol 2001; 13: 385-394. DOI: 10.1093/intimm/13.3.385
  60. 60. Carosella ED, HoWangYin KY, Favier B, LeMaoult J. HLA-G dependent suppressor cells: diverse by nature, function, and significance. Hum Immunol 2008; 69: 700-707. DOI: 10.1016/j.humimm.2008.08.280
  61. 61. Torres MI, López-Casado MA, Luque J, Peña J, Ríos A. New advances in coeliac disease: serum and intestinal expression of HLA-G.International Immunology, 2006; 18 (5): 713-718. DOI: 10.1093/intimm/dxl008
  62. 62. Torres MI, Lopez Casado MA, Rios A. New aspects in celiac disease. World J Gastroenterol 2007; 13: 1156-1161. DOI: 10.3748/wjg.v13.i8.1156
  63. 63. Fabris A, Segat L, Catamo E, Morgutti M, Vendramin A, Crovella S. HLA-G 14bp deletion/insertion polymorphism in celiac disease. Am J Gastroenterol 2011; 106: 139-144. DOI: 10.1038/ajg.2010.340
  64. 64. Catamo E, Zupin, Segat L, Celsi F, Crovella S. HLA-G and susceptibility to develop celiac disease. Human Immunology 2015; 76: 36-41. DOI: 10.1016/ j.humimm. 2014. 12.006
  65. 65. Larsen MH, Hviid TV. Human leukocyte antigen-G polymorphism in relation to expression, function, and disease. Hum Immunol 2009; 70: 1026-1034. DOI: 10.1016/j.humimm.2009.07.015
  66. 66. LeMaoult J, Krawice-Radanne I, Dausset J, Carosella ED. HLA-G1-expressing antigen-presenting cells induce immunosuppressive CD4+ T cells. Proc Natl Acad Sci USA 2004; 101: 7064-7069. DOI: 10.1073/pnas.0401922101
  67. 67. Fournel S, Aguerre-Girr M, Huc X, Lenfant F, Alam A, Toubert A, et al. Cutting edge: soluble HLA-G1 triggers CD95/CD95 ligand-mediated apoptosis in activated CD8+ cells by interacting with CD8. J Immunol 2000; 164: 6100-6104. DOI: 10.4049/jimmunol.164.12.6100
  68. 68. Donadi EA, Castelli EC, Arnaiz-Villena A, Roger M, Rey D, Moreau P. Implications of the polymorphism of HLA-G on its functions, regulation, evolution and disease association. Cell Mol Life Sci 2011; 68: 369-395. DOI: 10.1007/s00018-010-0580-7
  69. 69. López AS, Alegre E, LeMaoult J, Carosella E, Gonzalez A. Regulatory role of tryptophan degradation pathway in HLA-G expression by human monocyte-derived dendritic cells. Mol Immunol. 2006; 43: 2151-2160. DOI: 10.1016/ j.molimm.2006.01.007
  70. 70. Honing A, Rieger L, Kapp M, Sutterlin M, Dietl J, Kammerer U. Indoleamine 2, 3-dioxygenase (IDO) expression in invasive extravillous trophoblast supports role of the enzyme for maternal-tolerance. J Reprod Immunol. 2004; 61: 79-86. DOI: 10.1016/j.jri.2003.11.002
  71. 71. Uyttenhove C, Pilotte L, Theate I, Stroobant V, Colau D, Parmentier N, et al. Evidence for a tumoral immune resistance mechanism based on tryptophan degradation by indoleamine 2,3-dioxygenase. Nat Med. 2003; 9: 1269-1274. DOI: 10.1038/nm934
  72. 72. Takikawa O, Kuroiwa T, Yamazaki F, Kido R. Mechanism of interferon-gamma action. Characterization of indoleamine 2, 3-dioxygenase in cultured human cell induced by interferon-gamma and evaluation of the enzyme-mediated tryptophan degradation in its anticellular activity. J Biol Chem. 1998; 263: 2041-2048
  73. 73. Yang Y, Chu W, Geraghty D, Hunt J. Expression of HLA-G in human mononuclear phagocytes and selective induction of IFN-gamma. J Immunol. 1996; 156: 4224-4231
  74. 74. Moreau P, Adrian-Cabestre F, Menier C, Guiard V, Gourand L, Dausset J, et al. IL-10 selectively induces HLA-G expression in human trophoblast and monocytes. Int Immunol. 1999; 11: 803-811. DOI: 10.1093/ intimm/11.5.803
  75. 75. González-Hernandez A, LeMaoult J, Lopez A, Alegre E, Caumartin, Le Rond S, Daouya. Linking Two Immuno-Suppressive Molecules: Indoleamine 2,3 Dioxygenase Can Modify HLA-G Cell-Surface Expression. Biol Reprod 2005; 73 (3):571-578, DOI: 10.1095/biolreprod.105.040089
  76. 76. Fife BT, Bluestone JA. Control of peripheral T-cell tolerance and autoimmunity via the CTLA-4 and PD-1 pathways. Immunol Rev 2008; 224: 166-82. DOI: 10.1111/j.1600-065X.2008 .00662.x
  77. 77. Simone R, Brizzolara R, Chiappori A, Milintenda-Floriani F, Natale C, Greco L, et al. A functional soluble form of CTLA-4 is present in the serum of celiac patients and correlates with mucosal injury. International Immunology, 2009; 21 (9): 1037-1045. DOI: 10.1093/intimm /dxp069
  78. 78. Oaks MK, Hallett KM, Penwell R, Stauber EC, Warren SJ, Tector AJ. A native soluble form of CTLA-4. Cell Immunol. 2000; 201:144-153. DOI: 10.1006 /cimm. 2000.1649
  79. 79. Brophy, K., Ryan, A. W., Thornton, J. M. Abuzakouk M, Fitzgerald AP, McLoughlin RM et al. Haplotypes in the CTLA4 region are associated with coeliac disease in the Irish population. Gen Immun. 2006; 7:19. DOI: 10.1038/ sj.gene .6364265
  80. 80. Francisco LM, Sage PT, Sharpe AH. The PD-1 Pathway in Tolerance and Autoimmunity. Immunol Rev. 2010; 236: 219-242. DOI: 10.1111/j.1600065 X.2010. 00923.x
  81. 81. Keir ME, Butte MJ, Freeman GJ, Sharpe AH. PD-1 and its ligands in tolerance and immunity. Annu Rev Immunol. 2008; 26: 677-704. DOI: 10.1146/annurev. immunol.26.021607.090331
  82. 82. Agata Y, Kawasaki A, Nishimura H, Ishida Y, Tsubata T, Yagita H, Honjo T. Expression of the PD-1 antigen on the surface of stimulated mouse T and B lymphocytes. Int Immunol. 1996; 8: 765-772. DOI: 10.1093/intimm/8.5.765
  83. 83. Kinter AL, Godbout EJ, McNally JP, Sereti I, Roby GA, O’Shea MA, Fauci AS. The common gamma-chain cytokines IL-2, IL-7, IL-15, and IL-21 induce the expression of programmed death-1 and its ligands. J Immunol. 2008; 181: 6738-6746. DOI: 10.4049/jimmunol.181.10.6738
  84. 84. Gianchecchi E, Fierabracci A. Inhibitory Receptors and Pathways of Lymphocytes: The Role of PD-1 in Treg Development and Their Involvement in Autoimmunity Onset and Cancer Progression. Front Immunol. 2018; 9: 2374. DOI: 10.3389/fimmu.2018.02374
  85. 85. Sharpe AH, Pauken KE. The diverse functions of the PD1 inhibitory pathway. Nat Rev Immunol. 2018; 18(3): 153-167. DOI: 10.1038/nri.2017.108
  86. 86. Dezutter-Dambuyant C, Durand I, Alberti L, Bendriss-Vermare N, Valladeau-Guilemond J, Duc A. A novel regulation of PD-1 ligands on mesenchymal stromal cells through MMP-mediated proteolytic cleavage. Oncoimmunology. 2016; 5(3): e1091146. doi: 10.1080/ 2162402X. 2015.1091146
  87. 87. Prokunina L, Castillejo-Lopez C, Oberg F, Gunnarsson I, Berg L, Magnusson V, et al. A regulatory polymorphism in PDCD1 is associated with susceptibility to systemic lupus erythematosus in humans. Nat Genet. 2002; 32: 666-669. DOI: 10.1038/ng1020
  88. 88. Liu C, Jiang J, Gao L, Hu X, Wang F, Shen Y, et al. A promoter region polymorphism in PDCD-1 gene is associated with risk of rheumatoid arthritis in the Han Chinese population of southeastern China. Int J Genomics. 2014; 2014: 247637. DOI: 10.1155/ 2014/247637
  89. 89. Bertsias GK, Nakou M, Choulaki C, Raptopoulou A, Papadimitraki E, Goulielmos G et al. Genetic, immunologic, and immunohistochemical analysis of the programmed death 1/programmed death ligand 1 pathway in human systemic lupus erythematosus. Arthritis Rheum. 2009, 60(1): 207-218. DOI: 10.1002 /art.24227
  90. 90. Ponce de León C, López-Casado MA, Lorite P, Palomeque T, Torres MI. Dysregulation of the PD-1/PD-L1 pathway contributes to the pathogenesis of celiac disease. Cell Mol Immunol 2019; 16(9): 777-779. DOI: 10.1038/s41423-019-0256-7
  91. 91. Abiko K, Matsumura N, Hamanishi J, Horikawa N, Murakami R, Yamaguchi K et al. IFN-γ from lymphocytes induces PD-L1 expression and promotes progression of ovarian cancer. Br J Cancer. 2015; 112(9): 1501-1509. DOI: 10.1038/bjc.2015.101
  92. 92. Wu H, Miao M, Zhang G, Hu Y, Ming Z, Zhang X. Soluble PD-1 is associated with aberrant regulation of T cells activation in aplastic anemia. Immunol Invest. 2009;38:408-421. DOI: 10.1080/08820130902912332
  93. 93. Carosella ED, Rouas-Freiss N, Tronik-Le Roux D, Moreau P, LeMaoult J. HLA-G: an immune checkpoint molecule. Adv Immunol. 2015; 1 27: 33-144. DOI: 10.1016/bs.ai.2015.04.001
  94. 94. Zhang YH, Sun HX. Immune checkpoint molecules in pregnancy: Focus on regulatory T cells. Eur J Immunol. 2020; 50(2):160-169. DOI: 10.1002 /eji. 201948382

Written By

Isabel Torres, Miguel Ángel López Casado, Teresa Palomeque and Pedro Lorite

Submitted: 30 September 2020 Reviewed: 14 January 2021 Published: 16 March 2021