Skip to main content
  • Review
  • Published:

T cells in rheumatoid arthritis

Abstract

Over the past decade and a half, advances in our understanding of the pathogenesis of immune-mediated diseases such as rheumatoid arthritis (RA) have translated directly into benefit for patients. Much of this benefit has arisen through the introduction of targeted biological therapies. At the same time, technological advances have made it possible to define, at the cellular and molecular levels, the key pathways that influence the initiation and persistence of chronic inflammatory autoimmune reactions. As our understanding grows, it is likely that this knowledge will be translated into a second generation of biological therapies that are tailor-made for the patient. This review summarizes current perspectives on RA disease pathogenesis, with particular emphasis on what RA T cells look like, what they are likely to see, and how they contribute to persistence of the chronic inflammatory response.

Introduction

What is rheumatoid arthritis (RA)? Traditional teaching would have us believe that RA is one of a cluster of autoimmune diseases, to be included in the same category as type I diabetes, autoimmune thyroiditis, and celiac disease. There are some important distinctions, however. The term 'RA', when used in the clinical context, probably represents a spectrum of distinct clinical entities (for review [1]). This spectrum manifests not only at the clinical level – presenting as a range of specific clinical phenotypes – but also at the cellular, molecular, and genetic levels. This is best illustrated by the variations observed in patterns of inflammatory infiltrates in histological sections of RA synovial tissue, the patterns of disease expression defined by serological profiles, and the emerging subsets of patients stratified according to genotype [1]. Such an approach to disease stratification means that we are now in a position to redefine subtypes of disease based on objective laboratory parameters, independent of clinical manifestations. By defining disease according to perturbations in cellular and molecular phenotypes and genotypes, we should be better placed to predict disease outcome and to select the most appropriate therapy for our patients.

Diseases such as RA, but also inflammatory bowel disease and psoriasis, are better characterized as a group of prototypic chronic inflammatory diseases. The most convincing evidence for such a notion comes directly from the clinic, in particular the striking clinical responses to anticytokine biological therapy [2]. Finally, RA is somewhat distinct from other organ-specific autoimmune diseases in the sense that, rather than destroying the target tissue, the inflammatory process stimulates and induces proliferation of synovial stromal tissues, especially in the early phase of disease [3]. Thus, RA is perhaps better described as the prototype chronic inflammatory disease, with features of autoimmunity.

These distinctions should be taken into account when considering pathways of disease pathogenesis, especially immune-mediated pathways. This review highlights current concepts of RA pathogenesis, with a specific focus on the immunobiology of the RA T cell.

What should an RA T cell look like?

If the immunobiology of RA plays by the rules governed by traditional paradigms of autoimmunity, then we would predict that RA synovial T cells infiltrating affected synovial joints would express a cell surface phenotype that is compatible with prior antigen experience, is indicative of extensive proliferative activity, is suggestive of clonal expansions of subsets of antigen-specific T cells, is consistent with enhanced migratory competence, and favors survival in situ. Current paradigms of adaptive immunity would also predict differentiation of T-helper cell subsets along a distinct effector T-cell lineage, whereas histological analysis of synovial tissue explants might reveal the presence of T cells as components of diffuse inflammatory infiltrates, with a prevalence of CD4+ over CD8+ T cells, in close association with antigen-presenting cells (APCs). If RA were to fulfil the criteria for a bona fide organ-specific autoimmune disease, then elaboration of T cells derived from synovial tissue explants should identify T cells that are activated and maintained in vivo by tissue-specific autoantigens derived from synovium, matrix, bone, or a combination of the three. Finally, the persistence of the chronic inflammatory process should predict that there might exist a relative paucity of regulatory T-cell subsets in terms of either cell number or function. How do these predictions match up with what we have learnt from laboratory studies?

What do RA T cells really look like?

The most convincing evidence for the involvement of T cells in chronic immune activation comes from the analysis of RA synovial biopsies [4]. In keeping with the spectrum of clinical entities that we classify as RA, distinct patterns of synovitis have been identified, suggesting topographical organization of lymphoid microarchitecture at a cellular level [4–6]. We now know that this process is directed by expression of cytokines and chemokines, such as lymphotoxin-α1β2 and B-lymphocyte chemokine/CXC chemokine ligand (CXCL)13 [7]. Patterns of synovitis include diffuse inflammatory infiltrates of T cells, B cells, macrophages, and dendritic cells (about 50%); clusters of lymphoid follicular aggregates comprising T cells, B cells, and dendritic cells (about 20%); structures with germinal center (GC)-like reactions (about 25%); and, rarely, granulomatous lesions (<5%; which have features similar to those of rheumatoid nodules). Of most interest are those tissues comprising secondary follicles and GCs, because these are characterized by the presence of follicular dendritic cells, which are critical components in follicle formation in the synovium and for perpetuation of adaptive immunity and autoimmunity [6, 7]. Thus, the synovial microenvironment and spatial relationships between cell subsets appear optimal for supporting antigen requisition, storage, processing, and presentation by APCs to T cells.

Data suggest that patterns of inflammatory infiltrate are relatively stable over time and common to tissues from different joints from the same patient, including unaffected joints, suggesting that they may reflect true disease variants [6, 7]. Furthermore, recent support for this notion comes from a comparison of RA knee joint synovial tissue from anti-cyclic citrullinated peptide (CCP) positive and negative patients [8]. That study demonstrated higher numbers of infiltrating lymphocytes (CD3+, CD8+, CD45RO+, and CXCL12+) associated with ectopic GC reactions, less extensive fibrosis, and a thinner synovial lining layer in the anti-CCP+subset [8]. It is noteworthy that the direct contribution of CD8+ T cells to GC+ follicular dendritic cell networks has previously been demonstrated [9, 10]. The anti-CCP+ associated synovial phenotype is also associated with more radiographic damage [8]. Emerging data have uncovered a gene expression profile for GC+ synovial tissue (CD21L+), which complements immunohistochemical analysis, characterized by elevated expression of CXCL13, CXCL12, CC chemokine ligand (CCL)19, CCL21, CXC chemokine receptor (CXCR)4, CXCR5 and CC chemokine receptor (CCR)7 [11]. In this study, Gene Ontology pathway analysis suggested that many pathways active in T cells, including JAK/STAT (Janus kinase/signal transducer and activator of transcription), T-cell receptor (TCR), IL-2R, IL-7R, and co-stimulatory signalling pathways are important features that may promote the cellular interactions that are required to support the organization of these tertiary lymphoid structures.

What is the phenotype of RA T cells at the cellular level? There are a number of distinguishing features. First, synovial T cells express an extensive array of cell surface antigens that reflect prior antigen experience (Figure 1). Principle among these are markers of effector memory cells, including the expression of specific chemokine receptors and integrins. The profile of chemokine receptors suggests that there is an element of selective trafficking of T-cell subsets in response to local production of chemoattractants by synovial stroma.

Figure 1
figure 1

The phenotype of RA synovial T cells. The schematic depicts examples of cell surface antigens and receptors that confer distinct phenotypes according to antigen experience, differentiation, and acquisition of memory, migratory competence, terminal differentiation and immune senescence, and effector function. CCR, CC chemokine receptor; CX3CR, CX3C chemokine receptor; CXCR, CXC chemokine receptor; LFA, lymphocyte function-associated antigen; LT, lymphotoxin; mTNF, membrane-associated tumor necrosis factor; NK, natural killer; RA, rheumatoid arthritis; RANKL, receptor activator for nuclear factor-κB ligand; TCR, T-cell receptor; TNFR, tumor necrosis factor receptor; VLA, very late antigen.

Aging is an important risk factor for RA. How does aging manifest itself at the cellular level? The immune system of patients with RA, including those developing the disease at a relatively early age, exhibit features of premature aging, as determined by accelerated erosion of telomeres, which are repetitive DNA sequences that cap the ends of chromosomes that are lost with cell replication [12]; loss of telomere length provides an estimate of the number of completed cell cycles. In RA this process of telomere erosion is found in both the naïve and memory T-cell compartments, which cannot therefore be accounted for exclusively by a specific antigen-driven response. The data suggest an ongoing process of replicative stress that occurs very early or that may even predate disease onset. It could arise in part through reduced generation of new thymic emigrants and/or homeostatic expansion of the peripheral naïve T-cell repertoire to environmental stimuli (cytokines or cognate antigen), based on enumeration of TCR excision circles [12, 13]. Nonetheless, the aging phenotype is not confined to the T-cell compartment, because telomere shortening has been documented in hematopoietic stem cells, which may explain the features of senescence in both granulocyte and B-cell compartments. Of particular interest was the finding of an association between human leukocyte antigen (HLA)-DR4 status and premature telomere erosion in healthy donors as well as in patients with RA, in spite of high telomerase repair activity [14]. These data provide interesting, if still poorly understood, insights into the genetic determinants of immune senescence in RA. They also define a molecular basis for impaired T-cell responsiveness characteristic of RA, given the close relationship between telomere erosion and growth arrest.

As mentioned above, the synovial microenvironment provides an ideal setting in which to support adaptive immune responses. These cellular interactions, combined with an abundance of inflammatory cytokines and reactive oxygen intermediates in a hypoxic setting, are probably responsible for imposing distinct characteristics on synovial T cells. Specific characteristics of such cells include an intrinsic resistance to apoptosis in vivo [15]; loss of expression of CD25, CD28, CD27, CD40L, and the invariant TCR subunit TCR-ζ [16–18]; enhanced expression of inflammatory cytokines such as IFN-γ and tumor necrosis factor (TNF)-α [19]; and expression of CCR4, CCR5, CXCR3, and CXCR5 [20]. Expansion of CD4+CD28null T cells is of interest because it has been found to correlate with systemic co-morbidities associated with severe RA, such as vasculitis and acute coronary syndrome [21, 22]. Although this terminally differentiated CD28null subset may become relatively independent of co-stimulation through CD80/CD86 engagement, co-stimulation through other pathways such as OX40 may compensate [23]. Indeed, the acquisition of a distinct gene expression program resembling that of natural killer (NK) cells, including expression of the activating NK cell receptors NKG2D and KIR2DS2, and acquisition of DAP10 and DAP12 adaptor proteins for signalling, may explain how these cells are such potent producers of IFN-γ [24, 25]. Synovial T cells can receive signals in situ, because the counter-ligands for NKG2D, major histocompatibility complex (MHC) class I chain related (MIC)A and MICB, are expressed on synovial fibroblasts [26].

In summary, an RA synovial T cell is ideally equipped to survive in situ through interactions with APCs, macrophages, B cells, and stromal elements. However, RA T cells are likely to depend for their effector function on a more promiscuous range of stimuli, besides cognate antigen presented by MHC class II molecules. Indeed, aberrant or excessive pathways of costimulation may permit the T cell to respond to low-affinity self peptides in the periphery, in spite of both inherited and acquired defects in TCR signaling.

What do RA T cells see?

Classical paradigms of autoimmunity state that tissue specificity is determined largely by recognition of tissue-specific self antigens by autoreactive T and B cells. In the past there was much enthusiasm for the idea that RA was driven by joint-specific antigens, for instance type II collagen, the cartilage protein HCgp-39, and proteoglycans such as aggrecan, especially because these antigens were inducers of autoimmune arthritis in appropriate strains of rodent [27–29]. At the same time, antigens such as heat shock proteins and BiP (immunoglobulin binding protein) have also been suggested to be autoantigenic targets [30, 31]. These are ubiquitous proteins expressed at high levels, which are certainly not tissue specific. It should follow from this paradigm that in RA it should be possible to detect expansions of specific clones that express distinct TCRB gene recombinations that are indicative of antigen drive.

The aberrations in the TCR repertoire are perhaps best understood within the context of aging. Thus, as thymic function declines with age, homeostatic proliferation of naïve peripheral T cells is required to 'fill the space' - that is, to undertake the task of T-cell generation. As a consequence of aging, which is accelerated in RA [12], large clonal expansions of T cells are observed [32]. These expansions are global, affecting both CD4+ and CD8+ subsets, as well as naïve and memory T-cell populations. Because the process of homeostatic proliferation is dependent on the recognition by TCRs of complexes of MHC and self peptide, it will inevitably lead to the selection of higher affinity autoreactive TCRs with the capacity to compensate for the functional, proliferative defects of senescent T cells [33]. Accordingly, there is profound repertoire contraction associated with a lack of diversity. Indeed, by the age of 40 to 50 years, patients with RA have already lost 90% of their available TCR repertoires, whereas the naïve repertoire has expanded 10-fold to fill the space [16, 34–36].

Whereas the expression of distinct HLA-DR molecules can play an important role in repertoire development through loss of diversity and contraction of T-cell repertoires in both CD4+ and CD8+ T-cell compartments, as discussed above, there are also data pointing to antigen-independent pathways of activation. Thus, both naïve and memory T cells may proliferate and differentiate in response to the inflammatory milieu into effector T-cell populations with specific cell surface phenotypes and functional properties, including cytokine expression, cell contact dependent activation of monocytes, and B-cell help [37]. This pathway can be reproduced in vitro using IL-2, IL-6, and TNF-α, and would account for the dilution of TCR excision circles in the CD45RA+ naïve compartment in patients with RA [13].

What MHC/peptide complexes might these T cells see? In RA we know that T cells see the 'shared epitope', a stretch of DRβ chain α-helix sequence exposed to the TCR encoded by RA disease associated HLA-DRB1 alleles [38]. We also know from recent genetic and epidemiological studies that carrying shared epitope positive alleles is a risk factor not so much for RA but for immunity to citrullinated proteins [39], arising through post-translational conversion of arginine residues to citrulline, which is catalyzed by a family of enzymes known as peptidyl-arginyl deiminases (PADs) [40]. This manifests as the presence of autoantibodies to citrullinated protein antigens [41]. Because these are IgG autoantibodies, it can be inferred that these responses come about as a consequence of T-cell recognition of related T-cell epitopes and T-cell help for B-cell antibody production. The precise T-cell determinants are still being defined, but there are studies that suggest that T cells may recognize citrulli-nated peptide epitopes. For example, studies in HLA-DR4 transgenic mice demonstrate a requirement for deimination to generate immunogenic peptides from a range of candidate antigens, through the conversion of an epitope with arginine at P4 that binds poorly to HLA-DR4, to an epitope with citrulline with more favorable binding kinetics [42]. These important observations provide a molecular basis for the association between HLA-DRB1 and anti-citrulline immunity.

The anti-CCP2 test is a robust diagnostic screening tool for anti-citrulline immunity that is now established in clinical practice, but the CCP2 antigen is not an autoantigen per se. Emerging data have defined reproducible autoantibody responses to citrullinated fibrin, vimentin, α-enolase, histones, and collagen II [43]. We can speculate that T cells from patients with RA might therefore see at least a subset of these antigens. Consistent with this notion is the observation that autoantibody responses to these specific antigens are also found in patients who carry 'at risk' HLA-DRB1 alleles [39]. For example, anti-calpastatin immunity is associated with HLA-DRB1*0404 [44], whereas anti-citrullinated α-enolase antibody responses are associated with HLA-DRB1*0401 and *0404 (Venables P, Lundberg K, personal communication). Finally, the generation of a specific monoclonal antibody that recognizes the immunodominant epitope of HCgp-39 in a complex bound to HLA-DR4 has made it possible to demonstrate in situ the presence of MHC II-peptide complexes in the inflamed RA synovial joint [45]. This observation would be entirely in keeping with previous studies demonstrating through autologous mixed lymphocyte reactions the capacity for synovial APCs to present endogenous antigenic peptides to T cells [46].

Multiple T-cell effector pathways contribute to the chronic inflammatory process

By convention, it is acknowledged that cytokines secreted by lineage specific T-helper (Th) cells provide the 'flavor' of the effector phase of cell-mediated immunity. In RA many studies have provided evidence for a Th1 bias, with the cytokine profile expressed by synovial Th effector cells derived from patients with established disease dominated by expression of IFN-γ and TNF, but also IL-10 (for review [47]). Expression of IL-2, IL-4, IL-5, and IL-13 is low or absent. This contrasts with the profile of cytokines expressed by synovial T cells and stromal cells very early in disease, which is dominated by IL-2, IL-4, IL-13, IL-17, IL-15, basic fibroblast growth factor, and epidermal growth factor [48]. The mechanisms underlying this transition are unclear, although reported data have led to the proposal that there is an intrinsic defect in Th2 differentiation that is explained in part by allelic variants of the IL-4R, which attenuate activation of STAT6 and induction of GATA-binding protein 3 [49], both of which are requisites for commitment to Th2 cell lineage differentiation.

The recent excitement following the identification of a third Th cell subset, characterized by the expression of IL-17A and IL-17F (hence termed Th17), as well as IL-22, IL-21 and TNF, has raised the possibility that Th17 cells may be an important effector T-cell subset in diseases such as RA, multiple sclerosis, and inflammatory bowel disease [50]. Much of the evidence is derived from rodent models. Certainly, expression of IL-17, IL-17RA, and IL-17RC has been described in RA synovial tissue and juxta-articular bone by immunohistochemistry [51, 52], and IL-17 protein production has been detected in synovial fluid and culture supernatants of synovial mononuclear cell explants [48, 53]. Indeed, the cytokine milieu in the joint, with IL-1β, IL-6, and IL-23 in particular, would certainly support Th17 differentiation. What is unclear is the precise phenotype of fully differentiated Th17 cells, because - unlike murine Th17 cells - IL-17+IFN-γ+ cells occur in humans. Nonetheless, IL-17 is a strong candidate effector cytokine that probably contributes to the pathogenic features of established disease, such as the following [54]: induction of TNF, IL-1, IL-6, and CCL20; synergistic actions with TNF and IL-1; and - through stimulation of stromal cells, macrophages, osteoblasts and chondrocytes - promotion of cartilage and bone destruction. Recent studies have demonstrated low frequencies of CD4+CD45RO+ T cells expressing IL-17 and TNF-α, when compared with corresponding peripheral blood T cells, with their Th1 counterparts being more abundant [55]. More data are therefore required to establish the contribution that this subset makes to early and established phases of the disease.

The spatial context of chronically activated T cells within the inflamed synovium is important. This has led investigators to explore the consequences of cell-to-cell interactions, and to uncover robust induction of monocyte-derived inflammatory cytokines, chemokines, and matrix metalloproteinases upon contact with activated T cells [56–58]. The receptor-ligand pairs involved in these interactions are likely to be numerous, but integrin and membrane TNF-dependent signaling appear to have been demonstrated consistently [59], and the reciprocal interactions between NKG2D and its ligands MICA and MICB may well contribute to the persistence of autoreactive T cells within the joint [26], as suggested above. Whereas the survival of T cells may also be supported through stromal cell derived type I IFNs [15], reciprocal interactions are also important because T cells downregulate collagen synthesis by fibroblasts, and are potent inducers of fibroblast-derived IL-6, prostaglandin E2, chemokines such as CXCL12, CXCL16 and CCL20, and matrix metalloproteinases [60–62]. Joint and bone matrix destruction is also induced in part through RANKL (receptor activator for nuclear factor-κB ligand)-positive T cells that promote bone erosion through activation and differentiation of osteoclast precursors [63]. Examples of these cellular interactions are shown in Figure 2.

Figure 2
figure 2

Cell contact-dependent effector function. Spatial and temporal organization of synovial inflammatory infiltrates favor cell contact-dependent effector responses between T cells, B cells, macrophages, and resident stromal cells and adipocytes. The extent to which Tregs influence these effector pathways is presently unclear. ADAM, a disintegrin and metalloprotease; ADAM TS, ADAM with ThromboSpondin-like motifs; FLS, fibroblast-like synoviocytes; GM-CSF, granulocyte macrophage colony-stimulating factor; HMGB1, high-mobility group box-1 proteins; IL, interleukin; M-CSF, macrophage colony-stimulating factor; MΦ, macrophage; MMP, matrix metalloproteinase; PG, prostaglandin; RANKL, receptor activator for nuclear factor-κB ligand; TCR, T-cell receptor; Teff, effector T cell; TNF, tumor necrosis factor; Treg, T regulatory cell; VEGF, vascular endothelial growth factor.

Why do effector T-cell responses persist?

A major challenge to our understanding of the pathogenesis of immune-mediated diseases is to elucidate why pathways of peripheral tolerance might fail. To this end, efforts have been invested in evaluating the number and function of regulatory cell subsets, in order to test the hypothesis that at sites of synovial inflammation there may be defects in pathways of peripheral tolerance. Studies have been thwarted to an extent because until recently the enumeration of subsets of naturally occurring regulatory T cells was based purely on CD4 and nonselective high level expression of the T-cell activation antigen CD25 (hereafter referred to as regulatory T cells [Tregs]). Even the inclusion of Foxp3 into the phenotyping armamentarium has not solved the problem, because this master transcription factor, which plays a central role in the function of Tregs [64], is also induced upon TCR engagement of CD4+CD25- T cells [65]. Furthermore, there remains controversy as to whether there are genuine defects in the numbers of circulating Tregs in RA. This may have as much to do with their selective migration and accumulation in synovial joints compared with peripheral blood [66], given that Tregs are known to express CXCR4, CCR4 and CCR8, and a phenotype commensurate with fully differentiated CD45RO+HLA-DR+CD69+CD62L- T cells, besides the Treg markers CTLA4 (cytotoxic T-lymphocyte antigen 4) and GITR (glucocorticoid-induced TNF receptor) [67]. Whether the inflammatory milieu influences their survival and function remains a topic of great debate. This is because, despite expression of transforming growth factor-β, which might promote the expression and function of Treg subsets, inflammatory cytokines such as IL-7, IL-15, and TNF have been shown to suppress Treg activity [68]. One study reported that high levels of TNF downregulate Foxp3 expression [69], whereas a second proposed that TNF may attenuate the generation of adaptive or inducible Tregs in vitro and in vivo [70]. A third study proposed that, cell for cell, there did not appear to be a consistent intrinsic defect of RA synovial Treg suppressor function, but rather synovial effector T cells might acquire resistance to regulation by Tregs [71]. This concept is in keeping with recent reports of synovial T cells being resistant to the immunosuppressive effects of indoleamine 2,3-dioxygenase [72]. How this state of refractoriness comes about is a key question and requires further investigation.

A recent study of Tregs in the KRN serum transfer model of autoimmune arthritis demonstrated that more than 40% of CD4+ T cells were CD25+Foxp3+ in the synovial fluid of either arthritic KRN mice or of C57BL/6 mice 14 days after injection with arthritogenic serum, indicating preferential recruitment of Tregs to inflamed joints, rather than local expansion [73]. When the scurfy loss-of-function Foxp3 mutant was crossed onto the KRN background, onset of arthritis was faster and more aggressive. Unexpectedly, the spectrum of affected joints changed to include joints that were normally clinically unaffected, including the hip, shoulder, and temperomandibular joint. The authors of this study raise the intriguing possibility that Tregs actively suppress lymphoid and nonlymphoid effector cells, actively protecting larger, proximal joints and rendering them refractory to arthritis. Thus, pathways of peripheral tolerance appear to be operational in chronic inflammatory disease, but are clearly insufficient for resolution of the autoimmune response.

Inherited factors and T-cell-mediated immunity in RA

What has the immunogenetics of RA taught us about disease pathogenesis? We have acknowledged the contribution of the HLA-DRB1 locus for decades, but we have only relatively recently learnt to appreciate that RA-associated HLA-DR molecules are not disease susceptibility alleles, but rather are risk factors for immunity to citrullinated antigens and for accelerated telomere erosion [14, 39]. Using a candidate gene approach, investigators have identified other disease-associated single nucleotide polymorphisms that would be expected to influence T-cell activation, either by modifying expression of HLA-DR (MHC2TA) [74], predisposing the host to increased levels of the citrullinating PAD enzymes (PADI4) [75], or by perturbing intracellular signals, which serve to regulate TCR signal transduction (CTLA4 and PD1) or its downstream consequences (STAT4) [76–78]. RA-associated allelic variants that would have an impact on pathways of T-cell activation are shown in Figure 3.

Figure 3
figure 3

Pathway biology: the impact of RA-associated allelic variants. Multiple single nucleotide polymorphisms may contribute to the disease-prone state through subtle modifications in antigen presentation ('single input') and T-cell antigen receptor responsiveness ('signal output'). It is proposed that these variants may influence multiple pathways of T-cell activation, differentiation, and effector function. DC, dendritic cell; RA, rheumatoid arthritis; TCR, T-cell receptor; Teff, effector T cell; Th, T-helper cell.

Perhaps the allelic variant that has received the most attention is the 1858T mutation of the PTPN22 gene [79], which encodes a hematopoietic protein tyrosine phosphatase called Lyp, whose homolog in the mouse (called PEP) is a potent negative regulator of TCR signaling. Lyp is thought to switch off TCR signaling by dephosphorylating tyrosine residues in a range of substrates, including ZAP-70, CD3ε, TCR-ζ, Vav, and src kinases [80]. Targeting of the src kinase Lck arises through direct protein-protein associations between a complex of Lyp and the tyrosine kinase Csk (through the P1 domain of Lyp and the SH3 domain of the tyrosine kinase Csk) and Lck. Although the Lyp R620W mutation is thought to reduce this association [79], the disease-associated variant is in fact a gain-of-function phosphatase mutant that further attenuates TCR signaling [81]. Defining how this contributes to autoimmunity is crucial because this same mutant has been found to be associated with a wide range of autoimmune syndromes, including systemic lupus erythematosus, autoimmune thyroiditis, Grave's disease, and type I diabetes, suggesting that it imparts a fundamental autoimmune diathesis [82].

One possibility is that the attenuation of TCR signaling by the Lyp R620W mutant may lead to a selection shift in thymic development, favoring the generation in the periphery of a repertoire of high-avidity autoreactive T cells. This hypothesis would certainly be in keeping with observations made in the SKG mouse, a model of spontaneous autoimmune inflammatory arthritis that arises through a point mutation in the protein tyrosine kinase ZAP-70 [83]. In this model there is clear evidence of a thymic selection shift in favor of autoreactive T cells, which - in spite of impaired TCR signals - proliferate vigourously in vivo in response to self antigen/MHC complexes and differentiate into Th17 effector T cells [84]. Impaired function of both human Lyp R620W and mouse SKG regulatory T cells has also been proposed to contribute to the autoimmune diathesis, through similar mechanisms relating to TCR signaling defects. Data to date are sparse, but one recent study [85] confirmed that PTPN22 1858T carriers make less IL-2 and IL-10 in response to TCR stimulation, whereas TNF-α and IFN-γ are spared. Interestingly, the number of memory CD4+T cells is also expanded, which might arise as a direct consequence of enhanced self reactivity. It remains to be determined whether this mutation imposes a distinct T-cell phenotype, or whether the function of B cells, NK cells, and/or dendritic cells is also affected.

Conclusion

RA is perhaps best described as a prototype chronic inflammatory disease with features of autoimmunity. The target of autoreactivity remains unclear, but emerging data suggest that in RA there are immune reactions to inflammation arising as a consequence of tissue damage and cell death, a process that is clearly dependent on aging and immune senescence. There exist both acquired and inherited defects in antigen receptor signaling that probably have impacts on both central thymic and peripheral pathways of tolerance. In the inflamed synovial joint, persistence is probably mediated more through over-exuberant co-stimulatory pathways in T cells than through conventional pathways of antigen-specific activation. Clearly, this has implications for therapy, and may explain why the more selective approaches for targeting adaptive immunity have been more successful than those based on more conventional immunosuppression. Because we now have the tools to define disease subtypes at the tissue, cellular, and molecular levels, patient-tailored therapy could soon be a reality.

Key messages

  • Our understanding of the phenotype and function of RA peripheral blood and synovial T cells points to a key role for inflammation and aging in the pathogenic process.

  • Extensive phenotypic analysis has defined a number of cell surface molecules that can be exploited for therapeutic purposes, including co-stimulatory molecules.

  • The antigenic determinants recognized by autoreactive T cells remain poorly defined in RA. However, a growing appreciation of the specificity of serum autoantibodies is likely to underpin attempts to define the relevant pathogenic antigen reactive T cell clones in the near future.

  • T-cell effector pathways in RA are complex. The precise role and contribution of Th1 and Th17 T cells, and how they contribute to the chronicity of the inflammatory process, remain to be determined.

  • More detailed phenotypic analysis and characterisation of regulatory T cell subsets are required in RA to determine whether any functional deficiencies reside in the Treg or responding T effector cell populations (or both).

  • Advances in defining genetic susceptibility to RA suggest that aberrant pathways of T-cell activation, differentiation, and persistence are key. Apart from HLA-DRB1, disease-associated variants include PTPN22, CTLA4, IL2RA, IL2RB, and STAT4.

Abbreviations

APC:

antigen-presenting cell

CCL:

CC chemokine ligand

CCP:

cyclic citrullinated peptide

CCR:

CC chemokine receptor

CXCL:

CXC chemokine ligand

CXCR:

CXC chemokine receptor

GC:

germinal center

HLA:

human leukocyte antigen

IFN:

interferon

IL:

interleukin

MHC:

major histocompatibility complex

MIC:

MHC class I chain related

NK:

natural killer

PAD:

peptidyl-arginyl deiminase

RA:

rheumatoid arthritis

STAT:

signal transducer and activator of transcription

TCR:

T-cell receptor

Th:

T-helper

TNF:

tumor necrosis factor

Treg:

regulatory T cell.

References

  1. Cope AP: Rheumatoid arthritis. Clinical Immunology. Edited by: Rich RR, Fleischer TA, Shearer WT, Schroeder HW, Frew AJ, Weyard CM. 2008, New York, NY: Elsevier, 52.1-52.23. 3

    Google Scholar 

  2. Furst DE, Breedveld FC, Kalden JR, Smolen JS, Burmester GR, Sieper J, Emery P, Keystone EC, Schiff MH, Mease P, van Riel PL, Fleischmann R, Weisman MH, Weinblatt ME: Updated consensus statement on biological agents for the treatment of rheumatic diseases, 2007. Ann Rheum Dis. 2007, 66 (suppl): 2-22. 10.1136/ard.2007.081430.

    Google Scholar 

  3. Müller-Ladner U, Ospelt C, Gay S, Distler O, Pap T: Cells of the synovium in rheumatoid arthritis. Synovial fibroblasts. Arthritis Res Ther. 2007, 9: 223-10.1186/ar2337.

    Article  PubMed Central  PubMed  Google Scholar 

  4. Duke O, Panayi GS, Janossy G, Poulter LW: An immunohistological analysis of lymphocyte subpopulations and their microenvironment in the synovial membranes of patients with rheumatoid arthritis using monoclonal antibodies. Clin Exp Immunol. 1982, 49: 22-30.

    PubMed Central  CAS  PubMed  Google Scholar 

  5. Schröder AE, Greiner A, Seyfert C, Berek C: Differentiation of B cells in the nonlymphoid tissue of the synovial membrane of patients with rheumatoid arthritis. Proc Natl Acad Sci USA. 1996, 93: 221-225. 10.1073/pnas.93.1.221.

    Article  PubMed Central  PubMed  Google Scholar 

  6. Weyand CM, Kurtin PJ, Goronzy JJ: Ectopic lymphoid organogenesis: a fast track for autoimmunity. Am J Pathol. 2001, 159: 787-793.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  7. Takemura S, Braun A, Crowson C, Kurtin PJ, Cofield RH, O'Fallon WM, Goronzy JJ, Weyand CM: Lymphoid neogenesis in rheumatoid synovitis. J Immunol. 2001, 167: 1072-1080.

    Article  CAS  PubMed  Google Scholar 

  8. van Oosterhout M, Bajema I, Levarht EW, Toes RE, Huizinga TW, van Laar JM: Differences in synovial tissue infiltrates between anti-cyclic citrullinated peptide-positive rheumatoid arthritis and anti-cyclic citrullinated peptide-negative rheumatoid arthritis. Arthritis Rheum. 2008, 58: 53-60. 10.1002/art.23148.

    Article  CAS  PubMed  Google Scholar 

  9. Wagner UG, Kurtin PJ, Wahner A, Brackertz M, Berry DJ, Goronzy JJ, Weyand CM: The role of CD8+CD40L+ T cells in the formation of germinal centers in rheumatoid synovitis. J Immunol. 1998, 161: 6390-6397.

    CAS  PubMed  Google Scholar 

  10. Kang YM, Zhang X, Wagner UG, Yang H, Beckenbaugh RD, Kurtin PJ, Goronzy JJ, Weyand CM: CD8 T cells are required for the formation of ectopic germinal centers in rheumatoid synovitis. J Exp Med. 2002, 195: 1325-1336. 10.1084/jem.20011565.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. Timmer TC, Baltus B, Vondenhoff M, Huizinga TW, Tak PP, Verweij CL, Mebius RE, Pouw Kraan van der TC: Inflammation and ectopic lymphoid structures in rheumatoid arthritis synovial tissues dissected by genomics technology: identification of the interleukin-7 signaling pathway in tissues with lymphoid neogenesis. Arthritis Rheum. 2007, 56: 2492-2502. 10.1002/art.22748.

    Article  CAS  PubMed  Google Scholar 

  12. Koetz K, Bryl E, Spickschen K, O'Fallon WM, Goronzy JJ, Weyand CM: T cell homeostasis in patients with rheumatoid arthritis. Proc Natl Acad Sci USA. 2000, 97: 9203-9208. 10.1073/pnas.97.16.9203.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  13. Ponchel F, Morgan AW, Bingham SJ, Quinn M, Buch M, Verburg RJ, Henwood J, Douglas SH, Masurel A, Conaghan P, Gesinde M, Taylor J, Markham AF, Emery P, van Laar JM, Isaacs JD: Dys-regulated lymphocyte proliferation and differentiation in patients with rheumatoid arthritis. Blood. 2002, 100: 4550-4556. 10.1182/blood-2002-03-0671.

    Article  CAS  PubMed  Google Scholar 

  14. Schönland SO, Lopez C, Widmann T, Zimmer J, Bryl E, Goronzy JJ, Weyand CM: Premature telomeric loss in rheumatoid arthritis is genetically determined and involves both myeloid and lymphoid cell lineages. Proc Natl Acad Sci USA. 2003, 100: 13471-13476. 10.1073/pnas.2233561100.

    Article  PubMed Central  PubMed  Google Scholar 

  15. Salmon M, Scheel-Toellner D, Huissoon AP, Pilling D, Shamsadeen N, Hyde H, D'Angeac AD, Bacon PA, Emery P, Akbar AN: Inhibition of T cell apoptosis in the rheumatoid synovium. J Clin Invest. 1997, 99: 439-446. 10.1172/JCI119178.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  16. Schmidt D, Martens PB, Weyand CM, Goronzy JJ: The repertoire of CD4+ CD28- T cells in rheumatoid arthritis. Mol Med. 1996, 2: 608-618.

    PubMed Central  CAS  PubMed  Google Scholar 

  17. Maurice MM, Lankester AC, Bezemer AC, Geertsma MF, Tak PP, Breedveld FC, van Lier RA, Verweij CL: Defective TCR-mediated signaling in synovial T cells in rheumatoid arthritis. J Immunol. 1997, 159: 2973-2978.

    CAS  PubMed  Google Scholar 

  18. Zhang Z, Gorman CL, Vermi AC, Monaco C, Foey A, Owen S, Amjadi P, Vallance A, McClinton C, Marelli-Berg F, Isomäki P, Russell A, Dazzi F, Vyse TJ, Brennan FM, Cope AP: TCRzetadim lymphocytes define populations of circulating effector cells that migrate to inflamed tissues. Blood. 2007, 109: 4328-4335. 10.1182/blood-2006-12-064170.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  19. Dolhain RJ, Heiden van der AN, ter Haar NT, Breedveld FC, Miltenburg AM: Shift toward T lymphocytes with a T helper 1 cytokine-secretion profile in the joints of patients with rheumatoid arthritis. Arthritis Rheum. 1996, 12: 1961-1969. 10.1002/art.1780391204.

    Article  Google Scholar 

  20. Norii M, Yamamura M, Iwahashi M, Ueno A, Yamana J, Makino H: Selective recruitment of CXCR3+ and CCR5+CCR4+ T cells into synovial tissue in patients with rheumatoid arthritis. Acta Med Okayama. 2006, 60: 149-157.

    CAS  PubMed  Google Scholar 

  21. Weyand CM, Goronzy JJ: Medium- and large-vessel vasculitis. N Engl J Med. 2003, 349: 160-169. 10.1056/NEJMra022694.

    Article  CAS  PubMed  Google Scholar 

  22. Liuzzo G, Goronzy JJ, Yang H, Kopecky SL, Holmes DR, Frye RL, Weyand CM: Monoclonal T-cell proliferation and plaque instability in acute coronary syndromes. Circulation. 2000, 101: 2883-2888.

    Article  CAS  PubMed  Google Scholar 

  23. Giacomelli R, Passacantando A, Perricone R, Parzanese I, Rascente M, Minisola G, Tonietti G: T lymphocytes in the synovial fluid of patients with active rheumatoid arthritis display CD134-OX40 surface antigen. Clin Exp Rheumatol. 2001, 19: 317-320.

    CAS  PubMed  Google Scholar 

  24. Namekawa T, Snyder MR, Yen JH, Goehring BE, Leibson PJ, Weyand CM, Goronzy JJ: Killer cell activating receptors function as costimulatory molecules on CD4+CD28null T cells clonally expanded in rheumatoid arthritis. J Immunol. 2000, 165: 1138-1145.

    Article  CAS  PubMed  Google Scholar 

  25. Snyder MR, Nakajima T, Leibson PJ, Weyand CM, Goronzy JJ: Stimulatory killer Ig-like receptors modulate T cell activation through DAP12-dependent and DAP12-independent mechanisms. J Immunol. 2004, 173: 3725-3731.

    Article  CAS  PubMed  Google Scholar 

  26. Groh V, Bruhl A, El-Gabalawy H, Nelson JL, Spies T: Stimulation of T cell autoreactivity by anomalous expression of NKG2D and its MIC ligands in rheumatoid arthritis. Proc Natl Acad Sci USA. 2003, 100: 9452-9457. 10.1073/pnas.1632807100.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  27. Courtenay JS, Dallman MJ, Dayan AD, Martin A, Mosedale B: Immunisation against heterologous type II collagen induces arthritis in mice. Nature. 1980, 283: 666-668. 10.1038/283666a0.

    Article  CAS  PubMed  Google Scholar 

  28. Verheijden GF, Rijnders AW, Bos E, Coenen-de Roo CJ, van Staveren CJ, Miltenburg AM, Meijerink JH, Elewaut D, de Keyser F, Veys E, Boots AM: Human cartilage glycoprotein-39 as a candidate autoantigen in rheumatoid arthritis. Arthritis Rheum. 1997, 40: 1115-1125. 10.1002/art.1780400616.

    Article  CAS  PubMed  Google Scholar 

  29. Szántó S, Bárdos T, Szabó Z, David CS, Buzás EI, Mikecz K, Glant TT: Induction of arthritis in HLA-DR4-humanized and HLA-DQ8-humanized mice by human cartilage proteoglycan aggrecan but only in the presence of an appropriate (non-MHC) genetic background. Arthritis Rheum. 2004, 50: 1984-1995. 10.1002/art.20285.

    Article  PubMed  Google Scholar 

  30. Van Eden W, Wick G, Albani S, Cohen I: Stress, heat shock proteins, and autoimmunity: how immune responses to heat shock proteins are to be used for the control of chronic inflammatory diseases. Ann N Y Acad Sci. 2007, 1113: 217-237. 10.1196/annals.1391.020.

    Article  CAS  PubMed  Google Scholar 

  31. Bodman-Smith MD, Corrigall VM, Berglin E, Cornell HR, Tzioufas AG, Mavragani CP, Chan C, Rantapää-Dahlqvist S, Panayi GS: Antibody response to the human stress protein BiP in rheumatoid arthritis. Rheumatology. 2004, 43: 1283-1287. 10.1093/rheumatology/keh312.

    Article  CAS  PubMed  Google Scholar 

  32. Schwab R, Szabo P, Manavalan JS, Weksler ME, Posnett DN, Pannetier C, Kourilsky P, Even J: Expanded CD4+ and CD8+T cell clones in elderly humans. J Immunol. 1997, 158: 4493-4499.

    CAS  PubMed  Google Scholar 

  33. Goronzy JJ, Weyand CM: Aging, autoimmunity and arthritis: T-cell senescence and contraction of T-cell repertoire diversity-catalysts of autoimmunity and chronic inflammation. Arthritis Res Ther. 2003, 5: 225-234. 10.1186/ar974.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  34. Fitzgerald JE, Ricalton NS, Meyer AC, West SG, Kaplan H, Behrendt C, Kotzin BL: Analysis of clonal CD8+ T cell expansions in normal individuals and patients with rheumatoid arthritis. J Immunol. 1995, 154: 3538-3547.

    CAS  PubMed  Google Scholar 

  35. Hingorani R, Monteiro J, Furie R, Chartash E, Navarrete C, Pergolizzi R, Gregersen PK: Oligoclonality of V beta 3 TCR chains in the CD8+ T cell population of rheumatoid arthritis patients. J Immunol. 1996, 156: 852-858.

    CAS  PubMed  Google Scholar 

  36. Wagner UG, Koetz K, Weyand CM, Goronzy JJ: Perturbation of the T cell repertoire in rheumatoid arthritis. Proc Natl Acad Sci USA. 1998, 95: 14447-14452. 10.1073/pnas.95.24.14447.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  37. Unutmaz D, Pileri P, Abrignani S: Antigen-independent activation of naive and memory resting T cells by a cytokine combination. J Exp Med. 1994, 180: 1159-1164. 10.1084/jem.180.3.1159.

    Article  CAS  PubMed  Google Scholar 

  38. Gregersen PK, Silver J, Winchester RJ: The shared epitope hypothesis. An approach to understanding the molecular genetics of susceptibility to rheumatoid arthritis. Arthritis Rheum. 1987, 30: 1205-1213. 10.1002/art.1780301102.

    Article  CAS  PubMed  Google Scholar 

  39. Huizinga TW, Amos CI, Helm-van Mil van der AH, Chen W, van Gaalen FA, Jawaheer D, Schreuder GM, Wener M, Breedveld FC, Ahmad N, Lum RF, de Vries RR, Gregersen PK, Toes RE, Criswell LA: Refining the complex rheumatoid arthritis phenotype based on specificity of the HLA-DRB1 shared epitope for antibodies to citrullinated proteins. Arthritis Rheum. 2005, 52: 3433-3438. 10.1002/art.21385.

    Article  CAS  PubMed  Google Scholar 

  40. van Venrooij WJ, Zendman AJ, Pruijn GJ: Autoantibodies to citrullinated antigens in (early) rheumatoid arthritis. Autoimmun Rev. 2006, 6: 37-41. 10.1016/j.autrev.2006.03.008.

    Article  CAS  PubMed  Google Scholar 

  41. Schellekens GA, de Jong BA, Hoogen van den FH, Putte van de LB, van Venrooij WJ: Citrulline is an essential constituent of antigenic determinants recognized by rheumatoid arthritis-specific autoantibodies. J Clin Invest. 1998, 101: 273-281. 10.1172/JCI1316.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  42. Hill JA, Southwood S, Sette A, Jevnikar AM, Bell DA, Cairns E: Cutting edge: the conversion of arginine to citrulline allows for a high-affinity peptide interaction with the rheumatoid arthritis-associated HLA-DRB1*0401 MHC class II molecule. J Immunol. 2003, 171: 538-541.

    Article  CAS  PubMed  Google Scholar 

  43. Klareskog L, Rönnelid J, Lundberg K, Padyukov L, Alfredsson L: Immunity to citrullinated proteins in rheumatoid arthritis. Annu Rev Immunol. 2008, 26: 651-675. 10.1146/annurev.immunol.26.021607.090244.

    Article  CAS  PubMed  Google Scholar 

  44. Auger I, Roudier C, Guis S, Balandraud N, Roudier J: HLA-DRB1*0404 is strongly associated with anticalpastatin antibodies in rheumatoid arthritis. Ann Rheum Dis. 2007, 66: 1588-1593. 10.1136/ard.2006.067231.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  45. Steenbakkers PG, Baeten D, Rovers E, Veys EM, Rijnders AW, Meijerink J, De Keyser F, Boots AM: Localization of MHC class II/human cartilage glycoprotein-39 complexes in synovia of rheumatoid arthritis patients using complex-specific monoclonal antibodies. J Immunol. 2003, 170: 5719-5727.

    Article  CAS  PubMed  Google Scholar 

  46. Thomas R, Davis LS, Lipsky PE: Rheumatoid synovium is enriched in mature antigen-presenting dendritic cells. J Immunol. 1994, 152: 2613-2623.

    CAS  PubMed  Google Scholar 

  47. McInnes IB, Schett G: Cytokines in the pathogenesis of rheumatoid arthritis. Nat Rev Immunol. 2007, 7: 429-442. 10.1038/nri2094.

    Article  CAS  PubMed  Google Scholar 

  48. Raza K, Scheel-Toellner D, Lee CY, Pilling D, Curnow SJ, Falciani F, Trevino V, Kumar K, Assi LK, Lord JM, Gordon C, Buckley CD, Salmon M: Synovial fluid leukocyte apoptosis is inhibited in patients with very early rheumatoid arthritis. Arthritis Res Ther. 2006, 8: R120-10.1186/ar2009.

    Article  PubMed Central  PubMed  Google Scholar 

  49. Prots I, Skapenko A, Wendler J, Mattyasovszky S, Yoné CL, Spriewald B, Burkhardt H, Rau R, Kalden JR, Lipsky PE, Schulze-Koops H: Association of the IL4R single-nucleotide polymorphism I50V with rapidly erosive rheumatoid arthritis. Arthritis Rheum. 2006, 54: 1491-500. 10.1002/art.21832.

    Article  CAS  PubMed  Google Scholar 

  50. Weaver CT, Hatton RD, Mangan PR, Harrington LE: IL-17 family cytokines and the expanding diversity of effector T cell lineages. Annu Rev Immunol. 2007, 25: 821-852. 10.1146/annurev.immunol.25.022106.141557.

    Article  CAS  PubMed  Google Scholar 

  51. Chabaud M, Durand JM, Buchs N, Fossiez F, Page G, Frappart L, Miossec P: Human interleukin-17: A T cell-derived proinflammatory cytokine produced by the rheumatoid synovium. Arthritis Rheum. 1999, 42: 963-970. 10.1002/1529-0131(199905)42:5<963::AID-ANR15>3.0.CO;2-E.

    Article  CAS  PubMed  Google Scholar 

  52. Zrioual S, Toh ML, Tournadre A, Zhou Y, Cazalis MA, Pachot A, Miossec V, Miossec P: IL-17RA and IL-17RC receptors are essential for IL-17A-induced ELR+ CXC chemokine expression in synoviocytes and are overexpressed in rheumatoid blood. J Immunol. 2008, 180: 655-663.

    Article  CAS  PubMed  Google Scholar 

  53. Chabaud M, Garnero P, Dayer JM, Guerne PA, Fossiez F, Miossec P: Contribution of interleukin 17 to synovium matrix destruction in rheumatoid arthritis. Cytokine. 2000, 12: 1092-1099. 10.1006/cyto.2000.0681.

    Article  CAS  PubMed  Google Scholar 

  54. Miossec P: Interleukin-17 in fashion, at last: ten years after its description, its cellular source has been identified. Arthritis Rheum. 2007, 56: 2111-2115. 10.1002/art.22733.

    Article  CAS  PubMed  Google Scholar 

  55. Yamada H, Nakashima Y, Okazaki K, Mawatari T, Fukushi JI, Kaibara N, Hori A, Iwamoto Y, Yoshikai Y: Th1 but not Th17 cells predominate in the joints of patients with rheumatoid arthritis. Ann Rheum Dis. 2007

    Google Scholar 

  56. McInnes IB, Leung BP, Sturrock RD, Field M, Liew FY: Interleukin-15 mediates T cell-dependent regulation of tumor necrosis factor-alpha production in rheumatoid arthritis. Nat Med. 1997, 3: 189-195. 10.1038/nm0297-189.

    Article  CAS  PubMed  Google Scholar 

  57. Beech JT, Andreakos E, Ciesielski CJ, Green P, Foxwell BM, Brennan FM: T-cell contact-dependent regulation of CC and CXC chemokine production in monocytes through differential involvement of NFkappaB: implications for rheumatoid arthritis. Arthritis Res Ther. 2006, 8: R168-10.1186/ar2077.

    Article  PubMed Central  PubMed  Google Scholar 

  58. Lacraz S, Isler P, Vey E, Welgus HG, Dayer JM: Direct contact between T lymphocytes and monocytes is a major pathway for induction of metalloproteinase expression. J Biol Chem. 1994, 269: 22027-22033.

    CAS  PubMed  Google Scholar 

  59. Burger D, Dayer JM: The role of human T-lymphocyte-monocyte contact in inflammation and tissue destruction. Arthritis Res. 2002, 4 (suppl): S169-S176. 10.1186/ar558.

    Article  PubMed Central  PubMed  Google Scholar 

  60. Rezzonico R, Burger D, Dayer JM: Direct contact between T lymphocytes and human dermal fibroblasts or synoviocytes down-regulates types I and III collagen production via cell-associated cytokines. J Biol Chem. 1998, 273: 18720-18728. 10.1074/jbc.273.30.18720.

    Article  CAS  PubMed  Google Scholar 

  61. Yamamura Y, Gupta R, Morita Y, He X, Pai R, Endres J, Freiberg A, Chung K, Fox DA: Effector function of resting T cells: activation of synovial fibroblasts. J Immunol. 2001, 166: 2270-2275.

    Article  CAS  PubMed  Google Scholar 

  62. Tran CN, Lundy SK, White PT, Endres JL, Motyl CD, Gupta R, Wilke CM, Shelden EA, Chung KC, Urquhart AG, Fox DA: Molecular interactions between T cells and fibroblast-like synoviocytes: role of membrane tumor necrosis factor-alpha on cytokine-activated T cells. Am J Pathol. 2007, 171: 1588-1598. 10.2353/ajpath.2007.070004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  63. Udagawa N, Kotake S, Kamatani N, Takahashi N, Suda T: The molecular mechanism of osteoclastogenesis in rheumatoid arthritis. Arthritis Res. 2002, 4: 281-289. 10.1186/ar431.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  64. Zheng Y, Rudensky AY: Foxp3 in control of the regulatory T cell lineage. Nat Immunol. 2007, 8: 457-462. 10.1038/ni1455.

    Article  CAS  PubMed  Google Scholar 

  65. Wang J, Ioan-Facsinay A, Voort van der EI, Huizinga TW, Toes RE: Transient expression of FOXP3 in human activated non-regulatory CD4+ T cells. Eur J Immunol. 2007, 37: 129-138. 10.1002/eji.200636435.

    Article  CAS  PubMed  Google Scholar 

  66. Cao D, van Vollenhoven R, Klareskog L, Trollmo C, Malmström V: CD25brightCD4+ regulatory T cells are enriched in inflamed joints of patients with chronic rheumatic disease. Arthritis Res Ther. 2004, 6: R335-R346. 10.1186/ar1192.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  67. Baecher-Allan C, Hafler DA: Human regulatory T cells and their role in autoimmune disease. Immunol Rev. 2006, 212: 203-216. 10.1111/j.0105-2896.2006.00417.x.

    Article  CAS  PubMed  Google Scholar 

  68. van Amelsfort JM, van Roon JA, Noordegraaf M, Jacobs KM, Bijlsma JW, Lafeber FP, Taams LS: Proinflammatory mediator-induced reversal of CD4+, CD25+ regulatory T cell-mediated suppression in rheumatoid arthritis. Arthritis Rheum. 2007, 56: 732-742. 10.1002/art.22414.

    Article  CAS  PubMed  Google Scholar 

  69. Valencia X, Stephens G, Goldbach-Mansky R, Wilson M, Shevach EM, Lipsky PE: TNF downmodulates the function of human CD4+CD25hi T-regulatory cells. Blood. 2006, 108: 253-261. 10.1182/blood-2005-11-4567.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  70. Nadkarni S, Mauri C, Ehrenstein MR: Anti-TNF-alpha therapy induces a distinct regulatory T cell population in patients with rheumatoid arthritis via TGF-beta. J Exp Med. 2007, 204: 33-39. 10.1084/jem.20061531.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  71. van Amelsfort JM, Jacobs KM, Bijlsma JW, Lafeber FP, Taams LS: CD4+CD25+ regulatory T cells in rheumatoid arthritis: differences in the presence, phenotype, and function between peripheral blood and synovial fluid. Arthritis Rheum. 2004, 50: 2775-2785. 10.1002/art.20499.

    Article  PubMed  Google Scholar 

  72. Zhu L, Ji F, Wang Y, Zhang Y, Liu Q, Zhang JZ, Matsushima K, Cao Q, Zhang Y: Synovial autoreactive T cells in rheumatoid arthritis resist IDO-mediated inhibition. J Immunol. 2006, 177: 8226-8233.

    Article  CAS  PubMed  Google Scholar 

  73. Nguyen LT, Jacobs J, Mathis D, Benoist C: Where FoxP3-dependent regulatory T cells impinge on the development of inflammatory arthritis. Arthritis Rheum. 2007, 56: 509-520. 10.1002/art.22272.

    Article  CAS  PubMed  Google Scholar 

  74. Swanberg M, Lidman O, Padyukov L, Eriksson P, Akesson E, Jagodic M, Lobell A, Khademi M, Börjesson O, Lindgren CM, Lundman P, Brookes AJ, Kere J, Luthman H, Alfredsson L, Hillert J, Klareskog L, Hamsten A, Piehl F, Olsson T: MHC2TA is associated with differential MHC molecule expression and susceptibility to rheumatoid arthritis, multiple sclerosis and myocardial infarction. Nat Genet. 2005, 37: 486-494. 10.1038/ng1544.

    Article  CAS  PubMed  Google Scholar 

  75. Suzuki A, Yamada R, Chang X, Tokuhiro S, Sawada T, Suzuki M, Nagasaki M, Nakayama-Hamada M, Kawaida R, Ono M, Ohtsuki M, Furukawa H, Yoshino S, Yukioka M, Tohma S, Matsubara T, Wakitani S, Teshima R, Nishioka Y, Sekine A, Iida A, Takahashi A, Tsunoda T, Nakamura Y, Yamamoto K: Functional haplotypes of PADI4, encoding citrullinating enzyme peptidylarginine deiminase 4, are associated with rheumatoid arthritis. Nat Genet. 2003, 34: 395-402. 10.1038/ng1206.

    Article  CAS  PubMed  Google Scholar 

  76. Plenge RM, Padyukov L, Remmers EF, Purcell S, Lee AT, Karlson EW, Wolfe F, Kastner DL, Alfredsson L, Altshuler D, Gregersen PK, Klareskog L, Rioux JD: Replication of putative candidate-gene associations with rheumatoid arthritis in >4,000 samples from North America and Sweden: association of susceptibility with PTPN22, CTLA4, and PADI4. Am J Hum Genet. 2005, 1044-1060. 10.1086/498651.

    Google Scholar 

  77. Lin SC, Yen JH, Tsai JJ, Tsai WC, Ou TT, Liu HW, Chen CJ: Association of a programmed death 1 gene polymorphism with the development of rheumatoid arthritis, but not systemic lupus erythematosus. Arthritis Rheum. 2004, 50: 770-775. 10.1002/art.20040.

    Article  CAS  PubMed  Google Scholar 

  78. Remmers EF, Plenge RM, Lee AT, Graham RR, Hom G, Behrens TW, de Bakker PI, Le JM, Lee HS, Batliwalla F, Li W, Masters SL, Booty MG, Carulli JP, Padyukov L, Alfredsson L, Klareskog L, Chen WV, Amos CI, Criswell LA, Seldin MF, Kastner DL, Gregersen PK: STAT4 and the risk of rheumatoid arthritis and systemic lupus erythematosus. N Engl J Med. 2007, 357: 977-986. 10.1056/NEJMoa073003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  79. Begovich AB, Carlton VE, Honigberg LA, Schrodi SJ, Chokkalingam AP, Alexander HC, Ardlie KG, Huang Q, Smith AM, Spoerke JM, Conn MT, Chang M, Chang SY, Saiki RK, Catanese JJ, Leong DU, Garcia VE, McAllister LB, Jeffery DA, Lee AT, Batliwalla F, Remmers E, Criswell LA, Seldin MF, Kastner DL, Amos CI, Sninsky JJ, Gregersen PK: A missense single-nucleotide polymorphism in a gene encoding a protein tyrosine phosphatase (PTPN22) is associated with rheumatoid arthritis. Am J Hum Genet. 2004, 75: 330-337. 10.1086/422827.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  80. Wu J, Katrekar A, Honigberg LA, Smith AM, Conn MT, Tang J, Jeffery D, Mortara K, Sampang J, Williams SR, Buggy J, Clark JM: Identification of substrates of human protein-tyrosine phosphatase PTPN22. J Biol Chem. 2006, 281: 11002-11010. 10.1074/jbc.M600498200.

    Article  CAS  PubMed  Google Scholar 

  81. Vang T, Congia M, Macis MD, Musumeci L, Orrú V, Zavattari P, Nika K, Tautz L, Taskén K, Cucca F, Mustelin T, Bottini N: Autoimmune-associated lymphoid tyrosine phosphatase is a gain-of-function variant. Nat Genet. 2005, 37: 1317-1319. 10.1038/ng1673.

    Article  CAS  PubMed  Google Scholar 

  82. Bottini N, Vang T, Cucca F, Mustelin T: Role of PTPN22 in type 1 diabetes and other autoimmune diseases. Semin Immunol. 2006, 18: 207-213. 10.1016/j.smim.2006.03.008.

    Article  CAS  PubMed  Google Scholar 

  83. Sakaguchi N, Takahashi T, Hata H, Nomura T, Tagami T, Yamazaki S, Sakihama T, Matsutani T, Negishi I, Nakatsuru S, Sakaguchi S: Altered thymic T-cell selection due to a mutation of the ZAP-70 gene causes autoimmune arthritis in mice. Nature. 2003, 426: 454-460. 10.1038/nature02119.

    Article  CAS  PubMed  Google Scholar 

  84. Hirota K, Hashimoto M, Yoshitomi H, Tanaka S, Nomura T, Yamaguchi T, Iwakura Y, Sakaguchi N, Sakaguchi S: T cell self-reactivity forms a cytokine milieu for spontaneous development of IL-17+ Th cells that cause autoimmune arthritis. J Exp Med. 2007, 204: 41-47. 10.1084/jem.20062259.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  85. Rieck M, Arechiga A, Onengut-Gumuscu S, Greenbaum C, Concannon P, Buckner JH: Genetic variation in PTPN22 corresponds to altered function of T and B lymphocytes. J Immunol. 2007, 179: 4704-4710.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The author's laboratory research has been supported by grants from the Wellcome Trust UK, the Arthritis Research Campaign (arc), the Medical Research Council UK, and the European Commission.

This article is published as part of Arthritis Research & Therapy Volume 10 Supplement 1, 2008: Co-stimulation blockade: from bench to bedside. The full contents of the supplement are available online at http://arthritis-research.com/supplements/10/S1.

Publication of this supplement has been sponsored by Bristol-Myers Squibb Company.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andrew P Cope.

Additional information

Competing interests

The author has no competing financial or intellectual interests.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Cope, A.P. T cells in rheumatoid arthritis. Arthritis Res Ther 10 (Suppl 1), S1 (2008). https://doi.org/10.1186/ar2412

Download citation

  • Published:

  • DOI: https://doi.org/10.1186/ar2412

Keywords