fbpx
Wikipedia

Regulatory T cell

The regulatory T cells (Tregs /ˈtrɛɡ/ or Treg cells), formerly known as suppressor T cells, are a subpopulation of T cells that modulate the immune system, maintain tolerance to self-antigens, and prevent autoimmune disease. Treg cells are immunosuppressive and generally suppress or downregulate induction and proliferation of effector T cells.[1] Treg cells express the biomarkers CD4, FOXP3, and CD25 and are thought to be derived from the same lineage as naïve CD4+ cells.[2] Because effector T cells also express CD4 and CD25, Treg cells are very difficult to effectively discern from effector CD4+, making them difficult to study. Research has found that the cytokine transforming growth factor beta (TGF-β) is essential for Treg cells to differentiate from naïve CD4+ cells and is important in maintaining Treg cell homeostasis.[3]

Mouse models have suggested that modulation of Treg cells can treat autoimmune disease and cancer and can facilitate organ transplantation[4] and wound healing.[5] Their implications for cancer are complicated. Treg cells tend to be upregulated in individuals with cancer, and they seem to be recruited to the site of many tumors. Studies in both humans and animal models have implicated that high numbers of Treg cells in the tumor microenvironment is indicative of a poor prognosis, and Treg cells are thought to suppress tumor immunity, thus hindering the body's innate ability to control the growth of cancerous cells.[6] Immunotherapy research is studying how regulation of T cells could possibly be utilized in the treatment of cancer.[7]

Populations edit

T regulatory cells are a component of the immune system that suppress immune responses of other cells. This is an important "self-check" built into the immune system to prevent excessive reactions. Regulatory T cells come in many forms with the most well-understood being those that express CD4, CD25, and FOXP3 (CD4+CD25+ regulatory T cells). These Treg cells are different from helper T cells.[8] Another regulatory T cell subset is Treg17 cells.[9] Regulatory T cells are involved in shutting down immune responses after they have successfully eliminated invading organisms, and also in preventing autoimmunity.[10]

CD4+ FOXP3+ CD25(high) regulatory T cells have been called "naturally occurring" regulatory T cells[11] to distinguish them from "suppressor" T cell populations that are generated in vitro. Additional regulatory T cell populations include Tr1, Th3, CD8+CD28, and Qa-1 restricted T cells. The contribution of these populations to self-tolerance and immune homeostasis is less well defined. FOXP3 can be used as a good marker for mouse CD4+CD25+ T cells, although recent studies have also shown evidence for FOXP3 expression in CD4+CD25 T cells. In humans, FOXP3 is also expressed by recently activated conventional T cells and thus does not specifically identify human Tregs.[12]

Development edit

All T cells derive from progenitor cells in the bone marrow, which become committed to their lineage in the thymus. All T cells begin as CD4-CD8-TCR- cells at the DN (double-negative) stage, where an individual cell will rearrange its T cell receptor genes to form a unique, functional molecule, which they, in turn, test against cells in the thymic cortex for a minimal level of interaction with self-MHC. If they receive these signals, they proliferate and express both CD4 and CD8, becoming double-positive cells. The selection of Tregs occurs on radio-resistant hematopoietically derived MHC class II-expressing cells in the medulla or Hassall's corpuscles in the thymus. At the DP (double-positive) stage, they are selected by their interaction with the cells within the thymus, begin the transcription of Foxp3, and become Treg cells, although they may not begin to express Foxp3 until the single-positive stage, at which point they are functional Tregs. Tregs do not have the limited TCR expression of NKT or γδ T cells; Tregs have a larger TCR diversity than effector T cells, biased towards self-peptides.

The process of Treg selection is determined by the affinity of interaction with the self-peptide MHC complex. Selection to become a Treg is a "Goldilocks" process - i.e. not too high, not too low, but just right;[13] a T cell that receives very strong signals will undergo apoptotic death; a cell that receives a weak signal will survive and be selected to become an effector cell. If a T cell receives an intermediate signal, then it will become a regulatory cell. Due to the stochastic nature of the process of T cell activation, all T cell populations with a given TCR will end up with a mixture of Teff and Treg – the relative proportions determined by the affinities of the T cell for the self-peptide-MHC. Even in mouse models with TCR-transgenic cells selected on specific-antigen-secreting stroma, deletion or conversion is not complete.

After interaction with self-peptide MHC complex, T cell has to upregulate IL-2R, CD25 and TNFR superfamily members GITR, OX40 and TNFR2 to become CD25+FOXP3- Treg cell progenitor. To become mature Treg, FOXP3 transcription factor has to be upregulated which is driven by γ-chain (CD132) dependent cytokines, in particular IL-2 and/or IL-15.[14][15] Only IL-2 alone is not sufficient to stimulate Foxp3 expression, other cytokines are needed. Whereas IL-2 is produced by self-reactive thymocytes, IL-15 is produced by stromal cells of the thymus, mainly mTECs and cTECs.[14]

Recently, other subset of Treg precursors was identified. This subset lacks CD25 and has low expression of Foxp3. Its development is mainly dependent on IL-15. This subset has lower affinity for self antigens than CD25+Foxp3high subset.  Both subsets generate mature Treg cells after stimulation with IL-2 with comparable efficiency both in vitro and in vivo. CD25+Foxp3high progenitors exhibit increased apoptosis and develop into mature Treg cells with faster kinetics than Foxp3low progenitors.[16] Tregs derived from CD25+Foxp3high progenitors protect from experimental auto-immune encephalomyelitis, whereas those derived from CD25+Foxp3low progenitors protect from T-cell induced colitis.[14]

Mature CD25+Foxp3+ Tregs can be also divided into two different subsets based on the expression level of CD25, GITR, and PD-1. Tregs expressing low amounts of CD25, GITR and PD-1 limit the development of colitis by promoting the conversion of conventional CD4+ T cells into pTreg. Tregs highly expressing CD25, GITR and PD-1 are more self-reactive and control lymphoproliferation in peripheral lymph nodes - they may have the ability to protect against autoimmune disorders.[14]

Foxp3+ Treg generation in the thymus is delayed by several days compared to Teff cells and does not reach adult levels in either the thymus or periphery until around three weeks post-partum. Treg cells require CD28 co-stimulation and B7.2 expression is largely restricted to the medulla, the development of which seems to parallel the development of Foxp3+ cells. It has been suggested that the two are linked, but no definitive link between the processes has yet been shown. TGF-β is not required for Treg functionality, in the thymus, as thymic Tregs from TGF-β insensitive TGFβRII-DN mice are functional.

Thymic recirculation edit

There was observed, that some FOXP3+ Treg cells are recirculating back to thymus, where they have developed. This Treg were mainly present in thymic medulla, which is the main site of Treg cells differentiation.[17] The presence of this cells in thymus or addition into fetal thymic tissue culture suppress development of new Treg cells by 34–60%,[17] but Tconv cells are not affected. That means, that recirculating Treg to thymus inhibited just de novo development of Treg cells. Molecular mechanism of this process works due to the ability of Treg to adsorb IL-2 from the microenvironments, thus being able to induce apoptosis of other T cells which need IL-2 as main growth factor.[18] Recirculating T reg cells in thymus express high amount of high affinity IL-2 receptor α chain (CD25) encoded by Il2ra gene which gather IL-2 from thymic medulla, and decrease its concentration. New generated FOXP3+ Treg cells in thymus have not so high amount of Il2ra expression.[17] IL-2 is a cytokine necessary for the development of Treg cells in the thymus. It is important for T cells proliferation and survival, but in the case of its deficiency, IL-15 may be replaced. However, Treg cells' development is dependent on IL-2.[19] In humans, there was found population of CD31 negative Treg cells in thymus.[17] CD31 could be used as a marker of new generated Treg cells as same as other T lymphocytes. Mature and peripheral Treg cells have decreased its expression.[20] So it is possible that this regulatory mechanism of thymic Treg cells development is also functional in humans.

There is probably also positive regulation of thymic Treg cells development caused by recirculating Treg cells into thymus. There was found population of CD24 low FOXP3+ in thymus with increased expression of IL-1R2 (Il1r2) compared with peripheral Treg cells.[21][22] High concentration of IL-1β caused by inflammation decrease de novo development of Treg cells in thymus.[22] The presence of recirculating Treg cells in the thymus with high IL1R2 expression during inflammatory conditions helps to uptake IL-1β and reduce its concentration in the medulla microenvironment, thus they are helping to the development of de novo Treg cells.[22] High concentration of IL-1β caused by inflammation decrease de novo development of Treg cells in thymus.[22] Binding of IL-1β to IL1R2 on the surface of Treg cells does not cause any signal transduction because there is no present Intracellular (TIR) Toll interleukin-1 receptor domain, which is normally present in innate immune cells.[23]

Function edit

The immune system must be able to discriminate between self and non-self. When self/non-self discrimination fails, the immune system destroys cells and tissues of the body and as a result causes autoimmune diseases. Regulatory T cells actively suppress activation of the immune system and prevent pathological self-reactivity, i.e. autoimmune disease. The critical role regulatory T cells play within the immune system is evidenced by the severe autoimmune syndrome that results from a genetic deficiency in regulatory T cells (IPEX syndrome – see also below).

 
Diagram of regulatory T cell, effector T cells and dendritic cell showing putative mechanisms of suppression by regulatory T cells.

The molecular mechanism by which regulatory T cells exert their suppressor/regulatory activity has not been definitively characterized and is the subject of intense research. In vitro experiments have given mixed results regarding the requirement of cell-to-cell contact with the cell being suppressed. The following represent some of the proposed mechanisms of immune suppression:

  • Regulatory T cells produce a number of inhibitory cytokines. These include TGF-β,[24] Interleukin 35,[25] and Interleukin 10.[26] It also appears that regulatory T cells can induce other cell types to express interleukin-10.[27]
  • Regulatory T cells can produce Granzyme B, which in turn can induce apoptosis of effector cells. Regulatory T cells from Granzyme B deficient mice are reported to be less effective suppressors of the activation of effector T cells.[28]
  • Reverse signalling through direct interaction with dendritic cells and the induction of immunosuppressive indoleamine 2,3-dioxygenase.[29]
  • Signalling through the ectoenzymes CD39 and CD73 with the production of immunosuppressive adenosine.[30][31]
  • Through direct interactions with dendritic cells by LAG3 and by TIGIT.[32][33] This review of Treg interactions with dendritic cells provides distinction between mechanisms described for human cells versus mouse cells.[34]
  • Another control mechanism is through the IL-2 feedback loop. Antigen-activated T cells produce IL-2 which then acts on IL-2 receptors on regulatory T cells alerting them to the fact that high T cell activity is occurring in the region, and they mount a suppressory response against them. This is a negative feedback loop to ensure that overreaction is not occurring. If an actual infection is present other inflammatory factors downregulate the suppression. Disruption of the loop leads to hyperreactivity, regulation can modify the strength of the immune response.[35] A related suggestion with regard to interleukin 2 is that activated regulatory T cells take up interleukin 2 so avidly that they deprive effector T cells of sufficient to avoid apoptosis.[18]
  • A major mechanism of suppression by regulatory T cells is through the prevention of co-stimulation through CD28 on effector T cells by the action of the molecule CTLA-4.[36]

Natural and induced regulatory T cells edit

T regulatory lymphocytes develop during ontogeny either in the thymus or in the periphery. Accordingly, they are divided into natural and induced T regulatory cells.[37]

Natural T regulatory lymphocytes (tTregs, nTregs) are characterized by continuous expression of FoxP3 and T cell receptor (TCR) with relatively high autoaffinity. These cells are predominantly found in the body in the bloodstream or lymph nodes and serve mainly to confer tolerance to autoantigens.[37]

Induced (peripheral) T regulatory cells (iTregs, pTregs) arise under certain situations in the presence of IL-2 and TGF-b in the periphery and begin to express FoxP3 inducibly, thus becoming the functional equivalent of tTreg cells. iTregs, however, are found primarily in peripheral barrier tissues, where they are primarily involved in preventing inflammation in the presence of external antigens.[37]

The main features that differentiate tTreg and iTreg cells include Helios and Neuropilin-1, the presence of which suggests origin in the thymus. Another feature distinguishing these two Treg cell populations is the stability of FoxP3 expression in different settings.[37]

Induced T regulatory cells edit

Induced regulatory T (iTreg) cells (CD4+ CD25+ FOXP3+) are suppressive cells involved in tolerance. iTreg cells have been shown to suppress T cell proliferation and experimental autoimmune diseases. These cells include Treg17 cells. iTreg cells develop from mature CD4+ conventional T cells outside of the thymus: a defining distinction between natural regulatory T (nTreg) cells and iTreg cells. Though iTreg and nTreg cells share a similar function iTreg cells have recently been shown to be "an essential non-redundant regulatory subset that supplements nTreg cells, in part by expanding TCR diversity within regulatory responses".[38] Acute depletion of the iTreg cell pool in mouse models has resulted in inflammation and weight loss. The contribution of nTreg cells versus iTreg cells in maintaining tolerance is unknown, but both are important. Epigenetic differences have been observed between nTreg and iTreg cells, with the former having more stable FOXP3 expression and wider demethylation.

The small intestinal environment is high in vitamin A and is a location where retinoic acid is produced.[39] The retinoic acid and TGF-beta produced by dendritic cells within this area signal for production of regulatory T cells.[39] Vitamin A and TGF-beta promote T cell differentiation into regulatory T cells opposed to Th17 cells, even in the presence of IL-6.[40][41] The intestinal environment can lead to induced regulatory T cells with TGF-beta and retinoic acid,[42] some of which express the lectin-like receptor CD161 and are specialized to maintain barrier integrity by accelerating wound healing.[43] The Tregs within the gut are differentiated from naïve T cells after antigen is introduced.[44] It has recently been shown that human regulatory T cells can be induced from both naive and pre-committed Th1 cells and Th17 cells[45] using a parasite-derived TGF-β mimic, secreted by Heligmosomoides polygyrus and termed Hp-TGM (H. polygyrus TGF-β mimic).[46][47] Hp-TGM can induce murine FOXP3 expressing regulatory T cells that were stabile in presence of inflammation in vivo.[48] Hp-TGM-induced human FOXP3+ regulatory T cells were stable in the presence of inflammation and had increased levels of CD25, CTLA4 and decreased methylation in the FOXP3 Treg-Specific demethylated region compared to TGF-β-induced Tregs.[45]

RORγt+ regulatory T lymphocytes edit

The iTregs are able to differentiate into RORγt-expressing cells and thus acquire the phenotype of Th17 cells. These cells are associated with the functions of mucosal lymphoid tissues such as the intestinal barrier. In the intestinal lamina propria, 20-30% of Foxp3+ T regulatory cells expressing RORyt are found and this high proportion is strongly dependent on the presence of a complex gut microbiome. In germ-free (GF) mice, the population of RORγt+ T regulatory cells is strongly reduced, whereas recolonization by the specific pathogen-free (SPF) microbiota restores normal numbers of these lymphocytes in the gut. The mechanism by which the gut microbiota induces the formation of RORγt+ Treg cells involves the production of short-chain fatty acids (SCFAs), on which this induction is dependent. SCFAs are a by-product of fermentation and digestion of dietary fiber, therefore, microbial-free mice have very low concentrations of both SCFAs and RORγt Treg cells. Induction of RORγt Treg cells is also dependent on the presence of dendritic cells in adults, Thetis cells in neonatal and antigen presentation by MHC II.[49][50]

RORγt+ Treg cells are not present in the thymus and do not express Helios or Neuropilin-1, but have high expression of CD44, IL-10, ICOS, CTLA-4, and the nucleotidases CD39 and CD73, suggesting a strong regulatory function.[49]

Function of RORγt+ regulatory T lymphocytes edit

Induction of RORγt+ Treg cells in lymph nodes of the small intestine is crucial for the establishment of intestinal luminal antigen tolerance. These cells are particularly important in the prevention of food allergies. One mechanism is the production of suppressive molecules such as the cytokine IL-10. These cells also suppress the Th17 cell population and inhibit the production of IL-17, thus suppressing the pro-inflammatory response.[49]

Gata3+ regulatory T lymphocytes edit

Another important subset of Treg cells are Gata3+ Treg cells, which respond to IL-33 in the gut and influence the regulation of effector T cells during inflammation. Unlike RORγt+ Treg cells, these cells express Helios and are not dependent on the microbiome.[50][51]

Gata3+ T regs are major immunosuppressors during intestinal inflammation and T regs use Gata3 to limit tissue inflammation. This cell population also restrict Th17 T cells immunity in the intestine, because Gata3-deficient T regs express higher Rorc and IL-17a transcript.[52]

Disease edit

An important question in the field of immunology is how the immunosuppressive activity of regulatory T cells is modulated during the course of an ongoing immune response. While the immunosuppressive function of regulatory T cells prevents the development of autoimmune disease, it is not desirable during immune responses to infectious microorganisms. Current hypotheses suggest that, upon encounter with infectious microorganisms, the activity of regulatory T cells may be downregulated, either directly or indirectly, by other cells to facilitate elimination of the infection. Experimental evidence from mouse models suggests that some pathogens may have evolved to manipulate regulatory T cells to immunosuppress the host and so potentiate their own survival. For example, regulatory T cell activity has been reported to increase in several infectious contexts, such as retroviral infections (the most well-known of which is HIV), mycobacterial infections (e.g., tuberculosis[53]), and various parasitic infections including Leishmania and malaria.

Treg cells play major roles during HIV infection. They suppress the immune system, thus limiting target cells and reducing inflammation, but this simultaneously disrupts the clearance of virus by the cell-mediated immune response and enhances the reservoir by pushing CD4+ T cells to a resting state, including infected cells. Additionally, Treg cells can be infected by HIV, increasing the size of the HIV reservoir directly. Thus, Treg cells are being investigated as targets for HIV cure research.[54] Some Treg cell depletion strategies have been tested in SIV infected nonhuman primates, and shown to cause viral reactivation and enhanced SIV specific CD8+ T cell responses.[55]

Regulatory T cells have a large role in the pathology of visceral leishmaniasis and in preventing excess inflammation in patients cured of visceral leishmaniasis.

CD4+ regulatory T cells are often associated with solid tumours in both humans and murine models. Increased numbers of regulatory T cells in breast, colorectal and ovarian cancers is associated with a poorer prognosis.[56]

CD70+ non-Hodgkin lymphoma B cells induce FOXP3 expression and regulatory function in intratumoral CD4+CD25 T cells.[57]

There is some evidence that Treg cells may be dysfunctional and driving neuroinflammation in amyotrophic lateral sclerosis due to lower expression of FOXP3.[58] Ex vivo expansion of Treg cells for subsequent autologous transplant is currently being investigated after promising results were obtained in a phase I clinical trial.[59]

Additionally, while regulatory T cells have been shown to increase via polyclonal expansion both systemically and locally during healthy pregnancies to protect the fetus from the maternal immune response (a process called maternal immune tolerance), there is evidence that this polyclonal expansion is impaired in preeclamptic mothers and their offspring.[60] Research suggests reduced production and development of regulatory T cells during preeclampsia may degrade maternal immune tolerance, leading to the hyperactive immune response characteristic of preeclampsia.[61]

Cancer edit

 
The recruitment and maintenance of Treg cells in the tumor microenvironment

Most tumors elicit an immune response in the host that is mediated by tumor antigens, thus distinguishing the tumor from other non-cancerous cells. This causes large numbers of tumor-infiltrating lymphocytes (TILs) to be found in the tumor microenvironment.[62] Although it is not entirely understood, it is thought that these lymphocytes target cancerous cells and therefore slow or terminate the development of the tumor. However, this process is complicated because Treg cells seem to be preferentially trafficked to the tumor microenvironment. While Treg cells normally make up only about 4% of CD4+ T cells, they can make up as much as 20–30% of the total CD4+ population around the tumor microenvironment.[63]

Although high levels of TILs were initially thought to be important in determining an immune response against cancer, it is now widely recognized that the ratio of Treg to effector T cells in the tumor microenvironment is a determining factor in the success of the immune response against the cancer. High levels of Treg cells in the tumor microenvironment are associated with poor prognosis in many cancers,[64] such as ovarian, breast, renal, and pancreatic cancer.[63] This indicates that Treg cells suppress effector T cells and hinder the body's immune response against the cancer. However, in some types of cancer the opposite is true, and high levels of Treg cells are associated with a positive prognosis. This trend is seen in cancers such as colorectal carcinoma and follicular lymphoma. This could be due to Treg cells' ability to suppress general inflammation which is known to trigger cell proliferation and metastasis .[63] These opposite effects indicate that Treg cells' role in the development of cancer is highly dependent on both type and location of the tumor.

Although it is still not entirely understood how Treg cells are preferentially trafficked to the tumor microenvironment, the chemotaxis is probably driven by the production of chemokines by the tumor. Treg infiltration into the tumor microenvironment is facilitated by the binding of the chemokine receptor CCR4, which is expressed on Treg cells, to its ligand CCL22, which is secreted by many types of tumor cells.[65] Treg cell expansion at the site of the tumor could also explain the increased levels of Treg cells. The cytokine, TGF-β, which is commonly produced by tumor cells, is known to induce the differentiation and expansion of Treg cells.[65]

Forkhead box protein 3 (FOXP3) as a transcription factor is an essential molecular marker of Treg cells. FOXP3 polymorphism (rs3761548) might be involved in the gastric cancer progression through influencing Treg function and the secretion of immunomodulatory cytokines such as IL-10, IL-35, and TGF-β.[66]

Treg cells present in the tumor microenvironment (TME)  can be either induced Tregs or natural (thymic) Tregs which develop from naive precursors. However, tumor-associated Tregs may also originate from IL-17A+Foxp3+ Tregs which develop from Th17 cells.[67][68]

In general, the immunosuppression of the tumor microenvironment has largely contributed to the unsuccessful outcomes of many cancer immunotherapy treatments. Depletion of Treg cells in animal models has shown an increased efficacy of immunotherapy treatments, and therefore, many immunotherapy treatments are now incorporating Treg depletion.[2]

Cancer therapies targeting regulatory T lymphocytes edit

Tregs in the TME are abundantly effector Tregs which over-express immunosuppressive molecules such as CTLA-4. Anti-CTLA-4 antibodies cause depletion of Tregs and thus increase CD8+ T cells effective against the tumor. Anti-CTLA-4 antibody ipilimumab was approved for patients with advanced melanoma. Immune-checkpoint molecule PD-1 inhibits activation of both conventional T cells and Tregs and use of anti-PD-1 antibodies may lead to activation and immunosuppressive function of Tregs. Resistance to anti-PD-1-mAb treatment is probably caused by enhanced Treg cell activity. Rapid cancer progression upon PD-1 blockade is called hyperprogressive disease. Therapies targeting Treg suppression include anti-CD25 mAbs and anti-CCR4 mAbs. OX40 agonist and GITR agonists are currently being investigated.[67][69] Therapy targeting TCR signaling is also possible by blocking tyrosine kinases. For example, tyrosine-kinase inhibitor dasatinib is used for treatment of chronic myeloid leukemia and is associated with Treg inhibition.[70]

Molecular characterization edit

Similar to other T cells, regulatory T cells develop in the thymus. The latest research suggests that regulatory T cells are defined by expression of the forkhead family transcription factor FOXP3 (forkhead box p3). Expression of FOXP3 is required for regulatory T cell development and appears to control a genetic program specifying this cell's fate.[71] The large majority of Foxp3-expressing regulatory T cells are found within the major histocompatibility complex (MHC) class II restricted CD4-expressing (CD4+) population and express high levels of the interleukin-2 receptor alpha chain (CD25). In addition to the FOXP3-expressing CD4+ CD25+, there also appears to be a minor population of MHC class I restricted CD8+ FOXP3-expressing regulatory T cells. These FOXP3-expressing CD8+ T cells do not appear to be functional in healthy individuals but are induced in autoimmune disease states by T cell receptor stimulation to suppress IL-17-mediated immune responses.[72] Unlike conventional T cells, regulatory T cells do not produce IL-2 and are therefore anergic at baseline.

A number of different methods are employed in research to identify and monitor Treg cells. Originally, high expression of CD25 and CD4 surface markers was used (CD4+CD25+ cells). This is problematic as CD25 is also expressed on non-regulatory T cells in the setting of immune activation such as during an immune response to a pathogen. As defined by CD4 and CD25 expression, regulatory T cells comprise about 5–10% of the mature CD4+ T cell subpopulation in mice and humans, while about 1–2% of Treg can be measured in whole blood. The additional measurement of cellular expression of FOXP3 protein allowed a more specific analysis of Treg cells (CD4+CD25+FOXP3+ cells). However, FOXP3 is also transiently expressed in activated human effector T cells, thus complicating a correct Treg analysis using CD4, CD25 and FOXP3 as markers in humans. Therefore, the gold standard surface marker combination to defined Tregs within unactivated CD3+CD4+ T cells is high CD25 expression combined with the absent or low-level expression of the surface protein CD127 (IL-7RA). If viable cells are not required then the addition of FOXP3 to the CD25 and CD127 combination will provide further stringency. Several additional markers have been described, e.g., high levels of CTLA-4 (cytotoxic T-lymphocyte associated molecule-4) and GITR (glucocorticoid-induced TNF receptor) are also expressed on regulatory T cells, however the functional significance of this expression remains to be defined. There is a great interest in identifying cell surface markers that are uniquely and specifically expressed on all FOXP3-expressing regulatory T cells. However, to date no such molecule has been identified.

The identification of Tregs following cell activation is challenging as conventional T cells will express CD25, transiently express FOXP3 and lose CD127 expression upon activation. It has been shown that Tregs can be detected using an activation-induced marker assay by expression of CD39[73] in combination with co-expression of CD25 and OX40(CD134) which define antigen-specific cells following 24-48h stimulation with antigen.[74][75]

In addition to the search for novel protein markers, a different method to analyze and monitor Treg cells more accurately has been described in the literature. This method is based on DNA methylation analysis. Only in Treg cells, but not in any other cell type, including activated effector T cells, a certain region within the FOXP3 gene (TSDR, Treg-specific-demethylated region) is found demethylated, which allows to monitor Treg cells through a PCR reaction or other DNA-based analysis methods.[76] Interplay between the Th17 cells and regulatory T cells are important in many diseases like respiratory diseases.[77]

Recent evidence suggests that mast cells may be important mediators of Treg-dependent peripheral tolerance.[78]

Epitopes edit

Regulatory T cell epitopes ('Tregitopes') were discovered in 2008 and consist of linear sequences of amino acids contained within monoclonal antibodies and immunoglobulin G (IgG). Since their discovery, evidence has indicated Tregitopes may be crucial to the activation of natural regulatory T cells.[79][80][81]

Potential applications of regulatory T cell epitopes have been hypothesised: tolerisation to transplants, protein drugs, blood transfer therapies, and type I diabetes as well as reduction of immune response for the treatment of allergies.[82][83][84][85][86][87][81]

Genetic deficiency edit

Genetic mutations in the gene encoding FOXP3 have been identified in both humans and mice based on the heritable disease caused by these mutations. This disease provides the most striking evidence that regulatory T cells play a critical role in maintaining normal immune system function. Humans with mutations in FOXP3 develop a severe and rapidly fatal autoimmune disorder known as Immune dysregulation, Polyendocrinopathy, Enteropathy X-linked (IPEX) syndrome.[88][89]

The IPEX syndrome is characterized by the development of overwhelming systemic autoimmunity in the first year of life, resulting in the commonly observed triad of watery diarrhea, eczematous dermatitis, and endocrinopathy seen most commonly as insulin-dependent diabetes mellitus. Most individuals have other autoimmune phenomena including Coombs-positive hemolytic anemia, autoimmune thrombocytopenia, autoimmune neutropenia, and tubular nephropathy. The majority of affected males die within the first year of life of either metabolic derangements or sepsis. An analogous disease is also observed in a spontaneous FOXP3-mutant mouse known as "scurfy".

See also edit

References edit

  1. ^ Bettelli E, Carrier Y, Gao W, Korn T, Strom TB, Oukka M, et al. (May 2006). "Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells". Nature. 441 (7090): 235–8. Bibcode:2006Natur.441..235B. doi:10.1038/nature04753. PMID 16648838. S2CID 4391497.
  2. ^ a b Curiel TJ (May 2007). "Tregs and rethinking cancer immunotherapy". The Journal of Clinical Investigation. 117 (5): 1167–74. doi:10.1172/JCI31202. PMC 1857250. PMID 17476346.
  3. ^ Chen W (August 2011). "Tregs in immunotherapy: opportunities and challenges". Immunotherapy. 3 (8): 911–4. doi:10.2217/imt.11.79. PMID 21843075.
  4. ^ Miyara M, Gorochov G, Ehrenstein M, Musset L, Sakaguchi S, Amoura Z (October 2011). "Human FoxP3+ regulatory T cells in systemic autoimmune diseases". Autoimmunity Reviews. 10 (12): 744–55. doi:10.1016/j.autrev.2011.05.004. PMID 21621000.
  5. ^ Nosbaum A, Prevel N, Truong HA, Mehta P, Ettinger M, Scharschmidt TC, et al. (March 2016). "Cutting Edge: Regulatory T Cells Facilitate Cutaneous Wound Healing". Journal of Immunology. 196 (5): 2010–4. doi:10.4049/jimmunol.1502139. PMC 4761457. PMID 26826250.
  6. ^ Adeegbe DO, Nishikawa H (2013). "Natural and induced T regulatory cells in cancer". Frontiers in Immunology. 4: 190. doi:10.3389/fimmu.2013.00190. PMC 3708155. PMID 23874336.
  7. ^ Curiel TJ (April 2008). "Regulatory T cells and treatment of cancer". Current Opinion in Immunology. 20 (2): 241–6. doi:10.1016/j.coi.2008.04.008. PMC 3319305. PMID 18508251.
  8. ^ Hori S, Nomura T, Sakaguchi S (February 2003). "Control of regulatory T cell development by the transcription factor Foxp3". Science. 299 (5609): 1057–61. Bibcode:2003Sci...299.1057H. doi:10.1126/science.1079490. PMID 12522256. S2CID 9697928.
  9. ^ Singh B, Schwartz JA, Sandrock C, Bellemore SM, Nikoopour E (November 2013). "Modulation of autoimmune diseases by interleukin (IL)-17 producing regulatory T helper (Th17) cells". The Indian Journal of Medical Research. 138 (5): 591–4. PMC 3928692. PMID 24434314.
  10. ^ Shevach EM (2000). "Regulatory T cells in autoimmmunity*". Annual Review of Immunology. 18: 423–49. doi:10.1146/annurev.immunol.18.1.423. PMID 10837065. S2CID 15160752.
  11. ^ Schmetterer KG, Neunkirchner A, Pickl WF (June 2012). "Naturally occurring regulatory T cells: markers, mechanisms, and manipulation". FASEB Journal. 26 (6): 2253–76. doi:10.1096/fj.11-193672. PMID 22362896. S2CID 36277557.
  12. ^ Sakaguchi S (2004). "Naturally arising CD4+ regulatory t cells for immunologic self-tolerance and negative control of immune responses". Annual Review of Immunology. 22: 531–62. doi:10.1146/annurev.immunol.21.120601.141122. PMID 15032588.
  13. ^ Li MO, Rudensky AY (April 2016). "T cell receptor signalling in the control of regulatory T cell differentiation and function". Nature Reviews. Immunology. 16 (4): 220–33. doi:10.1038/nri.2016.26. PMC 4968889. PMID 27026074.
  14. ^ a b c d Santamaria, Jérémy C.; Borelli, Alexia; Irla, Magali (2021-02-11). "Regulatory T Cell Heterogeneity in the Thymus: Impact on Their Functional Activities". Frontiers in Immunology. 12: 643153. doi:10.3389/fimmu.2021.643153. ISSN 1664-3224. PMC 7904894. PMID 33643324.
  15. ^ Owen, David L.; Sjaastad, Louisa E.; Farrar, Michael A. (2019-10-15). "Regulatory T Cell Development in the Thymus". The Journal of Immunology. 203 (8): 2031–2041. doi:10.4049/jimmunol.1900662. ISSN 0022-1767. PMC 6910132. PMID 31591259.
  16. ^ Owen, David L.; Mahmud, Shawn A.; Sjaastad, Louisa E.; Williams, Jason B.; Spanier, Justin A.; Simeonov, Dimitre R.; Ruscher, Roland; Huang, Weishan; Proekt, Irina; Miller, Corey N.; Hekim, Can; Jeschke, Jonathan C.; Aggarwal, Praful; Broeckel, Ulrich; LaRue, Rebecca S. (February 2019). "Thymic regulatory T cells arise via two distinct developmental programs". Nature Immunology. 20 (2): 195–205. doi:10.1038/s41590-018-0289-6. ISSN 1529-2916. PMC 6650268. PMID 30643267.
  17. ^ a b c d Thiault N, Darrigues J, Adoue V, Gros M, Binet B, Perals C, et al. (June 2015). "Peripheral regulatory T lymphocytes recirculating to the thymus suppress the development of their precursors". Nature Immunology. 16 (6): 628–34. doi:10.1038/ni.3150. PMID 25939024. S2CID 7670443.
  18. ^ a b Pandiyan P, Zheng L, Ishihara S, Reed J, Lenardo MJ (December 2007). "CD4+CD25+Foxp3+ regulatory T cells induce cytokine deprivation-mediated apoptosis of effector CD4+ T cells". Nature Immunology. 8 (12): 1353–62. doi:10.1038/ni1536. PMID 17982458. S2CID 8925488.
  19. ^ Cheng G, Yu A, Malek TR (May 2011). "T-cell tolerance and the multi-functional role of IL-2R signaling in T-regulatory cells". Immunological Reviews. 241 (1): 63–76. doi:10.1111/j.1600-065X.2011.01004.x. PMC 3101713. PMID 21488890.
  20. ^ Kimmig S, Przybylski GK, Schmidt CA, Laurisch K, Möwes B, Radbruch A, Thiel A (March 2002). "Two subsets of naive T helper cells with distinct T cell receptor excision circle content in human adult peripheral blood". The Journal of Experimental Medicine. 195 (6): 789–94. doi:10.1084/jem.20011756. PMC 2193736. PMID 11901204.
  21. ^ Toker A, Engelbert D, Garg G, Polansky JK, Floess S, Miyao T, et al. (April 2013). "Active demethylation of the Foxp3 locus leads to the generation of stable regulatory T cells within the thymus". Journal of Immunology. 190 (7): 3180–8. doi:10.4049/jimmunol.1203473. PMID 23420886.
  22. ^ a b c d Nikolouli E, Elfaki Y, Herppich S, Schelmbauer C, Delacher M, Falk C, et al. (January 2021). "Recirculating IL-1R2+ Treg fine-tune intrathymic Treg development under inflammatory conditions". Cellular & Molecular Immunology. 18 (1): 182–193. doi:10.1038/s41423-019-0352-8. hdl:10033/622148. PMC 7853075. PMID 31988493. S2CID 210913733.
  23. ^ Peters VA, Joesting JJ, Freund GG (August 2013). "IL-1 receptor 2 (IL-1R2) and its role in immune regulation". Brain, Behavior, and Immunity. 32: 1–8. doi:10.1016/j.bbi.2012.11.006. PMC 3610842. PMID 23195532.
  24. ^ Read S, Malmström V, Powrie F (July 2000). "Cytotoxic T lymphocyte-associated antigen 4 plays an essential role in the function of CD25(+)CD4(+) regulatory cells that control intestinal inflammation". The Journal of Experimental Medicine. 192 (2): 295–302. doi:10.1084/jem.192.2.295. PMC 2193261. PMID 10899916.
  25. ^ Collison LW, Workman CJ, Kuo TT, Boyd K, Wang Y, Vignali KM, et al. (November 2007). "The inhibitory cytokine IL-35 contributes to regulatory T-cell function". Nature. 450 (7169): 566–9. Bibcode:2007Natur.450..566C. doi:10.1038/nature06306. PMID 18033300. S2CID 4425281.
  26. ^ Annacker O, Asseman C, Read S, Powrie F (June 2003). "Interleukin-10 in the regulation of T cell-induced colitis". Journal of Autoimmunity. 20 (4): 277–9. doi:10.1016/s0896-8411(03)00045-3. PMID 12791312.
  27. ^ Kearley J, Barker JE, Robinson DS, Lloyd CM (December 2005). "Resolution of airway inflammation and hyperreactivity after in vivo transfer of CD4+CD25+ regulatory T cells is interleukin 10 dependent". The Journal of Experimental Medicine. 202 (11): 1539–47. doi:10.1084/jem.20051166. PMC 1350743. PMID 16314435.
  28. ^ Gondek DC, Lu LF, Quezada SA, Sakaguchi S, Noelle RJ (February 2005). "Cutting edge: contact-mediated suppression by CD4+CD25+ regulatory cells involves a granzyme B-dependent, perforin-independent mechanism". Journal of Immunology. 174 (4): 1783–6. doi:10.4049/jimmunol.174.4.1783. PMID 15699103.
  29. ^ Puccetti P, Grohmann U (October 2007). "IDO and regulatory T cells: a role for reverse signalling and non-canonical NF-kappaB activation". Nature Reviews. Immunology. 7 (10): 817–23. doi:10.1038/nri2163. PMID 17767193. S2CID 5544429.
  30. ^ Borsellino G, Kleinewietfeld M, Di Mitri D, Sternjak A, Diamantini A, Giometto R, et al. (August 2007). "Expression of ectonucleotidase CD39 by Foxp3+ Treg cells: hydrolysis of extracellular ATP and immune suppression". Blood. 110 (4): 1225–32. doi:10.1182/blood-2006-12-064527. PMID 17449799.
  31. ^ Kobie JJ, Shah PR, Yang L, Rebhahn JA, Fowell DJ, Mosmann TR (November 2006). "T regulatory and primed uncommitted CD4 T cells express CD73, which suppresses effector CD4 T cells by converting 5'-adenosine monophosphate to adenosine". Journal of Immunology. 177 (10): 6780–6. doi:10.4049/jimmunol.177.10.6780. PMID 17082591.
  32. ^ Huang CT, Workman CJ, Flies D, Pan X, Marson AL, Zhou G, et al. (October 2004). "Role of LAG-3 in regulatory T cells". Immunity. 21 (4): 503–13. doi:10.1016/j.immuni.2004.08.010. PMID 15485628.
  33. ^ Yu X, Harden K, Gonzalez LC, Francesco M, Chiang E, Irving B, et al. (January 2009). "The surface protein TIGIT suppresses T cell activation by promoting the generation of mature immunoregulatory dendritic cells". Nature Immunology. 10 (1): 48–57. doi:10.1038/ni.1674. PMID 19011627. S2CID 205361984.
  34. ^ Wardell CM, MacDonald KN, Levings MK, Cook L (January 2021). "Cross talk between human regulatory T cells and antigen-presenting cells: Lessons for clinical applications". European Journal of Immunology. 51 (1): 27–38. doi:10.1002/eji.202048746. PMID 33301176.
  35. ^ Sakaguchi S, Yamaguchi T, Nomura T, Ono M (May 2008). "Regulatory T cells and immune tolerance". Cell. 133 (5): 775–87. doi:10.1016/j.cell.2008.05.009. PMID 18510923.
  36. ^ Walker LS, Sansom DM (November 2011). "The emerging role of CTLA4 as a cell-extrinsic regulator of T cell responses". Nature Reviews. Immunology. 11 (12): 852–63. doi:10.1038/nri3108. PMID 22116087. S2CID 9617595.
  37. ^ a b c d Shevyrev, Daniil; Tereshchenko, Valeriy (2020). "Treg Heterogeneity, Function, and Homeostasis". Frontiers in Immunology. 10: 3100. doi:10.3389/fimmu.2019.03100. ISSN 1664-3224. PMC 6971100. PMID 31993063.
  38. ^ Haribhai D, Williams JB, Jia S, Nickerson D, Schmitt EG, Edwards B, et al. (July 2011). "A requisite role for induced regulatory T cells in tolerance based on expanding antigen receptor diversity". Immunity. 35 (1): 109–22. doi:10.1016/j.immuni.2011.03.029. PMC 3295638. PMID 21723159.
  39. ^ a b Sun CM, Hall JA, Blank RB, Bouladoux N, Oukka M, Mora JR, Belkaid Y (August 2007). "Small intestine lamina propria dendritic cells promote de novo generation of Foxp3 T reg cells via retinoic acid". The Journal of Experimental Medicine. 204 (8): 1775–85. doi:10.1084/jem.20070602. PMC 2118682. PMID 17620362.
  40. ^ Mucida D, Park Y, Kim G, Turovskaya O, Scott I, Kronenberg M, Cheroutre H (July 2007). "Reciprocal TH17 and regulatory T cell differentiation mediated by retinoic acid". Science. 317 (5835): 256–60. Bibcode:2007Sci...317..256M. doi:10.1126/science.1145697. PMID 17569825. S2CID 24736012.
  41. ^ Erkelens MN, Mebius RE (March 2017). "Retinoic Acid and Immune Homeostasis: A Balancing Act". Trends in Immunology. 38 (3): 168–180. doi:10.1016/j.it.2016.12.006. PMID 28094101.
  42. ^ Ziegler SF, Buckner JH (April 2009). "FOXP3 and the regulation of Treg/Th17 differentiation". Microbes and Infection. 11 (5): 594–8. doi:10.1016/j.micinf.2009.04.002. PMC 2728495. PMID 19371792.
  43. ^ Povoleri GA, Nova-Lamperti E, Scottà C, Fanelli G, Chen YC, Becker PD, et al. (December 2018). "Human retinoic acid-regulated CD161+ regulatory T cells support wound repair in intestinal mucosa". Nature Immunology. 19 (12): 1403–1414. doi:10.1038/s41590-018-0230-z. PMC 6474659. PMID 30397350.
  44. ^ Coombes JL, Siddiqui KR, Arancibia-Cárcamo CV, Hall J, Sun CM, Belkaid Y, Powrie F (August 2007). "A functionally specialized population of mucosal CD103+ DCs induces Foxp3+ regulatory T cells via a TGF-beta and retinoic acid-dependent mechanism". The Journal of Experimental Medicine. 204 (8): 1757–64. doi:10.1084/jem.20070590. PMC 2118683. PMID 17620361.
  45. ^ a b Cook L, Reid KT, Häkkinen E, de Bie B, Tanaka S, Smyth DJ, et al. (September 2021). "Induction of stable human FOXP3+ Tregs by a parasite-derived TGF-β mimic". Immunology and Cell Biology. 99 (8): 833–847. doi:10.1111/IMCB.12475. PMC 8453874. PMID 33929751.
  46. ^ Johnston CJ, Smyth DJ, Kodali RB, White MP, Harcus Y, Filbey KJ, et al. (November 2017). "A structurally distinct TGF-β mimic from an intestinal helminth parasite potently induces regulatory T cells". Nature Communications. 8 (1): 1741. Bibcode:2017NatCo...8.1741J. doi:10.1038/s41467-017-01886-6. PMC 5701006. PMID 29170498.
  47. ^ Smyth DJ, Harcus Y, White MP, Gregory WF, Nahler J, Stephens I, et al. (April 2018). "TGF-β mimic proteins form an extended gene family in the murine parasite Heligmosomoides polygyrus". International Journal for Parasitology. 48 (5): 379–385. doi:10.1016/j.ijpara.2017.12.004. PMC 5904571. PMID 29510118.
  48. ^ White MP, Smyth DJ, Cook L, Ziegler SF, Levings MK, Maizels RM (September 2021). "The parasite cytokine mimic Hp-TGM potently replicates the regulatory effects of TGF-β on murine CD4+ T cells". Immunology and Cell Biology. 99 (8): 848–864. doi:10.1111/IMCB.12479. PMC 9214624. PMID 33988885.
  49. ^ a b c Ning, Xixi; Lei, Zengjie; Rui, Binqi; Li, Yuyuan; Li, Ming (2022-12-05). "Gut Microbiota Promotes Immune Tolerance by Regulating RORγt+ Treg Cells in Food Allergy". Advanced Gut & Microbiome Research. 2022: e8529578. doi:10.1155/2022/8529578.
  50. ^ a b Ohnmacht, Caspar; Park, Joo-Hong; Cording, Sascha; Wing, James B.; Atarashi, Koji; Obata, Yuuki; Gaboriau-Routhiau, Valérie; Marques, Rute; Dulauroy, Sophie; Fedoseeva, Maria; Busslinger, Meinrad; Cerf-Bensussan, Nadine; Boneca, Ivo G.; Voehringer, David; Hase, Koji (2015-08-28). "The microbiota regulates type 2 immunity through RORγt + T cells". Science. 349 (6251): 989–993. Bibcode:2015Sci...349..989O. doi:10.1126/science.aac4263. ISSN 0036-8075. PMID 26160380. S2CID 2663636.
  51. ^ Jacobse, Justin; Li, Jing; Rings, Edmond H. H. M.; Samsom, Janneke N.; Goettel, Jeremy A. (2021). "Intestinal Regulatory T Cells as Specialized Tissue-Restricted Immune Cells in Intestinal Immune Homeostasis and Disease". Frontiers in Immunology. 12: 716499. doi:10.3389/fimmu.2021.716499. ISSN 1664-3224. PMC 8371910. PMID 34421921.
  52. ^ Lui, Prudence PokWai; Cho, Inchul; Ali, Niwa (September 2020). "Tissue regulatory T cells". Immunology. 161 (1): 4–17. doi:10.1111/imm.13208. ISSN 0019-2805. PMC 7450170. PMID 32463116.
  53. ^ Stringari LL, Covre LP, da Silva FD, de Oliveira VL, Campana MC, Hadad DJ, et al. (July 2021). "Increase of CD4+CD25highFoxP3+ cells impairs in vitro human microbicidal activity against Mycobacterium tuberculosis during latent and acute pulmonary tuberculosis". PLOS Neglected Tropical Diseases. 15 (7): e0009605. doi:10.1371/journal.pntd.0009605. PMC 8321116. PMID 34324509.
  54. ^ Kleinman AJ, Sivanandham R, Pandrea I, Chougnet CA, Apetrei C (2018). "Regulatory T Cells As Potential Targets for HIV Cure Research". Frontiers in Immunology. 9: 734. doi:10.3389/fimmu.2018.00734. PMC 5908895. PMID 29706961.
  55. ^ Sivanandham R, Kleinman AJ, Sette P, Brocca-Cofano E, Kilapandal Venkatraman SM, Policicchio BB, et al. (September 2020). "Nonhuman Primate Testing of the Impact of Different Regulatory T Cell Depletion Strategies on Reactivation and Clearance of Latent Simian Immunodeficiency Virus". Journal of Virology. 94 (19): JVI.00533–20, jvi, JVI.00533–20v1. doi:10.1128/JVI.00533-20. PMC 7495362. PMID 32669326. S2CID 220579402.
  56. ^ Dranoff G (December 2005). "The therapeutic implications of intratumoral regulatory T cells". Clinical Cancer Research. 11 (23): 8226–9. doi:10.1158/1078-0432.CCR-05-2035. PMID 16322278. S2CID 18794337.
  57. ^ Yang ZZ, Novak AJ, Ziesmer SC, Witzig TE, Ansell SM (October 2007). "CD70+ non-Hodgkin lymphoma B cells induce Foxp3 expression and regulatory function in intratumoral CD4+CD25 T cells". Blood. 110 (7): 2537–44. doi:10.1182/blood-2007-03-082578. PMC 1988926. PMID 17615291.
  58. ^ Beers DR, Zhao W, Wang J, Zhang X, Wen S, Neal D, et al. (March 2017). "ALS patients' regulatory T lymphocytes are dysfunctional, and correlate with disease progression rate and severity". JCI Insight. 2 (5): e89530. doi:10.1172/jci.insight.89530. PMC 5333967. PMID 28289705.
  59. ^ Thonhoff JR, Beers DR, Zhao W, Pleitez M, Simpson EP, Berry JD, et al. (July 2018). "Expanded autologous regulatory T-lymphocyte infusions in ALS: A phase I, first-in-human study". Neurology. 5 (4): e465. doi:10.1212/NXI.0000000000000465. PMC 5961523. PMID 29845093.
  60. ^ Tsuda S, Nakashima A, Shima T, Saito S (2019). "New Paradigm in the Role of Regulatory T Cells During Pregnancy". Frontiers in Immunology. 10: 573. doi:10.3389/fimmu.2019.00573. PMC 6443934. PMID 30972068.
  61. ^ Hu M, Eviston D, Hsu P, Mariño E, Chidgey A, Santner-Nanan B, et al. (July 2019). "Decreased maternal serum acetate and impaired fetal thymic and regulatory T cell development in preeclampsia". Nature Communications. 10 (1): 3031. Bibcode:2019NatCo..10.3031H. doi:10.1038/s41467-019-10703-1. PMC 6620275. PMID 31292453.
  62. ^ Gooden MJ, de Bock GH, Leffers N, Daemen T, Nijman HW (June 2011). "The prognostic influence of tumour-infiltrating lymphocytes in cancer: a systematic review with meta-analysis". British Journal of Cancer. 105 (1): 93–103. doi:10.1038/bjc.2011.189. PMC 3137407. PMID 21629244.
  63. ^ a b c Oleinika K, Nibbs RJ, Graham GJ, Fraser AR (January 2013). "Suppression, subversion and escape: the role of regulatory T cells in cancer progression". Clinical and Experimental Immunology. 171 (1): 36–45. doi:10.1111/j.1365-2249.2012.04657.x. PMC 3530093. PMID 23199321.
  64. ^ Plitas G, Rudensky AY (2020-03-09). "Regulatory T Cells in Cancer". Annual Review of Cancer Biology. 4 (1): 459–477. doi:10.1146/annurev-cancerbio-030419-033428. ISSN 2472-3428.
  65. ^ a b Lippitz BE (May 2013). "Cytokine patterns in patients with cancer: a systematic review". The Lancet. Oncology. 14 (6): e218-28. doi:10.1016/s1470-2045(12)70582-x. PMID 23639322.
  66. ^ Ezzeddini R, Somi MH, Taghikhani M, Moaddab SY, Masnadi Shirazi K, Shirmohammadi M, Eftekharsadat AT, Sadighi Moghaddam B, Salek Farrokhi A (February 2021). "Association of Foxp3 rs3761548 polymorphism with cytokines concentration in gastric adenocarcinoma patients". Cytokine. 138: 155351. doi:10.1016/j.cyto.2020.155351. ISSN 1043-4666. PMID 33127257. S2CID 226218796.
  67. ^ a b Li, Chunxiao; Jiang, Ping; Wei, Shuhua; Xu, Xiaofei; Wang, Junjie (2020-07-17). "Regulatory T cells in tumor microenvironment: new mechanisms, potential therapeutic strategies and future prospects". Molecular Cancer. 19 (1): 116. doi:10.1186/s12943-020-01234-1. ISSN 1476-4598. PMC 7367382. PMID 32680511.
  68. ^ Downs-Canner, Stephanie; Berkey, Sara; Delgoffe, Greg M.; Edwards, Robert P.; Curiel, Tyler; Odunsi, Kunle; Bartlett, David L.; Obermajer, Nataša (2017-03-14). "Suppressive IL-17A+Foxp3+ and ex-Th17 IL-17AnegFoxp3+ Treg cells are a source of tumour-associated Treg cells". Nature Communications. 8 (1): 14649. Bibcode:2017NatCo...814649D. doi:10.1038/ncomms14649. ISSN 2041-1723. PMC 5355894. PMID 28290453.
  69. ^ Togashi, Yosuke; Shitara, Kohei; Nishikawa, Hiroyoshi (June 2019). "Regulatory T cells in cancer immunosuppression — implications for anticancer therapy". Nature Reviews Clinical Oncology. 16 (6): 356–371. doi:10.1038/s41571-019-0175-7. ISSN 1759-4782. PMID 30705439. S2CID 59526013.
  70. ^ Ohue, Yoshihiro; Nishikawa, Hiroyoshi (July 2019). "Regulatory T (Treg) cells in cancer: Can Treg cells be a new therapeutic target?". Cancer Science. 110 (7): 2080–2089. doi:10.1111/cas.14069. ISSN 1347-9032. PMC 6609813. PMID 31102428.
  71. ^ Marson A, Kretschmer K, Frampton GM, Jacobsen ES, Polansky JK, MacIsaac KD, et al. (February 2007). "Foxp3 occupancy and regulation of key target genes during T-cell stimulation". Nature. 445 (7130): 931–5. Bibcode:2007Natur.445..931M. doi:10.1038/nature05478. PMC 3008159. PMID 17237765.
  72. ^ Ellis SD, McGovern JL, van Maurik A, Howe D, Ehrenstein MR, Notley CA (October 2014). "Induced CD8+FoxP3+ Treg cells in rheumatoid arthritis are modulated by p38 phosphorylation and monocytes expressing membrane tumor necrosis factor α and CD86". Arthritis & Rheumatology. 66 (10): 2694–705. doi:10.1002/art.38761. PMID 24980778. S2CID 39984435.
  73. ^ Seddiki N, Cook L, Hsu DC, Phetsouphanh C, Brown K, Xu Y, et al. (June 2014). "Human antigen-specific CD4⁺ CD25⁺ CD134⁺ CD39⁺ T cells are enriched for regulatory T cells and comprise a substantial proportion of recall responses". European Journal of Immunology. 44 (6): 1644–61. doi:10.1002/eji.201344102. PMID 24752698. S2CID 24012204.
  74. ^ Zaunders JJ, Munier ML, Seddiki N, Pett S, Ip S, Bailey M, et al. (August 2009). "High levels of human antigen-specific CD4+ T cells in peripheral blood revealed by stimulated coexpression of CD25 and CD134 (OX40)". Journal of Immunology. 183 (4): 2827–36. doi:10.4049/jimmunol.0803548. PMID 19635903.
  75. ^ Poloni, Chad; Schonhofer, Cole; Ivison, Sabine; Levings, Megan K.; Steiner, Theodore S.; Cook, Laura (2023-02-24). "T-cell activation-induced marker assays in health and disease". Immunology and Cell Biology. 101 (6): 491–503. doi:10.1111/imcb.12636. ISSN 1440-1711. PMID 36825901. S2CID 257152898.
  76. ^ Wieczorek G, Asemissen A, Model F, Turbachova I, Floess S, Liebenberg V, et al. (January 2009). "Quantitative DNA methylation analysis of FOXP3 as a new method for counting regulatory T cells in peripheral blood and solid tissue". Cancer Research. 69 (2): 599–608. doi:10.1158/0008-5472.CAN-08-2361. PMID 19147574.
  77. ^ Agarwal A, Singh M, Chatterjee BP, Chauhan A, Chakraborti A (2014). "Interplay of T Helper 17 Cells with CD4(+)CD25(high) FOXP3(+) Tregs in Regulation of Allergic Asthma in Pediatric Patients". International Journal of Pediatrics. 2014: 636238. doi:10.1155/2014/636238. PMC 4065696. PMID 24995020.
  78. ^ Lu LF, Lind EF, Gondek DC, Bennett KA, Gleeson MW, Pino-Lagos K, et al. (August 2006). "Mast cells are essential intermediaries in regulatory T-cell tolerance". Nature. 442 (7106): 997–1002. Bibcode:2006Natur.442..997L. doi:10.1038/nature05010. PMID 16921386. S2CID 686654.
  79. ^ "Tregitope: Immunomodulation Power Tool". EpiVax. 2 August 2016.
  80. ^ Hui DJ, Basner-Tschakarjan E, Chen Y, Davidson RJ, Buchlis G, Yazicioglu M, et al. (September 2013). "Modulation of CD8+ T cell responses to AAV vectors with IgG-derived MHC class II epitopes". Molecular Therapy. 21 (9): 1727–37. doi:10.1038/mt.2013.166. PMC 3776637. PMID 23857231.
  81. ^ a b De Groot AS, Moise L, McMurry JA, Wambre E, Van Overtvelt L, Moingeon P, et al. (October 2008). "Activation of natural regulatory T cells by IgG Fc-derived peptide "Tregitopes"". Blood. 112 (8): 3303–11. doi:10.1182/blood-2008-02-138073. PMC 2569179. PMID 18660382.
  82. ^ "New $2.25M infusion of NIH funds for EpiVax' Tregitope, proposed "Paradigm-Shifting" Treatment". Fierce Biotech Research.
  83. ^ Su Y, Rossi R, De Groot AS, Scott DW (August 2013). "Regulatory T cell epitopes (Tregitopes) in IgG induce tolerance in vivo and lack immunogenicity per se". Journal of Leukocyte Biology. 94 (2): 377–83. doi:10.1189/jlb.0912441. PMC 3714563. PMID 23729499.
  84. ^ Cousens LP, Su Y, McClaine E, Li X, Terry F, Smith R, et al. (2013). "Application of IgG-derived natural Treg epitopes (IgG Tregitopes) to antigen-specific tolerance induction in a murine model of type 1 diabetes". Journal of Diabetes Research. 2013: 621693. doi:10.1155/2013/621693. PMC 3655598. PMID 23710469.
  85. ^ Cousens LP, Mingozzi F, van der Marel S, Su Y, Garman R, Ferreira V, et al. (October 2012). "Teaching tolerance: New approaches to enzyme replacement therapy for Pompe disease". Human Vaccines & Immunotherapeutics. 8 (10): 1459–64. doi:10.4161/hv.21405. PMC 3660767. PMID 23095864.
  86. ^ Cousens LP, Najafian N, Mingozzi F, Elyaman W, Mazer B, Moise L, et al. (January 2013). "In vitro and in vivo studies of IgG-derived Treg epitopes (Tregitopes): a promising new tool for tolerance induction and treatment of autoimmunity". Journal of Clinical Immunology. 33 (1): S43-9. doi:10.1007/s10875-012-9762-4. PMC 3538121. PMID 22941509.
  87. ^ Elyaman W, Khoury SJ, Scott DW, De Groot AS (2011). "Potential application of tregitopes as immunomodulating agents in multiple sclerosis". Neurology Research International. 2011: 256460. doi:10.1155/2011/256460. PMC 3175387. PMID 21941651.
  88. ^ Online Mendelian Inheritance in Man IPEX
  89. ^ ipex at NIH/UW GeneTests

External links edit

regulatory, cell, regulatory, cells, tregs, treg, cells, formerly, known, suppressor, cells, subpopulation, cells, that, modulate, immune, system, maintain, tolerance, self, antigens, prevent, autoimmune, disease, treg, cells, immunosuppressive, generally, sup. The regulatory T cells Tregs ˈ t iː r ɛ ɡ or Treg cells formerly known as suppressor T cells are a subpopulation of T cells that modulate the immune system maintain tolerance to self antigens and prevent autoimmune disease Treg cells are immunosuppressive and generally suppress or downregulate induction and proliferation of effector T cells 1 Treg cells express the biomarkers CD4 FOXP3 and CD25 and are thought to be derived from the same lineage as naive CD4 cells 2 Because effector T cells also express CD4 and CD25 Treg cells are very difficult to effectively discern from effector CD4 making them difficult to study Research has found that the cytokine transforming growth factor beta TGF b is essential for Treg cells to differentiate from naive CD4 cells and is important in maintaining Treg cell homeostasis 3 Mouse models have suggested that modulation of Treg cells can treat autoimmune disease and cancer and can facilitate organ transplantation 4 and wound healing 5 Their implications for cancer are complicated Treg cells tend to be upregulated in individuals with cancer and they seem to be recruited to the site of many tumors Studies in both humans and animal models have implicated that high numbers of Treg cells in the tumor microenvironment is indicative of a poor prognosis and Treg cells are thought to suppress tumor immunity thus hindering the body s innate ability to control the growth of cancerous cells 6 Immunotherapy research is studying how regulation of T cells could possibly be utilized in the treatment of cancer 7 Contents 1 Populations 2 Development 2 1 Thymic recirculation 3 Function 4 Natural and induced regulatory T cells 4 1 Induced T regulatory cells 4 2 RORgt regulatory T lymphocytes 4 2 1 Function of RORgt regulatory T lymphocytes 4 3 Gata3 regulatory T lymphocytes 5 Disease 5 1 Cancer 5 1 1 Cancer therapies targeting regulatory T lymphocytes 6 Molecular characterization 6 1 Epitopes 7 Genetic deficiency 8 See also 9 References 10 External linksPopulations editT regulatory cells are a component of the immune system that suppress immune responses of other cells This is an important self check built into the immune system to prevent excessive reactions Regulatory T cells come in many forms with the most well understood being those that express CD4 CD25 and FOXP3 CD4 CD25 regulatory T cells These Treg cells are different from helper T cells 8 Another regulatory T cell subset is Treg17 cells 9 Regulatory T cells are involved in shutting down immune responses after they have successfully eliminated invading organisms and also in preventing autoimmunity 10 CD4 FOXP3 CD25 high regulatory T cells have been called naturally occurring regulatory T cells 11 to distinguish them from suppressor T cell populations that are generated in vitro Additional regulatory T cell populations include Tr1 Th3 CD8 CD28 and Qa 1 restricted T cells The contribution of these populations to self tolerance and immune homeostasis is less well defined FOXP3 can be used as a good marker for mouse CD4 CD25 T cells although recent studies have also shown evidence for FOXP3 expression in CD4 CD25 T cells In humans FOXP3 is also expressed by recently activated conventional T cells and thus does not specifically identify human Tregs 12 Development editAll T cells derive from progenitor cells in the bone marrow which become committed to their lineage in the thymus All T cells begin as CD4 CD8 TCR cells at the DN double negative stage where an individual cell will rearrange its T cell receptor genes to form a unique functional molecule which they in turn test against cells in the thymic cortex for a minimal level of interaction with self MHC If they receive these signals they proliferate and express both CD4 and CD8 becoming double positive cells The selection of Tregs occurs on radio resistant hematopoietically derived MHC class II expressing cells in the medulla or Hassall s corpuscles in the thymus At the DP double positive stage they are selected by their interaction with the cells within the thymus begin the transcription of Foxp3 and become Treg cells although they may not begin to express Foxp3 until the single positive stage at which point they are functional Tregs Tregs do not have the limited TCR expression of NKT or gd T cells Tregs have a larger TCR diversity than effector T cells biased towards self peptides The process of Treg selection is determined by the affinity of interaction with the self peptide MHC complex Selection to become a Treg is a Goldilocks process i e not too high not too low but just right 13 a T cell that receives very strong signals will undergo apoptotic death a cell that receives a weak signal will survive and be selected to become an effector cell If a T cell receives an intermediate signal then it will become a regulatory cell Due to the stochastic nature of the process of T cell activation all T cell populations with a given TCR will end up with a mixture of Teff and Treg the relative proportions determined by the affinities of the T cell for the self peptide MHC Even in mouse models with TCR transgenic cells selected on specific antigen secreting stroma deletion or conversion is not complete After interaction with self peptide MHC complex T cell has to upregulate IL 2R CD25 and TNFR superfamily members GITR OX40 and TNFR2 to become CD25 FOXP3 Treg cell progenitor To become mature Treg FOXP3 transcription factor has to be upregulated which is driven by g chain CD132 dependent cytokines in particular IL 2 and or IL 15 14 15 Only IL 2 alone is not sufficient to stimulate Foxp3 expression other cytokines are needed Whereas IL 2 is produced by self reactive thymocytes IL 15 is produced by stromal cells of the thymus mainly mTECs and cTECs 14 Recently other subset of Treg precursors was identified This subset lacks CD25 and has low expression of Foxp3 Its development is mainly dependent on IL 15 This subset has lower affinity for self antigens than CD25 Foxp3high subset Both subsets generate mature Treg cells after stimulation with IL 2 with comparable efficiency both in vitro and in vivo CD25 Foxp3high progenitors exhibit increased apoptosis and develop into mature Treg cells with faster kinetics than Foxp3low progenitors 16 Tregs derived from CD25 Foxp3high progenitors protect from experimental auto immune encephalomyelitis whereas those derived from CD25 Foxp3low progenitors protect from T cell induced colitis 14 Mature CD25 Foxp3 Tregs can be also divided into two different subsets based on the expression level of CD25 GITR and PD 1 Tregs expressing low amounts of CD25 GITR and PD 1 limit the development of colitis by promoting the conversion of conventional CD4 T cells into pTreg Tregs highly expressing CD25 GITR and PD 1 are more self reactive and control lymphoproliferation in peripheral lymph nodes they may have the ability to protect against autoimmune disorders 14 Foxp3 Treg generation in the thymus is delayed by several days compared to Teff cells and does not reach adult levels in either the thymus or periphery until around three weeks post partum Treg cells require CD28 co stimulation and B7 2 expression is largely restricted to the medulla the development of which seems to parallel the development of Foxp3 cells It has been suggested that the two are linked but no definitive link between the processes has yet been shown TGF b is not required for Treg functionality in the thymus as thymic Tregs from TGF b insensitive TGFbRII DN mice are functional Thymic recirculation edit There was observed that some FOXP3 Treg cells are recirculating back to thymus where they have developed This Treg were mainly present in thymic medulla which is the main site of Treg cells differentiation 17 The presence of this cells in thymus or addition into fetal thymic tissue culture suppress development of new Treg cells by 34 60 17 but Tconv cells are not affected That means that recirculating Treg to thymus inhibited just de novo development of Treg cells Molecular mechanism of this process works due to the ability of Treg to adsorb IL 2 from the microenvironments thus being able to induce apoptosis of other T cells which need IL 2 as main growth factor 18 Recirculating T reg cells in thymus express high amount of high affinity IL 2 receptor a chain CD25 encoded by Il2ra gene which gather IL 2 from thymic medulla and decrease its concentration New generated FOXP3 Treg cells in thymus have not so high amount of Il2ra expression 17 IL 2 is a cytokine necessary for the development of Treg cells in the thymus It is important for T cells proliferation and survival but in the case of its deficiency IL 15 may be replaced However Treg cells development is dependent on IL 2 19 In humans there was found population of CD31 negative Treg cells in thymus 17 CD31 could be used as a marker of new generated Treg cells as same as other T lymphocytes Mature and peripheral Treg cells have decreased its expression 20 So it is possible that this regulatory mechanism of thymic Treg cells development is also functional in humans There is probably also positive regulation of thymic Treg cells development caused by recirculating Treg cells into thymus There was found population of CD24 low FOXP3 in thymus with increased expression of IL 1R2 Il1r2 compared with peripheral Treg cells 21 22 High concentration of IL 1b caused by inflammation decrease de novo development of Treg cells in thymus 22 The presence of recirculating Treg cells in the thymus with high IL1R2 expression during inflammatory conditions helps to uptake IL 1b and reduce its concentration in the medulla microenvironment thus they are helping to the development of de novo Treg cells 22 High concentration of IL 1b caused by inflammation decrease de novo development of Treg cells in thymus 22 Binding of IL 1b to IL1R2 on the surface of Treg cells does not cause any signal transduction because there is no present Intracellular TIR Toll interleukin 1 receptor domain which is normally present in innate immune cells 23 Function editThe immune system must be able to discriminate between self and non self When self non self discrimination fails the immune system destroys cells and tissues of the body and as a result causes autoimmune diseases Regulatory T cells actively suppress activation of the immune system and prevent pathological self reactivity i e autoimmune disease The critical role regulatory T cells play within the immune system is evidenced by the severe autoimmune syndrome that results from a genetic deficiency in regulatory T cells IPEX syndrome see also below nbsp Diagram of regulatory T cell effector T cells and dendritic cell showing putative mechanisms of suppression by regulatory T cells The molecular mechanism by which regulatory T cells exert their suppressor regulatory activity has not been definitively characterized and is the subject of intense research In vitro experiments have given mixed results regarding the requirement of cell to cell contact with the cell being suppressed The following represent some of the proposed mechanisms of immune suppression Regulatory T cells produce a number of inhibitory cytokines These include TGF b 24 Interleukin 35 25 and Interleukin 10 26 It also appears that regulatory T cells can induce other cell types to express interleukin 10 27 Regulatory T cells can produce Granzyme B which in turn can induce apoptosis of effector cells Regulatory T cells from Granzyme B deficient mice are reported to be less effective suppressors of the activation of effector T cells 28 Reverse signalling through direct interaction with dendritic cells and the induction of immunosuppressive indoleamine 2 3 dioxygenase 29 Signalling through the ectoenzymes CD39 and CD73 with the production of immunosuppressive adenosine 30 31 Through direct interactions with dendritic cells by LAG3 and by TIGIT 32 33 This review of Treg interactions with dendritic cells provides distinction between mechanisms described for human cells versus mouse cells 34 Another control mechanism is through the IL 2 feedback loop Antigen activated T cells produce IL 2 which then acts on IL 2 receptors on regulatory T cells alerting them to the fact that high T cell activity is occurring in the region and they mount a suppressory response against them This is a negative feedback loop to ensure that overreaction is not occurring If an actual infection is present other inflammatory factors downregulate the suppression Disruption of the loop leads to hyperreactivity regulation can modify the strength of the immune response 35 A related suggestion with regard to interleukin 2 is that activated regulatory T cells take up interleukin 2 so avidly that they deprive effector T cells of sufficient to avoid apoptosis 18 A major mechanism of suppression by regulatory T cells is through the prevention of co stimulation through CD28 on effector T cells by the action of the molecule CTLA 4 36 Natural and induced regulatory T cells editT regulatory lymphocytes develop during ontogeny either in the thymus or in the periphery Accordingly they are divided into natural and induced T regulatory cells 37 Natural T regulatory lymphocytes tTregs nTregs are characterized by continuous expression of FoxP3 and T cell receptor TCR with relatively high autoaffinity These cells are predominantly found in the body in the bloodstream or lymph nodes and serve mainly to confer tolerance to autoantigens 37 Induced peripheral T regulatory cells iTregs pTregs arise under certain situations in the presence of IL 2 and TGF b in the periphery and begin to express FoxP3 inducibly thus becoming the functional equivalent of tTreg cells iTregs however are found primarily in peripheral barrier tissues where they are primarily involved in preventing inflammation in the presence of external antigens 37 The main features that differentiate tTreg and iTreg cells include Helios and Neuropilin 1 the presence of which suggests origin in the thymus Another feature distinguishing these two Treg cell populations is the stability of FoxP3 expression in different settings 37 Induced T regulatory cells edit Induced regulatory T iTreg cells CD4 CD25 FOXP3 are suppressive cells involved in tolerance iTreg cells have been shown to suppress T cell proliferation and experimental autoimmune diseases These cells include Treg17 cells iTreg cells develop from mature CD4 conventional T cells outside of the thymus a defining distinction between natural regulatory T nTreg cells and iTreg cells Though iTreg and nTreg cells share a similar function iTreg cells have recently been shown to be an essential non redundant regulatory subset that supplements nTreg cells in part by expanding TCR diversity within regulatory responses 38 Acute depletion of the iTreg cell pool in mouse models has resulted in inflammation and weight loss The contribution of nTreg cells versus iTreg cells in maintaining tolerance is unknown but both are important Epigenetic differences have been observed between nTreg and iTreg cells with the former having more stable FOXP3 expression and wider demethylation The small intestinal environment is high in vitamin A and is a location where retinoic acid is produced 39 The retinoic acid and TGF beta produced by dendritic cells within this area signal for production of regulatory T cells 39 Vitamin A and TGF beta promote T cell differentiation into regulatory T cells opposed to Th17 cells even in the presence of IL 6 40 41 The intestinal environment can lead to induced regulatory T cells with TGF beta and retinoic acid 42 some of which express the lectin like receptor CD161 and are specialized to maintain barrier integrity by accelerating wound healing 43 The Tregs within the gut are differentiated from naive T cells after antigen is introduced 44 It has recently been shown that human regulatory T cells can be induced from both naive and pre committed Th1 cells and Th17 cells 45 using a parasite derived TGF b mimic secreted by Heligmosomoides polygyrus and termed Hp TGM H polygyrus TGF b mimic 46 47 Hp TGM can induce murine FOXP3 expressing regulatory T cells that were stabile in presence of inflammation in vivo 48 Hp TGM induced human FOXP3 regulatory T cells were stable in the presence of inflammation and had increased levels of CD25 CTLA4 and decreased methylation in the FOXP3 Treg Specific demethylated region compared to TGF b induced Tregs 45 RORgt regulatory T lymphocytes edit The iTregs are able to differentiate into RORgt expressing cells and thus acquire the phenotype of Th17 cells These cells are associated with the functions of mucosal lymphoid tissues such as the intestinal barrier In the intestinal lamina propria 20 30 of Foxp3 T regulatory cells expressing RORyt are found and this high proportion is strongly dependent on the presence of a complex gut microbiome In germ free GF mice the population of RORgt T regulatory cells is strongly reduced whereas recolonization by the specific pathogen free SPF microbiota restores normal numbers of these lymphocytes in the gut The mechanism by which the gut microbiota induces the formation of RORgt Treg cells involves the production of short chain fatty acids SCFAs on which this induction is dependent SCFAs are a by product of fermentation and digestion of dietary fiber therefore microbial free mice have very low concentrations of both SCFAs and RORgt Treg cells Induction of RORgt Treg cells is also dependent on the presence of dendritic cells in adults Thetis cells in neonatal and antigen presentation by MHC II 49 50 RORgt Treg cells are not present in the thymus and do not express Helios or Neuropilin 1 but have high expression of CD44 IL 10 ICOS CTLA 4 and the nucleotidases CD39 and CD73 suggesting a strong regulatory function 49 Function of RORgt regulatory T lymphocytes edit Induction of RORgt Treg cells in lymph nodes of the small intestine is crucial for the establishment of intestinal luminal antigen tolerance These cells are particularly important in the prevention of food allergies One mechanism is the production of suppressive molecules such as the cytokine IL 10 These cells also suppress the Th17 cell population and inhibit the production of IL 17 thus suppressing the pro inflammatory response 49 Gata3 regulatory T lymphocytes edit Another important subset of Treg cells are Gata3 Treg cells which respond to IL 33 in the gut and influence the regulation of effector T cells during inflammation Unlike RORgt Treg cells these cells express Helios and are not dependent on the microbiome 50 51 Gata3 T regs are major immunosuppressors during intestinal inflammation and T regs use Gata3 to limit tissue inflammation This cell population also restrict Th17 T cells immunity in the intestine because Gata3 deficient T regs express higher Rorc and IL 17a transcript 52 Disease editAn important question in the field of immunology is how the immunosuppressive activity of regulatory T cells is modulated during the course of an ongoing immune response While the immunosuppressive function of regulatory T cells prevents the development of autoimmune disease it is not desirable during immune responses to infectious microorganisms Current hypotheses suggest that upon encounter with infectious microorganisms the activity of regulatory T cells may be downregulated either directly or indirectly by other cells to facilitate elimination of the infection Experimental evidence from mouse models suggests that some pathogens may have evolved to manipulate regulatory T cells to immunosuppress the host and so potentiate their own survival For example regulatory T cell activity has been reported to increase in several infectious contexts such as retroviral infections the most well known of which is HIV mycobacterial infections e g tuberculosis 53 and various parasitic infections including Leishmania and malaria Treg cells play major roles during HIV infection They suppress the immune system thus limiting target cells and reducing inflammation but this simultaneously disrupts the clearance of virus by the cell mediated immune response and enhances the reservoir by pushing CD4 T cells to a resting state including infected cells Additionally Treg cells can be infected by HIV increasing the size of the HIV reservoir directly Thus Treg cells are being investigated as targets for HIV cure research 54 Some Treg cell depletion strategies have been tested in SIV infected nonhuman primates and shown to cause viral reactivation and enhanced SIV specific CD8 T cell responses 55 Regulatory T cells have a large role in the pathology of visceral leishmaniasis and in preventing excess inflammation in patients cured of visceral leishmaniasis CD4 regulatory T cells are often associated with solid tumours in both humans and murine models Increased numbers of regulatory T cells in breast colorectal and ovarian cancers is associated with a poorer prognosis 56 CD70 non Hodgkin lymphoma B cells induce FOXP3 expression and regulatory function in intratumoral CD4 CD25 T cells 57 There is some evidence that Treg cells may be dysfunctional and driving neuroinflammation in amyotrophic lateral sclerosis due to lower expression of FOXP3 58 Ex vivo expansion of Treg cells for subsequent autologous transplant is currently being investigated after promising results were obtained in a phase I clinical trial 59 Additionally while regulatory T cells have been shown to increase via polyclonal expansion both systemically and locally during healthy pregnancies to protect the fetus from the maternal immune response a process called maternal immune tolerance there is evidence that this polyclonal expansion is impaired in preeclamptic mothers and their offspring 60 Research suggests reduced production and development of regulatory T cells during preeclampsia may degrade maternal immune tolerance leading to the hyperactive immune response characteristic of preeclampsia 61 Cancer edit nbsp The recruitment and maintenance of Treg cells in the tumor microenvironmentMost tumors elicit an immune response in the host that is mediated by tumor antigens thus distinguishing the tumor from other non cancerous cells This causes large numbers of tumor infiltrating lymphocytes TILs to be found in the tumor microenvironment 62 Although it is not entirely understood it is thought that these lymphocytes target cancerous cells and therefore slow or terminate the development of the tumor However this process is complicated because Treg cells seem to be preferentially trafficked to the tumor microenvironment While Treg cells normally make up only about 4 of CD4 T cells they can make up as much as 20 30 of the total CD4 population around the tumor microenvironment 63 Although high levels of TILs were initially thought to be important in determining an immune response against cancer it is now widely recognized that the ratio of Treg to effector T cells in the tumor microenvironment is a determining factor in the success of the immune response against the cancer High levels of Treg cells in the tumor microenvironment are associated with poor prognosis in many cancers 64 such as ovarian breast renal and pancreatic cancer 63 This indicates that Treg cells suppress effector T cells and hinder the body s immune response against the cancer However in some types of cancer the opposite is true and high levels of Treg cells are associated with a positive prognosis This trend is seen in cancers such as colorectal carcinoma and follicular lymphoma This could be due to Treg cells ability to suppress general inflammation which is known to trigger cell proliferation and metastasis 63 These opposite effects indicate that Treg cells role in the development of cancer is highly dependent on both type and location of the tumor Although it is still not entirely understood how Treg cells are preferentially trafficked to the tumor microenvironment the chemotaxis is probably driven by the production of chemokines by the tumor Treg infiltration into the tumor microenvironment is facilitated by the binding of the chemokine receptor CCR4 which is expressed on Treg cells to its ligand CCL22 which is secreted by many types of tumor cells 65 Treg cell expansion at the site of the tumor could also explain the increased levels of Treg cells The cytokine TGF b which is commonly produced by tumor cells is known to induce the differentiation and expansion of Treg cells 65 Forkhead box protein 3 FOXP3 as a transcription factor is an essential molecular marker of Treg cells FOXP3 polymorphism rs3761548 might be involved in the gastric cancer progression through influencing Treg function and the secretion of immunomodulatory cytokines such as IL 10 IL 35 and TGF b 66 Treg cells present in the tumor microenvironment TME can be either induced Tregs or natural thymic Tregs which develop from naive precursors However tumor associated Tregs may also originate from IL 17A Foxp3 Tregs which develop from Th17 cells 67 68 In general the immunosuppression of the tumor microenvironment has largely contributed to the unsuccessful outcomes of many cancer immunotherapy treatments Depletion of Treg cells in animal models has shown an increased efficacy of immunotherapy treatments and therefore many immunotherapy treatments are now incorporating Treg depletion 2 Cancer therapies targeting regulatory T lymphocytes edit Tregs in the TME are abundantly effector Tregs which over express immunosuppressive molecules such as CTLA 4 Anti CTLA 4 antibodies cause depletion of Tregs and thus increase CD8 T cells effective against the tumor Anti CTLA 4 antibody ipilimumab was approved for patients with advanced melanoma Immune checkpoint molecule PD 1 inhibits activation of both conventional T cells and Tregs and use of anti PD 1 antibodies may lead to activation and immunosuppressive function of Tregs Resistance to anti PD 1 mAb treatment is probably caused by enhanced Treg cell activity Rapid cancer progression upon PD 1 blockade is called hyperprogressive disease Therapies targeting Treg suppression include anti CD25 mAbs and anti CCR4 mAbs OX40 agonist and GITR agonists are currently being investigated 67 69 Therapy targeting TCR signaling is also possible by blocking tyrosine kinases For example tyrosine kinase inhibitor dasatinib is used for treatment of chronic myeloid leukemia and is associated with Treg inhibition 70 Molecular characterization editSimilar to other T cells regulatory T cells develop in the thymus The latest research suggests that regulatory T cells are defined by expression of the forkhead family transcription factor FOXP3 forkhead box p3 Expression of FOXP3 is required for regulatory T cell development and appears to control a genetic program specifying this cell s fate 71 The large majority of Foxp3 expressing regulatory T cells are found within the major histocompatibility complex MHC class II restricted CD4 expressing CD4 population and express high levels of the interleukin 2 receptor alpha chain CD25 In addition to the FOXP3 expressing CD4 CD25 there also appears to be a minor population of MHC class I restricted CD8 FOXP3 expressing regulatory T cells These FOXP3 expressing CD8 T cells do not appear to be functional in healthy individuals but are induced in autoimmune disease states by T cell receptor stimulation to suppress IL 17 mediated immune responses 72 Unlike conventional T cells regulatory T cells do not produce IL 2 and are therefore anergic at baseline A number of different methods are employed in research to identify and monitor Treg cells Originally high expression of CD25 and CD4 surface markers was used CD4 CD25 cells This is problematic as CD25 is also expressed on non regulatory T cells in the setting of immune activation such as during an immune response to a pathogen As defined by CD4 and CD25 expression regulatory T cells comprise about 5 10 of the mature CD4 T cell subpopulation in mice and humans while about 1 2 of Treg can be measured in whole blood The additional measurement of cellular expression of FOXP3 protein allowed a more specific analysis of Treg cells CD4 CD25 FOXP3 cells However FOXP3 is also transiently expressed in activated human effector T cells thus complicating a correct Treg analysis using CD4 CD25 and FOXP3 as markers in humans Therefore the gold standard surface marker combination to defined Tregs within unactivated CD3 CD4 T cells is high CD25 expression combined with the absent or low level expression of the surface protein CD127 IL 7RA If viable cells are not required then the addition of FOXP3 to the CD25 and CD127 combination will provide further stringency Several additional markers have been described e g high levels of CTLA 4 cytotoxic T lymphocyte associated molecule 4 and GITR glucocorticoid induced TNF receptor are also expressed on regulatory T cells however the functional significance of this expression remains to be defined There is a great interest in identifying cell surface markers that are uniquely and specifically expressed on all FOXP3 expressing regulatory T cells However to date no such molecule has been identified The identification of Tregs following cell activation is challenging as conventional T cells will express CD25 transiently express FOXP3 and lose CD127 expression upon activation It has been shown that Tregs can be detected using an activation induced marker assay by expression of CD39 73 in combination with co expression of CD25 and OX40 CD134 which define antigen specific cells following 24 48h stimulation with antigen 74 75 In addition to the search for novel protein markers a different method to analyze and monitor Treg cells more accurately has been described in the literature This method is based on DNA methylation analysis Only in Treg cells but not in any other cell type including activated effector T cells a certain region within the FOXP3 gene TSDR Treg specific demethylated region is found demethylated which allows to monitor Treg cells through a PCR reaction or other DNA based analysis methods 76 Interplay between the Th17 cells and regulatory T cells are important in many diseases like respiratory diseases 77 Recent evidence suggests that mast cells may be important mediators of Treg dependent peripheral tolerance 78 Epitopes edit Regulatory T cell epitopes Tregitopes were discovered in 2008 and consist of linear sequences of amino acids contained within monoclonal antibodies and immunoglobulin G IgG Since their discovery evidence has indicated Tregitopes may be crucial to the activation of natural regulatory T cells 79 80 81 Potential applications of regulatory T cell epitopes have been hypothesised tolerisation to transplants protein drugs blood transfer therapies and type I diabetes as well as reduction of immune response for the treatment of allergies 82 83 84 85 86 87 81 Genetic deficiency editGenetic mutations in the gene encoding FOXP3 have been identified in both humans and mice based on the heritable disease caused by these mutations This disease provides the most striking evidence that regulatory T cells play a critical role in maintaining normal immune system function Humans with mutations in FOXP3 develop a severe and rapidly fatal autoimmune disorder known as Immune dysregulation Polyendocrinopathy Enteropathy X linked IPEX syndrome 88 89 The IPEX syndrome is characterized by the development of overwhelming systemic autoimmunity in the first year of life resulting in the commonly observed triad of watery diarrhea eczematous dermatitis and endocrinopathy seen most commonly as insulin dependent diabetes mellitus Most individuals have other autoimmune phenomena including Coombs positive hemolytic anemia autoimmune thrombocytopenia autoimmune neutropenia and tubular nephropathy The majority of affected males die within the first year of life of either metabolic derangements or sepsis An analogous disease is also observed in a spontaneous FOXP3 mutant mouse known as scurfy See also editList of distinct cell types in the adult human bodyReferences edit Bettelli E Carrier Y Gao W Korn T Strom TB Oukka M et al May 2006 Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells Nature 441 7090 235 8 Bibcode 2006Natur 441 235B doi 10 1038 nature04753 PMID 16648838 S2CID 4391497 a b Curiel TJ May 2007 Tregs and rethinking cancer immunotherapy The Journal of Clinical Investigation 117 5 1167 74 doi 10 1172 JCI31202 PMC 1857250 PMID 17476346 Chen W August 2011 Tregs in immunotherapy opportunities and challenges Immunotherapy 3 8 911 4 doi 10 2217 imt 11 79 PMID 21843075 Miyara M Gorochov G Ehrenstein M Musset L Sakaguchi S Amoura Z October 2011 Human FoxP3 regulatory T cells in systemic autoimmune diseases Autoimmunity Reviews 10 12 744 55 doi 10 1016 j autrev 2011 05 004 PMID 21621000 Nosbaum A Prevel N Truong HA Mehta P Ettinger M Scharschmidt TC et al March 2016 Cutting Edge Regulatory T Cells Facilitate Cutaneous Wound Healing Journal of Immunology 196 5 2010 4 doi 10 4049 jimmunol 1502139 PMC 4761457 PMID 26826250 Adeegbe DO Nishikawa H 2013 Natural and induced T regulatory cells in cancer Frontiers in Immunology 4 190 doi 10 3389 fimmu 2013 00190 PMC 3708155 PMID 23874336 Curiel TJ April 2008 Regulatory T cells and treatment of cancer Current Opinion in Immunology 20 2 241 6 doi 10 1016 j coi 2008 04 008 PMC 3319305 PMID 18508251 Hori S Nomura T Sakaguchi S February 2003 Control of regulatory T cell development by the transcription factor Foxp3 Science 299 5609 1057 61 Bibcode 2003Sci 299 1057H doi 10 1126 science 1079490 PMID 12522256 S2CID 9697928 Singh B Schwartz JA Sandrock C Bellemore SM Nikoopour E November 2013 Modulation of autoimmune diseases by interleukin IL 17 producing regulatory T helper Th17 cells The Indian Journal of Medical Research 138 5 591 4 PMC 3928692 PMID 24434314 Shevach EM 2000 Regulatory T cells in autoimmmunity Annual Review of Immunology 18 423 49 doi 10 1146 annurev immunol 18 1 423 PMID 10837065 S2CID 15160752 Schmetterer KG Neunkirchner A Pickl WF June 2012 Naturally occurring regulatory T cells markers mechanisms and manipulation FASEB Journal 26 6 2253 76 doi 10 1096 fj 11 193672 PMID 22362896 S2CID 36277557 Sakaguchi S 2004 Naturally arising CD4 regulatory t cells for immunologic self tolerance and negative control of immune responses Annual Review of Immunology 22 531 62 doi 10 1146 annurev immunol 21 120601 141122 PMID 15032588 Li MO Rudensky AY April 2016 T cell receptor signalling in the control of regulatory T cell differentiation and function Nature Reviews Immunology 16 4 220 33 doi 10 1038 nri 2016 26 PMC 4968889 PMID 27026074 a b c d Santamaria Jeremy C Borelli Alexia Irla Magali 2021 02 11 Regulatory T Cell Heterogeneity in the Thymus Impact on Their Functional Activities Frontiers in Immunology 12 643153 doi 10 3389 fimmu 2021 643153 ISSN 1664 3224 PMC 7904894 PMID 33643324 Owen David L Sjaastad Louisa E Farrar Michael A 2019 10 15 Regulatory T Cell Development in the Thymus The Journal of Immunology 203 8 2031 2041 doi 10 4049 jimmunol 1900662 ISSN 0022 1767 PMC 6910132 PMID 31591259 Owen David L Mahmud Shawn A Sjaastad Louisa E Williams Jason B Spanier Justin A Simeonov Dimitre R Ruscher Roland Huang Weishan Proekt Irina Miller Corey N Hekim Can Jeschke Jonathan C Aggarwal Praful Broeckel Ulrich LaRue Rebecca S February 2019 Thymic regulatory T cells arise via two distinct developmental programs Nature Immunology 20 2 195 205 doi 10 1038 s41590 018 0289 6 ISSN 1529 2916 PMC 6650268 PMID 30643267 a b c d Thiault N Darrigues J Adoue V Gros M Binet B Perals C et al June 2015 Peripheral regulatory T lymphocytes recirculating to the thymus suppress the development of their precursors Nature Immunology 16 6 628 34 doi 10 1038 ni 3150 PMID 25939024 S2CID 7670443 a b Pandiyan P Zheng L Ishihara S Reed J Lenardo MJ December 2007 CD4 CD25 Foxp3 regulatory T cells induce cytokine deprivation mediated apoptosis of effector CD4 T cells Nature Immunology 8 12 1353 62 doi 10 1038 ni1536 PMID 17982458 S2CID 8925488 Cheng G Yu A Malek TR May 2011 T cell tolerance and the multi functional role of IL 2R signaling in T regulatory cells Immunological Reviews 241 1 63 76 doi 10 1111 j 1600 065X 2011 01004 x PMC 3101713 PMID 21488890 Kimmig S Przybylski GK Schmidt CA Laurisch K Mowes B Radbruch A Thiel A March 2002 Two subsets of naive T helper cells with distinct T cell receptor excision circle content in human adult peripheral blood The Journal of Experimental Medicine 195 6 789 94 doi 10 1084 jem 20011756 PMC 2193736 PMID 11901204 Toker A Engelbert D Garg G Polansky JK Floess S Miyao T et al April 2013 Active demethylation of the Foxp3 locus leads to the generation of stable regulatory T cells within the thymus Journal of Immunology 190 7 3180 8 doi 10 4049 jimmunol 1203473 PMID 23420886 a b c d Nikolouli E Elfaki Y Herppich S Schelmbauer C Delacher M Falk C et al January 2021 Recirculating IL 1R2 Treg fine tune intrathymic Treg development under inflammatory conditions Cellular amp Molecular Immunology 18 1 182 193 doi 10 1038 s41423 019 0352 8 hdl 10033 622148 PMC 7853075 PMID 31988493 S2CID 210913733 Peters VA Joesting JJ Freund GG August 2013 IL 1 receptor 2 IL 1R2 and its role in immune regulation Brain Behavior and Immunity 32 1 8 doi 10 1016 j bbi 2012 11 006 PMC 3610842 PMID 23195532 Read S Malmstrom V Powrie F July 2000 Cytotoxic T lymphocyte associated antigen 4 plays an essential role in the function of CD25 CD4 regulatory cells that control intestinal inflammation The Journal of Experimental Medicine 192 2 295 302 doi 10 1084 jem 192 2 295 PMC 2193261 PMID 10899916 Collison LW Workman CJ Kuo TT Boyd K Wang Y Vignali KM et al November 2007 The inhibitory cytokine IL 35 contributes to regulatory T cell function Nature 450 7169 566 9 Bibcode 2007Natur 450 566C doi 10 1038 nature06306 PMID 18033300 S2CID 4425281 Annacker O Asseman C Read S Powrie F June 2003 Interleukin 10 in the regulation of T cell induced colitis Journal of Autoimmunity 20 4 277 9 doi 10 1016 s0896 8411 03 00045 3 PMID 12791312 Kearley J Barker JE Robinson DS Lloyd CM December 2005 Resolution of airway inflammation and hyperreactivity after in vivo transfer of CD4 CD25 regulatory T cells is interleukin 10 dependent The Journal of Experimental Medicine 202 11 1539 47 doi 10 1084 jem 20051166 PMC 1350743 PMID 16314435 Gondek DC Lu LF Quezada SA Sakaguchi S Noelle RJ February 2005 Cutting edge contact mediated suppression by CD4 CD25 regulatory cells involves a granzyme B dependent perforin independent mechanism Journal of Immunology 174 4 1783 6 doi 10 4049 jimmunol 174 4 1783 PMID 15699103 Puccetti P Grohmann U October 2007 IDO and regulatory T cells a role for reverse signalling and non canonical NF kappaB activation Nature Reviews Immunology 7 10 817 23 doi 10 1038 nri2163 PMID 17767193 S2CID 5544429 Borsellino G Kleinewietfeld M Di Mitri D Sternjak A Diamantini A Giometto R et al August 2007 Expression of ectonucleotidase CD39 by Foxp3 Treg cells hydrolysis of extracellular ATP and immune suppression Blood 110 4 1225 32 doi 10 1182 blood 2006 12 064527 PMID 17449799 Kobie JJ Shah PR Yang L Rebhahn JA Fowell DJ Mosmann TR November 2006 T regulatory and primed uncommitted CD4 T cells express CD73 which suppresses effector CD4 T cells by converting 5 adenosine monophosphate to adenosine Journal of Immunology 177 10 6780 6 doi 10 4049 jimmunol 177 10 6780 PMID 17082591 Huang CT Workman CJ Flies D Pan X Marson AL Zhou G et al October 2004 Role of LAG 3 in regulatory T cells Immunity 21 4 503 13 doi 10 1016 j immuni 2004 08 010 PMID 15485628 Yu X Harden K Gonzalez LC Francesco M Chiang E Irving B et al January 2009 The surface protein TIGIT suppresses T cell activation by promoting the generation of mature immunoregulatory dendritic cells Nature Immunology 10 1 48 57 doi 10 1038 ni 1674 PMID 19011627 S2CID 205361984 Wardell CM MacDonald KN Levings MK Cook L January 2021 Cross talk between human regulatory T cells and antigen presenting cells Lessons for clinical applications European Journal of Immunology 51 1 27 38 doi 10 1002 eji 202048746 PMID 33301176 Sakaguchi S Yamaguchi T Nomura T Ono M May 2008 Regulatory T cells and immune tolerance Cell 133 5 775 87 doi 10 1016 j cell 2008 05 009 PMID 18510923 Walker LS Sansom DM November 2011 The emerging role of CTLA4 as a cell extrinsic regulator of T cell responses Nature Reviews Immunology 11 12 852 63 doi 10 1038 nri3108 PMID 22116087 S2CID 9617595 a b c d Shevyrev Daniil Tereshchenko Valeriy 2020 Treg Heterogeneity Function and Homeostasis Frontiers in Immunology 10 3100 doi 10 3389 fimmu 2019 03100 ISSN 1664 3224 PMC 6971100 PMID 31993063 Haribhai D Williams JB Jia S Nickerson D Schmitt EG Edwards B et al July 2011 A requisite role for induced regulatory T cells in tolerance based on expanding antigen receptor diversity Immunity 35 1 109 22 doi 10 1016 j immuni 2011 03 029 PMC 3295638 PMID 21723159 a b Sun CM Hall JA Blank RB Bouladoux N Oukka M Mora JR Belkaid Y August 2007 Small intestine lamina propria dendritic cells promote de novo generation of Foxp3 T reg cells via retinoic acid The Journal of Experimental Medicine 204 8 1775 85 doi 10 1084 jem 20070602 PMC 2118682 PMID 17620362 Mucida D Park Y Kim G Turovskaya O Scott I Kronenberg M Cheroutre H July 2007 Reciprocal TH17 and regulatory T cell differentiation mediated by retinoic acid Science 317 5835 256 60 Bibcode 2007Sci 317 256M doi 10 1126 science 1145697 PMID 17569825 S2CID 24736012 Erkelens MN Mebius RE March 2017 Retinoic Acid and Immune Homeostasis A Balancing Act Trends in Immunology 38 3 168 180 doi 10 1016 j it 2016 12 006 PMID 28094101 Ziegler SF Buckner JH April 2009 FOXP3 and the regulation of Treg Th17 differentiation Microbes and Infection 11 5 594 8 doi 10 1016 j micinf 2009 04 002 PMC 2728495 PMID 19371792 Povoleri GA Nova Lamperti E Scotta C Fanelli G Chen YC Becker PD et al December 2018 Human retinoic acid regulated CD161 regulatory T cells support wound repair in intestinal mucosa Nature Immunology 19 12 1403 1414 doi 10 1038 s41590 018 0230 z PMC 6474659 PMID 30397350 Coombes JL Siddiqui KR Arancibia Carcamo CV Hall J Sun CM Belkaid Y Powrie F August 2007 A functionally specialized population of mucosal CD103 DCs induces Foxp3 regulatory T cells via a TGF beta and retinoic acid dependent mechanism The Journal of Experimental Medicine 204 8 1757 64 doi 10 1084 jem 20070590 PMC 2118683 PMID 17620361 a b Cook L Reid KT Hakkinen E de Bie B Tanaka S Smyth DJ et al September 2021 Induction of stable human FOXP3 Tregs by a parasite derived TGF b mimic Immunology and Cell Biology 99 8 833 847 doi 10 1111 IMCB 12475 PMC 8453874 PMID 33929751 Johnston CJ Smyth DJ Kodali RB White MP Harcus Y Filbey KJ et al November 2017 A structurally distinct TGF b mimic from an intestinal helminth parasite potently induces regulatory T cells Nature Communications 8 1 1741 Bibcode 2017NatCo 8 1741J doi 10 1038 s41467 017 01886 6 PMC 5701006 PMID 29170498 Smyth DJ Harcus Y White MP Gregory WF Nahler J Stephens I et al April 2018 TGF b mimic proteins form an extended gene family in the murine parasite Heligmosomoides polygyrus International Journal for Parasitology 48 5 379 385 doi 10 1016 j ijpara 2017 12 004 PMC 5904571 PMID 29510118 White MP Smyth DJ Cook L Ziegler SF Levings MK Maizels RM September 2021 The parasite cytokine mimic Hp TGM potently replicates the regulatory effects of TGF b on murine CD4 T cells Immunology and Cell Biology 99 8 848 864 doi 10 1111 IMCB 12479 PMC 9214624 PMID 33988885 a b c Ning Xixi Lei Zengjie Rui Binqi Li Yuyuan Li Ming 2022 12 05 Gut Microbiota Promotes Immune Tolerance by Regulating RORgt Treg Cells in Food Allergy Advanced Gut amp Microbiome Research 2022 e8529578 doi 10 1155 2022 8529578 a b Ohnmacht Caspar Park Joo Hong Cording Sascha Wing James B Atarashi Koji Obata Yuuki Gaboriau Routhiau Valerie Marques Rute Dulauroy Sophie Fedoseeva Maria Busslinger Meinrad Cerf Bensussan Nadine Boneca Ivo G Voehringer David Hase Koji 2015 08 28 The microbiota regulates type 2 immunity through RORgt T cells Science 349 6251 989 993 Bibcode 2015Sci 349 989O doi 10 1126 science aac4263 ISSN 0036 8075 PMID 26160380 S2CID 2663636 Jacobse Justin Li Jing Rings Edmond H H M Samsom Janneke N Goettel Jeremy A 2021 Intestinal Regulatory T Cells as Specialized Tissue Restricted Immune Cells in Intestinal Immune Homeostasis and Disease Frontiers in Immunology 12 716499 doi 10 3389 fimmu 2021 716499 ISSN 1664 3224 PMC 8371910 PMID 34421921 Lui Prudence PokWai Cho Inchul Ali Niwa September 2020 Tissue regulatory T cells Immunology 161 1 4 17 doi 10 1111 imm 13208 ISSN 0019 2805 PMC 7450170 PMID 32463116 Stringari LL Covre LP da Silva FD de Oliveira VL Campana MC Hadad DJ et al July 2021 Increase of CD4 CD25highFoxP3 cells impairs in vitro human microbicidal activity against Mycobacterium tuberculosis during latent and acute pulmonary tuberculosis PLOS Neglected Tropical Diseases 15 7 e0009605 doi 10 1371 journal pntd 0009605 PMC 8321116 PMID 34324509 Kleinman AJ Sivanandham R Pandrea I Chougnet CA Apetrei C 2018 Regulatory T Cells As Potential Targets for HIV Cure Research Frontiers in Immunology 9 734 doi 10 3389 fimmu 2018 00734 PMC 5908895 PMID 29706961 Sivanandham R Kleinman AJ Sette P Brocca Cofano E Kilapandal Venkatraman SM Policicchio BB et al September 2020 Nonhuman Primate Testing of the Impact of Different Regulatory T Cell Depletion Strategies on Reactivation and Clearance of Latent Simian Immunodeficiency Virus Journal of Virology 94 19 JVI 00533 20 jvi JVI 00533 20v1 doi 10 1128 JVI 00533 20 PMC 7495362 PMID 32669326 S2CID 220579402 Dranoff G December 2005 The therapeutic implications of intratumoral regulatory T cells Clinical Cancer Research 11 23 8226 9 doi 10 1158 1078 0432 CCR 05 2035 PMID 16322278 S2CID 18794337 Yang ZZ Novak AJ Ziesmer SC Witzig TE Ansell SM October 2007 CD70 non Hodgkin lymphoma B cells induce Foxp3 expression and regulatory function in intratumoral CD4 CD25 T cells Blood 110 7 2537 44 doi 10 1182 blood 2007 03 082578 PMC 1988926 PMID 17615291 Beers DR Zhao W Wang J Zhang X Wen S Neal D et al March 2017 ALS patients regulatory T lymphocytes are dysfunctional and correlate with disease progression rate and severity JCI Insight 2 5 e89530 doi 10 1172 jci insight 89530 PMC 5333967 PMID 28289705 Thonhoff JR Beers DR Zhao W Pleitez M Simpson EP Berry JD et al July 2018 Expanded autologous regulatory T lymphocyte infusions in ALS A phase I first in human study Neurology 5 4 e465 doi 10 1212 NXI 0000000000000465 PMC 5961523 PMID 29845093 Tsuda S Nakashima A Shima T Saito S 2019 New Paradigm in the Role of Regulatory T Cells During Pregnancy Frontiers in Immunology 10 573 doi 10 3389 fimmu 2019 00573 PMC 6443934 PMID 30972068 Hu M Eviston D Hsu P Marino E Chidgey A Santner Nanan B et al July 2019 Decreased maternal serum acetate and impaired fetal thymic and regulatory T cell development in preeclampsia Nature Communications 10 1 3031 Bibcode 2019NatCo 10 3031H doi 10 1038 s41467 019 10703 1 PMC 6620275 PMID 31292453 Gooden MJ de Bock GH Leffers N Daemen T Nijman HW June 2011 The prognostic influence of tumour infiltrating lymphocytes in cancer a systematic review with meta analysis British Journal of Cancer 105 1 93 103 doi 10 1038 bjc 2011 189 PMC 3137407 PMID 21629244 a b c Oleinika K Nibbs RJ Graham GJ Fraser AR January 2013 Suppression subversion and escape the role of regulatory T cells in cancer progression Clinical and Experimental Immunology 171 1 36 45 doi 10 1111 j 1365 2249 2012 04657 x PMC 3530093 PMID 23199321 Plitas G Rudensky AY 2020 03 09 Regulatory T Cells in Cancer Annual Review of Cancer Biology 4 1 459 477 doi 10 1146 annurev cancerbio 030419 033428 ISSN 2472 3428 a b Lippitz BE May 2013 Cytokine patterns in patients with cancer a systematic review The Lancet Oncology 14 6 e218 28 doi 10 1016 s1470 2045 12 70582 x PMID 23639322 Ezzeddini R Somi MH Taghikhani M Moaddab SY Masnadi Shirazi K Shirmohammadi M Eftekharsadat AT Sadighi Moghaddam B Salek Farrokhi A February 2021 Association of Foxp3 rs3761548 polymorphism with cytokines concentration in gastric adenocarcinoma patients Cytokine 138 155351 doi 10 1016 j cyto 2020 155351 ISSN 1043 4666 PMID 33127257 S2CID 226218796 a b Li Chunxiao Jiang Ping Wei Shuhua Xu Xiaofei Wang Junjie 2020 07 17 Regulatory T cells in tumor microenvironment new mechanisms potential therapeutic strategies and future prospects Molecular Cancer 19 1 116 doi 10 1186 s12943 020 01234 1 ISSN 1476 4598 PMC 7367382 PMID 32680511 Downs Canner Stephanie Berkey Sara Delgoffe Greg M Edwards Robert P Curiel Tyler Odunsi Kunle Bartlett David L Obermajer Natasa 2017 03 14 Suppressive IL 17A Foxp3 and ex Th17 IL 17AnegFoxp3 Treg cells are a source of tumour associated Treg cells Nature Communications 8 1 14649 Bibcode 2017NatCo 814649D doi 10 1038 ncomms14649 ISSN 2041 1723 PMC 5355894 PMID 28290453 Togashi Yosuke Shitara Kohei Nishikawa Hiroyoshi June 2019 Regulatory T cells in cancer immunosuppression implications for anticancer therapy Nature Reviews Clinical Oncology 16 6 356 371 doi 10 1038 s41571 019 0175 7 ISSN 1759 4782 PMID 30705439 S2CID 59526013 Ohue Yoshihiro Nishikawa Hiroyoshi July 2019 Regulatory T Treg cells in cancer Can Treg cells be a new therapeutic target Cancer Science 110 7 2080 2089 doi 10 1111 cas 14069 ISSN 1347 9032 PMC 6609813 PMID 31102428 Marson A Kretschmer K Frampton GM Jacobsen ES Polansky JK MacIsaac KD et al February 2007 Foxp3 occupancy and regulation of key target genes during T cell stimulation Nature 445 7130 931 5 Bibcode 2007Natur 445 931M doi 10 1038 nature05478 PMC 3008159 PMID 17237765 Ellis SD McGovern JL van Maurik A Howe D Ehrenstein MR Notley CA October 2014 Induced CD8 FoxP3 Treg cells in rheumatoid arthritis are modulated by p38 phosphorylation and monocytes expressing membrane tumor necrosis factor a and CD86 Arthritis amp Rheumatology 66 10 2694 705 doi 10 1002 art 38761 PMID 24980778 S2CID 39984435 Seddiki N Cook L Hsu DC Phetsouphanh C Brown K Xu Y et al June 2014 Human antigen specific CD4 CD25 CD134 CD39 T cells are enriched for regulatory T cells and comprise a substantial proportion of recall responses European Journal of Immunology 44 6 1644 61 doi 10 1002 eji 201344102 PMID 24752698 S2CID 24012204 Zaunders JJ Munier ML Seddiki N Pett S Ip S Bailey M et al August 2009 High levels of human antigen specific CD4 T cells in peripheral blood revealed by stimulated coexpression of CD25 and CD134 OX40 Journal of Immunology 183 4 2827 36 doi 10 4049 jimmunol 0803548 PMID 19635903 Poloni Chad Schonhofer Cole Ivison Sabine Levings Megan K Steiner Theodore S Cook Laura 2023 02 24 T cell activation induced marker assays in health and disease Immunology and Cell Biology 101 6 491 503 doi 10 1111 imcb 12636 ISSN 1440 1711 PMID 36825901 S2CID 257152898 Wieczorek G Asemissen A Model F Turbachova I Floess S Liebenberg V et al January 2009 Quantitative DNA methylation analysis of FOXP3 as a new method for counting regulatory T cells in peripheral blood and solid tissue Cancer Research 69 2 599 608 doi 10 1158 0008 5472 CAN 08 2361 PMID 19147574 Agarwal A Singh M Chatterjee BP Chauhan A Chakraborti A 2014 Interplay of T Helper 17 Cells with CD4 CD25 high FOXP3 Tregs in Regulation of Allergic Asthma in Pediatric Patients International Journal of Pediatrics 2014 636238 doi 10 1155 2014 636238 PMC 4065696 PMID 24995020 Lu LF Lind EF Gondek DC Bennett KA Gleeson MW Pino Lagos K et al August 2006 Mast cells are essential intermediaries in regulatory T cell tolerance Nature 442 7106 997 1002 Bibcode 2006Natur 442 997L doi 10 1038 nature05010 PMID 16921386 S2CID 686654 Tregitope Immunomodulation Power Tool EpiVax 2 August 2016 Hui DJ Basner Tschakarjan E Chen Y Davidson RJ Buchlis G Yazicioglu M et al September 2013 Modulation of CD8 T cell responses to AAV vectors with IgG derived MHC class II epitopes Molecular Therapy 21 9 1727 37 doi 10 1038 mt 2013 166 PMC 3776637 PMID 23857231 a b De Groot AS Moise L McMurry JA Wambre E Van Overtvelt L Moingeon P et al October 2008 Activation of natural regulatory T cells by IgG Fc derived peptide Tregitopes Blood 112 8 3303 11 doi 10 1182 blood 2008 02 138073 PMC 2569179 PMID 18660382 New 2 25M infusion of NIH funds for EpiVax Tregitope proposed Paradigm Shifting Treatment Fierce Biotech Research Su Y Rossi R De Groot AS Scott DW August 2013 Regulatory T cell epitopes Tregitopes in IgG induce tolerance in vivo and lack immunogenicity per se Journal of Leukocyte Biology 94 2 377 83 doi 10 1189 jlb 0912441 PMC 3714563 PMID 23729499 Cousens LP Su Y McClaine E Li X Terry F Smith R et al 2013 Application of IgG derived natural Treg epitopes IgG Tregitopes to antigen specific tolerance induction in a murine model of type 1 diabetes Journal of Diabetes Research 2013 621693 doi 10 1155 2013 621693 PMC 3655598 PMID 23710469 Cousens LP Mingozzi F van der Marel S Su Y Garman R Ferreira V et al October 2012 Teaching tolerance New approaches to enzyme replacement therapy for Pompe disease Human Vaccines amp Immunotherapeutics 8 10 1459 64 doi 10 4161 hv 21405 PMC 3660767 PMID 23095864 Cousens LP Najafian N Mingozzi F Elyaman W Mazer B Moise L et al January 2013 In vitro and in vivo studies of IgG derived Treg epitopes Tregitopes a promising new tool for tolerance induction and treatment of autoimmunity Journal of Clinical Immunology 33 1 S43 9 doi 10 1007 s10875 012 9762 4 PMC 3538121 PMID 22941509 Elyaman W Khoury SJ Scott DW De Groot AS 2011 Potential application of tregitopes as immunomodulating agents in multiple sclerosis Neurology Research International 2011 256460 doi 10 1155 2011 256460 PMC 3175387 PMID 21941651 Online Mendelian Inheritance in Man IPEX ipex at NIH UW GeneTestsExternal links editRegulatory T Cells at the U S National Library of Medicine Medical Subject Headings MeSH Portals nbsp Biology nbsp Medicine Retrieved from https en wikipedia org w index php title Regulatory T cell amp oldid 1187176150, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.