Skip to main content
Immunology logoLink to Immunology
. 2011 Nov;134(3):235–245. doi: 10.1111/j.1365-2567.2011.03483.x

Helper T-cell differentiation and plasticity: insights from epigenetics

Kiyoshi Hirahara 1,*, Golnaz Vahedi 1,*, Kamran Ghoreschi 1, Xiang-Ping Yang 1, Shingo Nakayamada 1, Yuka Kanno 1, John J O'Shea 1, Arian Laurence 1
PMCID: PMC3209564  PMID: 21977994

Abstract

CD4+ T cells have critical roles in orchestrating immune responses to diverse microbial pathogens. This is accomplished through the differentiation of CD4+ T helper cells to specialized subsets in response to microbial pathogens, which evoke a distinct cytokine milieu. Signal transducer and activator of transcription family transcription factors sense these cytokines and they in turn regulate expression of lineage-defining master regulators that programme selective gene expression, resulting in distinctive phenotypes. However, phenotype and restricted gene expression are determined not only by the action of transcription factors; chromatin accessibility is required for these factors to exert their effect. Technical advances have greatly expanded our understanding of transcription factor action and dynamic changes in the epigenome that accompany cellular differentiation. In this review, we will discuss recent progress in the understanding of how cytokines influence gene expression and epigenetic modifications, and the impact of these findings on our views of helper cell lineage commitment and plasticity.

Keywords: epigenetics, histone modification, signal transducer and activator of transcription, T-cell differentiation, T-cell plasticity

Introduction

The critical role of CD4+ T cells in host defence and immunoregulation is well-appreciated. Initially, the fates of CD4+ T helper (Th) cells were thought to be limited to Th1 and Th2 cells.1,2 Th1 cells, which express the transcription factor T-bet (encoded by the Tbx21 gene) and produce interferon-γ (IFN-γ) exclusively, protect the host against intracellular infections including viruses and Toxoplasma.35 Th2 cells, which express GATA-3 and secrete interleukin-4 (IL-4), IL-5 and IL-13, mediate host defence against helminths.6 More recently, cells that selectively produce IL-17 have been proposed to represent a distinct Th-cell lineage, Th17 cells.712 Th17 cells can also produce other cytokines including IL-9, IL-10, IL-21, IL-22 and, in humans, IL-26; however, these cytokines can also be produced by other subsets.1316 Th17 cells express a splice variant of the retinoid orphan receptor-γ, designated RORγt. Th17 cells contribute to host defence against extracellular bacteria and fungi.17,18 Other putative lineages have also been identified including Th9 and Th22, which produce their namesake cytokines IL-9 and IL-22, respectively.13,19,20

Follicular helper T (Tfh) cells have recently been proposed as yet another lineage, with specialized function in helping B cells make antibody responses.21,22 As with other helper cell lineages, extrinsic cytokine cues, namely IL-6 and IL-21, acting via signal transducer and activator of transcription 3 (STAT3), have been proposed to be the principal drivers of Tfh-cell differentiation.23 These factors induce expression of Bcl6 and current models posit that Tfh cells arise from naive T cells and through acquisition of this transcription factor, which has come to be viewed as the master regulator of Tfh cells.2426 Blimp-1 and Bcl6 are antagonistic and reciprocal regulators of Tfh differentiation, and it has been argued that balance of expression of these factors is important for Tfh versus effector cell differentiation.25,27 These findings have been invoked to support the lineage sovereignty of Tfh cells.

Regulatory T (Treg) cells are another CD4+ lineage with essential immunosuppressive functions and expression of the master regulator transcription factor FoxP3. Thymically derived Treg cells are referred to as natural Treg (nTreg) cells, whereas Treg cells that arise from naive CD4+ T cells in the periphery are denoted induced Treg (iTreg) cells.2830

In addition to their role in host defence, different T-cell subsets are also associated with immune-mediated disease. For example, Th1 and Th17 cell responses have been associated with a number of autoimmune diseases,3136 whereas Th2 cytokines are associated with atopic disease.3740 Absence of regulatory T cells because of mutations of FOXP3 results in severe, systemic autoimmunity.4143

All of the subsets can be generated in vitro by activating naive CD4+ cells in specific cocktails of cytokines. That is, T-cell receptor engagement in the presence of IL-12 and IFN-γ and the absence of IL-4 yields Th1 cells.44 In contrast, the presence of IL-4 and absence of IFN-γ results in Th2 polarization.45 The remaining two lineages can be generated in the presence of transforming growth factor-β (TGF-β) with the combination of TGF-β and IL-2 producing iTreg cells and the combination of TGF-β and IL-6 producing Th17 cells9,46; although Th17 cells can also be made in the absence of TGF-β signalling.47,48

Many of the aforementioned cytokines that promote helper cell specification bind to receptors that are members of the Type I/II cytokine receptor superfamily and signal by the JAK/STAT pathway. There are seven STAT proteins (STAT1-5a, 5b and 6). Mice that are deficient in either STAT3 or STAT5 die in utero because of broad, critical functions.49,50 However, these factors also have essential, non-redundant functions in helper cell differentiation, with STAT4 and STAT1, STAT6, STAT3, and STAT5a/b being required for the differentiation of Th1, Th2, Th17 and Treg cells, respectively.49,5158

Hence, characteristics of various Th subsets include distinct immunological functions, the expression of a unique repertoire of cytokines, and the expression of a master regulator of transcription factors induced by exogenous factors. These subsets have been viewed as ‘lineages’ in that the phenotypes are often stable and hereditable; however, there are also more examples of flexibility. In this review, we describe a few of the many examples of Th phenotypic plasticity, the epigenetic changes that occur in CD4+ T-cell differentiation and how these epigenetic modifications contribute to CD4+ T-cell heritability and plasticity. Moreover, we will briefly review how STAT proteins function as sensors of the cytokine environment and promote the acquisition of epigenetic changes.

Helper cell lineage commitment versus plasticity

The standard model of Th1, Th2 and Th17 differentiation implies that these subsets behave like terminally differentiated cells with a heritable expression of specific cytokines and transcription factors. To some extent this is correct – generally speaking, Th1-polarized cells make IFN-γ and do not make IL-4 or IL-17. Similarly, nTreg cells behave as a reasonably stable lineage in vivo.59 However, other findings indicate more flexibility than envisioned and recent findings provide mechanisms for flexibility in expression of key transcription factors.

Flexibility of Th2 cells

Although the established Th1/Th2 dogma would imply that these subsets conform to a lineage commitment model, recent data have provided some rather dramatic surprises that challenge this view. Hegazy et al.60 showed that committed GATA-3+ Th2 cells could be reprogrammed to adopt GATA-3+ T-bet+ and IL-4+ IFN-γ+ phenotype by lymphocytic choriomeningitis virus (LCMV) infection, and these GATA-3 T-bet double-positive cells were critical to LCMV exclusion.

Intrinsic instability of Th17 cells

Initially, IL-17-secreting Th cells were suggested to represent a new lineage, as the cells did not make other lineage-defining cytokines like IFN-γ and IL-4.7 Like Th1 and Th2 cells, Th17 cells could be induced in vitro from naive T cells and required a unique combination of cytokines (TGF-β and IL-6) 61 and the intracellular signalling molecule, STAT3,57,62,63 factors distinct from those required for classical Th1/Th2 polarization. Th17 cells, like their Th1/2 counterparts expressed unique master regulator transcription factors (RORγt, RORα).11,64 Finally, cytokines that drove Th1 and Th2 lineage commitment were potent inhibitors of Th17 development.7 Interestingly, IL-2 shows the reciprocal effects on Treg and Th17 differentiation; IL-2 promotes Treg-cell differentiation, while IL-2 negatively regulates Th17 differentiation; both of these effects being dependent upon STAT5.58,6567

Once polarized, relatively rapid extinction of IL-17 expression occurred with acquisition of IFN-γ expression.68 This was initially suspected in models of experimental autoimmune uveitis where transferred antigen-specific in-vitro-generated Th17 cells subsequently became IFN-γ producers.69 Similarly, tumour immunity mediated by Th17 cells required the expression of IFN-γ.70 These initial results were confirmed in vitro using T cells from IL-17F reporter animals, where pure populations of IL-17-secreting cells could be isolated on the basis of expressing fluorescent proteins and were subsequently found to convert to IFN-γ-producing cells.68 More recently, Stockinger and colleagues, using a fate mapping reporter strategy, demonstrated that this conversion occurs physiologically.71 As indicated, T-bet is a key Th1 transcription factor,72 and recent work indicates that T-bet suppresses Th17 differentiation by preventing Runx1-dependent transactivation of Rorc.73 We and others have identified the rapid appearance of RORγt T-bet double-positive cells.47,74 In a murine experimental autoimmune encephalomyelitis model, RORγt T-bet double-positive cells showed greater pathogenic potential compared with RORγt+, T-bet cells.47

In contrast to the multiple lines of evidence pointing to conversion of Th17 to Th1 cells, until recently there was relatively little evidence of overlap between Th17 and Th2 cells; although, the two lineages are both dependent on the transcription factor, Interferon regulatory factor 4 (IRF4). Recently, a subset of GATA3 RORγt double-positive cells were identified and these cells were reported to play a role in the development of experimental asthma.75 In this regard, it is of interest that severity of asthma correlates with serum IL-17 concentration.76,77

Tfh cells – lineage or state?

As the newest ‘lineage’ it is perhaps not surprising that there is considerable controversy surrounding the origin and fate of Tfh. Activated CD4+ T cells are not devoid of Bcl6 and the signature Tfh cytokine, IL-21, is not exclusively produced by Tfh cells.62,78,79 An added complication is that Tfh cells have the capacity to produce cytokines characteristic of Th1, Th2 and Th17.14,80 Furthermore, helminths infection results in GATA3 Bcl6 double-positive Tfh cells.81 Moreover, FoxP3+ T cells to Peyer's patches can convert to Tfh cells.82 It remains an open question whether Tfh cells represent a distinct lineage parallel to other subsets of helper T cells, a temporary ‘state’ of differentiation or a phenotype that can be superimposed upon Th1, Th2, Th17 or Treg cells.

Stability of Treg cells – ongoing controversy

Treg cells are an essential subset, whose function is to inhibit immune responses.65 They are characterized by the expression of the inhibitory transcription factor, FoxP3, mutation of which results in a rapidly fatal inflammatory disease in both mice (Scurfy) and humans (immune dysregulation, polyendocrinopathy, enteropathy, X-linked syndrome; IPEX).83 Conversely, administration of Treg cells has been proposed as cellular therapy in transplantation and autoimmune disease.84 This notion though is predicated on the assumption that Treg cells injected into patients will maintain their immunosuppressive phenotype and not become effector cells. However, using a fate mapping approach, it was argued that loss of FoxP3 expression was not infrequent and that Treg cells had the capacity to become IL-17-secreting cells that could induce tissue destruction.85 Furthermore, in response to IL-12 in vitro or in the setting of Toxoplasma infection or colitis, Treg cells can produce IFN-γ.8688 The loss of suppressive function is not necessarily seen upon acquisition of IFN-γ production. Interestingly, a unique Treg subset exists at sites of inflammation, which expresses both FoxP3 and T-bet and inhibits Th1 inflammation.89 Similarly, mice that lack the Th2 and Tfh-associated transcription factor IRF4 in Treg cells develop Th2-associated inflammation.90 Despite these findings, other work has argued that nTreg cells maintain their phenotype in the setting of Listeria infection.59

In contrast to mouse cells, FoxP3 is a less reliable marker of suppressive activity, being induced in activated human T cells. Both of CD45RA+ FoxP3low and CD45RA FoxP3high cells have suppressive activity and CD45RA+ FoxP3low cells have the capacity to convert into CD45RA FoxP3high cells.91

Collectively, the data indicate that although helper cells have features of phenotypic stability, it is clear that they also have the capacity for considerable plasticity. This then raises the broader question of what factors preserve the phenotype of specialized cells.

Epigenetics and cellular phenotype

The issue of flexibility versus stability of phenotype is certainly not a concept that is unique to T cells. On the contrary, it has long been appreciated by developmental biologists that once differentiated, cells maintain their phenotype, even in the absence of exogenous signals. That is, terminally differentiated cells autonomously maintain their distinctive features even through subsequent mitotic division. Though the term ‘epigenetics’ was coined by Waddington before the era of modern molecular biology, it has come to denote hereditable changes in phenotype or gene expression without changes in DNA sequence. Increasingly, epigenetics has come to encompass various aspects of nucleosome biology and DNA methylation, which help to explain selective gene expression.92,93 Indeed, it is now well accepted that cellular phenotype is the result of the action of transcription factors regulating gene expression, but that the accessibility of genes to activation is also highly regulated through epigenetic modifications.

The basic structural unit of chromatin is the nucleosome, which consists of two copies of histones H2A, H2B, H3 and H4. Each nucleosome is encircled by 146 base pairs of DNA and interconnected by linker DNA. Selective positioning of nucleosomes by ATP-dependent chromatin remodelling complexes is one important aspect of genomic regulation.94,95 DNA methylation is another modality that limits the accessibility of DNA to cellular machinery, thereby affecting the transcriptional potential of the underlying DNA sequence.96

Additionally, the N-terminal tails of histone proteins can be covalently modified through acetylation, methylation and phosphorylation.97 Histone modifications are associated with the regulation of gene expression by condensing or relaxing the chromatin structure to repress or activate transcription, respectively.98 For example, trimethylation of H3K4 (H3K4me3) is associated with gene activation.99,100 In contrast, polycomb complexes catalyse trimethylation of H3K27 (H3K27me3), which serves to repress gene expression. Initially discovered as regulators of the Drosophila homeotic genes, polycomb-dependent H3K27me3 is associated with a wide range of targets in mammals.101106

As we learn more about the biochemistry of histone modifications and the enzymes responsible for the plethora of covalent modifications, it is recognized that the regulation of chromatin is a highly dynamic process. Indeed, some modifications can occur rapidly in response to exogenous signals and hence the regulation of nucleosome biology can be viewed as an extension of signal transduction.107,108 Yet, it remains unclear what signals a cell must receive and send to organize the epigenome with respect to its roles in developmental biology, cell differentiation, or stem cell renewal. It is worth emphasizing that epigenetic modifications also have the ability to persist in the absence of continued exogenous signals and even the transcription factors that initially induced them. This allows a cell to ‘remember’ its distinctive transcriptional profile and, by extension, its cellular identity.

Recent studies have demonstrated that epigenetic modifications are also important factors for the regulation of Th-cell differentiation. DNase hypersensitive sites, DNA methylation and chromatin modifications at candidate loci in Th cells such as IFNG, IL-4 and Foxp3 have been analysed.109 For instance, DNase I hypersensitive sites in the Ifng locus are induced upon Th1, but not Th2, differentiation.110 Similarly, a conserved CpG-rich region denoted as a Treg-specific demethylated region resides upstream of the FoxP3 promoter111; this region is demethylated in nTreg cells and is associated with stable expression of FoxP3.112 In contrast, there is residual methylation of this region in iTreg cells, consistent with less stability in expression.113,114

Alternatively, the potential importance of epigenetic modifications in Th-cell differentiation has been inferred through the use of different gene knockout mice. For instance, haplo-insufficiency of the H3K4 methyltransferase MLL (MLL+/−) resulted in defective maintenance but not initial induction of Il4, Il5, Il13 and Gata3 expression in Th2 cells.115 EZH2, a major H3K27 methyltransferase, binds to the Il4/Il13 locus and has been reported to be responsible for the suppression of the Il4/Il13 locus in Th1 cells.116 However, a complicating factor is that Ezh2 also plays an essential role in actin polymerization-dependent processes such as antigen receptor signalling in T cells.117 Chromatin remodelling complexes, which contain Brahma-related gene 1 (BRG1), displace nucleosomes and remodel chromatin at the Ifng promoter region in a Th1-specific manner 118 and knockdown of Brg1 results in decreased IFN-γ production. Originally thought to principally mediate sister chromatid cohesion during mitosis, cohesin and CCCTC-binding factor (CTCF) also occupy a subset of DNase I hypersensitive sites in the Ifng locus.119 Knocking down the expression of RAD21, one component of the cohesin complex, reduces IFNG transcript levels.120

In recent years, there has been a remarkable increase in our ability to study the structure of chromosomes, particularly through the development of chromosome conformation capture (3C) technology.121 3C employs restriction enzymes to fragment DNA and regular PCR to detect ligation products. Studying the chromatin configurations revealed that chromatin is folded into loops bringing distal regulatory elements into intimate contact with the genes that they regulate.122

One of the first systems for which the intrachromosomal and interchromosomal associations have been reported is the differentiation of naive CD4+ Th cells into Th1 and Th2 subsets. Using the 3C technique, it was found that in CD4+ cells, but not natural killer cells, B cells and fibroblasts, the promoters for the genes encoding Th2 cytokines are located in close spatial proximity with the Th2 locus control region.123 A similar study revealed that depending on the expression profile of the cell, the Ifng locus adopts a secondary structure that brings the enhancer elements and Ifng promoter into close proximity.124 It was also shown that deletion of Rad50 hypersensitive site 7 (RHS7) did not affect the expression of the Rad50 gene, but it reduced long-range intrachromosomal interactions between the Th2 locus control region and promoters of the Th2 cytokine genes.125 In addition to these intrachromosomal interactions, the interchromosomal interactions between the promoter region of the IFN-gamma gene on chromosome 10 and the regulatory regions of the Th2 cytokine locus on chromosome 11 was also described.126

Recent studies indicate that microRNA (miRNA) molecules may also have important regulatory roles in transcription. The miRNAs are small RNAs that bind to target messenger RNAs (mRNAs) and lead to inhibition of translation or mRNA degradation. A central role for miRNAs in maintaining the stability of differentiated Treg cell function in vivo and homeostasis of the adaptive immune system has also been shown.127 Moreover, it was shown that lymphocyte-specific miRNAs are either tightly controlled by polycomb group-mediated H3K27me3 or maintained in a semi-activated epigenetic state before full expression.128 Collectively, these data argue that epigenetic changes influence fate determination of helper T cells.

Technical advances: next generation sequencing, ChIP-seq

Until recently, analysis of epigenetic changes has necessarily aimed at very limited regions of the genome, focusing on pre-selected elements (promoters or recognized enhancers). The problem with such an approach is the tacit assumption that we know the key players. Obviously though, this may or may not be the case. With the advent of next-generation sequencing methods, an unbiased genome-wide view of epigenetics modifications has become a reality.

This technique involves cross-linking of DNA/protein complexes and immunoprecipitation using antibodies directed at transcription factors or other DNA-binding proteins. Antibodies directed at the many specific histone modifications can also be employed. After reversing the cross-links, the isolated DNA can be sequenced using next-generation sequencing technology to produce digital maps of enrichment. This technique, referred to as chromatin immunoprecipitation and massive parallel sequencing (ChIP-seq) permits genome-wide profiling of DNA-binding proteins, DNA methylation, histone modifications or nucleosome positioning.129 Comprehensive mapping of 13 transcription factors in embryonic stem cells has now been obtained.130 Another example of using deep sequencing is the genome-wide mapping of histone acetyltransferases (HATs) and deacetylases (HDACs) binding on chromatin where it was shown that both are found at active genes with acetylated histones.131

Several methods have been developed to profile DNA methylation. Such regions can be identified based on the enrichment of methylated genomic DNA using an antibody specific for methylated cytosines.132 This method can provide relatively inexpensive and comprehensive data but the resolution is limited. Alternatively, sequencing of bisulphite-converted DNA, which provides base-resolution but at a higher cost, can also be employed.132 Complete DNA methylomes for several organisms have now been obtained, including a genome-wide map of methylated DNA in a mammalian genome from both human embryonic stem cells and fetal fibroblasts.133 High-throughput sequencing of 5-hydroxymethylcytosine, a modified base present at low levels in diverse cell types, from mouse embryonic stem cells, has also been completed.134 These data should clarify the evolutionary significance of this epigenetic mark and its distribution in key genomic elements.132

The development of high-throughput DNA sequencing methods can also be used for both mapping and quantifying transcriptomes. This method, termed RNA-Seq (RNA sequencing), has clear advantages over existing approaches and is expected to revolutionize the manner in which eukaryotic transcriptomes are analysed.135

The field of chromosome conformation capture made tremendous progress in the ability to map the three-dimensional organization of chromosomes at the scale of the complete genome. The high-throughput modifications of chromosome conformation capture (3C) assay include several related 4C and Hi-C methods.139 The 4C approach, which can be based on hybridization or sequencing technology, employs inverse PCR to detect all fragments ligated to a locus of choice. Finally, Hi-C, can map the three-dimensional organization of chromosomes at a resolution of several kilobase pairs at the scale of the complete genome. This technology employs restriction enzymes to fragment chromatin followed by filling in of the staggered ends using biotinylated nucleotides before DNA ligation. DNA fragments containing ligation junctions are purified using streptavidin-coated beads and are then directly sequenced.139

In summary, these new approaches permit the shift away from gene-centric paradigms to the genome-wide views of the actions of transcription factors and their impact on the global epigenetic and transcriptional status of cells.

Insights into CD4+ T-cell plasticity from epigenetic analysis

The profiling of 39 different histone marks in human CD4+ T cells has been now been accomplished.106 These maps have associated particular modifications with gene activation or repression and with various genomic features, including promoters, transcribed regions, enhancers and insulators. In addition, genome-wide H3K4me3 and H3K27me3 mapping of naive, Th1, Th2, Th17, iTreg and nTreg cells have also been obtained. This information provided some surprises but has also helped to explain some of the behaviour of helper T cells.86 Consistent with the standard ‘lineage commitment’ model of helper-cell differentiation, proximal promoters of cytokine genes show permissive (H3K4me3) marks on lineage-defining cytokines in the respective lineages (e.g. Ifng in Th1 cells), which are accompanied by repressive (H3K27me3) marks in other subsets (Fig. 1). A surprise though, was that the histone methylation patterns of genes encoding key transcriptional factors that drive lineage specification exhibit both repressive and accessible marks in subsets of cells in which the master regulators were not expressed.86 For example, the Tbx21 promoter is associated with H3K4me3 marks in Th1 cells. While H3K27me3 marks are present in Th2 and Th17 cells, H3K4me3 marks are also present even though the transcription factor is not expressed (Fig. 2). Such regions, marked by both chromatin modifications, are termed bivalent domains and have been seen in genes poised for expression in stem cells.136 This then can help explain how T-bet might be induced in Treg, Th17 and even Th2 cells. In this context, it is perhaps not surprising that Th17 cells generated in vitro are intrinsically unstable.69,110,137 Indeed this instability is correlated with the rapid changes of the epigenetic modifications when Th17 cells are stimulated in the presence of IL-12.110 In contrast, isolated memory Th17 cells appear to be more stable and although it was not shown, their stability was hypothesized to relate to the epigenetic landscape of memory cells.138 The Gata3 locus was also noted to exhibit bivalent domains and so helps to explain how a novel subset of Th2 memory/effector cells, which express GATA3 and RORγt and produce Th17 and Th2 cytokines, might arise.75 Similarly, Foxp3 was associated with strong H3K4me3 marks in Treg cells and H3K27me3 was detected in Th1 and Th2 cells as expected. However, no H3K27me3 repressive marks were present in the Foxp3 locus in Th17 cells, consistent with the evidence that Foxp3 can be expressed in Th17 cells.139

Figure 1.

Figure 1

Epigenetic modifications of cytokine loci promote stability of phenotype. H3K4me3 (red) and H3K27me3 (blue) modifications on Ifng and Il17a loci in T helper type 1 (Th1), Th2 and Th17 cells are shown. In Th1 cells, proximal promoter of Ifng shows strong permissive (H3K4me3) marks in the absence of repressive (H3K27me3) marks, while Th2 and Th17 cells demonstrate clearly opposite patterns of epigenetic modifications (strong H3K27me3 marks without H3K4me3 marks) in Ifng locus. In the Il17a locus, the pattern of H3K4me3 and H3K27me3 are clearly distinguishable between Th17 cells and other cells with accessible marks in Th17 cells and repressive marks elsewhere.

Figure 2.

Figure 2

Bivalent epigenetic modifications of master regulator may contribute to T helper (Th) lineage plasticity. H3K4me3 (red) and H3K27me3 (blue) modifications on Tbx21 locus in Th1, Th2 and Th17 cells are shown. In Th1 cells, promoter of Tbx21 is enriched with strong permissive (H3K4me3) marks and depleted of repressive (H3K27me3) modifications. In Th2 and Th17 cells, both H3K4me3 and H3K27me3 modification are present across the gene-body. This is referred to as bivalent chromatin structure.

STATs and the control of the epigenetic landscape of differentiating helper T cells

Since the global epigenetic landscape of the different helper cell subsets is distinct, the question immediately arises, what factor(s) are responsible for the unique epigenetic profiles? As STAT family transcription factors are critical for sensing the cytokine milieu and promoting Th-cell differentiation, it seemed logical to consider that they might contribute to the distinctive modifications seen in helper cell subsets. However, the extent to which STATs are drivers or followers of epigenetic modifications in T cells was by no means clear, and in fact, initial studies analysing IFN-γ signalling and STAT1 implied that epigenetic changes largely preceded STAT1 binding.140

The first issue to tackle was to define STAT target genes in helper cell subsets. Though it has widely been accepted that STATs are critical mediators of gene transcription,141 transcriptional profiling did not distinguish between directly and indirectly regulated genes. However, ChIP-Seq technology has the capacity to map DNA-protein binding sites genome-wide. In addition, through the use of gene-targeted mice and microarray analysis, one can relate transcription factor binding with factor-dependent gene regulation. In addition, STAT-dependent global epigenetic modifications can also be ascertained. Using ChIP-seq, STAT4 and STAT6 were found to have around 12 000 and 13 000 binding sites in Th1 and Th2 condition, respectively.142 Comparative epigenetic analysis of wild-type versus STAT4-deficient Th1 cells provided evidence that of the approximately 4000 genes bound by STAT4, nearly 1000 had STAT4-dependent alterations in epigenetic modifications. And of these 1000 genes, 200 had highly STAT4-dependent gene expression, as determined by microarray analysis of wild-type versus STAT4-deficient cells. Similarly, of around 4000 genes bound by STAT6, the epigenetic modifications in 970 of them were highly dependent on STAT6. Many of the expected target genes fell into this category. For instance, Ifng, Tbx21, Furin and Il18r in Th1 cells and Il4, Gata3, Il24 and Zbtb32 in Th2 cells. Using the same approach, we mapped STAT3 binding sites in Th17 cells. We found that STAT3 bound to more than 3000 genes and among its targets were many genes implicated in Th17-cell differentiation including Il17a, Il17f, Il21 and Il6ra.143 Interestingly, STAT3 binds to multiple sites in the Il17 locus,67 the most prominent of which are intergenic regions that coincide with conserved non-coding sequences.144

There appear to be subsets of genes that are highly dependent on STATs for promoting both gene expression and the local epigenetic environment.142 Interestingly though, there were subtle differences in how STAT4 and STAT6 affected the epigenetic landscape. Specifically, in a significant portion of genes STAT4 served to enhance H3K4me3 marks. In contrast, the global effect of STAT6 was to predominantly regulate H3K27me3 modifications.

Of particular interest is a subset of genes that are bound by STAT4 in Th1 cells but by STAT6 in Th2 cells, for which the two STATs have opposing effects on local epigenetic patterns. A notable example is the Il18r1–Il18rap locus (Fig. 3).142 Whereas one STAT (in this case STAT4) promotes permissive marks in Th1, the other STAT (STAT6) is critical for establishing repressive marks on the same locus in Th2 cells. This divergent action of STAT4 and STAT6 on the same gene provides a mechanism for preserving gene expression in one lineage while ensuring repression in the opposing lineage.

Figure 3.

Figure 3

Opposing effects of signal transducer and activator of transcription 4 (STAT4) and STAT6 on epigenetic modifications in the Il18rap locus. Both STAT4 and STAT6 bind to the promoter region of the Il18rap locus in T helper type 1 (Th1) and Th2 cells, respectively. In this locus, STAT4 is required for permissive (H3K4me3) marks in Th1 (upper panel), as these marks are absent in Stat4-deficient Th1 cells. On the other hand, STAT6 is critical for establishing repressive (H3K27me3) marks on the same locus in Th2 cells (lower panel), as evidenced by the absence of these marks in Stat6-deficient Th2 cells. This is an illustrative example of the ways that two STATs may have opposing effects on local epigenetic patterns.

These new data revealed yet another mechanism of opposing regulation. As discussed, IL-2 acting through STAT5 inhibits IL-17 production. Using genome-wide ChIP-seq analysis, we found that STAT3 and STAT5 bound the same binding sites in the Il17a-Il17f locus in CD4+ T cells.67 These data suggested the possibility that STAT5 could directly compete with STAT3 for binding to enhancer elements within the Il17a-Il17f genetic locus. In fact, that was the case; IL-2-dependent STAT5 binding by IL-2 was associated with a reduction of STAT3 binding at multiple sites throughout the Il17a–Il17f locus in Th17 cells (Fig. 4). STAT5 recruitment by IL-2 is associated with a reduction in positive epigenetic modifications at these sites, including histone H3 acetylation and H3K4me3 but did not enhance H3K27me3 accumulation (Fig. 4). These data suggest a novel mechanism whereby STAT proteins can repress gene expression not just by promoting the accumulation of repressive marks, but also by direct competition with one another to regulate gene transcription.

Figure 4.

Figure 4

Reciprocal actions of signal transducer and activator of transcription 4 (STAT3) and STAT5 proteins on chromatin in Il17 locus. Interleukin-6 (IL-6) along with other cytokines induces T helper type 17 (Th17) differentiation through STAT3 activation. However, IL-2 inhibits Th17 differentiation in a STAT5-dependent manner. In Th17 cells, activated STAT3 by IL-6 positively regulates H3K4me3 deposition in the Il17 locus. The presence of IL-2 causes activation of STAT5, which inhibits the STAT3 binding in the Il17 locus. Decreased STAT3 binding is associated with a reduction in positive epigenetic marks (H3K4me3) in the Il17 locus.

Conclusions and prospects for the future

In the last few years, there has been an avalanche of new information pertaining to novel Th-cell ‘lineages’. These discoveries have enhanced our understanding of how a key player, the CD4+ T cell, can attain different fates and thereby orchestrate the magnitude and character of immune responses. However, the flexibility versus stability of these phenotypes and the extent to which these fates represent specification and commitment or meta-stable states remain to be determined.

It is well-appreciated that specific cellular phenotypes are associated with distinctive transcriptional profiles and that gene expression is controlled by both trans-acting factors and cis-elements, but in addition a deeper understanding of epigenetics is rapidly emerging. With increasing evidence indicating a close relationship between the gene transcription, transcription factor binding and the state of chromatin in CD4+ T cells, we will hopefully gain insight into how extrinsic cytokines provide signals that induce expression of key transcription factors and how these factors work in concert to change and maintain the helper cell phenotype.

Driven by technological advances, new methods have allowed us to investigate the actions of transcription factors and the ways they modify the epigenome on a genome-wide scale. In addition to better understanding how histone modifications change during the course of helper cell differentiation, a genome-wide view of DNA methylation in CD4+ T cells also remains to be obtained. How histone modifications relate to the gain and loss of DNA methylation eventually need to be sorted out, with emphasis on defining the functional significance of these changes.

In addition to promoters, distal enhancer elements are recognized as a major means of controlling tissue-specific gene expression and functioning in a cell-type-specific manner.145,146 Importantly, a chromatin signature for enhancers is now recognized 145,146 and enhancers have been found to be preferentially occupied by the co-activator, p300.147 Relatively few enhancer elements have been defined in helper T cells; however, with ChIP-seq technology, it will be possible to comprehensively map the dynamic enhancer landscape of differentiating CD4+ T cells and begin to define factors responsible for the organization of these elements.

Recent studies have shown that various types of RNA also play an important regulatory role in transcription and epigenetic modifications. In addition to miRNAs, long non-coding RNAs (lncRNAs) influence the expression of neighbouring protein-coding genes in multiple cell lines. Such lncRNAs can serve as a scaffold for the assembly of transcription factors or other chromatin remodelling enzymes at the promoter.148 In the future, it will be important to identify and characterize RNAs that play fundamental roles in the differentiation of Th cells.

The field of chromosome conformation capture is now capable of mapping the three-dimensional organization of chromosomes at a resolution of several kilobase pairs and at the scale of the complete genome. Future chromatin interaction mapping of CD4+ T cells and its distinct subsets remains to be charted. Such data could provide insights into dynamic properties of chromatin in Th cells and reveal whether the chromatin has a distinct structure in stable Th subsets.

In summary, an assorted array of techniques is now available that permits genome-wide views of transcription factor action, epigenetic regulation and gene expression. What remains unclear is how the cellular microenvironment governs the changing of the Th-cell epigenome and resultant regulation of gene expression. Over the next few years, we will rapidly obtain a far more sophisticated and comprehensive view of Th-cell differentiation and a better understanding of what are the drivers. Ideally, this avalanche of information will provide increasingly nuanced views of lineage stability and plasticity and how this relates to the transition of naive cells to specialized effector/memory cells.

Acknowledgments

We would like to acknowledge the assistance of Gustavo Gutierrez-Cruz and the NIAMS Solexa Sequencing Core, and Dr. Hongwei Sun and the NIAMS Biodata Mining Section. This work was supported by NIH/NIAMS Intramural Research Program (IRP).

Disclosures

The authors have no competing financial interests nor conflicts of interest to disclose.

References

  • 1.Mosmann TR, Cherwinski H, Bond MW, Giedlin MA, Coffman RL. Two types of murine helper T cell clone. I. Definition according to profiles of lymphokine activities and secreted proteins. J Immunol. 1986;136:2348–57. [PubMed] [Google Scholar]
  • 2.Mosmann TR, Coffman RL. TH1 and TH2 cells: different patterns of lymphokine secretion lead to different functional properties. Annu Rev Immunol. 1989;7:145–73. doi: 10.1146/annurev.iy.07.040189.001045. [DOI] [PubMed] [Google Scholar]
  • 3.Abbas AK, Murphy KM, Sher A. Functional diversity of helper T lymphocytes. Nature. 1996;383:787–93. doi: 10.1038/383787a0. [DOI] [PubMed] [Google Scholar]
  • 4.Aliberti J, Jankovic D, Sher A. Turning it on and off: regulation of dendritic cell function in Toxoplasma gondii infection. Immunol Rev. 2004;201:26–34. doi: 10.1111/j.0105-2896.2004.00179.x. [DOI] [PubMed] [Google Scholar]
  • 5.Agnello D, Lankford CS, Bream J, Morinobu A, Gadina M, O'Shea JJ, Frucht DM. Cytokines and transcription factors that regulate T helper cell differentiation: new players and new insights. J Clin Immunol. 2003;23:147–61. doi: 10.1023/a:1023381027062. [DOI] [PubMed] [Google Scholar]
  • 6.Mowen KA, Glimcher LH. Signaling pathways in Th2 development. Immunol Rev. 2004;202:203–22. doi: 10.1111/j.0105-2896.2004.00209.x. [DOI] [PubMed] [Google Scholar]
  • 7.Harrington LE, Hatton RD, Mangan PR, Turner H, Murphy TL, Murphy KM, Weaver CT. Interleukin 17-producing CD4+ effector T cells develop via a lineage distinct from the T helper type 1 and 2 lineages. Nat Immunol. 2005;6:1123–32. doi: 10.1038/ni1254. [DOI] [PubMed] [Google Scholar]
  • 8.Park H, Li Z, Yang XO, et al. A distinct lineage of CD4 T cells regulates tissue inflammation by producing interleukin 17. Nat Immunol. 2005;6:1133–41. doi: 10.1038/ni1261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Bettelli E, Carrier Y, Gao W, Korn T, Strom TB, Oukka M, Weiner HL, Kuchroo VK. Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature. 2006;441:235–8. doi: 10.1038/nature04753. [DOI] [PubMed] [Google Scholar]
  • 10.Mangan PR, Harrington LE, O'Quinn DB, et al. Transforming growth factor-β induces development of the TH17 lineage. Nature. 2006;441:231–4. doi: 10.1038/nature04754. [DOI] [PubMed] [Google Scholar]
  • 11.Ivanov II, McKenzie BS, Zhou L, Tadokoro CE, Lepelley A, Lafaille JJ, Cua DJ, Littman DR. The orphan nuclear receptor RORγt directs the differentiation program of proinflammatory IL-17+ T helper cells. Cell. 2006;126:1121–33. doi: 10.1016/j.cell.2006.07.035. [DOI] [PubMed] [Google Scholar]
  • 12.Liang SC, Tan XY, Luxenberg DP, Karim R, Dunussi-Joannopoulos K, Collins M, Fouser LA. Interleukin (IL)-22 and IL-17 are coexpressed by Th17 cells and cooperatively enhance expression of antimicrobial peptides. J Exp Med. 2006;203:2271–9. doi: 10.1084/jem.20061308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Veldhoen M, Uyttenhove C, van Snick J, et al. Transforming growth factor-β‘reprograms’ the differentiation of T helper 2 cells and promotes an interleukin 9-producing subset. Nat Immunol. 2008;9:1341–6. doi: 10.1038/ni.1659. [DOI] [PubMed] [Google Scholar]
  • 14.Nurieva RI, Chung Y, Hwang D, et al. Generation of T follicular helper cells is mediated by interleukin-21 but independent of T helper 1, 2, or 17 cell lineages. Immunity. 2008;29:138–49. doi: 10.1016/j.immuni.2008.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Duhen T, Geiger R, Jarrossay D, Lanzavecchia A, Sallusto F. Production of interleukin 22 but not interleukin 17 by a subset of human skin-homing memory T cells. Nat Immunol. 2009;10:857–63. doi: 10.1038/ni.1767. [DOI] [PubMed] [Google Scholar]
  • 16.McGeachy MJ, Bak-Jensen KS, Chen Y, Tato CM, Blumenschein W, McClanahan T, Cua DJ. TGF-β and IL-6 drive the production of IL-17 and IL-10 by T cells and restrain T(H)-17 cell-mediated pathology. Nat Immunol. 2007;8:1390–7. doi: 10.1038/ni1539. [DOI] [PubMed] [Google Scholar]
  • 17.Ye P, Rodriguez FH, Kanaly S, et al. Requirement of interleukin 17 receptor signaling for lung CXC chemokine and granulocyte colony-stimulating factor expression, neutrophil recruitment, and host defense. J Exp Med. 2001;194:519–27. doi: 10.1084/jem.194.4.519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Puel A, Cypowyj S, Bustamante J, et al. Chronic mucocutaneous candidiasis in humans with inborn errors of interleukin-17 immunity. Science. 2011;332:65–8. doi: 10.1126/science.1200439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Dardalhon V, Awasthi A, Kwon H, et al. IL-4 inhibits TGF-β-induced Foxp3+ T cells and, together with TGF-β, generates IL-9+ IL-10+ Foxp3– effector T cells. Nat Immunol. 2008;9:1347–55. doi: 10.1038/ni.1677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Eyerich S, Eyerich K, Pennino D, et al. Th22 cells represent a distinct human T cell subset involved in epidermal immunity and remodeling. J Clin Invest. 2009;119:3573–85. doi: 10.1172/JCI40202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Crotty S. Follicular Helper CD4 T Cells (TFH) Annu Rev Immunol. 2011;29:621–63. doi: 10.1146/annurev-immunol-031210-101400. [DOI] [PubMed] [Google Scholar]
  • 22.Fazilleau N, Mark L, McHeyzer-Williams LJ, McHeyzer-Williams MG. Follicular helper T cells: lineage and location. Immunity. 2009;30:324–35. doi: 10.1016/j.immuni.2009.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.King C, Tangye SG, Mackay CR. T follicular helper (TFH) cells in normal and dysregulated immune responses. Annu Rev Immunol. 2008;26:741–66. doi: 10.1146/annurev.immunol.26.021607.090344. [DOI] [PubMed] [Google Scholar]
  • 24.Nurieva RI, Chung Y, Martinez GJ, Yang XO, Tanaka S, Matskevitch TD, Wang YH, Dong C. Bcl6 mediates the development of T follicular helper cells. Science. 2009;325:1001–5. doi: 10.1126/science.1176676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Johnston RJ, Poholek AC, DiToro D, et al. Bcl6 and Blimp-1 are reciprocal and antagonistic regulators of T follicular helper cell differentiation. Science. 2009;325:1006–10. doi: 10.1126/science.1175870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Yu D, Rao S, Tsai LM, et al. The transcriptional repressor Bcl-6 directs T follicular helper cell lineage commitment. Immunity. 2009;31:457–68. doi: 10.1016/j.immuni.2009.07.002. [DOI] [PubMed] [Google Scholar]
  • 27.Crotty S, Johnston RJ, Schoenberger SP. Effectors and memories: Bcl-6 and Blimp-1 in T and B lymphocyte differentiation. Nat Immunol. 2010;11:114–20. doi: 10.1038/ni.1837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Shevach EM. CD4+ CD25+ suppressor T cells: more questions than answers. Nat Rev Immunol. 2002;2:389–400. doi: 10.1038/nri821. [DOI] [PubMed] [Google Scholar]
  • 29.Sakaguchi S, Yamaguchi T, Nomura T, Ono M. Regulatory T cells and immune tolerance. Cell. 2008;133:775–87. doi: 10.1016/j.cell.2008.05.009. [DOI] [PubMed] [Google Scholar]
  • 30.Josefowicz SZ, Rudensky A. Control of regulatory T cell lineage commitment and maintenance. Immunity. 2009;30:616–25. doi: 10.1016/j.immuni.2009.04.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Amadi-Obi A, Yu CR, Liu X, et al. TH17 cells contribute to uveitis and scleritis and are expanded by IL-2 and inhibited by IL-27/STAT1. Nat Med. 2007;13:711–8. doi: 10.1038/nm1585. [DOI] [PubMed] [Google Scholar]
  • 32.Annunziato F, Cosmi L, Liotta F, Maggi E, Romagnani S. Type 17 T helper cells – origins, features and possible roles in rheumatic disease. Nat Rev Rheumatol. 2009;5:325–31. doi: 10.1038/nrrheum.2009.80. [DOI] [PubMed] [Google Scholar]
  • 33.Steinman L. Mixed results with modulation of TH-17 cells in human autoimmune diseases. Nat Immunol. 2010;11:41–4. doi: 10.1038/ni.1803. [DOI] [PubMed] [Google Scholar]
  • 34.Saijo S, Ikeda S, Yamabe K, et al. Dectin-2 recognition of α-mannans and induction of Th17 cell differentiation is essential for host defense against Candida albicans. Immunity. 2010;32:681–91. doi: 10.1016/j.immuni.2010.05.001. [DOI] [PubMed] [Google Scholar]
  • 35.Kyttaris VC, Zhang Z, Kuchroo VK, Oukka M, Tsokos GC. Cutting edge: IL-23 receptor deficiency prevents the development of lupus nephritis in C57BL/6-lpr/lpr mice. J Immunol. 2010;184:4605–9. doi: 10.4049/jimmunol.0903595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Murphy KM, Reiner SL. The lineage decisions of helper T cells. Nat Rev Immunol. 2002;2:933–44. doi: 10.1038/nri954. [DOI] [PubMed] [Google Scholar]
  • 37.Hamelmann E, Gelfand EW. IL-5-induced airway eosinophilia – the key to asthma? Immunol Rev. 2001;179:182–91. doi: 10.1034/j.1600-065x.2001.790118.x. [DOI] [PubMed] [Google Scholar]
  • 38.Wills-Karp M. Interleukin-13 in asthma pathogenesis. Immunol Rev. 2004;202:175–90. doi: 10.1111/j.0105-2896.2004.00215.x. [DOI] [PubMed] [Google Scholar]
  • 39.Umetsu DT, DeKruyff RH. The regulation of allergy and asthma. Immunol Rev. 2006;212:238–55. doi: 10.1111/j.0105-2896.2006.00413.x. [DOI] [PubMed] [Google Scholar]
  • 40.Hirahara K, Yamashita M, Iwamura C, et al. Repressor of GATA regulates TH2-driven allergic airway inflammation and airway hyperresponsiveness. J Allergy Clin Immunol. 2008;122:512–20. doi: 10.1016/j.jaci.2008.06.004. e11. [DOI] [PubMed] [Google Scholar]
  • 41.Brunkow ME, Jeffery EW, Hjerrild KA, et al. Disruption of a new forkhead/winged-helix protein, scurfin, results in the fatal lymphoproliferative disorder of the scurfy mouse. Nat Genet. 2001;27:68–73. doi: 10.1038/83784. [DOI] [PubMed] [Google Scholar]
  • 42.Bennett CL, Christie J, Ramsdell F, et al. The immune dysregulation, polyendocrinopathy, enteropathy, X-linked syndrome (IPEX) is caused by mutations of FOXP3. Nat Genet. 2001;27:20–1. doi: 10.1038/83713. [DOI] [PubMed] [Google Scholar]
  • 43.Wildin RS, Ramsdell F, Peake J, et al. X-linked neonatal diabetes mellitus, enteropathy and endocrinopathy syndrome is the human equivalent of mouse scurfy. Nat Genet. 2001;27:18–20. doi: 10.1038/83707. [DOI] [PubMed] [Google Scholar]
  • 44.Reiner SL, Locksley RM. The regulation of immunity to Leishmania major. Annu Rev Immunol. 1995;13:151–77. doi: 10.1146/annurev.iy.13.040195.001055. [DOI] [PubMed] [Google Scholar]
  • 45.Murphy KM, Ouyang W, Farrar JD, Yang J, Ranganath S, Asnagli H, Afkarian M, Murphy TL. Signaling and transcription in T helper development. Annu Rev Immunol. 2000;18:451–94. doi: 10.1146/annurev.immunol.18.1.451. [DOI] [PubMed] [Google Scholar]
  • 46.Marie JC, Letterio JJ, Gavin M, Rudensky AY. TGF-β1 maintains suppressor function and Foxp3 expression in CD4+ CD25+ regulatory T cells. J Exp Med. 2005;201:1061–7. doi: 10.1084/jem.20042276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Ghoreschi K, Laurence A, Yang XP, et al. Generation of pathogenic TH17 cells in the absence of TGF-β signalling. Nature. 2010;467:967–71. doi: 10.1038/nature09447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Das J, Ren G, Zhang L, et al. Transforming growth factor β is dispensable for the molecular orchestration of Th17 cell differentiation. J Exp Med. 2009;206:2407–16. doi: 10.1084/jem.20082286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Yao Z, Kanno Y, Kerenyi M, et al. Nonredundant roles for Stat5a/b in directly regulating Foxp3. Blood. 2007;109:4368–75. doi: 10.1182/blood-2006-11-055756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Takeda K, Noguchi K, Shi W, Tanaka T, Matsumoto M, Yoshida N, Kishimoto T, Akira S. Targeted disruption of the mouse Stat3 gene leads to early embryonic lethality. Proc Natl Acad Sci U S A. 1997;94:3801–4. doi: 10.1073/pnas.94.8.3801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Bacon CM, Petricoin EF, III, Ortaldo JR, Rees RC, Larner AC, Johnston JA, O'Shea JJ. Interleukin 12 induces tyrosine phosphorylation and activation of STAT4 in human lymphocytes. Proc Natl Acad Sci U S A. 1995;92:7307–11. doi: 10.1073/pnas.92.16.7307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Thierfelder WE, van Deursen JM, Yamamoto K, et al. Requirement for Stat4 in interleukin-12-mediated responses of natural killer and T cells. Nature. 1996;382:171–4. doi: 10.1038/382171a0. [DOI] [PubMed] [Google Scholar]
  • 53.Kaplan MH, Sun YL, Hoey T, Grusby MJ. Impaired IL-12 responses and enhanced development of Th2 cells in Stat4-deficient mice. Nature. 1996;382:174–7. doi: 10.1038/382174a0. [DOI] [PubMed] [Google Scholar]
  • 54.Afkarian M, Sedy JR, Yang J, Jacobson NG, Cereb N, Yang SY, Murphy TL, Murphy KM. T-bet is a STAT1-induced regulator of IL-12R expression in naive CD4+ T cells. Nat Immunol. 2002;3:549–57. doi: 10.1038/ni794. [DOI] [PubMed] [Google Scholar]
  • 55.Takeda K, Tanaka T, Shi W, et al. Essential role of Stat6 in IL-4 signalling. Nature. 1996;380:627–30. doi: 10.1038/380627a0. [DOI] [PubMed] [Google Scholar]
  • 56.Zhu J, Paul WE. CD4 T cells: fates, functions, and faults. Blood. 2008;112:1557–69. doi: 10.1182/blood-2008-05-078154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Yang XO, Panopoulos AD, Nurieva R, Chang SH, Wang D, Watowich SS, Dong C. STAT3 regulates cytokine-mediated generation of inflammatory helper T cells. J Biol Chem. 2007;282:9358–63. doi: 10.1074/jbc.C600321200. [DOI] [PubMed] [Google Scholar]
  • 58.Laurence A, Tato CM, Davidson TS, et al. Interleukin-2 signaling via STAT5 constrains T helper 17 cell generation. Immunity. 2007;26:371–81. doi: 10.1016/j.immuni.2007.02.009. [DOI] [PubMed] [Google Scholar]
  • 59.Rubtsov YP, Niec RE, Josefowicz S, Li L, Darce J, Mathis D, Benoist C, Rudensky AY. Stability of the regulatory T cell lineage in vivo. Science. 2010;329:1667–71. doi: 10.1126/science.1191996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hegazy AN, Peine M, Helmstetter C, et al. Interferons direct Th2 cell reprogramming to generate a stable GATA-3+ T-bet+ cell subset with combined Th2 and Th1 cell functions. Immunity. 2010;32:116–28. doi: 10.1016/j.immuni.2009.12.004. [DOI] [PubMed] [Google Scholar]
  • 61.Veldhoen M, Hocking RJ, Atkins CJ, Locksley RM, Stockinger B. TGFβ in the context of an inflammatory cytokine milieu supports de novo differentiation of IL-17-producing T cells. Immunity. 2006;24:179–89. doi: 10.1016/j.immuni.2006.01.001. [DOI] [PubMed] [Google Scholar]
  • 62.Wei L, Laurence A, Elias KM, O'Shea JJ. IL-21 is produced by Th17 cells and drives IL-17 production in a STAT3-dependent manner. J Biol Chem. 2007;282:34605–10. doi: 10.1074/jbc.M705100200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Mathur AN, Chang HC, Zisoulis DG, et al. Stat3 and Stat4 direct development of IL-17-secreting Th cells. J Immunol. 2007;178:4901–7. doi: 10.4049/jimmunol.178.8.4901. [DOI] [PubMed] [Google Scholar]
  • 64.Yang XO, Pappu BP, Nurieva R, et al. T helper 17 lineage differentiation is programmed by orphan nuclear receptors ROR α and ROR γ. Immunity. 2008;28:29–39. doi: 10.1016/j.immuni.2007.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Shevach EM. Mechanisms of foxp3+ T regulatory cell-mediated suppression. Immunity. 2009;30:636–45. doi: 10.1016/j.immuni.2009.04.010. [DOI] [PubMed] [Google Scholar]
  • 66.Yao Z, Cui Y, Watford WT, et al. Stat5a/b are essential for normal lymphoid development and differentiation. Proc Natl Acad Sci U S A. 2006;103:1000–5. doi: 10.1073/pnas.0507350103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Yang XP, Ghoreschi K, Steward-Tharp SM, et al. Opposing regulation of the locus encoding IL-17 through direct, reciprocal actions of STAT3 and STAT5. Nat Immunol. 2011;12:247–54. doi: 10.1038/ni.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Lee YK, Turner H, Maynard CL, Oliver JR, Chen D, Elson CO, Weaver CT. Late developmental plasticity in the T helper 17 lineage. Immunity. 2009;30:92–107. doi: 10.1016/j.immuni.2008.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Shi G, Cox CA, Vistica BP, Tan C, Wawrousek EF, Gery I. Phenotype switching by inflammation-inducing polarized Th17 cells, but not by Th1 cells. J Immunol. 2008;181:7205–13. doi: 10.4049/jimmunol.181.10.7205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Muranski P, Boni A, Antony PA, et al. Tumor-specific Th17-polarized cells eradicate large established melanoma. Blood. 2008;112:362–73. doi: 10.1182/blood-2007-11-120998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Hirota K, Duarte JH, Veldhoen M, et al. Fate mapping of IL-17-producing T cells in inflammatory responses. Nat Immunol. 2011;12:255–63. doi: 10.1038/ni.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Szabo SJ, Kim ST, Costa GL, Zhang X, Fathman CG, Glimcher LH. A novel transcription factor, T-bet, directs Th1 lineage commitment. Cell. 2000;100:655–69. doi: 10.1016/s0092-8674(00)80702-3. [DOI] [PubMed] [Google Scholar]
  • 73.Lazarevic V, Chen X, Shim JH, et al. T-bet represses TH17 differentiation by preventing Runx1-mediated activation of the gene encoding RORγt. Nat Immunol. 2011;12:96–104. doi: 10.1038/ni.1969. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Nistala K, Adams S, Cambrook H, et al. Th17 plasticity in human autoimmune arthritis is driven by the inflammatory environment. Proc Natl Acad Sci U S A. 2010;107:14751–6. doi: 10.1073/pnas.1003852107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Wang YH, Voo KS, Liu B, et al. A novel subset of CD4+ TH2 memory/effector cells that produce inflammatory IL-17 cytokine and promote the exacerbation of chronic allergic asthma. J Exp Med. 2010;207:2479–91. doi: 10.1084/jem.20101376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Molet S, Hamid Q, Davoine F, et al. IL-17 is increased in asthmatic airways and induces human bronchial fibroblasts to produce cytokines. J Allergy Clin Immunol. 2001;108:430–8. doi: 10.1067/mai.2001.117929. [DOI] [PubMed] [Google Scholar]
  • 77.Chakir J, Shannon J, Molet S, Fukakusa M, Elias J, Laviolette M, Boulet LP, Hamid Q. Airway remodeling-associated mediators in moderate to severe asthma: effect of steroids on TGF-β, IL-11, IL-17, and type I and type III collagen expression. J Allergy Clin Immunol. 2003;111:1293–8. doi: 10.1067/mai.2003.1557. [DOI] [PubMed] [Google Scholar]
  • 78.Lund R, Ahlfors H, Kainonen E, Lahesmaa AM, Dixon C, Lahesmaa R. Identification of genes involved in the initiation of human Th1 or Th2 cell commitment. Eur J Immunol. 2005;35:3307–19. doi: 10.1002/eji.200526079. [DOI] [PubMed] [Google Scholar]
  • 79.Wurster AL, Rodgers VL, Satoskar AR, Whitters MJ, Young DA, Collins M, Grusby MJ. Interleukin 21 is a T helper (Th) cell 2 cytokine that specifically inhibits the differentiation of naive Th cells into interferon γ-producing Th1 cells. J Exp Med. 2002;196:969–77. doi: 10.1084/jem.20020620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Reinhardt RL, Liang HE, Locksley RM. Cytokine-secreting follicular T cells shape the antibody repertoire. Nat Immunol. 2009;10:385–93. doi: 10.1038/ni.1715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Zaretsky AG, Taylor JJ, King IL, Marshall FA, Mohrs M, Pearce EJ. T follicular helper cells differentiate from Th2 cells in response to helminth antigens. J Exp Med. 2009;206:991–9. doi: 10.1084/jem.20090303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Tsuji M, Komatsu N, Kawamoto S, Suzuki K, Kanagawa O, Honjo T, Hori S, Fagarasan S. Preferential generation of follicular B helper T cells from Foxp3+ T cells in gut Peyer's patches. Science. 2009;323:1488–92. doi: 10.1126/science.1169152. [DOI] [PubMed] [Google Scholar]
  • 83.Ziegler SF. FOXP3: of mice and men. Annu Rev Immunol. 2006;24:209–26. doi: 10.1146/annurev.immunol.24.021605.090547. [DOI] [PubMed] [Google Scholar]
  • 84.Welniak LA, Blazar BR, Murphy WJ. Immunobiology of allogeneic hematopoietic stem cell transplantation. Annu Rev Immunol. 2007;25:139–70. doi: 10.1146/annurev.immunol.25.022106.141606. [DOI] [PubMed] [Google Scholar]
  • 85.Zhou X, Bailey-Bucktrout SL, Jeker LT, et al. Instability of the transcription factor Foxp3 leads to the generation of pathogenic memory T cells in vivo. Nat Immunol. 2009;10:1000–7. doi: 10.1038/ni.1774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Wei G, Wei L, Zhu J, et al. Global mapping of H3K4me3 and H3K27me3 reveals specificity and plasticity in lineage fate determination of differentiating CD4+ T cells. Immunity. 2009;30:155–67. doi: 10.1016/j.immuni.2008.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Oldenhove G, Bouladoux N, Wohlfert EA, et al. Decrease of Foxp3+ Treg cell number and acquisition of effector cell phenotype during lethal infection. Immunity. 2009;31:772–86. doi: 10.1016/j.immuni.2009.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Feng T, Cao AT, Weaver CT, Elson CO, Cong Y. IL-12 converts Foxp3+ regulatory T Cells to Foxp3+ IFN-γ+ T cells with inhibitory functions during induction of colitis. Gastroenterology. 2011;140:2031–43. doi: 10.1053/j.gastro.2011.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Koch MA, Tucker-Heard G, Perdue NR, Killebrew JR, Urdahl KB, Campbell DJ. The transcription factor T-bet controls regulatory T cell homeostasis and function during type 1 inflammation. Nat Immunol. 2009;10:595–602. doi: 10.1038/ni.1731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Zheng Y, Chaudhry A, Kas A, et al. Regulatory T-cell suppressor program co-opts transcription factor IRF4 to control TH2 responses. Nature. 2009;458:351–6. doi: 10.1038/nature07674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Miyara M, Yoshioka Y, Kitoh A, et al. Functional delineation and differentiation dynamics of human CD4+ T cells expressing the FoxP3 transcription factor. Immunity. 2009;30:899–911. doi: 10.1016/j.immuni.2009.03.019. [DOI] [PubMed] [Google Scholar]
  • 92.Jiang C, Pugh BF. Nucleosome positioning and gene regulation: advances through genomics. Nat Rev Genet. 2009;10:161–72. doi: 10.1038/nrg2522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Zhou VW, Goren A, Bernstein BE. Charting histone modifications and the functional organization of mammalian genomes. Nat Rev Genet. 2011;12:7–18. doi: 10.1038/nrg2905. [DOI] [PubMed] [Google Scholar]
  • 94.Eberharter A, Becker PB. ATP-dependent nucleosome remodelling: factors and functions. J Cell Sci. 2004;17:3707–11. doi: 10.1242/jcs.01175. [DOI] [PubMed] [Google Scholar]
  • 95.Saha A, Wittmeyer J, Cairns BR. Chromatin remodelling: the industrial revolution of DNA around histones. Nat Rev Mol Cell Biol. 2006;7:437–47. doi: 10.1038/nrm1945. [DOI] [PubMed] [Google Scholar]
  • 96.Cedar H, Bergman Y. Linking DNA methylation and histone modification: patterns and paradigms. Nat Rev Genet. 2009;10:295–304. doi: 10.1038/nrg2540. [DOI] [PubMed] [Google Scholar]
  • 97.Kouzarides T. Chromatin modifications and their function. Cell. 2007;128:693–705. doi: 10.1016/j.cell.2007.02.005. [DOI] [PubMed] [Google Scholar]
  • 98.Goldberg AD, Allis CD, Bernstein E. Epigenetics: a landscape takes shape. Cell. 2007;128:635–8. doi: 10.1016/j.cell.2007.02.006. [DOI] [PubMed] [Google Scholar]
  • 99.Heintzman ND, Stuart RK, Hon G, et al. Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the human genome. Nat Genet. 2007;39:311–8. doi: 10.1038/ng1966. [DOI] [PubMed] [Google Scholar]
  • 100.Roh TY, Cuddapah S, Cui K, Zhao K. The genomic landscape of histone modifications in human T cells. Proc Natl Acad Sci U S A. 2006;103:15782–7. doi: 10.1073/pnas.0607617103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Jürgens G. A group of genes controlling the spatial expression of the bithorax complex in Drosophila. Nature. 1985;316:153–5. [Google Scholar]
  • 102.Lewis EB. A gene complex controlling segmentation in Drosophila. Nature. 1978;276:565–70. doi: 10.1038/276565a0. [DOI] [PubMed] [Google Scholar]
  • 103.Kirmizis A, Bartley SM, Kuzmichev A, Margueron R, Reinberg D, Green R, Farnham PJ. Silencing of human polycomb target genes is associated with methylation of histone H3 Lys 27. Genes Dev. 2004;18:1592–605. doi: 10.1101/gad.1200204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Schwartz YB, Pirrotta V. Polycomb silencing mechanisms and the management of genomic programmes. Nat Rev Genet. 2007;8:9–22. doi: 10.1038/nrg1981. [DOI] [PubMed] [Google Scholar]
  • 105.Squazzo SL, O'Geen H, Komashko VM, et al. Suz12 binds to silenced regions of the genome in a cell-type-specific manner. Genome Res. 2006;16:890–900. doi: 10.1101/gr.5306606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Barski A, Cuddapah S, Cui K, et al. High-resolution profiling of histone methylations in the human genome. Cell. 2007;129:823–37. doi: 10.1016/j.cell.2007.05.009. [DOI] [PubMed] [Google Scholar]
  • 107.Mohammad HP, Baylin SB. Linking cell signaling and the epigenetic machinery. Nat Biotechnol. 2010;28:1033–8. doi: 10.1038/nbt1010-1033. [DOI] [PubMed] [Google Scholar]
  • 108.Liefke R, Oswald F, Alvarado C, Ferres-Marco D, Mittler G, Rodriguez P, Dominguez M, Borggrefe T. Histone demethylase KDM5A is an integral part of the core Notch-RBP-J repressor complex. Genes Dev. 2010;24:590–601. doi: 10.1101/gad.563210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Morinobu A, Kanno Y, O'Shea JJ. Discrete roles for histone acetylation in human T helper 1 cell-specific gene expression. J Biol Chem. 2004;279:40640–6. doi: 10.1074/jbc.M407576200. [DOI] [PubMed] [Google Scholar]
  • 110.Mukasa R, Balasubramani A, Lee YK, et al. Epigenetic instability of cytokine and transcription factor gene loci underlies plasticity of the T helper 17 cell lineage. Immunity. 2010;32:616–27. doi: 10.1016/j.immuni.2010.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Kim HP, Leonard WJ. CREB/ATF-dependent T cell receptor-induced FoxP3 gene expression: a role for DNA methylation. J Exp Med. 2007;204:1543–51. doi: 10.1084/jem.20070109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Floess S, Freyer J, Siewert C, et al. Epigenetic control of the foxp3 locus in regulatory T cells. PLoS Biol. 2007;5:e38. doi: 10.1371/journal.pbio.0050038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Huehn J, Polansky JK, Hamann A. Epigenetic control of FOXP3 expression: the key to a stable regulatory T-cell lineage? Nat Rev Immunol. 2009;9:83–9. doi: 10.1038/nri2474. [DOI] [PubMed] [Google Scholar]
  • 114.Xu L, Kitani A, Stuelten C, McGrady G, Fuss I, Strober W. Positive and negative transcriptional regulation of the Foxp3 gene is mediated by access and binding of the Smad3 protein to enhancer I. Immunity. 2010;33:313–25. doi: 10.1016/j.immuni.2010.09.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Yamashita M, Hirahara K, Shinnakasu R, Hosokawa H, Norikane S, Kimura MY, Hasegawa A, Nakayama T. Crucial role of MLL for the maintenance of memory T helper type 2 cell responses. Immunity. 2006;24:611–22. doi: 10.1016/j.immuni.2006.03.017. [DOI] [PubMed] [Google Scholar]
  • 116.Koyanagi M, Baguet A, Martens J, Margueron R, Jenuwein T, Bix M. EZH2 and histone 3 trimethyl lysine 27 associated with Il4 and Il13 gene silencing in Th1 cells. J Biol Chem. 2005;280:31470–7. doi: 10.1074/jbc.M504766200. [DOI] [PubMed] [Google Scholar]
  • 117.Su IH, Dobenecker MW, Dickinson E, et al. Polycomb group protein ezh2 controls actin polymerization and cell signaling. Cell. 2005;121:425–36. doi: 10.1016/j.cell.2005.02.029. [DOI] [PubMed] [Google Scholar]
  • 118.Zhang F, Boothby M. T helper type 1-specific Brg1 recruitment and remodeling of nucleosomes positioned at the IFN-γ promoter are Stat4 dependent. J Exp Med. 2006;203:1493–505. doi: 10.1084/jem.20060066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Parelho V, Hadjur S, Spivakov M, et al. Cohesins functionally associate with CTCF on mammalian chromosome arms. Cell. 2008;132:422–33. doi: 10.1016/j.cell.2008.01.011. [DOI] [PubMed] [Google Scholar]
  • 120.Hadjur S, Williams LM, Ryan NK, Cobb BS, Sexton T, Fraser P, Fisher AG, Merkenschlager M. Cohesins form chromosomal cis-interactions at the developmentally regulated IFNG locus. Nature. 2009;460:410–3. doi: 10.1038/nature08079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Sanyal A, Bau D, Marti-Renom MA, Dekker J. Chromatin globules: a common motif of higher order chromosome structure? Curr Opin Cell Biol. 2011;23:325–31. doi: 10.1016/j.ceb.2011.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Williams A, Spilianakis CG, Flavell RA. Interchromosomal association and gene regulation in trans. Trends Genet. 2010;26:188–97. doi: 10.1016/j.tig.2010.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Spilianakis CG, Flavell RA. Long-range intrachromosomal interactions in the T helper type 2 cytokine locus. Nat Immunol. 2004;5:1017–27. doi: 10.1038/ni1115. [DOI] [PubMed] [Google Scholar]
  • 124.Eivazova ER, Aune TM. Dynamic alterations in the conformation of the Ifng gene region during T helper cell differentiation. Proc Natl Acad Sci U S A. 2004;101:251–6. doi: 10.1073/pnas.0303919101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Lee GR, Spilianakis CG, Flavell RA. Hypersensitive site 7 of the TH2 locus control region is essential for expressing TH2 cytokine genes and for long-range intrachromosomal interactions. Nat Immunol. 2005;6:42–8. doi: 10.1038/ni1148. [DOI] [PubMed] [Google Scholar]
  • 126.Spilianakis CG, Lalioti MD, Town T, Lee GR, Flavell RA. Interchromosomal associations between alternatively expressed loci. Nature. 2005;435:637–45. doi: 10.1038/nature03574. [DOI] [PubMed] [Google Scholar]
  • 127.Zhou X, Jeker LT, Fife BT, Zhu S, Anderson MS, McManus MT, Bluestone JA. Selective miRNA disruption in T reg cells leads to uncontrolled autoimmunity. J Exp Med. 2008;205:1983–91. doi: 10.1084/jem.20080707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Kuchen S, Resch W, Yamane A, et al. Regulation of microRNA expression and abundance during lymphopoiesis. Immunity. 2010;32:828–39. doi: 10.1016/j.immuni.2010.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Park PJ. ChIP-seq: advantages and challenges of a maturing technology. Nat Rev Genet. 2009;10:669–80. doi: 10.1038/nrg2641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Chen X, Xu H, Yuan P, et al. Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell. 2008;133:1106–17. doi: 10.1016/j.cell.2008.04.043. [DOI] [PubMed] [Google Scholar]
  • 131.Wang Z, Zang C, Cui K, Schones DE, Barski A, Peng W, Zhao K. Genome-wide mapping of HATs and HDACs reveals distinct functions in active and inactive genes. Cell. 2009;138:1019–31. doi: 10.1016/j.cell.2009.06.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Pelizzola M, Ecker JR. The DNA methylome. FEBS Lett. 2010;585:1994–2000. doi: 10.1016/j.febslet.2010.10.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Lister R, Pelizzola M, Dowen RH, et al. Human DNA methylomes at base resolution show widespread epigenomic differences. Nature. 2009;462:315–22. doi: 10.1038/nature08514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Pastor WA, Pape UJ, Huang Y, et al. Genome-wide mapping of 5-hydroxymethylcytosine in embryonic stem cells. Nature. 2011;473:394–7. doi: 10.1038/nature10102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Wang Z, Gerstein M, Snyder M. RNA-Seq: a revolutionary tool for transcriptomics. Nat Rev Genet. 2009;10:57–63. doi: 10.1038/nrg2484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Bernstein BE, Meissner A, Lander ES. The mammalian epigenome. Cell. 2007;128:669–81. doi: 10.1016/j.cell.2007.01.033. [DOI] [PubMed] [Google Scholar]
  • 137.Bending D, De La Pena H, Veldhoen M, Phillips JM, Uyttenhove C, Stockinger B, Cooke A. Highly purified Th17 cells from BDC2.5NOD mice convert into Th1-like cells in NOD/SCID recipient mice. J Clin Invest. 2009;119:565–72. doi: 10.1172/JCI37865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Lexberg MH, Taubner A, Forster A, Albrecht I, Richter A, Kamradt T, Radbruch A, Chang HD. Th memory for interleukin-17 expression is stable in vivo. Eur J Immunol. 2008;38:2654–64. doi: 10.1002/eji.200838541. [DOI] [PubMed] [Google Scholar]
  • 139.Zhou L, Lopes JE, Chong MM, et al. TGF-β-induced Foxp3 inhibits TH17 cell differentiation by antagonizing RORγt function. Nature. 2008;453:236–40. doi: 10.1038/nature06878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Robertson AG, Bilenky M, Tam A, et al. Genome-wide relationship between histone H3 lysine 4 mono- and tri-methylation and transcription factor binding. Genome Res. 2008;18:1906–17. doi: 10.1101/gr.078519.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Levy DE, Darnell JE., Jr Stats: transcriptional control and biological impact. Nat Rev Mol Cell Biol. 2002;3:651–62. doi: 10.1038/nrm909. [DOI] [PubMed] [Google Scholar]
  • 142.Wei L, Vahedi G, Sun HW, et al. Discrete roles of STAT4 and STAT6 transcription factors in tuning epigenetic modifications and transcription during T helper cell differentiation. Immunity. 2010;32:840–51. doi: 10.1016/j.immuni.2010.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Durant L, Watford WT, Ramos HL, et al. Diverse targets of the transcription factor STAT3 contribute to T cell pathogenicity and homeostasis. Immunity. 2010;32:605–15. doi: 10.1016/j.immuni.2010.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Akimzhanov AM, Yang XO, Dong C. Chromatin remodeling of interleukin-17 (IL-17)-IL-17F cytokine gene locus during inflammatory helper T cell differentiation. J Biol Chem. 2007;282:5969–72. doi: 10.1074/jbc.C600322200. [DOI] [PubMed] [Google Scholar]
  • 145.Visel A, Rubin EM, Pennacchio LA. Genomic views of distant-acting enhancers. Nature. 2009;461:199–205. doi: 10.1038/nature08451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Heintzman ND, Hon GC, Hawkins RD, et al. Histone modifications at human enhancers reflect global cell-type-specific gene expression. Nature. 2009;459:108–12. doi: 10.1038/nature07829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Visel A, Blow MJ, Li Z, et al. ChIP-seq accurately predicts tissue-specific activity of enhancers. Nature. 2009;457:854–8. doi: 10.1038/nature07730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Orom UA, Derrien T, Beringer M, et al. Long noncoding RNAs with enhancer-like function in human cells. Cell. 2010;143:46–58. doi: 10.1016/j.cell.2010.09.001. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Immunology are provided here courtesy of British Society for Immunology

RESOURCES