Advances in Bioscience and Biotechnology
Vol.09 No.09(2018), Article ID:87366,26 pages
10.4236/abb.2018.99031

Dissecting Emerging Aspects of Regulatory Circuitry in Man and Mice: Regulatory T Cell Biology

Tania Rahman1,2*, Md. Ferdous Seraj3, Annelise Casellato4

1Department of Biochemistry and Molecular Biology, Bio21 Molecular Science and Biotechnology Institute, University of Melbourne, Melbourne, Australia

2Department of Microbiology and Immunology, University of Melbourne at the Peter Doherty Institute for Infection and Immunity, Melbourne, Australia

3School of Civil, Environmental and Chemical Engineering, RMIT University, Melbourne, Australia

4Departamento de Química Inorgânica, Instituto de Química, Universidade Federal do Rio de Janeiro, Rio de Janeiro, Brazil

Copyright © 2018 by authors and Scientific Research Publishing Inc.

This work is licensed under the Creative Commons Attribution International License (CC BY 4.0).

http://creativecommons.org/licenses/by/4.0/

Received: June 30, 2018; Accepted: August 20, 2018; Published: September 18, 2018

ABSTRACT

Regulatory T cells (Treg), a component of adaptive immunity, are well known for their immunosuppressive roles and their ability to maintain the balance between the immunological and pathological reactions. Treg have been shown to provide protective responses and their depletion has resulted severe pathology in some pathogen infections. The work presented here has unravelled the potential of regulatory cells in the immune system including different repertoir of Treg cell subsets, markers to distinguish them, Treg suppression mechanisms in the pathogenesis of various infections and summarize different mouse models depleting Tregs. These findings would help set up future avenues of research to elucidate a key mechanism of action of these cells and provide new therapeutic insights for pathogenesis and also for broader antibacterial/antiviral/antiproliferative immunity.

Keywords:

Regulatory T Cells, Foxp3, tTreg, pTreg, IL-10, TGF-β, Helios, Neuropilin-1

1. Introduction

A hallmark of the immune system is to maintain a harmonious balance and selection among its different components. B and T cells are selected rigorously for survival during maturation phase. Among the different repertoire of cell subsets, regulatory T cells (Treg) have been established as a developmentally and functionally distinct group that has been recognised as vital for keeping the immune system in check, aiding it to escape from self-targeted pathology and unhindered selection of both T and B cell populations [1][2]. In order to maintain a delicate balance, immune system is required to respond to pathogens, nevertheless is must tolerate beneficial microbes. It is widely acknowledged that Treg play a major role in minimizing deleterious immune mediated pathology caused by harmful microbes and self-antigens and thus making a significant impact to this balance.

2. Treg

Treg are currently thoroughly investigated for their key role in the maintenance of balance between the immunological and pathological reactions over the body [3][4]. They are crucial for the retainment of self-tolerance and the control of immune responses against pathogenic organisms, tumour antigens as well as allergens [5][6][7][8][9]. Treg are developmentally and functionally different from conventional T cells. Treg constitute 5% to 10% of CD4+ T cells. In the steady state, they are generated in the thymus and can be induced from naïve CD4+ T cells in the periphery.

Treg are initially characterized as expressing a CD4+CD25high phenotype [10][11]. However, an increasing number of markers have been identified that express constitutively on Treg. These include: cytotoxic T lymphocyte antigen 4 (CTLA-4), glucocorticoid-induced tumour necrosis factor receptor family related protein (GITR); TNFRSF18 (GITR), CD39, HLA-DR, CD45RA, OX40, CD127lo, and CD73 [12]- [17]. In fact, none of these surface markers are expressed exclusively on Treg. Moreover, as CD25 is expressed on other activated T cells and there are some Treg in the peripheral tissues which do not express CD25 limiting the use of this marker for Treg [18][19][20][21]. To date the most specific marker identified for the classification of Treg is expression of the transcription factor recognized as forkhead box P3 (Foxp3) [22], which has been exhibited to be expressed specifically in CD4+ T cells. The Foxp3 gene encodes Scurffin, which is identified as a member of forkhead-winged-helix family of transcriptional regulators and is vastly conserved in humans [23]. In mice neither activated CD4+ T cells nor differentiated Th1/Th2 cells express Foxp3 [24][25]. Foxp3 is found to be expressed almost exclusively by CD4+CD25+ T cells in both thymus and periphery [26]. Mutation of Foxp3 causes an aggressive X-linked autoimmune disease in scurfy mice [23]and the human equivalent, immune dysregulation, polyendocrinopathy, enteropathy, X-linked syndrome (IPEX) [27][28]. Moreover, forced expression of Foxp3 can transform naive T cells to Treg [24][29]. Foxp3-sufficient or Foxp3-deficient mixed bone marrow chimeras study demonstrated that CD4+CD25+ cells only developed from Foxp3 sufficient bone marrow suggesting that Foxp3 is indispensible for the development of Treg [24][30]. Importantly, expression of green fluorescence protein (GFP) under the control of Foxp3 gene showed that Foxp3 expression is limited to CD4+ T cells that exhibit suppressive potency [31]. Also, ablation of Foxp3 in Treg results loss of suppressive function and phenotype [32][33]. Thus, Foxp3 appeared to be a lineage-specification factor of Treg and its expression is indispensable for the development and function of Treg.

2.1. Functions of Treg

Preliminary studies with Treg were based on their role in dominant tolerance and development of autoimmune disease. However, a handful of studies indicate that Treg play roles in the development of allergic diseases (reviewed in [34]), in the suppression of anti-tumour immunity [35], during pathogen infection (reviewed in [4]) and in controlling responses to commensal microbes in inflammatory diseases [36]. Treg are well known for their immunosuppressive role of varying immune cells including non-Treg CD4+ T cells [37], CD8+ T cells [38], dendritic cells (DC) [39], B cells [40], Th17 cells [41], natural killer (NK) cells [42], macrophages [43]and mast cells [44]which are activated in response to pathogen (Figure 1).

2.2. Different Subsets of Treg

Studies of Treg have identified several lineages of cells with different sites of induction, characterization and to a degree with various mechanisms of action. Treg are largely divided in to two major groups: thymus-derived Treg cell

Figure 1. Diverse functions of Treg. Treg control Th1 type cells which though secretion of IFNγ activate antimicrobial activity of macrophages and cytotoxic activity of CD8 T cells which lyse host cells infected with pathogen. Treg suppress Th2 type activity which through secretion of IL-4, 5 and 6 activates naïve B cells to produce antibodies as well as controlled recruitment of eosinophils. In addition, Treg alter neutrophil activity by inhibiting their recruitment on the epithelial cell membrane.

(tTreg) and peripherally derived Treg cell (pTreg) (Table 1) [45][46][47][48].

2.2.1. Thymus Derived Treg (tTreg)

Treg that develop within the thymus were often referred to as natural Treg cells as opposed to peripherally derived Treg [46]. tTreg arise during the normal process of T cell development in the thymus and are imprinted with regulatory function before being released into the periphery. They help prevent autoimmunity and are resistant to thymic deletion and are non-redundantly required for the establishment of self-tolerance [49][50]. tTreg are known to suppress effector T cell proliferation in vitro through cytokine independent fashion that is mostly cell contact dependent [51]. The proportion of tTreg appears to be small when any part of this population reacts with antigens [52]. In thymus, tTreg can develop from progenitors to tTreg through tTreg precursor cells [53][54]. Resting tTreg express CD45RA+Foxp3low, while activated tTreg express CD45RO+Foxp3high

Table 1. Different subsets of regulatory T cells.

[50][55]. tTreg constitute about 5% - 10% of mouse peripheral CD4 cells which is corresponding to 1% - 2% of human counterpart [1].

tTreg develop as a consequence of high-affinity interactions with MHC class II-peptide and their TCR repertoire is primarily self-reactive [56][57]. They require CD28 co-stimulation for their development in thymus as mice lacking CD28 have an approximately 80% decrease in the frequency of tTreg [58][59][60]. IL-2, a member of γc cytokine family is important for thymic induction of tTreg as mice lacking IL-2Rβ (CD122) develop spontaneous autoimmune diseases which can be blocked by administration of donor Treg [61][62]. Signal transducer and activator of transcription 5 (STAT5) has been demonstrated to increase the in vivo frequency of tTreg and expression of STAT5 is regulated by IL-2 [63]. IL-15 is also essential for induction of Foxp3 expression in thymocytes [64]. Transcription factor c-Rel, a member of NF-κB family has been shown to be highly expressed by tTreg. c-Rel binds to the conserved non-coding sequence 3 (CNS3) region of Foxp3 locus and in mice deficient in c-Rel, tTreg numbers are markedly reduced [65][66]. TGF-β has been suggested to be absolutely engaged in Foxp3 induction during tTreg development. However, its role has been controversial in tTreg development [67][68].

Medullary thymic epithelial cells (mTECs) are responsible for inducing selection of thymic tTreg, also bone marrow derived APCs can facilitate tTreg differentiation [2]. Either mTECs or bone derived APCs alone may be sufficient for the generation of tTreg numbers and both of these subsets can present self-antigens in order to induce development of tTReg [69].

2.2.2. Peripherally Derived Treg Cell

pTreg develop extra-thymically when they are exposed to certain regulatory cytokines which are released during inflammatory conditions, upon encountering cognate antigens and costimulation [70][71]. The antigen-specific Treg repertoire is more abundant than tTreg population and in contrast to tTreg, they migrate towards the sites of inflammation [72]. pTreg play an indispensible role in establishing peripheral tolerance to commensal microbes in gut and non-pathogenic environmental antigens derived from food and they express predominantly in mucosa associated lymphoid tissues (MALT) including peyer’s patches and lamina propria of small and large intestines [73][74]. Moreover, pTreg which present only in the placental mammals are involved in the establishment of maternal-foetal tolerance [75]. Their suppressive actions are mostly cell contact independent and depend on the availability of immunosuppressive cytokines [76][77]. They are known to develop from CD4+Foxp3 effector T cells in the periphery and they apparently have a similar TCR repertoire to that of the effector T cells [78][79]. In particular, pTreg can convert from CD4+Foxp3 to CD4+Foxp3, thereby expanding the range of Treg specificities to exogenous antigens [80]. Conversely, pTreg can revert to effector T cells, losing expression of Foxp3 under certain situations and thus pTreg are not irreversibly programmed [81].

TGF-β and its receptor signal pathway is essential for the generation of pTreg as diminishing TGF-β receptor signaling blocks the induction of Foxp3 expression and the resulting operational suppressive capability [82][83][84]. TGF-β downregulates the expression of growth factor independent 1, a transcriptional repressor that inhibits Treg differentiation [85]. TGF-β also antagonizes DNA methyltransferase 1 (Dnmt1), which inhibits the expression of Foxp3 [86]. CNS-1, a non-coding region of Foxp3 locus are a crucial regulatory element in the generation of pTreg [87]. CNS1, which possesses a TGFβ-NFAT response element are involved in pTreg differentiation in GALT [88].

IL-2 play a vital part in the differentiation of Foxp3+ pTreg as TGF-β fails to induce Foxp3+ pTreg from naïve CD4+CD25 cells in mice deficient in IL-2 [89][90][91]. Downstream STAT5-dependent IL2 signaling is also essential for the differentiation of pTreg [92][93][94].

Retinoic acid (RA), a vitamin A metabolite plays a significant role in enhancing the generation of Foxp3 Treg cells in the GALT [95]. However, the molecular mechanism by which RA promotes TGFβ-mediated Foxp3 induction has controversies. RA has been demonstrated to enhance Foxp3 expression by promoting the expression of Smad 3, thereby amplifying Foxp3 transcription [96][97]. However, another study has suggested that even though RA appears to increase Smad 3, it augments Treg cell translation independently of Smad 3 [98].

Smad 2 and Smad 3 participate in the pTreg differentiation process and they are activated through TGF-β signaling pathways by inducing Foxp3 [99][100]. Smad 2 and 3 induces the expression of TGF-β induced transcription factor, TGF-β-inducible early gene 1 product (TIEG1), which induces Foxp3 and its transcription [96]. In addition, Smad 2 and Smad 3 are also appeared to stimulate differentiation of pTreg though Foxp3-independent pathway [101].

Besides pTreg, there are other types of Treg including Tr1 cells (T regulatory type 1) and Th3 Treg [114][115]. Tr1 population are disturbed in individuals with prolonged inflammatory conditions like colitis, arthritis and asthma. IL-10 is the major cytokine involved in Tr1 cells differentiation pathway [102]. Once mature, they exert their suppressive activity through the production of large amounts of IL-10 [46]. Usually Tr1 cells do not express Foxp3 constitutively, however, when activated, they can induce Foxp3 expression [116]. There is accumulating evidence that Tr1 cells do not require Foxp3 expression in order to exert their suppressive potency, as Tr1 cells suppress conventional T cells independent of Foxp3 expression [117][118]. In addition, it is revealed that Tr1 cells may be differentiated from naive T cells in patients with IPEX disease [119]. An essential growth factor for the development of Tr1 is IL-15, which can mediate Tr1 cell differentiation without TCR triggering [103][104].

IL10 and IL27 are the major cytokines involved in the differentiation of Tr1 cells [105][106]. IL27 signaling results activation of transcription factors including c-Maf, IL21, and the costimulatory receptor ICOS [105]. c-Maf is the key factor, which activates IL21 production. IL21, an autocrine growth factor drives the proliferation of Tr1 cells [107]. ICOS stimulates the IL27-induced differentiation of Tr1. Currently, the aryl hydrocarbon receptor (AhR), also induced by IL27, was shown to be involved in the differentiation of Tr1 [108]. c-Maf and AhR perform synergistically to facilitate proliferation of Tr1 [109]. In the case of Tr1, as IL-10 can also be produced by both Th1 (IFNγ+) and Th2 (IL-4+) effectors, definition of this subset is relatively fluid [120].

Th3 cells were first recognized due to their role in immune tolerance after oral ingestion of antigens [110]. Th3 mediate their suppressive activity by the production of TGF-β. Mice deficient in Th3 often develop spontaneous autoimmunity and Th3 exhibit a mutual relationship with Th17 cells [111][112]. Th17 being proinflammatory are essential in autoimmune related disorders [121][122]. No specific surface marker has been identified so far for Th3 cells, though Th3 is induced by Foxp3 expression.

Apart from Foxp3-expressing Treg there are other functional regulatory cells, which are involved in IL-35 production (Tr35) [113]. While Treg are commonly found to be CD4+, CD8+ T cells might express Foxp3 and produce the same suppressive cytokines [123].

2.2.3. Markers to Distinguish tTreg from pTreg

The tTreg vary in functionality from pTreg, however, the major problem in exploring the relative function of these two subsets is the lack of markers to distinguish the two populations.

Helios, a T cell restricted member of Ikaros family transcription factor has been demonstrated to be useful for distinguishing these two Treg populations [124][125]. Helios is preferentially expressed by tTreg, whereas Treg generated in vitro and in vivo are negative for Helios expression [124]. However, others have described that pTreg can express Helios as when T cells were stimulated with irradiated splenocytes, more than half of the pTreg expressed Helios [126]. In addition, Helios is also expressed on Th2 and T follicular helper cells and is related to the differentiation of these cells [127]. It is reported that activated Foxp3 T cells also express helios and in fact a marker for Treg activation [128]. Also expression of Helios has been shown to be associated with T cell tolerance in both thymus and periphery [129]. Thus Helios does not appear to be a marker to distinguish pTreg from tTreg.

Neuropilin-1 (Nrp1) expression, a neuronal receptor of the class 3 semaphorin subfamily and a co-receptor for vascular endothelial growth factor A, offers another biological marker to distinguish tTreg from pTreg [130][131]. In contrast to low levels of Nrp1 expressed on pTreg, majority of tTreg express Nrp1 [130]. However, others have demonstrated that Treg generated extrathymically in the central nervous system during a spontaneous model of EAE were Nrp+ [125]. In addition, Nrp1 is not a marker for human Treg [132]. Thus Nrp1 appeared to be an imperfect marker to distinguish pTreg from tTreg.

Lap, a component of latent TGFβ and Garp (Lrrc32), a membrane anchoring molecule that binds to latent TGFβ, were found to be expressed on the surface of Treg [133]. While Lap and Garp expression selectively identifies tTreg that represent a stable subset with highly potent suppressive ability, pTreg fail to express surface Lap or Garp [134]. Recently transcription factor Krϋppel-like factor 2 (KLF2) has been demonstrated for the generation of pTreg [135]. KLF2 is not required for the generation of tTreg and is only necessary for pTreg development. In addition, it has been exhibited that drugs that block KLF2 proteolysis during T cell activation augment pTreg development.

2.3. Mechanism of Suppression of Treg

To date a number of mechanisms have been associated with the suppressive action of Treg suppression. Treg execute their suppressive function rapidly when they are activated via the T cell receptor, either specifically by its natural class II ligand, or by foreign antigens that are cross-reactive to self-antigen receptors in the periphery [136].

IL-10 and TGF-β are potent immune-suppressants that facilitate Treg to inhibit Th1 inflammatory responses. IL-10 has been illustrated as the key components in the suppressive function of Treg (Figure 2) [137]. Secretion of IL-10 by Treg cells has been shown to be essential for the prevention of experimental

Figure 2. Potential mechanisms used by Treg cells. Illustration of various Treg cell suppressive mechanisms based on three basic modes of action: A. Inhibitory cytokines consists of TGFβ, IL-35 and IL-10. B. Modulation of DC includes mechanisms that influences DC maturation and function such as LAG3, MHC-II-mediated suppression of DC maturation, and CTLA4?CD80/CD86-mediated induction of IDO, which is an immunosuppressive molecule made by DCs. C. Cytolysis consists of granzyme A and granzyme B dependent and perforin pore dependent killing mechanisms. The concept of the figure emerged from [168].

autoimmune encephalomyelitis [138], colitis [139][140], airway allergic response in lung [141]and skin hypersensitivity [142].

TGF-β, is a well-recognized inhibitory cytokine, however, the function of TGF-β in Treg mediated suppression has some controversies (Figure 2). Mice with a T-cell specific deletion of tgfb1 gene developed lethal immunopathology and could not inhibit inflammatory bowel disease in a transfer model [143]. TGF-β was found to be the major mechanism of suppression of prostate tumour infiltrating CD8 cells [144].

Membrane-tethered TGF-β expressed by Treg cells has also been demonstrated by Treg cell mediated suppression [145][146]. However, another study has demonstrated conflicting findings that membrane-tethered TGF-β is dispensable for Treg function [147].

Another cytokine, IL-35, a member of IL-12 family has also been attributed to the in vitro and in vivo suppressive potency of Treg [148]. IL-35 consists of Epstein-Barr virus induced gene 3 (Ebi3) and IL-12α/p35, both are highly expressed by Treg but not by naïve or activated T cells [113][149]. Treg cells from Ebi3 and IL-12α-deficient mice have a reduced suppressive capability in vitro and fail to cure inflammatory bowel disease in vivo [148].

There is evidence that Treg can indirectly suppress other cells through modulation of DC function (Figure 2). Treg cells suppress the capacity of DCs to activate T cells by down-regulating costimulatory molecules CD80 and CD86 [150][151]. A handful of molecules expressed at high levels in Treg have been identified which facilitate the blockade of DC maturation and DC mediated activation of effector cells. Among them the most well-known is CTLA4. CTLA4 a member of CD28/B7 family and CTLA4 mediated immunosuppression has been demonstrated in several settings including autoimmune diseases and different tumour types [152][153]. Mice with a Treg cell specific deletion of CTLA-4 develop systemic autoimmunity and CTLA4-deficient mice inhibit upregulation of CD80 and CD86 even in the presence of strong DC maturation stimuli [152][154][155][156]. Moreover, anti-CTLA4 inhibits the ability of Treg cells to suppress colitis [154].

Treg cells can induce the expression of indoleamine 2, 3-dioxygenase (IDO) in DCs in a CTLA4-dependent manner [157]. IDO is a potent immunosuppressive enzyme which induces catabolism of tryptophan in to pro-apoptotic metabolites, resulting in the suppression of effector cells [158]. Ligation of CTLA-4 to CD80 and CD86 induce DCs by activating transcription factor Foxo3, which downregulate production of IL-6 and TNFα by DCs [159].

Other molecules expressed by Treg that can affect the function of DCs are: Lymphocyte activation gene-3 (LAG-3) [160][161], immunoreceptor tyrosine-based activation motif (ITAM) [162]and Fibrinogen-like protein 2 (FGL2) [163]. Treg cells from LAG-3 deficient mice had a decreased suppressive activity and anti-LAG-3 blocked suppressive activity of Treg cells both in vitro and in vivo [160][164].

Treg cells can suppress DC maturation by inhibiting down-regulation of MARCH1 and upregulation of CD83 [165]. MARCH1, a membrane bound E3 ubiquitin ligase, was found to degrade CD86 and MHC-II on DC by directing them to the late endosomal or lysosomal compartment [166]. CD83 is accompanied by DC maturation and it inhibits the action of MARCH1 [167]. Treg use IL-10 to influence MARCH1-CD83 pathway.

There is growing evidence that Programmed death 1 (PD-1) pathway plays a role in Treg mediated suppression [169][170]. PD-1 is expressed predominantly on exhausted CD8 cells and PD-L1, the ligand for PD-1is expressed on Treg and tumour cells [171]. In a BL/6 model, Treg from tumour draining lymph node suppress via PD-1/PD-L1 pathway. In samples of T cells taken from melanoma patients, PD-1 blockade was found to enhance effector T cell proliferation and inhibit the suppressive function of PD-L1 expressing Treg [172]. Treg have also been shown to mediate their suppressive activity through Fas-FasL interaction resulting in apoptosis of B cells [173].

Treg are able to mediate immune suppression by triggering direct cytolysis of target cells [174][175][176]. Granzyme and perforin pathway are two suggested mechanisms for Treg mediated immune suppression (Figure 2). Treg cells from mice deficient in granzyme B had a reduced suppressive activity [177]. Also granzyme B deficient mice were able to clear tumour more effectively than wild type mice since Treg cells inhibit anti-tumour immunity in a granzyme B and perforin dependent manner [178].

Another mechanism of suppression is the induction of metabolic disruption in target cells. Treg can be multiplied significantly by TCR stimulation in presence of high concentrations of IL-2 [179][180]. As IL-2 is secreted by Th1 type cells, and elevated level of IL-2 acts as a negative-feedback mechanism thereby halts the effector T cell activity by the expanding Treg [181].

Treg has been found to exert metabolic disruption in target cells by their expression of ectoenzymes CD39 and CD73. Treg from CD39 deficient mice have reduced suppressive functions both in vitro and in vivo [182]. CD39 and CD73 convert proinflammatory nucleotides to anti-inflammatory adenosine and these ectoenzymes catalyse the generation of perinuclear adenosine from extracellular ATP or ADP [183]. Adenosine generated in this process then suppresses effector T cells via binding to adenosine A2A receptor.

IL-17 and IFN-γ has been shown to modulate the suppressive capacity of Treg on Th2 immune responses [184][185].

IRF4 a transcription factor responsible for the differentiation of Th2 cells has been shown to play a role in the in Treg mediated suppression of Th2 responses. Mice wherein irf4 is depleted in Treg developed a lymphoproliferative disease that occurs due to a selective dysregulation of Th2 cells [186]. T-bet, a transcription factor is indispensible for Th1 cell differentiation. It has been demonstrated that Treg mediated suppression of Th1 is regulated by Treg expression of T-bet [187]. Treg mediated suppression of Th17 responses is regulated by Treg expression of STAT3. Selective ablation of STAT3 in Treg cells results selective uncontrolled Th17 dependent pathology of intestinal mucosa [41]. Other researchers have described a subset of Foxp3+Treg that express the Tfh cell master transcriptional regulator, Bcl-6/Blimp-1 and accumulation at sites of B cell germinal centre responses where they function to control these reactions [188][189][190][191].

2.4. Depletion of Treg Using Mouse Model

A number of different methods to deplete Treg depletion have been used to study their function and variable degrees of depletion have been seen. The two commonly used methods are: antibody depletion method and transgenic mouse model depletion. With the recognition of the IL-2 receptor CD25 on Treg, antibodies specific to CD25 have been developed. About 65% - 70% of Treg populations are depleted following introduction of anti-CD25 antibody clone i.e., PC-61 [192]. Also, a population of CD25Foxp3+ Treg cells cannot be depleted using PC61. Importantly, expression of CD25 is induced on activated conventional CD4+T cells [18][19][20][21].

To deal with the issue of in vivo function of Treg in immunopathology, mice with fully depleted Treg have been engineered. These mice permit selective ablation of Foxp3+ Treg while without impairing CD25+ effector T cells. The genetic introduction of human diphtheria toxin receptor (DTR) targeted to a specific murine cell type has become a prevailing tool to selectively ablate Treg upon DT injection [193]. Rudensky and colleagues generated a knock-in mice called Foxp3DTR mice in which human DTR is introduced in to the 3́ untranslated region of Foxp3 which allow specific depletion of Treg by injecting DT. This model showed more than 97% depletion of Foxp3+ cells after DT administration [194]. However, these mice subjected to DT succumb to catastrophic autoimmune disease within 2 - 3 weeks, as a result of massive expansion of diverse array of immune cells.

Suffner et al generated a novel set of BAC (Bacterial Artificial Chromosome) transgenic mouse, called Foxp3.LuciDTR mouse, in which Treg express luciferase and the human DTR [195]. They developed several founder lines with different degrees of Treg depletion such as Foxp3.LuciDTR-3 and Foxp3.LuciDTR-4 and Foxp3.LuciDTR-5. Mice from lines Foxp3.LuciDTR-3, Foxp3.LuciDTR-4 and Foxp3.LuciDTR-5 exhibited ~75%, 90% and >95% depletion of Tregs, respectively, while Foxp3.LuciDTR-5 mice did suffer from wasting disease due to autoimmunity. Lahl and Sparwasser [193]produced another BAC technology based transgenic mouse model known as DEpletion of REGulatory T cell (DEREG) mice, which express a DTR fused with GFP protein under the control of an additional Foxp3 promoter, allowing efficient depletion of Foxp3+ Treg by DT injection. Depletion in DEREG mice led to ablation of Foxp3+ Treg to 95% - 98% [196].

3. Concluding Remarks

There are complex dynamic suppression processes controlling other cells by Treg and that involve various inhibitory cytokines, DC function and cytolysis. These mechanisms are different depending upon Treg cell subtypes and the nature of the pathogens. Also they can act together or independently, according to the requirements of the immune system and homeostasis maintenance, or during the progression of various pathological processes. This gives two layers of reinforcement of Treg function: the suppressive effects of Treg and protective potential of Treg against different pathogens and they also alter the functional Treg pool in response to tissue and inflammatory cues. Together, these data suggest Treg mediated suppression might be a valuable component in different pathogens, which might provide novel therapeutic approaches for vaccination against them. Further investigations are required on Treg in different type of infection models in order to extend the current understanding on the salient features of these cells, their either beneficial or detrimental role during infection and their mechanisms participating to the immunity against various pathogens.

Conflicts of Interest

No potential conflicts of interest relevant to this article were reported.

Cite this paper

Rahman, T., Seraj, Md.F. and Casellato, A. (2018) Dissecting Emerging Aspects of Regulatory Circuitry in Man and Mice: Regulatory T Cell Biology. Advances in Bioscience and Biotechnology, 9, 443-468. https://doi.org/10.4236/abb.2018.99031

References

  1. 1. Belkaid, Y. (2007) Regulatory T Cells and Infection: A Dangerous Necessity. Nature Reviews Immunology, 7, 875-888. https://doi.org/10.1038/nri2189

  2. 2. Povoleri, G.A., et al. (2013) Thymic versus Induced Regulatory T Cells—Who Regulates the Regulators? Frontiers in Immunology, 4, 169. https://doi.org/10.3389/fimmu.2013.00169

  3. 3. Liston, A. and Gray, D.H. (2014) Homeostatic Control of Regulatory T Cell Diversity. Nature Reviews Immunology, 14, 154-165. https://doi.org/10.1038/nri3605

  4. 4. Maizels, R.M. and Smith, K.A. (2011) Regulatory T Cells in Infection. Advances in Immunology, 112, 73-136. https://doi.org/10.1016/B978-0-12-387827-4.00003-6

  5. 5. Mittrucker, H.W. and Kaufmann, S.H. (2004) Mini-Review: Regulatory T Cells and Infection: Suppression Revisited. European Journal of Immunology, 34, 306-312. https://doi.org/10.1002/eji.200324578

  6. 6. Sakaguchi, S., et al. (2001) Immunologic Tolerance Maintained by CD25+ CD4+ Regulatory T Cells: Their Common Role in Controlling Autoimmunity, Tumor Immunity, and Transplantation Tolerance. Immunological Reviews, 182, 18-32. https://doi.org/10.1034/j.1600-065X.2001.1820102.x

  7. 7. Belkaid, Y. and Rouse, B.T. (2005) Natural Regulatory T Cells in Infectious Disease. Nature Immunology, 6, 353-360. https://doi.org/10.1038/ni1181

  8. 8. Umetsu, D.T. and De Kruyff, R.H. (2006) The Regulation of Allergy and Asthma. Immunological Reviews, 212, 238-255. https://doi.org/10.1111/j.0105-2896.2006.00413.x

  9. 9. Fehervari, Z. and Sakaguchi, S. (2004) CD4+ Tregs and Immune Control. Journal of Clinical Investigation, 114, 1209-1217. https://doi.org/10.1172/JCI200423395

  10. 10. Baecher-Allan, C., et al. (2003) CD4+CD25+ Regulatory Cells from Human Peripheral Blood Express Very High Levels of CD25 ex Vivo. Novartis Foundation Symposium, 252, 67-91, 106-114.

  11. 11. Baecher-Allan, C., et al. (2001) CD4+CD25 High Regulatory Cells in Human Peripheral Blood. The Journal of Immunology, 167, 1245-1253. https://doi.org/10.4049/jimmunol.167.3.1245

  12. 12. Itoh, M., et al. (1999) Thymus and Autoimmunity: Production of CD25+CD4+ Naturally Anergic and Suppressive T Cells as a Key Function of the Thymus in Maintaining Immunologic Self-Tolerance. The Journal of Immunology, 162, 5317-5326.

  13. 13. Yi, H., et al. (2006) The Phenotypic Characterization of Naturally Occurring Regulatory CD4+CD25+ T Cells. Cellular & Molecular Immunology, 3, 189-195.

  14. 14. Deaglio, S., et al. (2007) Adenosine Generation Catalyzed by CD39 and CD73 Expressed on Regulatory T Cells Mediates Immune Suppression. The Journal of Experimental Medicine, 204, 1257-1265. https://doi.org/10.1084/jem.20062512

  15. 15. Fontenot, J.D. and Rudensky, A.Y. (2004) Molecular Aspects of Regulatory T Cell Development. Seminars in Immunology, 16, 73-80. https://doi.org/10.1016/j.smim.2003.12.002

  16. 16. Ustun, C., Miller, J.S., Munn, D.H., Weisdorf, D.J. and Blazar, B.R. (2011) Regulatory T Cells in Acute Myelogenous Leukemia: Is It Time for Immunomodulation? Blood, 118, 5084-5095. https://doi.org/10.1182/blood-2011-07-365817

  17. 17. Piccirillo, C.A. and Shevach, E.M. (2004) Naturally-Occurring CD4+ CD25+ Immunoregulatory T Cells: Central Players in the Arena of Peripheral Tolerance. Seminars in Immunology, 16, 81-88. https://doi.org/10.1016/j.smim.2003.12.003

  18. 18. Leithauser, F., Meinhardt-Krajina, T., Fink, K., Wotschke, B., Moller, P. and Reimann, J. (2006) Foxp3-Expressing CD103+ Regulatory T Cells Accumulate in Dendritic Cell Aggregates of the Colonic Mucosa in Murine Transfer Colitis. American Journal of Pathology, 168, 1898-1909. https://doi.org/10.2353/ajpath.2006.050228

  19. 19. Needham, D.J., Lee, J.X. and Beilharz, M.W. (2006) Intra-Tumoural Regulatory T Cells: A Potential New Target in Cancer Immunotherapy. Biochemical and Biophysical Research Communications, 343, 684-691. https://doi.org/10.1016/j.bbrc.2006.03.018

  20. 20. Sakaguchi, S. (2005) Naturally Arising Foxp3-Expressing CD25+ CD4+ Regulatory T Cells in Immunological Tolerance to Self and Non-Self. Nature Immunology, 6, 345-352. https://doi.org/10.1038/ni1178

  21. 21. Buckner, J.H. (2010) Mechanisms of Impaired Regulation by CD4(+) CD25(+) FOXP3(+) Regulatory T Cells in Human Autoimmune Diseases. Nature Reviews Immunology, 10, 849-859. https://doi.org/10.1038/nri2889

  22. 22. Fontenot, J.D. and Rudensky, A.Y. (2005) A Well Adapted Regulatory Contrivance: Regulatory T Cell Development and the Forkhead Family Transcription Factor Foxp3. Nature Immunology, 6, 331-337. https://doi.org/10.1038/ni1179

  23. 23. Brunkow, M.E., et al. (2001) Disruption of a New Forkhead/Winged-Helix Protein, Scurfin, Results in the Fatal Lymphoproliferative Disorder of the Scurfy Mouse. Nature Genetics, 27, 68-73. https://doi.org/10.1038/83784

  24. 24. Fontenot, J.D., Gavin, M.A. and Rudensky, A.Y. (2003) Foxp3 Programs the Development and Function of CD4+ CD25+ Regulatory T Cells. Nature Immunology, 4, 330-336. https://doi.org/10.1038/ni904

  25. 25. Hori, S., Nomura, T. and Sakaguchi, S. (2003) Control of Regulatory T Cell Development by the Transcription Factor Foxp3. Science, 299, 1057-1061. https://doi.org/10.1126/science.1079490

  26. 26. Fontenot, J.D., Rasmussen, J.P., Williams, L.M., Dooley, J.L., Farr, A.G. and Rudensky, A.Y. (2005) Regulatory T Cell Lineage Specification by the Forkhead Transcription Factor Foxp3. Immunity, 22, 329-341. https://doi.org/10.1016/j.immuni.2005.01.016

  27. 27. Bennett, C.L., et al. (2001) The Immune Dysregulation, Polyendocrinopathy, Enteropathy, X-Linked Syndrome (IPEX) Is Caused by Mutations of FOXP3. Nature Genetics, 27, 20-21. https://doi.org/10.1038/83713

  28. 28. Wildin, R.S., et al. (2001) X-Linked Neonatal Diabetes Mellitus, Enteropathy and Endocrinopathy Syndrome Is the Human Equivalent of Mouse Scurfy. Nature Genetics, 27, 18-20. https://doi.org/10.1038/83707

  29. 29. Khattri, R., Cox, T., Yasayko, S.A. and Ramsdell, F. (2003) An Essential Role for Scurfin in CD4+ CD25+ T Regulatory Cells. Nature Immunology, 4, 337-342. https://doi.org/10.1038/ni909

  30. 30. Lu, L.F. and Rudensky, A. (2009) Molecular Orchestration of Differentiation and Function of Regulatory T Cells. Genes & Development, 23, 1270-1282. https://doi.org/10.1101/gad.1791009

  31. 31. Wan, Y.Y. and Flavell, R.A. (2005) Identifying Foxp3-Expressing Suppressor T Cells with a Bicistronic Reporter. Proceedings of the National Academy of Sciences, 102, 5126-5131. https://doi.org/10.1073/pnas.0501701102

  32. 32. Williams, L.M. and Rudensky, A.Y. (2007) Maintenance of the Foxp3-Dependent Developmental Program in Mature Regulatory T Cells Requires Continued Expression of Foxp3. Nature Immunology, 8, 277-284. https://doi.org/10.1038/ni1437

  33. 33. Rudensky, A.Y. (2011) Regulatory T Cells and Foxp3. Immunological Reviews, 241, 260-268. https://doi.org/10.1111/j.1600-065X.2011.01018.x

  34. 34. Seroogy, C.M. and Gern, J.E. (2005) The Role of T Regulatory Cells in Asthma. Journal of Allergy and Clinical Immunology, 116, 996-999. https://doi.org/10.1016/j.jaci.2005.07.015

  35. 35. Nishikawa, H. and Sakaguchi, S. (2010) Regulatory T Cells in Tumor Immunity. International Journal of Cancer, 127, 759-767. https://doi.org/10.1002/ijc.25429

  36. 36. Arpaia, N., et al. (2013) Metabolites Produced by Commensal Bacteria Promote Peripheral Regulatory T-Cell Generation. Nature, 504, 451-455. https://doi.org/10.1038/nature12726

  37. 37. Aandahl, E.M., Michaelsson, J., Moretto, W.J., Hecht, F.M. and Nixon, D.F. (2004) Human CD4+ CD25+ Regulatory T Cells Control T-Cell Responses to Human Immunodeficiency Virus and Cytomegalovirus Antigens. Journal of Virology, 78, 2454-2459. https://doi.org/10.1128/JVI.78.5.2454-2459.2004

  38. 38. McNally, A., Hill, G.R., Sparwasser, T., Thomas, R. and Steptoe, R.J. (2011) CD4+ CD25+ Regulatory T Cells Control CD8+ T-Cell Effector Differentiation by Modulating IL-2 Homeostasis. Proceedings of the National Academy of Sciences, 108, 7529-7534. https://doi.org/10.1073/pnas.1103782108

  39. 39. Terme, M., Chaput, N., Combadiere, B., Ma, A., Ohteki, T. and Zitvogel, L. (2008) Regulatory T Cells Control Dendritic Cell/NK Cell Cross-Talk in Lymph Nodes at the Steady State by Inhibiting CD4+ Self-Reactive T Cells. The Journal of Immunology, 180, 4679-4686. https://doi.org/10.4049/jimmunol.180.7.4679

  40. 40. Iikuni, N., Lourenco, E.V., Hahn, B.H. and La Cava, A. (2009) Cutting Edge: Regulatory T Cells Directly Suppress B Cells in Systemic Lupus Erythematosus. The Journal of Immunology, 183, 1518-1522. https://doi.org/10.4049/jimmunol.0901163

  41. 41. Chaudhry, A., et al. (2009) CD4+ Regulatory T Cells Control TH17 Responses in a Stat3-Dependent Manner. Science, 326, 986-991. https://doi.org/10.1126/science.1172702

  42. 42. Ghiringhelli, F., Ménard, C., Martin, F. and Zitvogel, L. (2006) The Role of Regulatory T Cells in the Control of Natural Killer Cells: Relevance during Tumor Progression. Immunological Reviews, 214, 229-238. https://doi.org/10.1111/j.1600-065X.2006.00445.x

  43. 43. Taams, L.S., et al. (2005) Modulation of Monocyte/Macrophage Function by Human CD4+ CD25+ Regulatory T Cells. Human Immunology, 66, 222-230. https://doi.org/10.1016/j.humimm.2004.12.006

  44. 44. Gri, G., et al. (2008) CD4+ CD25+ Regulatory T Cells Suppress Mast Cell Degranulation and Allergic Responses through OX40-OX40L Interaction. Immunity, 29, 771-781. https://doi.org/10.1016/j.immuni.2008.08.018

  45. 45. Jonuleit, H. and Schmitt, E. (2003) The Regulatory T Cell Family: Distinct Subsets and Their Interrelations. The Journal of Immunology, 171, 6323-6327. https://doi.org/10.4049/jimmunol.171.12.6323

  46. 46. Curotto de Lafaille, M.A. and Lafaille, J.J. (2009) Natural and Adaptive Foxp3+ Regulatory T Cells: More of the Same or a Division of Labor? Immunity, 30, 626-635. https://doi.org/10.1016/j.immuni.2009.05.002

  47. 47. Abbas, A.K., et al. (2013) Regulatory T Cells: Recommendations to Simplify the Nomenclature. Nature Immunology, 14, 307-308. https://doi.org/10.1038/ni.2554

  48. 48. Shevach, E.M. and Thornton, A.M. (2014) tTregs, pTregs, and iTregs: Similarities and Differences. Immunological Reviews, 259, 88-102. https://doi.org/10.1111/imr.12160

  49. 49. Bayer, A.L., Yu, A., Adeegbe, D. and Malek, T.R. (2005) Essential Role for Interleukin-2 for CD4(+) CD25(+) T Regulatory Cell Development during the Neonatal Period. The Journal of Experimental Medicine, 201, 769-777. https://doi.org/10.1084/jem.20041179

  50. 50. Lim, H.W., Broxmeyer, H.E. and Kim, C.H. (2006) Regulation of Trafficking Receptor Expression in Human Forkhead Box P3+ Regulatory T Cells. The Journal of Immunology, 177, 840-851. https://doi.org/10.4049/jimmunol.177.2.840

  51. 51. Shevach, E.M. (2009) Mechanisms of Foxp3+ T Regulatory Cell-Mediated Suppression. Immunity, 30, 636-645. https://doi.org/10.1016/j.immuni.2009.04.010

  52. 52. Rubtsov, Y.P., et al. (2010) Stability of the Regulatory T Cell Lineage in Vivo. Science, 329, 1667-1671. https://doi.org/10.1126/science.1191996

  53. 53. Engel, M., Sidwell, T., Vasanthakumar, A., Grigoriadis, G. and Banerjee, A. (2013) Thymic Regulatory T Cell Development: Role of Signalling Pathways and Transcription Factors. Clinical & Developmental Immunology, 2013, Article ID: 617595. https://doi.org/10.1155/2013/617595

  54. 54. Caramalho, I., Nunes-Cabaco, H., Foxall, R.B. and Sousa, A.E. (2015) Regulatory T-Cell Development in the Human Thymus. Frontiers in Immunology, 6, 395. https://doi.org/10.3389/fimmu.2015.00395

  55. 55. Beissert, S., Schwarz, A. and Schwarz, T. (2006) Regulatory T Cells. Journal of Investigative Dermatology, 126, 15-24. https://doi.org/10.1038/sj.jid.5700004

  56. 56. Jordan, M.S., et al. (2001) Thymic Selection of CD4+ CD25+ Regulatory T Cells Induced by an Agonist Self-Peptide. Nature Immunology, 2, 301-306. https://doi.org/10.1038/86302

  57. 57. Horwitz, D.A., Zheng, S.G. and Gray, J.D. (2008) Natural and TGF-Beta-Induced Foxp3(+) CD4(+) CD25(+) Regulatory T Cells Are Not Mirror Images of Each Other. Trends in Immunology, 29, 429-435. https://doi.org/10.1016/j.it.2008.06.005

  58. 58. Salomon, B., et al. (2000) B7/CD28 Costimulation Is Essential for the Homeostasis of the CD4+ CD25+ Immunoregulatory T Cells That Control Autoimmune Diabetes. Immunity, 12, 431-440. https://doi.org/10.1016/S1074-7613(00)80195-8

  59. 59. Vang, K.B., et al. (2010) Cutting Edge: CD28 and c-Rel-Dependent Pathways Initiate Regulatory T Cell Development. The Journal of Immunology, 184, 4074-4077. https://doi.org/10.4049/jimmunol.0903933

  60. 60. Lio, C.W., Dodson, L.F., Deppong, C.M., Hsieh, C.S. and Green, J.M. (2010) CD28 Facilitates the Generation of Foxp3(-) Cytokine Responsive Regulatory T Cell Precursors. The Journal of Immunology, 184, 6007-6013. https://doi.org/10.4049/jimmunol.1000019

  61. 61. Malek, T.R., Yu, A., Vincek, V., Scibelli, P. and Kong, L. (2002) CD4 Regulatory T Cells Prevent Lethal Autoimmunity in IL-2Rbeta-Deficient Mice. Implications for the Nonredundant Function of IL-2. Immunity, 17, 167-178. https://doi.org/10.1016/S1074-7613(02)00367-9

  62. 62. Suzuki, H., et al. (1995) Deregulated T Cell Activation and Autoimmunity in Mice Lacking Interleukin-2 Receptor Beta. Science, 268, 1472-1476. https://doi.org/10.1126/science.7770771

  63. 63. Zorn, E., et al. (2006) IL-2 Regulates FOXP3 Expression in Human CD4+ CD25+ Regulatory T Cells through a STAT-Dependent Mechanism and Induces the Expansion of These Cells in Vivo. Blood, 108, 1571-1579. https://doi.org/10.1182/blood-2006-02-004747

  64. 64. Wirnsberger, G., Mair, F. and Klein, L. (2009) Regulatory T Cell Differentiation of Thymocytes Does Not Require a Dedicated Antigen-Presenting Cell But Is under T Cell-Intrinsic Developmental Control. Proceedings of the National Academy of Sciences, 106, 10278-10283. https://doi.org/10.1073/pnas.0901877106

  65. 65. Isomura, I., et al. (2009) c-Rel Is Required for the Development of Thymic Foxp3+ CD4 Regulatory T Cells. The Journal of Experimental Medicine, 206, 3001-3014. https://doi.org/10.1084/jem.20091411

  66. 66. Hori, S. (2010) c-Rel: A Pioneer in Directing Regulatory T-Cell Lineage Commitment? European Journal of Immunology, 40, 664-667. https://doi.org/10.1002/eji.201040372

  67. 67. Marie, J.C., Letterio, J.J., Gavin, M. and Rudensky, A.Y. (2005) TGF-beta1 Maintains Suppressor Function and Foxp3 Expression in CD4+ CD25+ Regulatory T Cells. The Journal of Experimental Medicine, 201, 1061-1067. https://doi.org/10.1084/jem.20042276

  68. 68. Li, M.O., Sanjabi, S. and Flavell, R.A. (2006) Transforming Growth Factor-Beta Controls Development, Homeostasis, and Tolerance of T Cells by Regulatory T Cell-Dependent and -Independent Mechanisms. Immunity, 25, 455-471. https://doi.org/10.1016/j.immuni.2006.07.011

  69. 69. Nazzal, D., Gradolatto, A., Truffault, F., Bismuth, J. and Berrih-Aknin, S. (2014) Human Thymus Medullary Epithelial Cells Promote Regulatory T-Cell Generation by Stimulating Interleukin-2 Production via ICOS Ligand. Cell Death & Disease, 5, e1420. https://doi.org/10.1038/cddis.2014.377

  70. 70. Chen, W., et al. (2003) Conversion of Peripheral CD4+ CD25-Naive T Cells to CD4+ CD25+ Regulatory T Cells by TGF-Beta Induction of Transcription Factor Foxp3. The Journal of Experimental Medicine, 198, 1875-1886. https://doi.org/10.1084/jem.20030152

  71. 71. Lan, R.Y., Ansari, A.A., Lian, Z.X. and Gershwin, M.E. (2005) Regulatory T Cells: Development, Function and Role in Autoimmunity. Autoimmunity Reviews, 4, 351-363. https://doi.org/10.1016/j.autrev.2005.01.007

  72. 72. Zheng, S.G. (2013) Regulatory T Cells vs. Th17: Differentiation of Th17 versus Treg, Are the Mutually Exclusive? American Journal of Clinical and Experimental Immunology, 2, 94-106.

  73. 73. Chai, J.N., Zhou, Y.W. and Hsieh, C.S. (2014) T Cells and Intestinal Commensal Bacteria—Ignorance, Rejection, and Acceptance. FEBS Letters, 588, 4167-4175. https://doi.org/10.1016/j.febslet.2014.06.040

  74. 74. Yuan, X., Cheng, G. and Malek, T.R. (2014) The Importance of Regulatory T-Cell Heterogeneity in Maintaining Self-Tolerance. Immunological Reviews, 259, 103-114. https://doi.org/10.1111/imr.12163

  75. 75. Gobert, M. and Lafaille, J.J. (2012) Maternal-Fetal Immune Tolerance, Block by Block. Cell, 150, 7-9. https://doi.org/10.1016/j.cell.2012.06.020

  76. 76. Fu, S., et al. (2004) TGF-Beta Induces Foxp3+ T-Regulatory Cells from CD4+ CD25-Precursors. American Journal of Transplantation, 4, 1614-1627. https://doi.org/10.1111/j.1600-6143.2004.00566.x

  77. 77. Horwitz, D.A., Zheng, S.G. and Gray, J.D. (2003) The Role of the Combination of IL-2 and TGF-Beta or IL-10 in the Generation and Function of CD4+ CD25+ and CD8+ Regulatory T Cell Subsets. Journal of Leukocyte Biology, 74, 471-478. https://doi.org/10.1189/jlb.0503228

  78. 78. Yamagiwa, S., Gray, J.D., Hashimoto, S. and Horwitz, D.A. (2001) A Role for TGF-Beta in the Generation and Expansion of CD4+ CD25+ Regulatory T Cells from Human Peripheral Blood. The Journal of Immunology, 166, 7282-7289. https://doi.org/10.4049/jimmunol.166.12.7282

  79. 79. Rudensky, A.Y. and Campbell, D.J. (2006) In Vivo Sites and Cellular Mechanisms of T Reg Cell-Mediated Suppression. The Journal of Experimental Medicine, 203, 489-492. https://doi.org/10.1084/jem.20060214

  80. 80. Bluestone, J.A. and Abbas, A.K. (2003) Natural versus Adaptive Regulatory T Cells. Nature Reviews Immunology, 3, 253-257. https://doi.org/10.1038/nri1032

  81. 81. Whiteside, T.L. (2014) Induced Regulatory T Cells in Inhibitory Microenvironments Created by Cancer. Expert Opinion on Biological Therapy, 14, 1411-1425. https://doi.org/10.1517/14712598.2014.927432

  82. 82. Lu, L., et al. (2010) Synergistic Effect of TGF-Beta Superfamily Members on the Induction of Foxp3+ Treg. European Journal of Immunology, 40, 142-152. https://doi.org/10.1002/eji.200939618

  83. 83. Lu, L., et al. (2010) Role of SMAD and Non-SMAD Signals in the Development of Th17 and Regulatory T Cells. The Journal of Immunology, 184, 4295-4306. https://doi.org/10.4049/jimmunol.0903418

  84. 84. Zhou, X., et al. (2010) Isolation of Purified and Live Foxp3+ Regulatory T Cells Using FACS Sorting on Scatter Plot. Journal of Molecular Cell Biology, 2, 164-169. https://doi.org/10.1093/jmcb/mjq007

  85. 85. Chalmin, F., et al. (2012) Stat3 and Gfi-1 Transcription Factors Control Th17 Cell Immunosuppressive Activity via the Regulation of Ectonucleotidase Expression. Immunity, 36, 362-373. https://doi.org/10.1016/j.immuni.2011.12.019

  86. 86. Li, C., Ebert, P.J. and Li, Q.J. (2013) T Cell Receptor (TCR) and Transforming Growth Factor Beta (TGF-Beta) Signaling Converge on DNA (Cytosine-5)-Methyltransferase to Control Forkhead Box Protein 3 (Foxp3) Locus Methylation and Inducible Regulatory T Cell Differentiation. The Journal of Biological Chemistry, 288, 19127-19139. https://doi.org/10.1074/jbc.M113.453357

  87. 87. Yadav, M., Stephan, S. and Bluestone, J.A. (2013) Peripherally Induced Tregs-Role in Immune Homeostasis and Autoimmunity. Frontiers in Immunology, 4, 232. https://doi.org/10.3389/fimmu.2013.00232

  88. 88. Zheng, Y., Josefowicz, S., Chaudhry, A., Peng, X.P., Forbush, K. and Rudensky, A.Y. (2010) Role of Conserved Non-Coding DNA Elements in the Foxp3 Gene in Regulatory T-Cell Fate. Nature, 463, 808-812. https://doi.org/10.1038/nature08750

  89. 89. Zheng, S.G., Wang, J., Wang, P., Gray, J.D. and Horwitz, D.A. (2007) IL-2 Is Essential for TGF-Beta to Convert Naive CD4+ CD25-Cells to CD25+ Foxp3+ Regulatory T Cells and for Expansion of These Cells. The Journal of Immunology, 178, 2018-2027. https://doi.org/10.4049/jimmunol.178.4.2018

  90. 90. Lu, L., et al. (2011) All-Trans Retinoic Acid Promotes TGF-Beta-Induced Tregs via Histone Modification But Not DNA Demethylation on Foxp3 Gene Locus. PLoS ONE, 6, e24590. https://doi.org/10.1371/journal.pone.0024590

  91. 91. Shi, Q., et al. (2011) CD4+ Foxp3+ Regulatory T Cells Induced by TGF-Beta, IL-2 and All-Trans Retinoic Acid Attenuate Obliterative Bronchiolitis in Rat Trachea Transplantation. International Immunopharmacology, 11, 1887-1894. https://doi.org/10.1016/j.intimp.2011.07.020

  92. 92. Laurence, A., et al. (2007) Interleukin-2 Signaling via STAT5 Constrains T Helper 17 Cell Generation. Immunity, 26, 371-381. https://doi.org/10.1016/j.immuni.2007.02.009

  93. 93. Burchill, M.A., Yang, J., Vogtenhuber, C., Blazar, B.R. and Farrar, M.A. (2007) IL-2 Receptor Beta-Dependent STAT5 Activation Is Required for the Development of Foxp3+ Regulatory T Cells. The Journal of Immunology, 178, 280-290. https://doi.org/10.4049/jimmunol.178.1.280

  94. 94. Brandenburg, S., et al. (2008) IL-2 Induces in Vivo Suppression by CD4(+) CD25(+) Foxp3(+) Regulatory T Cells. European Journal of Immunology, 38, 1643-1653. https://doi.org/10.1002/eji.200737791

  95. 95. Hall, J.A., Grainger, J.R., Spencer, S.P. and Belkaid, Y. (2011) The Role of Retinoic Acid in Tolerance and Immunity. Immunity, 35, 13-22. https://doi.org/10.1016/j.immuni.2011.07.002

  96. 96. Maruyama, T., Konkel, J.E., Zamarron, B.F. and Chen, W. (2011) The Molecular Mechanisms of Foxp3 Gene Regulation. Seminars in Immunology, 23, 418-423. https://doi.org/10.1016/j.smim.2011.06.005

  97. 97. Xiao, S., et al. (2008) Retinoic Acid Increases Foxp3+ Regulatory T Cells and Inhibits Development of Th17 Cells by Enhancing TGF-Beta-Driven Smad3 Signaling and Inhibiting IL-6 and IL-23 Receptor Expression. The Journal of Immunology, 181, 2277-2284. https://doi.org/10.4049/jimmunol.181.4.2277

  98. 98. Nolting, J., et al. (2009) Retinoic Acid Can Enhance Conversion of Naive into Regulatory T Cells Independently of Secreted Cytokines. The Journal of Experimental Medicine, 206, 2131-2139. https://doi.org/10.1084/jem.20090639

  99. 99. Yoshimura, A. and Muto, G. (2011) TGF-Beta Function in Immune Suppression. Current Topics in Microbiology and Immunology, 350, 127-147. https://doi.org/10.1007/82_2010_87

  100. 100. Takimoto, T., et al. (2010) Smad2 and Smad3 Are Redundantly Essential for the TGF-Beta-Mediated Regulation of Regulatory T Plasticity and Th1 Development. The Journal of Immunology, 185, 842-855. https://doi.org/10.4049/jimmunol.0904100

  101. 101. Luckheeram, R.V., Zhou, R., Verma, A.D. and Xia, B. (2012) CD4(+) T Cells: Differentiation and Functions. Clinical & Developmental Immunology, 2012, Article ID: 925135. https://doi.org/10.1155/2012/925135

  102. 102. Couper, K.N., et al. (2008) IL-10 from CD4CD25Foxp3CD127 Adaptive Regulatory T Cells Modulates Parasite Clearance and Pathology during Malaria Infection. PLOS Pathogens, 4, e1000004. https://doi.org/10.1371/journal.ppat.1000004

  103. 103. Fujio, K., Okamura, T. and Yamamoto, K. (2010) The Family of IL-10-Secreting CD4+ T Cells. Advances in Immunology, 105, 99-130. https://doi.org/10.1016/S0065-2776(10)05004-2

  104. 104. Roncarolo, M.G., et al. (2006) Interleukin-10-Secreting Type 1 Regulatory T Cells in Rodents and Humans. Immunological Reviews, 212, 28-50. https://doi.org/10.1111/j.0105-2896.2006.00420.x

  105. 105. Awasthi, A., et al. (2007) A Dominant Function for Interleukin 27 in Generating Interleukin 10-Producing Anti-Inflammatory T Cells. Nature Immunology, 8, 1380-1389. https://doi.org/10.1038/ni1541

  106. 106. Gregori, S., et al. (2010) Differentiation of Type 1 T Regulatory Cells (Tr1) by Tolerogenic DC-10 Requires the IL-10-Dependent ILT4/HLA-G Pathway. Blood, 116, 935-944. https://doi.org/10.1182/blood-2009-07-234872

  107. 107. Pot, C., et al. (2009) Cutting Edge: IL-27 Induces the Transcription Factor c-Maf, Cytokine IL-21, and the Costimulatory Receptor ICOS That Coordinately Act Together to Promote Differentiation of IL-10-Producing Tr1 Cells. The Journal of Immunology, 183, 797-801. https://doi.org/10.4049/jimmunol.0901233

  108. 108. Quintana, F.J. and Sherr, D.H. (2013) Aryl Hydrocarbon Receptor Control of Adaptive Immunity. Pharmacological Reviews, 65, 1148-1161. https://doi.org/10.1124/pr.113.007823

  109. 109. Apetoh, L., et al. (2010) The Aryl Hydrocarbon Receptor Interacts with c-Maf to Promote the Differentiation of Type 1 Regulatory T Cells Induced by IL-27. Nature Immunology, 11, 854-861. https://doi.org/10.1038/ni.1912

  110. 110. Walsh, P.T., Taylor, D.K. and Turka, L.A. (2004) Tregs and Transplantation Tolerance. Journal of Clinical Investigation, 114, 1398-1403. https://doi.org/10.1172/JCI200423238

  111. 111. Korn, T., Bettelli, E., Oukka, M. and Kuchroo, V.K. (2009) IL-17 and Th17 Cells. Annual Review of Immunology, 27, 485-517. https://doi.org/10.1146/annurev.immunol.021908.132710

  112. 112. Nistala, K. and Wedderburn, L.R. (2009) Th17 and Regulatory T Cells: Rebalancing Pro- and Anti-Inflammatory Forces in Autoimmune Arthritis. Rheumatology (Oxford), 48, 602-606. https://doi.org/10.1093/rheumatology/kep028

  113. 113. Collison, L.W., et al. (2010) IL-35-Mediated Induction of a Potent Regulatory T Cell Population. Nature Immunology, 11, 1093-1101. https://doi.org/10.1038/ni.1952

  114. 114. Roncarolo, M.G., Bacchetta, R., Bordignon, C., Narula, S. and Levings, M.K. (2001) Type 1 T Regulatory Cells. Immunological Reviews, 182, 68-79. https://doi.org/10.1034/j.1600-065X.2001.1820105.x

  115. 115. Weiner, H.L. (2001) Induction and Mechanism of Action of Transforming Growth Factor-Beta-Secreting Th3 Regulatory Cells. Immunological Reviews, 182, 207-214. https://doi.org/10.1034/j.1600-065X.2001.1820117.x

  116. 116. Vieira, P.L., et al. (2004) IL-10-Secreting Regulatory T Cells Do Not Express Foxp3 But Have Comparable Regulatory Function to Naturally Occurring CD4+ CD25+ Regulatory T Cells. The Journal of Immunology, 172, 5986-5993. https://doi.org/10.4049/jimmunol.172.10.5986

  117. 117. Passerini, L., et al. (2011) Functional Type 1 Regulatory T Cells Develop Regardless of FOXP3 Mutations in Patients with IPEX Syndrome. European Journal of Immunology, 41, 1120-1131. https://doi.org/10.1002/eji.201040909

  118. 118. Levings, M.K. and Roncarolo, M.G. (2005) Phenotypic and Functional Differences between Human CD4+ CD25+ and Type 1 Regulatory T Cells. Current Topics in Microbiology and Immunology, 293, 303-326. https://doi.org/10.1007/3-540-27702-1_14

  119. 119. Gol-Ara, M., Jadidi-Niaragh, F., Sadria, R., Azizi, G. and Mirshafiey, A. (2012) The Role of Different Subsets of Regulatory T Cells in Immunopathogenesis of Rheumatoid Arthritis. Arthritis, 2012, Article ID: 805875. https://doi.org/10.1155/2012/805875

  120. 120. Jankovic, D., et al. (2007) Conventional T-bet(+)Foxp3(-) Th1 Cells Are the Major Source of Host-Protective Regulatory IL-10 during Intracellular Protozoan Infection. The Journal of Experimental Medicine, 204, 273-283. https://doi.org/10.1084/jem.20062175

  121. 121. Peterson, R.A. (2012) Regulatory T-Cells: Diverse Phenotypes Integral to Immune Homeostasis and Suppression. Toxicologic Pathology, 40, 186-204. https://doi.org/10.1177/0192623311430693

  122. 122. Crome, S.Q., Wang, A.Y. and Levings, M.K. (2010) Translational Mini-Review Series on Th17 Cells: Function and Regulation of Human T Helper 17 Cells in Health and Disease. Clinical & Experimental Immunology, 159, 109-119. https://doi.org/10.1111/j.1365-2249.2009.04037.x

  123. 123. Nakagawa, T., et al. (2010) IL-6 Positively Regulates Foxp3+ CD8+ T Cells in Vivo. International Immunology, 22, 129-139. https://doi.org/10.1093/intimm/dxp119

  124. 124. Thornton, A.M., et al. (2010) Expression of Helios, an Ikaros Transcription Factor Family Member, Differentiates Thymic-Derived from Peripherally Induced Foxp3+ T Regulatory Cells. The Journal of Immunology, 184, 3433-3441. https://doi.org/10.4049/jimmunol.0904028

  125. 125. Lin, X., et al. (2013) Advances in Distinguishing Natural from Induced Foxp3(+) Regulatory T Cells. International Journal of Clinical and Experimental Pathology, 6, 116-123.

  126. 126. Akimova, T., Beier, U.H., Wang, L., Levine, M.H. and Hancock, W.W. (2011) Helios Expression Is a Marker of T Cell Activation and Proliferation. PLoS ONE, 6, e24226. https://doi.org/10.1371/journal.pone.0024226

  127. 127. Serre, K., et al. (2011) Helios Is Associated with CD4 T Cells Differentiating to T Helper 2 and Follicular Helper T Cells in Vivo Independently of Foxp3 Expression. PLoS ONE, 6, e20731. https://doi.org/10.1371/journal.pone.0020731

  128. 128. Gottschalk, R.A., Corse, E. and Allison, J.P. (2012) Expression of Helios in Peripherally Induced Foxp3+ Regulatory T Cells. The Journal of Immunology, 188, 976-980. https://doi.org/10.4049/jimmunol.1102964

  129. 129. Ross, E.M., et al. (2014) Helios Defines T Cells Being Driven to Tolerance in the Periphery and Thymus. European Journal of Immunology, 44, 2048-2058. https://doi.org/10.1002/eji.201343999

  130. 130. Yadav, M., et al. (2012) Neuropilin-1 Distinguishes Natural and Inducible Regulatory T Cells among Regulatory T Cell Subsets in Vivo. The Journal of Experimental Medicine, 209, 1713-1722. https://doi.org/10.1084/jem.20120822

  131. 131. Hansen, W., et al. (2012) Neuropilin 1 Deficiency on CD4+ Foxp3+ Regulatory T Cells Impairs Mouse Melanoma Growth. The Journal of Experimental Medicine, 209, 2001-2016. https://doi.org/10.1084/jem.20111497

  132. 132. Milpied, P., et al. (2009) Neuropilin-1 Is Not a Marker of Human Foxp3+ Treg. European Journal of Immunology, 39, 1466-1471. https://doi.org/10.1002/eji.200839040

  133. 133. Dhamne, C., Chung, Y., Alousi, A.M., Cooper, L.J.N. and Tran, D.Q. (2013) Peripheral and Thymic Foxp3(+) Regulatory T Cells in Search of Origin, Distinction, and Function. Frontiers in Immunology, 4, 253. https://doi.org/10.3389/fimmu.2013.00253

  134. 134. Tran, D.Q., et al. (2009) Selective Expression of Latency-Associated Peptide (LAP) and IL-1 Receptor Type I/II (CD121a/CD121b) on Activated Human FOXP3+ Regulatory T Cells Allows for Their Purification from Expansion Cultures. Blood, 113, 5125-5133. https://doi.org/10.1182/blood-2009-01-199950

  135. 135. Pabbisetty, S.K., et al. (2014) KLF2 Is a Rate-Limiting Transcription Factor That Can Be Targeted to Enhance Regulatory T-Cell Production. Proceedings of the National Academy of Sciences, 111, 9579-9584. https://doi.org/10.1073/pnas.1323493111

  136. 136. Suffia, I.J., Reckling, S.K., Piccirillo, C.A., Goldszmid, R.S. and Belkaid, Y. (2006) Infected Site-Restricted Foxp3+ Natural Regulatory T Cells Are Specific for Microbial Antigens. The Journal of Experimental Medicine, 203, 777-788. https://doi.org/10.1084/jem.20052056

  137. 137. Bettini, M. and Vignali, D.A. (2009) Regulatory T Cells and Inhibitory Cytokines in Autoimmunity. Current Opinion in Immunology, 21, 612-618. https://doi.org/10.1016/j.coi.2009.09.011

  138. 138. McGeachy, M.J., Stephens, L.A. and Anderton, S.M. (2005) Natural Recovery and Protection from Autoimmune Encephalomyelitis: Contribution of CD4+ CD25+ Regulatory Cells within the Central Nervous System. The Journal of Immunology, 175, 3025-3032. https://doi.org/10.4049/jimmunol.175.5.3025

  139. 139. Asseman, C., Mauze, S., Leach, M.W., Coffman, R.L. and Powrie, F. (1999) An Essential Role for Interleukin 10 in the Function of Regulatory T Cells That Inhibit Intestinal Inflammation. The Journal of Experimental Medicine, 190, 995-1004. https://doi.org/10.1084/jem.190.7.995

  140. 140. Annacker, O., Asseman, C., Read, S. and Powrie, F. (2003) Interleukin-10 in the Regulation of T Cell-Induced Colitis. Journal of Autoimmunity, 20, 277-279. https://doi.org/10.1016/S0896-8411(03)00045-3

  141. 141. Joetham, A., et al. (2007) Naturally Occurring Lung CD4(+)CD25(+) T Cell Regulation of Airway Allergic Responses Depends on IL-10 Induction of TGF-Beta. The Journal of Immunology, 178, 1433-1442. https://doi.org/10.4049/jimmunol.178.3.1433

  142. 142. Rubtsov, Y.P., et al. (2008) Regulatory T Cell-Derived Interleukin-10 Limits Inflammation at Environmental Interfaces. Immunity, 28, 546-558. https://doi.org/10.1016/j.immuni.2008.02.017

  143. 143. Li, M.O., Wan, Y.Y. and Flavell, R.A. (2007) T Cell-Produced Transforming Growth Factor-Beta1 Controls T Cell Tolerance and Regulates Th1- and Th17-Cell Differentiation. Immunity, 26, 579-591. https://doi.org/10.1016/j.immuni.2007.03.014

  144. 144. Zhang, Q., et al. (2006) Blockade of Transforming Growth Factor-Beta Signaling in Tumor-Reactive CD8(+) T Cells Activates the Antitumor Immune Response Cycle. Molecular Cancer Therapeutics, 5, 1733-1743. https://doi.org/10.1158/1535-7163.MCT-06-0109

  145. 145. Shevach, E.M., Tran, D.Q., Davidson, T.S. and Andersson, J. (2008) The Critical Contribution of TGF-Beta to the Induction of Foxp3 Expression and Regulatory T Cell Function. European Journal of Immunology, 38, 915-917. https://doi.org/10.1002/eji.200738111

  146. 146. Andersson, J., et al. (2008) CD4+ FoxP3+ Regulatory T Cells Confer Infectious Tolerance in a TGF-Beta-Dependent Manner. The Journal of Experimental Medicine, 205, 1975-1981. https://doi.org/10.1084/jem.20080308

  147. 147. Piccirillo, C.A., et al. (2002) CD4(+)CD25(+) Regulatory T Cells Can Mediate Suppressor Function in the Absence of Transforming Growth Factor Beta1 Production and Responsiveness. The Journal of Experimental Medicine, 196, 237-246. https://doi.org/10.1084/jem.20020590

  148. 148. Collison, L.W., et al. (2007) The Inhibitory Cytokine IL-35 Contributes to Regulatory T-Cell Function. Nature, 450, 566-569. https://doi.org/10.1038/nature06306

  149. 149. Chaturvedi, V., Collison, L.W., Guy, C.S., Workman, C.J. and Vignali, D.A. (2011) Cutting Edge: Human Regulatory T Cells Require IL-35 to Mediate Suppression and Infectious Tolerance. The Journal of Immunology, 186, 6661-6666. https://doi.org/10.4049/jimmunol.1100315

  150. 150. Veldhoen, M., et al. (2006) Modulation of Dendritic Cell Function by Naive and Regulatory CD4+ T Cells. The Journal of Immunology, 176, 6202-6210. https://doi.org/10.4049/jimmunol.176.10.6202

  151. 151. Houot, R., Perrot, I., Garcia, E., Durand, I. and Lebecque, S. (2006) Human CD4+ CD25 High Regulatory T Cells Modulate Myeloid But Not Plasmacytoid Dendritic Cells Activation. The Journal of Immunology, 176, 5293-5298. https://doi.org/10.4049/jimmunol.176.9.5293

  152. 152. Wing, K., et al. (2008) CTLA-4 Control over Foxp3+ Regulatory T Cell Function. Science, 322, 271-275. https://doi.org/10.1126/science.1160062

  153. 153. Oderup, C., Cederbom, L., Makowska, A., Cilio, C.M. and Ivars, F. (2006) Cytotoxic T Lymphocyte Antigen-4-Dependent Down-Modulation of Costimulatory Molecules on Dendritic Cells in CD4+ CD25+ Regulatory T-Cell-Mediated Suppression. Immunology, 118, 240-249. https://doi.org/10.1111/j.1365-2567.2006.02362.x

  154. 154. Read, S., et al. (2006) Blockade of CTLA-4 on CD4+ CD25+ Regulatory T Cells Abrogates Their Function in Vivo. The Journal of Immunology, 177, 4376-4383. https://doi.org/10.4049/jimmunol.177.7.4376

  155. 155. Takahashi, T., et al. (2000) Immunologic Self-Tolerance Maintained by CD25(+) CD4(+) Regulatory T Cells Constitutively Expressing Cytotoxic T Lymphocyte-Associated Antigen 4. The Journal of Experimental Medicine, 192, 303-310. https://doi.org/10.1084/jem.192.2.303

  156. 156. Friedline, R.H., et al. (2009) CD4+ Regulatory T Cells Require CTLA-4 for the Maintenance of Systemic Tolerance. The Journal of Experimental Medicine, 206, 421-434. https://doi.org/10.1084/jem.20081811

  157. 157. Puccetti, P. and Grohmann, U. (2007) IDO and Regulatory T Cells: A Role for Reverse Signalling and Non-Canonical NF-kappaB Activation. Nature Reviews Immunology, 7, 817-823. https://doi.org/10.1038/nri2163

  158. 158. Mbongue, J.C., et al. (2015) The Role of Indoleamine 2, 3-Dioxygenase in Immune Suppression and Autoimmunity. Vaccines (Basel), 3, 703-729. https://doi.org/10.3390/vaccines3030703

  159. 159. Dejean, A.S., Hedrick, S.M. and Kerdiles, Y.M. (2011) Highly Specialized Role of Forkhead Box O Transcription Factors in the Immune System. Antioxid Redox Signal, 14, 663-674. https://doi.org/10.1089/ars.2010.3414

  160. 160. Liang, B., et al. (2008) Regulatory T Cells Inhibit Dendritic Cells by Lymphocyte Activation Gene-3 Engagement of MHC Class II. The Journal of Immunology, 180, 5916-5926. https://doi.org/10.4049/jimmunol.180.9.5916

  161. 161. Goldberg, M.V. and Drake, C.G. (2011) LAG-3 in Cancer Immunotherapy. Current Topics in Microbiology and Immunology, 344, 269-278. https://doi.org/10.1007/82_2010_114

  162. 162. Holst, J., et al. (2008) Scalable Signaling Mediated by T Cell Antigen Receptor-CD3 ITAMs Ensures Effective Negative Selection and Prevents Autoimmunity. Nature Immunology, 9, 658-666. https://doi.org/10.1038/ni.1611

  163. 163. Shalev, I., et al. (2008) Targeted Deletion of fgl2 Leads to Impaired Regulatory T Cell Activity and Development of Autoimmune Glomerulonephritis. The Journal of Immunology, 180, 249-260. https://doi.org/10.4049/jimmunol.180.1.249

  164. 164. Huang, C.T., et al. (2004) Role of LAG-3 in Regulatory T Cells. Immunity, 21, 503-513. https://doi.org/10.1016/j.immuni.2004.08.010

  165. 165. Chattopadhyay, G. and Shevach, E.M. (2013) Antigen-Specific Induced T Regulatory Cells Impair Dendritic Cell Function via an IL-10/MARCH1-Dependent Mechanism. The Journal of Immunology, 191, 5875-5884. https://doi.org/10.4049/jimmunol.1301693

  166. 166. Bartee, E., Mansouri, M., Hovey Nerenberg, B.T., Gouveia, K. and Früh, K. (2004) Down-Regulation of Major Histocompatibility Complex Class I by Human Ubiquitin Ligases Related to Viral Immune Evasion Proteins. Journal of Virology, 78, 1109-1120. https://doi.org/10.1128/JVI.78.3.1109-1120.2004

  167. 167. Tze, L.E., et al. (2011) CD83 Increases MHC II and CD86 on Dendritic Cells by Opposing IL-10-Driven MARCH1-Mediated Ubiquitination and Degradation. The Journal of Experimental Medicine, 208, 149-165. https://doi.org/10.1084/jem.20092203

  168. 168. Vignali, D.A., Collison, L.W. and Workman, C.J. (2008) How Regulatory T Cells Work. Nature Reviews Immunology, 8, 523-532. https://doi.org/10.1038/nri2343

  169. 169. Woo, S.R., et al. (2012) Immune Inhibitory Molecules LAG-3 and PD-1 Synergistically Regulate T-Cell Function to Promote Tumoral Immune Escape. Cancer Research, 72, 917-927. https://doi.org/10.1158/0008-5472.CAN-11-1620

  170. 170. Nguyen, L.T. and Ohashi, P.S. (2015) Clinical Blockade of PD1 and LAG3—Potential Mechanisms of Action. Nature Reviews Immunology, 15, 45-56. https://doi.org/10.1038/nri3790

  171. 171. Riella, L.V., Paterson, A.M., Sharpe, A.H. and Chandraker, A. (2012) Role of the PD-1 Pathway in the Immune Response. American Journal of Transplantation, 12, 2575-2587. https://doi.org/10.1111/j.1600-6143.2012.04224.x

  172. 172. Pardoll, D.M. (2012) The Blockade of Immune Checkpoints in Cancer Immunotherapy. Nature Reviews Cancer, 12, 252-264. https://doi.org/10.1038/nrc3239

  173. 173. Janssens, W., et al. (2003) CD4+ CD25+ T Cells Lyse Antigen-Presenting B Cells by Fas-Fas Ligand Interaction in an Epitope-Specific Manner. The Journal of Immunology, 171, 4604-4612. https://doi.org/10.4049/jimmunol.171.9.4604

  174. 174. Grossman, W.J., Verbsky, J.W., Barchet, W., Colonna, M., Atkinson, J.P. and Ley, T.J. (2004) Human T Regulatory Cells Can Use the Perforin Pathway to Cause Autologous Target Cell Death. Immunity, 21, 589-601. https://doi.org/10.1016/j.immuni.2004.09.002

  175. 175. Grossman, W.J., Verbsky, J.W., Tollefsen, B.L., Kemper, C., Atkinson, J.P. and Ley, T.J. (2004) Differential Expression of Granzymes A and B in Human Cytotoxic Lymphocyte Subsets and T Regulatory Cells. Blood, 104, 2840-2848. https://doi.org/10.1182/blood-2004-03-0859

  176. 176. Gondek, D.C., Lu, L.F., Quezada, S.A., Sakaguchi, S. and Noelle, R.J. (2005) Cutting Edge: Contact-Mediated Suppression by CD4+ CD25+ Regulatory Cells Involves a Granzyme B-Dependent, Perforin-Independent Mechanism. The Journal of Immunology, 174, 1783-1786.

  177. 177. Sojka, D.K., Huang, Y.H. and Fowell, D.J. (2008) Mechanisms of Regulatory T-Cell Suppression—A Diverse Arsenal for a Moving Target. Immunology, 124, 13-22. https://doi.org/10.1111/j.1365-2567.2008.02813.x

  178. 178. Cao, X., et al. (2007) Granzyme B and Perforin Are Important for Regulatory T Cell-Mediated Suppression of Tumor Clearance. Immunity, 27, 635-646. https://doi.org/10.1016/j.immuni.2007.08.014

  179. 179. Fontenot, J.D., Rasmussen, J.P., Gavin, M.A. and Rudensky, A.Y. (2005) A Function for Interleukin 2 in Foxp3-Expressing Regulatory T Cells. Nature Immunology, 6, 1142-1151. https://doi.org/10.1038/ni1263

  180. 180. Huibregtse, I.L., van Lent, A.U. and van Deventer, S.J. (2007) Immunopathogenesis of IBD: Insufficient Suppressor Function in the Gut? Gut, 56, 584-592. https://doi.org/10.1136/gut.2006.103523

  181. 181. Sakaguchi, S., Wing, K., Onishi, Y., Prieto-Martin, P. and Yamaguchi, T. (2009) Regulatory T Cells: How Do They Suppress Immune Responses? International Immunology, 21, 1105-1111. https://doi.org/10.1093/intimm/dxp095

  182. 182. Schmidt, A., Oberle, N. and Krammer, P.H. (2012) Molecular Mechanisms of Treg-Mediated T Cell Suppression. Frontiers in Immunology, 3, 51. https://doi.org/10.3389/fimmu.2012.00051

  183. 183. Longhi, M.S., Robson, S.C., Bernstein, S.H., Serra, S. and Deaglio, S. (2013) Biological Functions of Ecto-Enzymes in Regulating Extracellular Adenosine Levels in Neoplastic and Inflammatory Disease States. Journal of Molecular Medicine, 91, 165-172. https://doi.org/10.1007/s00109-012-0991-z

  184. 184. Ayyoub, M., et al. (2009) Human Memory FOXP3+ Tregs Secrete IL-17 ex Vivo and Constitutively Express the T(H)17 Lineage-Specific Transcription Factor RORgamma t. Proceedings of the National Academy of Sciences, 106, 8635-8640. https://doi.org/10.1073/pnas.0900621106

  185. 185. Beriou, G., et al. (2009) IL-17-Producing Human Peripheral Regulatory T Cells Retain Suppressive Function. Blood, 113, 4240-4249. https://doi.org/10.1182/blood-2008-10-183251

  186. 186. Zheng, Y., et al. (2009) Regulatory T-Cell Suppressor Program Co-Opts Transcription Factor IRF4 to Control T(H)2 Responses. Nature, 458, 351-356. https://doi.org/10.1038/nature07674

  187. 187. Koch, M.A., Thomas, K.R., Perdue, N.R., Smigiel, K.S., Srivastava, S. and Campbell, D.J. (2012) T-bet(+) Treg Cells Undergo Abortive Th1 Cell Differentiation Due to Impaired Expression of IL-12 Receptor beta2. Immunity, 37, 501-510. https://doi.org/10.1016/j.immuni.2012.05.031

  188. 188. Linterman, M.A., Liston, A. and Vinuesa, C.G. (2012) T-Follicular Helper Cell Differentiation and the Co-Option of This Pathway by Non-Helper Cells. Immunological Reviews, 247, 143-159. https://doi.org/10.1111/j.1600-065X.2012.01121.x

  189. 189. Kroenke, M.A., et al. (2012) Bcl6 and Maf Cooperate to Instruct Human Follicular Helper CD4 T Cell Differentiation. The Journal of Immunology, 188, 3734-3744. https://doi.org/10.4049/jimmunol.1103246

  190. 190. Liu, X., et al. (2012) Bcl6 Expression Specifies the T Follicular Helper Cell Program in Vivo. The Journal of Experimental Medicine, 209, 1841-1852. https://doi.org/10.1084/jem.20120219

  191. 191. Perreau, M., et al. (2013) Follicular Helper T Cells Serve as the Major CD4 T Cell Compartment for HIV-1 Infection, Replication, and Production. The Journal of Experimental Medicine, 210, 143-156. https://doi.org/10.1084/jem.20121932

  192. 192. Matsushita, N., et al. (2008) Comparative Methodologies of Regulatory T Cell Depletion in a Murine Melanoma Model. The Journal of Immunology Methods, 333, 167-179. https://doi.org/10.1016/j.jim.2008.01.012

  193. 193. Lahl, K. and Sparwasser, T. (2011) In Vivo Depletion of FoxP3+ Tregs Using the DEREG Mouse Model. Methods in Molecular Biology, 707, 157-172. https://doi.org/10.1007/978-1-61737-979-6_10

  194. 194. Kim, J.M., Rasmussen, J.P. and Rudensky, A.Y. (2007) Regulatory T Cells Prevent Catastrophic Autoimmunity throughout the Lifespan of Mice. Nature Immunology, 8, 191-197. https://doi.org/10.1038/ni1428

  195. 195. Suffner, J., et al. (2010) Dendritic Cells Support Homeostatic Expansion of Foxp3+ Regulatory T Cells in Foxp3.LuciDTR Mice. The Journal of Immunology, 184, 1810-1820. https://doi.org/10.4049/jimmunol.0902420

  196. 196. Lahl, K., et al. (2007) Selective Depletion of Foxp3+ Regulatory T Cells Induces a Scurfy-Like Disease. The Journal of Experimental Medicine, 204, 57-63. https://doi.org/10.1084/jem.20061852

Abbreviations

CTLA-4, Cytotoxic T lymphocyte antigen 4

DEREG, Depletion of regulatory T cells

DTR, Diphtheria toxin receptor

DT, Diphtheria toxin

EAE, Experimental autoimmune encephalomyelitis

Ebi3, Epstein-Barr virus induced gene 3

Foxp3, Forkhead box P3 protein

Foxo, Forkhead box O transcription factor

ICOS, Inducible T cell costimulatory

IDO, Indoleamine 2, 3-dioxygenase

IPEX, Immune dysregulation, polyendocrinopathy, enteropathy X-linked

IRF, Interferon regulatory factor

ITAM, Immunoreceptor tyrosine-based activation motif

KLF2, Krϋppel-like factor 2

LAG-3, Lymphocyte activation gene 3

LAP, Latency associated peptide

MARCH1, Membrane-associated E3 ubiquitin ligase RING-CH1

mTECs, Medullary thymic epithelial cells

NFAT, Nuclear factor of activated T cells

Nrp1, Neuropilin-1

nTreg cells, Naturally occurring regulatory T cells

pTreg, Peripherally derived Treg cell

PD-1, Programmed death 1

PD-L, Programmed death 1 ligand

RA, Retinoic acid

STAT, Signal transducer and activator of transcription

T-bet, T-box expressed in T cells

TCR, T cell receptor

Tfh cell, T follicular helper cell

TGFβ, Transforming growth factor β

TIEG1, TGF-β-inducible early gene 1 product

TNF, Tumour necrosis factor

TNFRSF18, Tumour necrosis factor receptor superfamily 18

Tr1, T regulatory type 1

Treg, CD4+Foxp3+ Regulatory T cell

tTreg, Thymus derived Treg cell