GET THE APP

TEC and MAPK Kinase Signalling Pathways in T helper (TH) cell Dev
Journal of Clinical and Cellular Immunology

Journal of Clinical and Cellular Immunology
Open Access

ISSN: 2155-9899

+44 1223 790975

Research Article - (2012) Volume 0, Issue 0

TEC and MAPK Kinase Signalling Pathways in T helper (TH) cell Development, TH2 Differentiation and Allergic Asthma

Yashaswini Kannan and Mark S. Wilson*
Division of Molecular Immunology, National Institute for Medical Research, MRC, London, NW7 1AA, UK
*Corresponding Author: Mark S. Wilson, Division of Molecular Immunology, National Institute for Medical Research, MRC, London, NW7 1AA, UK, Tel: +44 (0)208 816 2189, Fax: +44 (0)208 816 2085 Email:

Abstract

Significant advances in our understanding of the signalling events during T cell development and differentiation have been made in the past few decades. It is clear that ligation of the T cell receptor (TCR) triggers a series of proximal signalling cascades regulated by an array of protein kinases. These orchestrated and highly regulated series of events, with differential requirements of particular kinases, highlight the disparities between αβ+CD4+ T cells. Throughout this review we summarise both new and old studies, highlighting the role of Tec and MAPK in T cell development and differentiation with particular focus on T helper 2 (TH2) cells. Finally, as the allergy epidemic continues, we feature the role played by TH2 cells in the development of allergy and provide a brief update on promising kinase inhibitors that have been tested in vitro, in pre-clinical disease models in vivo and into clinical studies.

Introduction

Peripheral αβ+CD4+ T cell differentiation

Antigen-restricted effector and effector-memory αβ+CD4+ T cells are a central component of adaptive immunity. Following successful development and activation in the periphery, effector αβ+CD4+ T cells function to rally innate cells, provide help to B cells for antigen-specific immunoglobulin production and stimulate local tissue

responses; while regulatory αβ+CD4+ T cells control the proliferation of effector αβ+CD4+ T cells, suppress innate cell activation and prevent autoimmune reactions. These divergent functions are tightly regulated through cell-intrinsic and extrinsic mechanisms. When these regulatory checkpoints fail, effector αβ+CD4+ T cells can

cause lethal lymphomas or hyper-inflammatory conditions such as autoimmune, allergy and leukaemia. Conversely, if effector αβ+CD4+ T cells fail to develop mature or differentiate, individuals can be left with insufficient immunological protection with equally catastrophic outcomes, such as life-threatening severe immunodeficiency. Similarly, failure of regulatory αβ+CD4+ T cell development can allow the activation of auto-reactive T cells and uncontrolled inflammation [1]. Thus, throughout the development, differentiation, activation, effector/ regulatory function and long-term survival, multiple feedback loops are in place regulating αβ+CD4+ T cell responses.

Once in the periphery, αβ+CD4+ T cells can reversibly differentiate into a variety of helper (TH) / effector (TH1, TH2, TH17) T follicular helper (TFH) and regulatory (TREG) populations, often characterized by their cytokine expression profile and up-stream transcription factors (reviewed elsewhere [2-5]). With occasional exceptions, the molecular programs involved in the differentiation of TH, TFH or TREG cells are mostly well defined. For example, IFNγ and IL-12 stimulate Tbx21 (T-bet) through activation of STAT-1 and STAT-4 for TH1 differentiation and IL-4- and IL-2-induce GATA-3 / STAT-6 and STAT-5 for TH2 differentiation. Similarly, IL-6 and TGFβ promote RORγt and STAT-3 for TH17 differentiation and IL-4 in combination with TGFβ induces PU-1 for TH9 differentiation (thoroughly reviewed [6,7]). While, the precise signals required for TFH cell differentiation are unclear, BCL6 has been demonstrated to orchestrate part of the TFH cell developmental program [8,9]. Finally, IL-2 and TGFβ induce STAT-5 and Foxp3, which prescribe natural TREG (nTREG) development in the thymus or inducible TREG (iTREG) development in the periphery [10]. Foxp3, a transcription factor restricted to TREG cells is essential for TREG development, maintenance and function [11-14]. Despite their importance in specifying TH cell lineage commitment, many of these transcription factors are not self sufficient in coordinating complete transcriptional programs; for example, Bcl6 and PU-1 for TFH and TH9 cell differentiation respectively [7,8]. This suggests a role for multiple transcriptional regulators functioning together in TH cell differentiation. Although differentiated αβ+CD4+ T cells can adopt different effector/ regulatory characteristics, there is significant flexibility between their phenotypes [15-18]. In addition to being phenotypically flexible, different αβ+CD4+ phenotypes share a common activation step via the T cell receptor (TCR).

T cell receptor (TCR) complex and proximal signalling events

The αβ+TCR functions as a biological bottleneck, translating peptide-loaded MHCII-delivered messages into cellular responses via signalling modules and a series of inter-dependent signalling cascades. Signals transmitted via these pathways influence T cell fate decisions in the thymus, differentiation and proliferation in the periphery and antigen-induced cell death. The α and β subunits of the TCR, similar to the γ and δ subunits, undergo a series of selection processes during T cell ontogeny in the thymus. Pairing of β subunits with α chain occurs during the double negative 3 (DN3) stage with the emergence of CD4+CD8+ (double positive, DP) thymocytes. At this stage, αβ+T cells are again selected in the thymus by their ability to respond, or not, to antigen. Just as a response generated in the absence of an antigen leads to cell death by neglect, strong antigen-induced responses also result in cell death by positive selection. Thymocytes that respond weakly to the antigen undergo further selection into CD4+ or CD8+ single positive cells that mature further to contribute to the peripheral T lymphocyte pool (a detailed review of this process can be found at [19]).

The α and β chains of the TCR encode short cytoplasmic tails and do not signal. Instead, they transduce TCR signals through the noncovalently associated CD3 proteins that exist as a series of hetero (CD3γε and CD3δε) and homodimers (CD3ζζ) along with the αβ chains [20,21]. The interaction between the αβ chains and CD3 is facilitated by positively charged amino acids in the αβ transmembrane domain and the negatively charged residues in the transmembrane regions of the CD3 complex [22]. Successful selection of a mature T lymphocyte involves the expression of all the components of the TCR, with unassembled protein components ubiquitinated and proteolytically degraded by the endoplasmic reticulum [23]. The cytoplasmic domain of CD3 encode immunoreceptor tyrosine-based activation motif (ITAM) that play a crucial role in initiating signalling downstream of the TCR.

Briefly, in response to TCR engagement by antigen, clustering of the TCR and the co-receptors induces protein tyrosine kinases (PTKs) such as Lck, a member of the Src family of tyrosine kinases, to phosphorylate the tyrosine residues in the ITAMs in the cytoplasmic tails of CD3. These phosphorylated ITAMs serve as docking sites for the Src homology 2 (SH2) domain containing tyrosine kinase ZAP- 70 (ζ-associated protein of 70kD). Recruited ZAP-70 is then able to phosphorylate adaptor proteins such as LAT and SLP-76, which brings about downstream activation of different signalling pathways like the PI3-kinase pathway, Ras-MAPK pathway, Ca2+-mediated signalling and Vav-1 mediated signalling leading to actin-cytoskeletal reorganization (Figure 1).

clinical-cellular-immunology-TCR-signaling

Figure 1: Proximal TCR signaling
Following TCR engagement protein tyrosine kinases (PTKs) such as Lck, phosphorylate the tyrosine residues in the ITAMs in the cytoplasmic tails of CD3ζζ, which serve as docking sites for the Src homology 2 (SH2) domain containing tyrosine kinase ZAP-70. ZAP-70 phosphorylates adaptor proteins, including LAT and SLP-76, activating PI3-kinase pathway, Ras- MAPK pathways and Ca2+-mediated signalling. In the appropriate cytokine environment (red cytokines), CD4+ T cells differentiate into reversible TH1, TH2, TH9, TH17 and Treg cells and secrete a suite of characteristic cytokines (blue, bold cytokines).

The kinase kingdom

Over 500 kinases regulate many aspects of a cell function. By the addition of phosphate groups to substrate proteins, kinases coordinate protein movement, localisation and activity. With regard to T lymphocytes, three PTK families namely, the Src family kinases (Lck, Fyn), the Syk family kinases (Syk, Zap-70) and the Tec family kinases (Tec, Itk, Btk and Rlk), are essential for peripheral T cell development. Spontaneous mutations in any of these kinase families, or genetic ablation in murine models, results in dramatically dysregulated T cell development [24-30]. In addition to regulating the function of these protein kinases, the T cell kinome also regulates processes ranging from metabolism, actin re-organisation, anergy/ tolerance and cytokine secretion [31-36].

Kinases are subdivided into eight conventional groups and four atypical groups based on sequence similarity, accessory domains and known modes of regulation (Table 1) [37]. The different kinase families share divergent roles in T lymphocyte development, differentiation of TH cells and their effector functions. Although other kinase families serve important functions in T cells, reviewed elsewhere [38-40], Tec kinase and MAPK (ERK, p38 and JNK) signalling pathways have been intensively studied in αβ+CD4+ T cells and are the main focus of this review. We have highlighted the unique features of these kinases in TH2 cells in the context of allergic diseases, and provide an update on current therapeutic strategies using inhibitors to interfere with these signalling pathways in allergic diseases.

Family Description Reference
Conventional Family
AGC Named after Protein Kinase A, G and C families, consisting of 60 members across 16 families of cytoplasmic serine/threonine kinases, regulated by secondary messengers. E.g.: PKC and PKA. Mutation and dysregulation linked to many conditions such as diabetes and cancer. [41,42]
CaMK Named after Calmodulin/Calcium regulated kinases, the CaMK group consists of both calcium and non-calcium regulated kinases. E.g.: CaMK I, CaMK II and CaMK IV. CaMK II is essential for NF-κB activation following TCR ligation. [43,44]
CK1 Named after Casein kinase 1 (CK1), these monomeric serine/threonine selective kinases are evolutionary conserved within eukaryotes. CK1 kinases regulate diverse functions form circadian rhythm, Wnt signalling, DNA replication and RNA metabolism. [45]
CMGC The CMGC group includes the evolutionarily conserved, cyclin-dependent kinases (CDKs), mitogen-activated protein kinases (MAPKs), glycogen synthase kinases (GSKs) and CDK-like kinases (CLKs). Their roles include, cell cycle regulation, immune signalling mediated downstream gene expression, regulating cell proliferation and apoptosis. [32,46-49]
RGC The receptor guanylate cyclases (RGCs) synthesize cyclic GMP from GTP. RGCs are membrane bound and are activated by calcium or the binding of atrial naturetic peptide or related peptides. E.g.: GC-A and GC-G. [50,51]
STE Serine threonine kinases (STE) are subdivided into 3 main families Ste7, Ste11 and Ste20, which sequentially activate each other and MAPK members. Ste20 kinases (MAP4K) act on Ste11 kinases (MAP3K), which activate Ste7 kinases (MAP2K, MEK, MKKs). These families have distinct roles in T lymphocyte activation. [32,52]
TK Tyrosine kinases (TK) evolutionarily conserved in mammals, phosphorylate tyrosine residues. The TK are divided into receptor and non-receptor (cytoplasmic) tyrosine kinases (CTK). The receptor tyrosine kinases are activated by extracellular signals such as growth factors at the cell surface. The CTK include, Src kinases such as Tec and janus kinases (JAK). [53,54]
TKL The tyrosine kinases-like (TKL) group has close sequence similarity to the TKs. These diverse, serine/threonine kinases, are subdivided into 7 major families, including the RAF, IRAK and RIPK families. The TKL kinases are involved in a wide range of immune cell processes including cell growth, Toll like receptor (TLR) and IL-1 signalling and cell death. [55-57]
                                                                                    Atypical Family
Alpha α-Kinases have low sequence homology with other conventional kinases and have known functions in protein translation, intracellular transport, cell migration and proliferation. E.g.: elongation factor 2 kinase and Dictyostelium myosin heavy chain kinases [58]
PIKK The phosphoinositide 3-kinase-related kinases (PIKK) are involved in DNA damage, nutrient dependent signalling mRNA decay. E.g.: the mammalian target of rapamycin (mTOR), has been well studied in the context of T cell metabolism and differentiation. [33,59]
PDK The mitochondrial pyruvate dehydrogenase kinases (PDKs), required for oxidative metabolism mediate the phosphorylation of pyruvate dehydrogenase. There are four known mammalian isoforms; PDHK1-4, and these regulate the balance between glycolysis and lipid metabolism. [60,61]
RIO The right open reading frame (RIO) kinases are expressed in archea to humans. Their functions include, ribosome biogenesis and chromosome maintenance, although the precise substrate of the RIO kinases is unknown. RIO kinases from the ruminant nematode, Haemonchus contortus have been targeted for nematocidal drugs. [62,63]

Table 1: Kinase families and their functions.

Involvement of Kinases in αβ+CD4+ T helper cell Development and Differentiation

Tec-family Kinases

The Tec family of non receptor tyrosine kinases came into focus when Btk (Bruton’s tyrosine kinase), one of the members of the family was identified in severe B cell immunodeficiency, X-linked agammaglobulinemia (XLA) and X-linked immunodeficiency in humans and mice, respectively [64-67]. These reports provided the first evidence of how mutations in tyrosine kinases affect lymphocyte signaling and hence could be linked to primary immunodeficiencies. The Tec family of kinases, of which Tec (tyrosine kinase expressed in hepatocellular carcinoma) was the first to be identified, also consists of: Btk, Itk (interleukin-2 (IL-2)-inducible T-cell kinase; also known as Tsk or Emt), Rlk (resting lymphocyte kinase; also known as Txk) and Bmx (bone-marrow tyrosine kinase gene on chromosome X; also known as Etk) [68]. These proteins expressed primarily in hematopoietic cells are characterized by a common domain organization (Figure 2A). With the exception of Rlk, these proteins encode a pleckstrin homology (PH) domain that binds to phosphatidylinositol 3, 4, 5- triphosphate (PtdIns(3,4,5)P3) on their amino-terminal end [68]. Rlk is an atypical Tec kinase that has a palmitoylated string of cysteine residues at its amino-terminal end [69,70]. This is followed by one or two proline rich regions (PRRs), a SRC homology 3 (SH3), SH2 and a kinase domain at the carboxy terminal end [68]. The presence of the PH domain allows for these proteins to be regulated by phosphatidylinositol 3- kinase (PI3K) mediated signalling, while the cysteine string in Rlk allows for its constitutive association with the plasma membrane independent of PI3K signalling [71-73].

clinical-cellular-immunology-kinase-family

Figure 2: Tec kinases structure and their differential roles
The different members of the Tec kinase family are shown in a schematic. The proteins typically encode a N terminal pleckstrin homology domain (PH), followed by a Btk homology domain (BH), a proline rich region (PRR), a Src homology 3 (SH3) domain, a SH2 domain and a carboxy terminal kinase domain. Rlk, unlike the other members, encodes a palmitoylated string of cysteine residues on its amino terminal domain. Itk is expressed maximally in and is important for TH2 differentiation, while Rlk has a significant role in regulating TH1 cells. The role of Itk in TH1 cells is unclear, while Rlk may compensate for Itk function in Itk-/- cells. BMS-509744 is a pharmacological inhibitor of Itk.

Typically, recruitment of Tec kinases to the plasma membrane through interaction with PtdIns(3,4,5)P3 (generated upon PI3K action on phosphatidylinositol-4,5-bisphosphate (PtdIns(4,5)P2)), is followed by the phosphorylation of a tyrosine residue in the kinase domain by Src-family kinases, which brings about conformational changes that allow for the recruitment of these proteins into antigenreceptor signalling complexes [68,74]. Downstream phosphorylation of adaptor molecules, LAT and SLP-76 brings about the recruitment of the protein tyrosine kinase Itk. Itk in turn phosphorylates and activates PLC-γ, which catabolises inositol-1,4,5-triphosphate (Ins(1,4,5)P3) and diacylglycerol (DAG). This activates DAG binding proteins like protein kinase C (PKC) and Ras guanyl releasing proteins (RasGRP), which activate the MAPK family leading to downstream effector responses (Figure 1). Increases in the intracellular Ca2+ levels due to Ins(1,4,5)P3 signalling also activates the nuclear factor of activated T cells (NFAT) family of transcription factors, which bring about changes in the gene expression levels.

In addition to being primarily activated in response to TCR signalling, Tec kinases are also involved in chemokine, cytokine, growth factor and co-stimulatory signaling pathways [74]. For example, Itk and Tec interact with the T cell co-stimulatory receptor CD28, where they become phosphorylated and mediate downstream signalling [75-77]. Itk is involved in CD2, CXCR4 and the chemokine, stromal cell derived factor 1a (SDF-1a)/(CXCL12) mediated signalling [78-81]. Tec regulates signalling downstream of cytokine receptors, IL-3, IL-5, prolactin, erythropoietin, GCSF, c-kit and the IL-6 family, while Rlk is involved in CXCL12 and MIP3-β-mediated signalling [74,81].

Tec kinases and T lymphocytes

T lymphocytes express Itk, Rlk and Tec differentially; suggesting their importance in different sub-populations of T lymphocytes, with Itk maximally expressed in naïve T cells [68,74]. T cell activation further increases the expression of Itk, with higher expression of the protein in TH2 cells compared to TH1 cells. In this regard, Itk has been implicated in the development of TH2 responses [82]. In addition to T cell activation, IL-2R signalling can also induce Itk expression [82,83]. Rlk, akin to Itk, is also expressed in thymocytes and mature naïve T lymphocytes, though its expression levels are 3-10 fold lower in naïve T lymphocytes compared to Itk [84]. Unlike Itk, Rlk is preferentially expressed in TH1 cells implicating its importance in TH1 responses [85,86]. Tec has relatively low expression in T lymphocytes compared to Itk and Rlk with 100-fold lower expression than Itk. Tec is increased upon stimulation and, similar to Itk, is expressed higher in TH2 cells than in TH1 cells [84,87]. These observations suggest that Tec kinases play an important role in regulating effector T cell differentiation.

Function of Tec kinases in T lymphocytes

In accordance with their roles in TCR activation, mice deficient in Tec kinases demonstrate defective PLC-γ phosphorylation and activation. Consequently, this leads to defective production of Ins(1,4,5)P3, reduced Ca2+ influx and activation of downstream transcription factors. Impaired activation of MAPKs like extracellular signal regulated kinase (ERK), JUN N-terminal kinase (JNK) and activated protein-1 (AP-1) is also observed in these cells [68]. Mutations affecting Tec kinases lead to impaired TCR stimulation, proliferation and IL-2 production [27,88,89]. Itk was first demonstrated to be important for the development of conventional T lymphocytes [27]. This was followed by studies demonstrating the role of Itk and Rlk in positive selection and negative selection in response to both MHC class I and class II restricted transgenes [27,28,90]. These roles of Itk and Rlk were attributed to decreased ERK activation and defects in cellular adhesion in the kinase deficient cells [28,91,92]. Among other important roles, Tec kinases have been implicated in the development and differentiation of T lymphocytes. Specifically, Itk deficiency led to impaired recruitment of Vav-1, a guanine nucleotide exchange factor, at the TCR, resulting in disruptions in the nucleation of actin polymerization [68]. Itk also regulates integrin expression and adhesion, following TCR engagement, with Itk-deficiency resulting in reduced TCR activation [93]. Collectively both Itk and Rlk regulate the strength of signal originating from the TCR during T cell development. It is also important to note that beyond αβ+ T cells, Itk is also required for the development and function of iNKT cells and γδ T cells (reviewed in [94]).

Itk and Rlk have been analysed for their roles in cytokine production, immunity to infection and inflammatory diseases, such as allergy and autoimmunity. These studies have established clear roles for Itk in the differentiation of TH2 cells (Figure 2B), with CD4+ T cells from Itk-/- mice impaired in their ability to produce TH2 cytokines in vitro [82,83,95,96]. Concordant with this, Itk-/- mice failed to mount TH2 responses following infection with Leishmania major, Nippostrongylus brasiliensis, Schistosoma mansoni, or allergen-induced airway inflammation [82,83,95]. In contrast, Itk-/- mice were capable of generating effective TH1 responses in these settings. Consistent with its role in regulating TH2 responses, enhanced Itk expression was observed in patients with atopic dermatitis, a TH2-associated inflammatory disease of the skin [97]. These defects have been attributed to impaired recruitment of Vav-1 and PKC-θ following TCR activation, with reduced NFAT activation, as well as the requirement of Itk to support the maintenance and amplification of the TH2 effector responses [98].

Rlk, distinct from Itk, is not only expressed in high levels in TH1 cells but also bound to the Ifng promoter and was required for the production of IFN-γ [86] (Figure 2B). Despite this preferential regulation in TH1 cells, Rlk-/- mice show only minor defects in response to infection by Toxoplasma gondii, a TH1 inducing pathogen [83,89]. Though this observation is yet to be completely understood, it is currently thought that the existence of compensatory mechanisms in these mice could contribute to this effect [68].

Despite the distinct roles of Itk and Rlk in TH cell differentiation, further examination of Itk/Rlk double knockout mice under different TH cell regulatory conditions suggests that this dichotomy might be an oversimplified model. Itk/Rlk double knockout (DKO) mice upon initial characterization demonstrated severely impaired TH1 responses following T. gondii infection [89]. In contrast, Itk/Rlk DKO mice mounted an effective TH2 response following S. mansoni infection, despite impaired NFAT activation [83]. These differences have lead to many different ideas about the roles of the individual Tec kinases in TH cells [68].

The models put forth to explain these differences in the Itk-/-, Rlk-/- and the Itk/Rlk DKO mice include, the idea of “differential expression of Tec kinases in TH cells regulate their distinct functions”. This idea is supportive of the individual roles of Itk and Rlk in TH1 and TH2 cells respectively, and the compensatory roles of Itk and Rlk, as the expression of an Rlk transgene in the Itk-/- mice rescues the impaired TH2 response upon challenge with TH2-inducing Schistosome egg antigen (SEA) or in a mouse model of allergic asthma [99].

Another model draws its basis from, “the TCR signal strength dictates TH-cell development”. This model was put forth to explain the TH2 biased nature of Itk/Rlk DKO mice upon S. mansoni infection. This notion would explain how weak signals, such as low antigen doses or altered peptide ligands, promotes a TH2 biased response [100], as activated cells down regulate GATA-3 and differentiate into other lineages [83]. Concordant with this, Itk/Rlk DKO mice were severely compromised in signalling downstream of their TCR and T cells from these mice were defective in their ability to repress the expression of GATA-3 [83]. This suggests, that regulation of GATA-3 expression by TCR signalling in Itk/Rlk wild type cells is precedent to the differentiation of the TH cells to the TH2 lineage.

Defective activation of the NFAT family members in Itk/Rlk DKO mice might also be responsible for the intact TH2 responses [83,89]. Inadequate NFAT activation, due to a lack of, or impaired activation of, NFATc2 and NFATc3, members of the NFAT family, is also associated with increased development and activation of TH2 cells [101,102]. Additionally, the potency of the TCR signal contributes to the differential expression of NFATc1, another NFAT family member, which enhances TH2 polarization, and transcription of il4 [103]. Despite these observations, the differential expression of NFAT proteins in Itk/Rlk DKO mice is debated [68,83]. Since impaired Ca2+ flux is associated with the development of a TH2 phenotype and tyrosine mutant LAT is associated with impaired Ca2+ flux and a TH2 bias, the possibility of a similar LAT mutant in Itk/Rlk DKO mice is a plausible hypothesis worth evaluation [104]. Moreover, the role of T cell extrinsic mechanisms such as contributions from Natural killer (NK) cells, Mast cells, Basophils and Eosinophil’s in Itk/Rlk DKO mice is another aspect yet to be explored [68].

Although the precise roles of Itk and Rlk in modulating TH2 development with contributions from the above-mentioned models is yet to be confirmed, the role of Itk in modulating TH2 responses remains undisputed. These observations have prompted several studies targeting Itk as a potential therapeutic target for asthma and other inflammatory disorders, described in more detail below [105-107].

Itk also regulates TH17-associated cytokines [108]. CD4+ T lymphocytes from Itk-/- mice expressed very low levels of IL-17A when polarized under TH17 conditions in vitro. This effect was even more dramatic in T lymphocytes from Itk/Rlk DKO mice, supporting the functional redundancy between the Tec kinases. This role of Itk was once again coupled to its ability to activate NFAT in response to TCR signalling, suggesting activated NFAT binds to the Il17a promoter in wild type cells but not in Itk-/- cells [108]. Collectively, Tec kinases regulate distinct T lymphocyte development and functions at multiple levels and provide an ‘Achilles heel’ to target T cell-mediated diseases.

Mitogen Activated Protein Kinases (MAPKs)

MAPKs are one of the most ancient and evolutionarily conserved families of kinases. These proteins are activated in response to cytokines, growth factors, and environmental stresses and are involved in cell proliferation, differentiation and death. The MAPK family includes three major groups in mammalian cells: the extracellular signal regulated kinases (ERKs), the p38 MAPK family and the c-Jun Nterminal kinases (JNK) (Figure 3) [109-111]. Each family of MAPK is activated by dual phosphorylation of threonine and tyrosine residues in their activation loop (Thr- X-Tyr) by evolutionarily conserved upstream kinases, MAPK kinases (MAPKK or MAP2Ks) and MAPK kinase kinases (MAPKKK or MAP3Ks). The MAPKKK are serine/threonine kinases that are typically activated by RAS superfamily GTPases or signalling intermediates in response to cytokines and growth factors. This brings about the phosphorylation of the downstream MAPKK, which activate the MAPKs by dual phosphorylation of the threonine and tyrosine residues in the activation loop. Once activated, the MAPKs phosphorylate their substrates such as transcription factors, phospholipases, cytoskeletal proteins, as well as several additional protein kinase families such as the ribosomal S6 kinases, mitogen and stress activated kinases and MAPK-interacting kinases [49]. This serine/threonine specific phosphorylation of the target substrate is marked by a preceding proline residue. Once phosphorylated, the substrates mediate downstream signalling ranging from regulation of gene expression and cellular differentiation to apoptosis.

clinical-cellular-immunology-molecular-patterns

Figure 3: MAPK activation cascade
MAPKs are activated in response to G-protein coupled receptors (GPCRs), growth factors, mitogens, pathogen associated molecular patterns (PAMPs) like LPS, inflammatory cytokines and environmental stresses. The kinase signalling cascade is organized into a three-tier system, with the activated MAP3Ks activating MAP2Ks, which in turn activate the MAPKs. There is significant crosstalk between the pathways and different MAPKs can activate different TH cell specific transcription factors, that signal the T cells to differentiate accordingly. ERKs and p38 family members are specific for TH2 and TH1 cells respectively. While the role of p38 in TH2 cells remains debated, JNKs are specific inhibitors of TH2 cells. The different pharmacological inhibitors of the MAPKs are also shown.

The ERK family, the first family of mammalian MAPK to be identified, encodes two classical members, ERK1 and ERK2. These kinases are dual phosphorylated at the activation loop Thr-Glu- Tyr and are classically activated in response to mitogens or insulin resulting in the activation of the proto-oncogene, Ras. Ras recruits and activates the Raf family of MAP3Ks. Raf in turn phosphorylates the MAP2K members, MEK1 and MEK2 which activates the downstream ERK family of kinases [112]. The ERK families of kinases are also activated in response to pro-inflammatory cytokines, such as TNFα, IL-1β and pathogen associated molecular patterns (PAMPs) like lipopolysaccharide (LPS) [49]. The ribosomal S6 kinases (RSKs), mitogen and stress activated kinase family members (MSKs), MSK1 and MSK2 and the MAPK-interacting kinase (MNK), MNK1 and MNK2 are selective substrates for the ERK family [113-117]. ERK can also regulate the activation of Elk1, member of the ternary complex factor (TCF) family, which is an important regulator of the AP-1 family of transcription factors [118].

The p38 family, another group of proteins in the MAPK family, has 4 members; p38α, p38β, p38γ and p38δ. These are activated in response to environmental stresses like ionizing radiation and osmotic shock, inflammatory cytokines and PAMPs. Upon activation, the MAP3Ks phosphorylate the MAP2Ks, MKK3/MKK4/MKK6, which in turn phosphorylate the p38 family members in their activation loop, Thr-Gly-Tyr. The p38 family in addition to regulating the MAPKinteracting kinases (MNK1/2) and the mitogen and stress activated protein kinases (MSK-1/2) also regulate the MAPK-activated protein kinases (MK2) [119-122]. The p38 family members also regulate the activation of the transcription factors, C/EBP homologous protein (CHOP/GADD), myocyte enhancer factor 2 (MEF2) family members MEF2A/C and ATF2 of the AP-1 family [123-125]. The p38s also phosphorylate Sap-1a, a member of the TCF family, which together with serum response factor (SRF) binds to the serum response element (SRE) on the c-fos promoter and mediates c-fos induction [126,127].

The last group of the MAPK family, the JNK family, has 3 main members in eukaryotic cells; JNK1, JNK2 and JNK3 [111]. The JNK family, in addition to being activated by cytokines and growth factors, is also activated by environmental stresses, genotoxins, mechanical stress, pro-inflammatory cytokines, PAMPs and danger associated molecular patterns (DAMPs). The JNK family members are subject to differential pre-mRNA splicing thus giving rise to 12 different JNK polypeptides. While the functional significance of these isoforms remains unclear, it has been reported some of these isoforms (α and β JNK) differ in their affinities for their substrate [128-130]. A typical signalling cascade brings about the phosphorylation of the MAP2Ks, MKK4 and MKK7 by MAP3Ks. This in turn phosphorylates the JNK family at the activation loop Thr-Pro-Tyr to activate downstream signalling. A major group of JNK substrates are the AP-1 family of transcription factors including, c-Jun, JunD and ATF2 [128,131,132]. JNKs also phosphorylate Elk-1 at the carboxy terminus, contributing to the regulation of c-fos [133], NFAT4 and NFAT2 [134,135]. Since NFAT2 activity is important for the polarization of TH cells to TH2 cells, JNK mediated inhibition of NFAT2 has been shown to negatively regulate TH2 differentiation [136,137].

MAPKs in T cell Development, Differentiation and Function

ERK and T lymphocytes

The role of MAPKs, specifically the MEK-ERK pathway, in early thymocyte development was initially established in fetal thymic organ cultures (FTOC) exhibiting compromised differentiation of immature CD25+CD44- thymocytes to the pre-TCR expressing thymocytes [138]. Further support for the functional role of ERK in pre-TCR development came from studies that demonstrated that transfection with a TCR-β chain in Rag2-/- FTOC resulted in ERK activation. Furthermore, constitutive activation of ERK using a mutant form of c-Raf-1 induced maturation and expansion of CD25-CD44- thymocytes from Rag2-/- mice [139,140].

The importance of the Ras-ERK pathway in late thymocyte development was also demonstrated when Erk1-/- mice had reduced percentages of mature CD4+/CD8+ thymocytes [141]. The role of the Ras-ERK pathway in positive selection was substantiated when Delgado and colleagues reported the inability of Cd3δ-/- DP thymocytes to undergo positive selection [142]. This report held important implications for the role of ERK in thymocyte maturation and TCR signalling, as the positive selection defect was coupled to impaired ERK activation, due to defective tyrosine phosphorylation of the upstream adaptor LAT and the CD3ζ chain upon engagement of the TCR. This suggested that activation of the LAT-Ras-ERK pathway was important for positive selection. ERK function in TCR signalling was also validated by studies linking the adaptor proteins, SLP76 and Grb2 to the downstream activation of the SOS-Ras-MEK-ERK pathway [143-145]. Simultaneously, studies demonstrated the functional significance of clonal anergy linked to incomplete Ras and ERK activation [146,147].

In addition to TCR signalling, the Ras-ERK pathway also regulates TH2 differentiation. ERK activation, mediated by MEK, induced phosphorylation of STAT-6 and IL-4Rα in response to IL-4 [148]. Furthermore, TCR activation of the Ras-ERK pathway stabilized the TH2 specific transcription factor GATA-3 through inhibition of its degradation by the ubiquitin proteasome pathway [149]. ERK activation has also been linked to IL-4 expression during TCR induced TH differentiation [150,151]. The mechanism mediating this regulation was governed by the level of ERK activation, which in turn, was governed by the strength of signal originating from the TCR. Mechanistically, low strength TCR activation induced IL-4 production by naïve CD4+ T cells through activation of ERK weakly and transiently. Intense and sustained ERK activation in response to strong TCR signals inhibited the expression of early IL-4 from CD4+ T cells and led to TH1 differentiation [100,150,152].

In addition to the direct reports of ERK regulation in T lymphocytes, the MAP3K, tumour progression locus-2 (TPL-2), has also been linked to T cell differentiation. TPL-2, initially identified as an oncoprotein kinase, is an important regulator of ERK activation in macrophages [153-157]. Deficiency of tpl2 led to impaired IFN-γ production by TH1 cells, with compromised control of T. gondii [158]. The impaired TH1 responses in T. gondii infected Tpl2-/- mice was suggested to be T cell intrinsic. Furthermore, Tpl2-/- mice develop an enhanced inflammatory TH2 phenotype upon OVA challenge [159]. This exaggerated TH2 response was linked to the intrinsic inability of the T lymphocytes to differentiate to IFN-γ producing TH1 cells in vitro, as a consequence of poor induction of T-bet and STAT-4 [158,159]. This functional role of TPL-2 in influencing TH-polarization was further attributed to TPL-2 dependent ERK activation. Contrastingly in a separate study, Tpl2-/- mice develop skewed TH1 responses upon L. major infection and OVA immunization, suggesting TPL-2 functions as a negative regulator of TH1 responses by inhibiting IL-12 production in innate cells [160]. These studies conflicting observations on the role of TPL-2 in T cell differentiation and function may be complicated further as TPL-2 regulates the proteolytic degradation of the NFκB inhibitory protein, p105, which functions to retain the NFκB subunits inactive in the cytosol [161]. Thus, defects in TPL-2 signalling pathway would also encode defects in NFκB signalling.

p38 family in T cell development and differentiation

The p38 kinase is required for the early expansion of immature thymocytes [162]. High levels of p38 kinase activity measured by an in vitro kinase assay, are associated with CD4/CD8 DN thymocytes, which are CD25-/CD44+, CD25+/CD44+ and CD25+/CD44-. However, low levels of p38 kinase activity are associated with the CD25-/CD44- preTCR thymocytes as well as CD4/CD8 DP thymocytes [162]. While p38 MAPK regulates the early stages of thymocyte development, constitutive activation of p38 MAPK in a constitutively active mkk6 transgenic mouse, blocks the differentiation of immature thymocytes to the DP stage, and results in a lack of T lymphocytes in the peripheral immune system [162]. The role of p38 MAPK in positive and negative selection was initially addressed using FTOC assays and a pharmacological inhibitor of p38 MAPK. This initial study demonstrated that p38 was not required for positive selection. Instead, forced expression of a constitutively active p38 MAPK caused the deletion of DP thymocytes [163]. Meanwhile, a contrasting study documented an important regulatory role for the p38 family in positive selection, where inhibition of p38 kinase using the pharmacological inhibitor SB 203580 prevented the differentiation of DP cells into single positive cells in vitro [164]. This observation was further supported by studies using transgenic mice with dominant negative mutations in mkk3 and mkk6 that suggested, complete inhibition of p38 kinase activity lead to impaired positive selection with a decrease in single positive CD4 or CD8 thymocytes [165]. This study provided a definitive role for p38 kinase in positive selection, unlike the previous observations with incomplete inhibition of p38 kinase activity- either due to an ineffective dominant negative mutation or due to compensatory mechanisms from alternative signalling pathways [162,166].

With regard to T cell differentiation, inhibition of p38 MAPK in CD4+ T cells from a transgenic mouse expressing a dominant negative p38, resulted in decreased IFN-γ production [167]. The regulatory role of p38 in IFN-γ production was further supported by the observation that CD4+ T cells from a mkk6 transgenic mouse with a constitutively active p38 kinase also produced increased levels of IFN-γ [167]. These observations, supported by other independent studies, suggest an important role for the p38 MAPK pathway in regulating TH1 responses both in vitro and in vivo [168-170]. Additionally, the role of p38 kinase, specifically its α isoform, p38α in TH1 responses was demonstrated to be cytokine dependent with p38α-/- CD4+ TH1 cells defective in IFNγ secretion upon IL-12 and IL-18 stimulation compared to TCR induced IFNγ production [171]. This suggested that p38 kinase played an important role in maintaining TH1 cytokine production. Contrasting observations were made with respect to TH2 cells. While, the CD4+ T cells from the dominant negative p38 transgenic model showed no change in IL-4 production, CD4+ T cells from the mkk6 transgenic model demonstrated decreased IL-4 production [167]. In an independent study, inhibition of p38 MAPK kinase activity using the inhibitor SB203850, demonstrated a role for p38 MAPK in GATA-3 phosphorylation and positive regulation of TH2 cytokines [172]. Exactly how this differential regulation by p38 MAPK occurs in TH1 and TH2 cells is not completely understood. Common upstream activators like Rac2 and GADD45γ (the stress inducible protein), whose expression is specific to, or higher in, TH1 cells than TH2 cells, have been suggested to potentially activate one pathway over the other [170]. In this regard, expression of GADD45γ or GADD45β in the stress activated MAP3K, MEKK4 sufficient cells, mediates p38 activation and up regulates IFN-γ production in CD4+ T cells [173].

In addition to the canonical pathway of p38 activation by MAP2Ks and MAP3Ks, there exists a non-canonical pathway of p38 activation, which is mediated by phosphorylation of Y323 of the p38 MAPK, catalysed by the TCR signalling adaptor Zap70 [174]. This phosphorylation is followed by the auto phosphorylation at the activation loop T180/Y182 that brings about the activation of the p38 MAPK [174]. The non-canonical pathway has important implications in T cell differentiation as a mutant form of p38α in T cells, which is incapable of undergoing non-canonical activation due to a Y323F mutation, has been linked to reduced IFN-γ production [175]. All these studies indicate the p38 MAPK family has important regulatory functions in early thymocyte maturation and differentiation of CD4+ T cells into IFN-γ producing TH1 cells.

JNK family in T cell development and differentiation

Though there is no strong support for the role of JNKs in the development of CD4+ or CD8+ T cells, several studies have implicated the JNK pathway in the deletion of CD4/CD8 DP thymocytes [176-178]. These studies demonstrated that the JNK pathway is activated in response to signals initiating negative selection in response to MKK7 activation. Experiments with a dominant negative jnk transgenic mouse demonstrated that inhibition of the JNK pathway resulted in the deletion of CD4/CD8 DP thymocytes [176]. Of the different isoforms of JNK, JNK2 was demonstrated to have a dominant effect on TCR induced thymocyte apoptosis, which was in turn related to its ability to modulate c-Jun and NFAT [179].

JNKs are very weakly expressed in peripheral lymphoid tissues and their activity is also very low [180]. This low level of activity is also supported by the low level of expression of their kinases MKK4 and MKK7 [170]. However, their activity is significantly upregulated upon TCR activation and peaks 36 - 60h post activation. Despite comparable expression of the JNK proteins in the effector phase of TH1 and TH2 cells, JNK kinase activity is dominant in TH1 cells rather than TH2 cells [181].

Several groups analysed the role of JNK1 and JNK2 in T cell activation and cytokine production [136,137,177,178,182]. JNK2 deficient ‘TH1’ cells produced significantly lower levels of IFN-γ and CD4+ T cells from these mice failed to differentiate into TH1 cells [182]. Reduced IFN-γ production in Jnk2-/- mice was due to reduced expression of the IL-12 receptor β chain (IL-12Rβ2) [182]. JNK1, unlike JNK2, seemed to have a more pronounced role in freshly activated T cells despite comparable protein expression of the two isoforms in these cells [136]. This suggested that total JNK kinase activity in T cells might not always be a result of the cumulative effects of the functional JNK1 and JNK2 isoforms. Indeed, in support of this observation kinase activity by JNK2 is more pronounced in TH1 effector cells [182]. The role of JNK1 and JNK2 in IL-2 expression was resolved by Dong and colleagues, who generated JNK1/JNK2 double knockout mice and analysed the effect of JNK1/JNK2 absence TCR activation and IL-2 expression [137]. This study revealed JNK1 and JNK2 were not required for the activation of T cells and IL-2 expression but were important for TH cell, particularly TH1 differentiation and its corresponding effector cytokine production, beyond IL-2.

The role of JNK1 in TH cell differentiation comes from studies using Jnk1-/- mice, which exhibit exaggerated TH2 responses in addition to impaired TH1 responses [136]. This was true even for CD4+ T cells cultured under TH1 conditions. The enhanced TH2 response observed in Jnk1-/- mice was consistent with their inability to heal skin lesions and ulcers upon infection with L. major [183]. Dong and colleagues examined the negative regulatory effect by JNK1 on TH2 differentiation and showed Jnk1-/- mice had elevated levels of NFATc, whose nuclear retention in wild type cells is regulated by JNK mediated phosphorylation [136]. JNK phosphorylation of NFATc prevents it from being dephosphorylated by calcineurin phosphatase that is responsible for the nuclear retention of NFATc [135]. In the absence of JNK signalling, NFATc susceptible to dephosphorylation by calcineurin phosphatase, accumulates in the nucleus and mediates exaggerated transcriptional activity by increasing the expression of TH2 cytokines.

Additionally, it has been demonstrated that JNK1 mediated suppression of TH2 effector cytokine production was dependent on the proteolytic turnover of the AP-1 family member, JunB [184-186]. JunB is post-translationally regulated by ubiquitination in CD4+ T cells [187]. This is mediated by the E3 ubiquitin ligase, Itch, which is activated following TCR co-stimulation by JNK1 phosphorylation [187,188]. These observations were corroborated by Enzler and colleagues who demonstrated that the stress induced MAP3K, MEKK1 upon TCR activation recruits Itch to its signalling complex and activates JNK1, thus regulating JunB mediated TH2 cytokine expression [189]. Itch induced ubiquitination of JunB, mediated by MEKK1-JNK1 signalling has also been shown to be involved in peripheral TH2 tolerance [190].

The stress activated MAP3Ks and T lymphocytes

The stress activated MAP3Ks consist of a family of five members, MEKK1-4 and the NF-κB-inducing kinase (NIK). This group of proteins can activate multiple members of the MAPK family. Of these family members, MEKK2 and MEKK3 are promiscuous activators as they are non-selective in their target substrates and equally activate ERK, JNK as well as the p38 MAPK family [49]. Their function in T cell activation and homeostasis has been demonstrated using biochemical and genetic studies, where they share the downstream activation of MAPK [191,192]. MEKK3 has also been recently implicated in T cell responsiveness wherein conditional knockout of mekk3 in T cells led to impaired IFN-γ expression in T cells [193]. This impaired TH1 phenotype correlates with decreased clearance of L. monocytogenes– OVA infection and is associated with reduced TCR mediated activation of ERK, JNK and p38 MAPKs.

Both MEKK2 and MEKK3 are co-expressed in T cells suggesting they might have overlapping functions. Chang and colleagues generated MEKK2/MEKK3 DKO mice to investigate the role of MEKK2 and MEKK3 in TH cell differentiation [194]. These mice exhibited increased number of TREG as well as TH17 cells. This was due to a cell intrinsic phenomena as mixed bone marrow chimeras with equal ratios of wild type or MEKK2/MEKK3 DKO T cells in Rag2-/- mice maintained the TREG and TH17 accumulation in the periphery, identical to the parent MEKK2/MEKK3 DKO mice. The increased numbers of TREG and TH17 cells was suspected to be due to the enhanced sensitivity of the naïve T cells from the MEKK2/MEKK3 DKO mice to TGF-β (required for TREG and TH17 differentiation) as pharmacological inhibition of TGF-β receptor signalling in vivo reduced this increase in TREG and TH17 cells. This study also linked the enhanced sensitivity to TGF-β to defective activation of ERK1 and ERK2, which in turn brought about the hypophosphorylation of the SMAD2/3 proteins in their linker regions, which is important for suppression of TGF-β signalling [195-198]. Thus this study provided evidence to support the mechanistic role of MEKK2 and MEKK3 mediated inhibition of TGF-β signalling and regulating TH cell differentiation.

Therapeutically Targeting Kinases in Allergy

Atopy and the manifestation of allergic asthma presents as a heterogeneous clinical disease, mediated in part by CD4+ TH1 and TH17 cells, in addition to a dominant role played by TH2 cells [199]. Briefly, the development of IL-4-secreting allergen-specific TH2 cells, which promote the class-switching and clonal expansion of allergen- reactive B cells, leads to allergen-reactive IgE and an atopic or sensitized state. Following subsequent exposure to the same allergen, cross-linking of mast cell- and basophil-bound IgE molecules, trigger acute early phase reactions with histamine, prostaglandin and leukotriene release. Reactivation of allergen-specific T cells typify the late phase response, reenforcing antibody class switching but also mobilizing, maturing and activating eosinophils and promoting goblet cell hyperplasia and mucus hyper secretion, through the actions of IL-5 and IL-13. In chronic cases, resolution of repeated inflammatory events within the tissue can lead to excessive wound healing and the development of fibrotic plaques and smooth muscle thickening. Many cellular events contribute to this inflammatory disease, and therefore targeting of any particular kinase, or kinase family is fraught with complexities. Nevertheless, several kinase inhibitors have been proposed, tested or moved into pre-clinical and clinical trials to treat inflammatory diseases, including asthma. Below we summarize some of the many inhibitors being developed and refer to more focused reviews, where possible, for further reading.

Tec kinase inhibitors

PI3kinase inhibitors: PI3kδ, one member of the PI3K family, is predominantly expressed in hematopoietic cells including mast cells, B cells, T cells and neutrophils. Although not a Tec kinase, PI3kδ converts the membrane phospholipid PtdIns(4,5)P2 into PtdIns(3,4,5) P3, allowing the recruitment of Tec family kinases. Concordant with its expression, PI3kδ-dead knock-in mice have significant defects in B and T cell signalling, high affinity IgE receptor, FcεR1, signalling and neutrophil activation [200-202]. IC87114, an inhibitor of PI3Kδ (and to a lesser extent PI3Kγ and β), significantly inhibits the TH2-associated inflammatory cascade with diminished allergen-induced airway inflammation and hyper-responsiveness (AHR) in a mouse model [203]. Additionally, enhanced TREG cell development has been reported in vitro upon PI3K interference, due to premature termination of TCR activation. Despite this observation, this effect does not fully translate in-vivo, as PI3Kδ-knock-in mice develop inflammatory bowels disease (IBD), a condition primarily due to a lack of intestinal TREG cells [200]. Hence, despite an array of PI3K inhibitors in development and phase 2 trials, few seem to have been applied to the allergy arena [204].

Btk inhibitors: Btk is crucial for B cell and IgE receptor (FcεR1) signalling but also appears to play important roles in myeloid cells [205-207]. Btk is neither expressed in T cells, nor capable of activating downstream pathways when over expressed in T cell lines [87]. Nevertheless, inhibiting Btk could disrupt the B cell / IgE / mast cell axis, which is responsible for allergen sensitisation. Indeed, inhibiting Btk with the selective irreversible inhibitor, PCI-32765, significantly blocked IgE-mediated basophil activation, cytokine secretion and degranulation [208]. Many Btk inhibitors (PCI-32765, terreic acid and LFM-A13, GDC-0834, RN486) have been tested in a variety of in vivo settings, with minimal toxicity and significant reduction in arthritis and leukaemia [209-213]. However, despite suitable B cell /IgE / mast cell-dependent models of allergic inflammation and an array of Btk inhibitors, there are no studies reporting the impact of Btk inhibitors in allergy models. Currently 27 Phase 1 and 2 clinical studies are investigating the use of Btk inhibitors in a variety of diseases.

Itk inhibitors: Itk, as mentioned above, is significantly expressed in T cells and regulates TH2 cell development [27,83]. Itk also regulates mast cell cytokine secretion, but not degranulation [214,215]. In animal models, Itk-deficient mice and transgenic mice lacking Itk activity are protected from allergen-induced airway inflammation, supporting the case for therapeutic Itk inhibition [95,216-218]. A significant amount of work has been done optimizing inhibitors of Itk, however very few have made it through to pre-clinical trials, or at least been reported [106,107,219-222]. An Itk inhibitor, BMS-509744, was shown to inhibit airway eosinophilia in a murine model of ovalbumininduced inflammation, however many of the other airway allergyassociated parameters were not reported in this study [223]. Many other pharmaceutical companies, including Boehringer Ingelheim, AstraZeneca and Sanofi-Aventis have products in the pipeline, but have not published their findings yet.

MAP Kinase inhibitors

The three-tier MAP kinases that phosphorylate serine/threonine MAP3Ks, which in turn phosphorylate MAP2Ks, that phosphorylate MAPKs, have been widely studied and interrogated at every level with small molecule inhibitors [224].

MEK/ERK inhibitors: Of the several MEK/ERK inhibitors, PD098059 and U0126 have received significant attention inhibiting MEK1 and MEK2 with a high degree of specificity [225,226]. Using in vitro guinea pig bronchial rings to test the contraction of airways as a model of airway contraction, PD098059 was shown to inhibit peptide-leukotriene’s from mast cells, allow a quicker recovery time post contraction and block IL-1β-induced prostaglandin D2 release from primary and immortalised airway epithelial cells [227-230]. In vivo, U0126 inhibited allergen-induced airway inflammation, although it also increased steroid resistance [231-233]. To date and to our knowledge, neither PD098059 nor U0126 have moved into clinical studies, but have remained invaluable tools for researchers to dissect molecular pathways. Eleven other MEK/ERK inhibitors, including PD184352, have proved promising in vitro and in vivo, and have been moved into phase 1 and 2 in various clinical trails in cancer patients [234-237]. Although cancer treatment has driven the development of PD184352 and other first and second-generation MEK/ERK inhibitors, their application to other inflammatory diseases, including allergy, will be of significant interest.

p38 inhibitors: p38 regulates many inflammatory pathways involved in respiratory diseases (reviewed by [238-240]). It has been extensively pursued, and unlike many MEK/ERK inhibitors to date, several p38 inhibitors have been tested in allergy models and allergic patients [241,242]. The p38 inhibitors, SD282, SB239063 and a respirable p38α antisense oligonucleotide have all been shown to inhibit experimental allergen-induced airway eosinophilia, IgE and airway hyper-responsiveness in mice [243-246]. TH2 cell-derived IL-5 mobilises eosinophils from the bone marrow, matures and activates them in the tissue [247]. In this context, the p38 inhibitor SB203580 inhibits IL-5, reduces IL-13 synthesis from human T cells and induces eosinophil apoptosis, thus suggesting that the IL-5/eosinophil axis is sensitive to p38 inhibition [231,248,249]. Additionally, it has been observed that many asthmatic patients are resistant, or insensitive, to inhaled corticosteroids (ICS) [250-252]. It appears that this insensitivity is mediated, in part, by p38 signalling, and that inhibition of p38 can restore steroid sensitivity [253-256]. Thus, p38 inhibitors hold great promise in interfering with allergen-induced inflammatory axis (IL- 5/eosinophilia), and also in hard-to-treat allergic patients. There are currently 3 experimental clinical studies investigating the role of p38 in steroid sensitivity in severe asthmatics, with one phase 2 clinical trial testing the safety of the p38 inhibitor SB-681323.

JNK inhibitors: Several JNK inhibitors have been tested in a variety of in vitro and in vivo disease models (reviewed in [257]). In particular XG-102 and SP600125 have been successfully tested in pre-clinical in vivo models of IBD and have been shown to reduce TNFα expression and disease severity [258-260]. With respect to asthma, SP600125 has been tested in ozone and allergen-induced airway inflammation models in mice and rats, with both studies reporting positive outcomes [261-263]. In mice, 30 mg/kg of SP600125 reduced airway eosinophilia, goblet cell and mucus secretion and airway hyper responsiveness, following ovalbumin airway challenge [262]. However in rats, 90 mg/ kg of SP600125 reduced cellular infiltration into the air spaces, but did not impact inflammatory cytokine secretion, IgE or airway hyper responsiveness [263]. To date, several JNK inhibitors (CC-401, CNI- 1493, AM-111, XG-102, CC359, CC930 and CEP-1347) have moved into phase 1 and phase 2 clinical trials, but are yet to be tested in allergic individuals [257].

Concluding Comments

In summary, TEC kinase-regulated pathways, which differentially control TH1, TH2, and TH17 responses through regulation of PLC-γ activation, Ca2+ influx and transcription factor translocation are attractive therapeutic targets to disable T cell-mediated responses. Specifically in the allergy field, the TEC kinase Itk that is required for Th2 cells, may forestall atopy and Th2-reactivation. Similarly, the MAPKs, which integrate into proximal TCR signalling cascades, required for downstream transcription factor activation, including the phosphorylation of STAT6 and stabilisation of GATA3, are an attractive family of kinases to therapeutically target.

However, despite significant advances in our understanding of the signalling events in T cells, the complexities and abundance of kinase-regulated signalling pathways and a better grasp of the pathogenesis of allergic asthma, broad spectrum non-specific inhaled corticosteroids are still the mainstay of current treatment. This is due to many factors, including; the complexities of CD4 T cell development and differentiation, the apparent differential kinase requirements in different T cells and the emerging appreciation of T cell plasticity. For example, recent studies have identified that fully differentiated T cells maintain a degree of plasticity with the ability to change between different TH phenotypes and even between helper and regulatory cells [264]. If the kinase-regulated signalling pathways associated with the different TH phenotypes are also plastic, allowing for dynamic re-organisation, then this would pose a significant challenge for therapeutically targeting any specific TH population with inhibitors. Furthermore, allergic asthma is now viewed as a heterogeneous disease consisting of TH1-, TH2- and TH17-associated responses, with differential kinase requirements in different TH cells, as described above [228,245]. This poses an additional obstacle as it considerably expands both the number and expression of the potential targets. For example, targeting Itk, which is the most abundantly expressed TEC kinase and required for TH2 and TH17, but not TH1, cell development, may skew allergen-reactivity to TH1 cells that may be equally as damaging. The potential off-target effects of kinase inhibitors, whether in non-targeted cells or closely related pathways in the appropriate cell, pose another great difficulty which is yet to be overcome [265]. Thus, despite our current understanding of the substrates and phosphorylated residues involved in these signalling pathways, there are many unknown gaps in the chain, which need to be investigated and identified. Nevertheless, the requirement for kinase signalling pathways in various TH cells and at various stages is still an appealing therapeutic target. Finally, to tackle the complexities of allergic diseases and in a bid to move our efforts forward, a cross-disciplinary approach between chemists, biochemists, immunologists and in-vivo disease biologists may be more fruitful and greater than the sum of our individual parts.

Acknowledgements

YK and MSW would like to thank the MRC (File Reference number MC_UP_ A253_1028) and the Lady TATA foundation for financial support. We would also like to apologise to our many colleagues whose important work in this area we did not mention due to space limitations.

References

  1. Barzaghi F, Passerini L, Bacchetta R (2012) Immune dysregulation, polyendocrinopathy, enteropathy, x-linked syndrome: a paradigm of immunodeficiency with autoimmunity. Front Immunol 3: 211.
  2. Okoye IS, Wilson MS (2011) CD4+ T helper 2 cells--microbial triggers, differentiation requirements and effector functions. Immunology 134: 368-377.
  3. Zhu J, Yamane H, Paul WE (2010) Differentiation of effector CD4 T cell populations (*). Annu Rev Immunol 28: 445-489.
  4. Crotty S (2011) Follicular helper CD4 T cells (TFH). Annu Rev Immunol 29: 621-663.
  5. Josefowicz SZ, Lu LF, Rudensky AY (2012) Regulatory T cells: mechanisms of differentiation and function. Annu Rev Immunol 30: 531-564.
  6. Zygmunt B, Veldhoen M (2011) T helper cell differentiation more than just cytokines. Adv Immunol 109: 159-196.
  7. Chang HC, Sehra S, Goswami R, Yao W, Yu Q, et al. (2010) The transcription factor PU.1 is required for the development of IL-9-producing T cells and allergic inflammation. Nat Immunol 11: 527-534.
  8. Yu D, Rao S, Tsai LM, Lee SK, He Y, et al. (2009) The transcriptional repressor Bcl-6 directs T follicular helper cell lineage commitment. Immunity 31: 457-468.
  9. Liu X, Yan X, Zhong B, Nurieva RI, Wang A, et al. (2012) Bcl6 expression specifies the T follicular helper cell program in vivo. J Exp Med 209: 1841-1852, S1-24.
  10. Maruyama T, Konkel JE, Zamarron BF, Chen W (2011) The molecular mechanisms of Foxp3 gene regulation. Semin Immunol 23: 418-423.
  11. Fontenot JD, Gavin MA, Rudensky AY (2003) Foxp3 programs the development and function of CD4+CD25+ regulatory T cells. Nat Immunol 4: 330-336.
  12. Khattri R, Cox T, Yasayko SA, Ramsdell F (2003) An essential role for Scurfin in CD4+CD25+ T regulatory cells. Nat Immunol 4: 337-342.
  13. Hori S, Nomura T, Sakaguchi S (2003) Control of regulatory T cell development by the transcription factor Foxp3. Science 299: 1057-1061.
  14. Zheng Y, Rudensky AY (2007) Foxp3 in control of the regulatory T cell lineage. Nat Immunol 8: 457-462.
  15. Murphy KM, Stockinger B (2011) Effector T cell plasticity: flexibility in the face of changing circumstances. Nat Immunol 11: 674-680.
  16. Hegazy AN, Peine M, Helmstetter C, Panse I, Fröhlich A, et al. (2010) Interferons direct Th2 cell reprogramming to generate a stable GATA-3(+)T-bet(+) cell subset with combined Th2 and Th1 cell functions. Immunity 32: 116-128.
  17. Wang YH, Voo KS, Liu B, Chen CY, Uygungil B, et al. (2010) A novel subset of CD4(+) T(H)2 memory/effector cells that produce inflammatory IL-17 cytokine and promote the exacerbation of chronic allergic asthma. J Exp Med 207: 2479-2491.
  18. Veldhoen M, Uyttenhove C, van Snick J, Helmby H, Westendorf A, et al. (2008) Transforming growth factor-beta 'reprograms' the differentiation of T helper 2 cells and promotes an interleukin 9-producing subset. Nat Immunol 9: 1341-1346.
  19. Starr TK, Jameson SC, Hogquist KA (2003) Positive and negative selection of T cells. Annu Rev Immunol 21: 139-176.
  20. Irving BA, Weiss A (1991) The cytoplasmic domain of the T cell receptor zeta chain is sufficient to couple to receptor-associated signal transduction pathways. Cell 64: 891-901.
  21. Iwashima M, Irving BA, van Oers NS, Chan AC, Weiss A (1994) Sequential interactions of the TCR with two distinct cytoplasmic tyrosine kinases. Science 263: 1136-1139.
  22. Manolios N, Kemp O, Li ZG (1994) The T cell antigen receptor alpha and beta chains interact via distinct regions with CD3 chains. Eur J Immunol 24: 84-92.
  23. Ishikura S, Weissman AM, Bonifacino JS (2010) Serine residues in the cytosolic tail of the T-cell antigen receptor alpha-chain mediate ubiquitination and endoplasmic reticulum-associated degradation of the unassembled protein. J Biol Chem 285: 23916-23924.
  24. Molina TJ, Kishihara K, Siderovski DP, van Ewijk W, Narendran A, et al. (1992) Profound block in thymocyte development in mice lacking p56lck. Nature 357: 161-164.
  25. Wallace VA, Kawai K, Levelt CN, Kishihara K, Molina T, et al. (1995) T lymphocyte development in p56lck deficient mice: allelic exclusion of the TcR beta locus is incomplete but thymocyte development is not restored by TcR beta or TcR alpha beta transgenes. Eur J Immunol 25: 1312-1318.
  26. van Oers NS, Lowin-Kropf B, Finlay D, Connolly K, Weiss A (1996) alpha beta T cell development is abolished in mice lacking both Lck and Fyn protein tyrosine kinases. Immunity 5: 429-436.
  27. Liao XC, Littman DR (1995) Altered T cell receptor signaling and disrupted T cell development in mice lacking Itk. Immunity 3: 757-769.
  28. Schaeffer EM, Broussard C, Debnath J, Anderson S, McVicar DW, et al. (2000) Tec family kinases modulate thresholds for thymocyte development and selection. J Exp Med 192: 987-1000.
  29. Roifman CM, Dadi H, Somech R, Nahum A, Sharfe N (2010) Characterization of ζ-associated protein, 70 kd (ZAP70)-deficient human lymphocytes. J Allergy Clin Immunol 126: 1226-1233.
  30. Hauck F, Randriamampita C, Martin E, Gerart S, Lambert N, et al. (2012) Primary T-cell immunodeficiency with immunodysregulation caused by autosomal recessive LCK deficiency. J Allergy Clin Immunol 130: 1144-1152.
  31. König S, Probst-Kepper M, Reinl T, Jeron A, Huehn J, et al. (2012) First insight into the kinome of human regulatory T cells. PLoS One 7: e40896.
  32. Mack KD, Von Goetz M, Lin M, Venegas M, Barnhart J, et al. (2005) Functional identification of kinases essential for T-cell activation through a genetic suppression screen. Immunol Lett 96: 129-145.
  33. Chi H (2012) Regulation and function of mTOR signalling in T cell fate decisions. Nat Rev Immunol 12: 325-338.
  34. Xu X, Guo J, Vorster P, Wu Y (2012) Involvement of LIM kinase 1 in actin polarization in human CD4 T cells. Commun Integr Biol 5: 381-383.
  35. Sestero CM, McGuire DJ, De Sarno P, Brantley EC, Soldevila G, et al. (2012) CD5-dependent CK2 activation pathway regulates threshold for T cell anergy. J Immunol 189: 2918-2930.
  36. Dorn A, Zoellner A, Follo M, Martin S, Weber F, et al. (2012) Rap1a deficiency modifies cytokine responses and MAPK-signaling in vitro and impairs the in vivo inflammatory response. Cell Immunol 276: 187-195.
  37. Hanks SK, Quinn AM, Hunter T (1988) The protein kinase family: conserved features and deduced phylogeny of the catalytic domains. Science 241: 42-52.
  38. Racioppi L, Means AR (2008) Calcium/calmodulin-dependent kinase IV in immune and inflammatory responses: novel routes for an ancient traveller. Trends Immunol 29: 600-607.
  39. Baier G (2007) PKC isotype functions in T lymphocytes. Ernst Schering Found Symp Proc 7: 29-41.
  40. Katoh M, Katoh M (2007) WNT signaling pathway and stem cell signaling network. Clin Cancer Res 13: 4042-4045.
  41. Pearce LR, Komander D, Alessi DR (2010) The nuts and bolts of AGC protein kinases. Nat Rev Mol Cell Biol 11: 9-22.
  42. Fu G, Gascoigne NR (2012) The role of protein kinase ceta in T cell biology. Front Immunol 3: 177.
  43. Wayman GA, Tokumitsu H, Davare MA, Soderling TR (2011) Analysis of CaM-kinase signaling in cells. Cell Calcium 50: 1-8.
  44. Hughes K, Edin S, Antonsson A, Grundström T (2001) Calmodulin-dependent kinase II mediates T cell receptor/CD3- and phorbol ester-induced activation of IkappaB kinase. J Biol Chem 276: 36008-36013.
  45. Cheong JK, Virshup DM (2011) Casein kinase 1: Complexity in the family. Int J Biochem Cell Biol 43: 465-469.
  46. Kannan N, Neuwald AF (2004) Evolutionary constraints associated with functional specificity of the CMGC protein kinases MAPK, CDK, GSK, SRPK, DYRK, and CK2alpha. Protein Sci 13: 2059-2077.
  47. Hochegger H, Takeda S, Hunt T (2008) Cyclin-dependent kinases and cell-cycle transitions: does one fit all? Nat Rev Mol Cell Biol 9: 910-916.
  48. Forde JE, Dale TC (2007) Glycogen synthase kinase 3: a key regulator of cellular fate. Cell Mol Life Sci 64: 1930-1944.
  49. Kyriakis JM, Avruch J (2012) Mammalian MAPK signal transduction pathways activated by stress and inflammation: a 10-year update. Physiol Rev 92: 689-737.
  50. Pugh EN Jr, Duda T, Sitaramayya A, Sharma RK (1997) Photoreceptor guanylate cyclases: a review. Biosci Rep 17: 429-473.
  51. Lucas KA, Pitari GM, Kazerounian S, Ruiz-Stewart I, Park J, et al. (2000) Guanylyl cyclases and signaling by cyclic GMP. Pharmacol Rev 52: 375-414.
  52. Dan I, Watanabe NM, Kusumi A (2001) The Ste20 group kinases as regulators of MAP kinase cascades. Trends Cell Biol 11: 220-230.
  53. Shiu SH, Li WH (2004) Origins, lineage-specific expansions, and multiple losses of tyrosine kinases in eukaryotes. Mol Biol Evol 21: 828-840.
  54. Pao LI, Badour K, Siminovitch KA, Neel BG (2007) Nonreceptor protein-tyrosine phosphatases in immune cell signaling. Annu Rev Immunol 25: 473-523.
  55. Rusconi P, Caiola E, Broggini M (2012) RAS/RAF/MEK inhibitors in oncology. Curr Med Chem 19: 1164-1176.
  56. Flannery S, Bowie AG (2010) The interleukin-1 receptor-associated kinases: critical regulators of innate immune signalling. Biochem Pharmacol 80: 1981-1991.
  57. Moquin D, Chan FK (2010) The molecular regulation of programmed necrotic cell injury. Trends Biochem Sci 35: 434-441.
  58. Middelbeek J, Clark K, Venselaar H, Huynen MA, van Leeuwen FN (2010) The alpha-kinase family: an exceptional branch on the protein kinase tree. Cell Mol Life Sci 67: 875-890.
  59. Bakkenist CJ, Kastan MB (2004) Initiating cellular stress responses. Cell 118: 9-17.
  60. Mayers RM, Leighton B, Kilgour E (2005) PDH kinase inhibitors: a novel therapy for Type II diabetes? Biochem Soc Trans 33: 367-370.
  61. Roche TE, Hiromasa Y (2007) Pyruvate dehydrogenase kinase regulatory mechanisms and inhibition in treating diabetes, heart ischemia, and cancer. Cell Mol Life Sci 64: 830-849.
  62. LaRonde-LeBlanc N, Wlodawer A (2005) A family portrait of the RIO kinases. J Biol Chem 280: 37297-37300.
  63. Campbell BE, Boag PR, Hofmann A, Cantacessi C, Wang CK, et al. (2011) Atypical (RIO) protein kinases from Haemonchus contortus--promise as new targets for nematocidal drugs. Biotechnol Adv 29: 338-350.
  64. Thomas JD, Sideras P, Smith CI, Vorechovský I, Chapman V, et al. (1993) Colocalization of X-linked agammaglobulinemia and X-linked immunodeficiency genes. Science 261: 355-358.
  65. Rawlings DJ, Saffran DC, Tsukada S, Largaespada DA, Grimaldi JC, et al. (1993) Mutation of unique region of Bruton's tyrosine kinase in immunodeficient XID mice. Science 261: 358-361.
  66. Tsukada S, Saffran DC, Rawlings DJ, Parolini O, Allen RC, et al. (1993) Deficient expression of a B cell cytoplasmic tyrosine kinase in human X-linked agammaglobulinemia. Cell 72: 279-290.
  67. Vetrie D, Vorechovský I, Sideras P, Holland J, Davies A, et al. (1993) The gene involved in X-linked agammaglobulinaemia is a member of the src family of protein-tyrosine kinases. Nature 361: 226-233.
  68. Schwartzberg PL, Finkelstein LD, Readinger JA (2005) TEC-family kinases: regulators of T-helper-cell differentiation. Nature reviews. Immunology 5: 284-295.
  69. Debnath J, Chamorro M, Czar MJ, Schaeffer EM, Lenardo MJ, et al. (1999) rlk/TXK encodes two forms of a novel cysteine string tyrosine kinase activated by Src family kinases. Mol Cell Biol 19: 1498-1507.
  70. Chamorro M, Czar MJ, Debnath J, Cheng G, Lenardo MJ, et al. (2001) Requirements for activation and RAFT localization of the T-lymphocyte kinase Rlk/Txk. BMC Immunol 2: 3.
  71. Scharenberg AM, Kinet JP (1998) PtdIns-3,4,5-P3: a regulatory nexus between tyrosine kinases and sustained calcium signals. Cell 94: 5-8.
  72. Tomlinson MG, Heath VL, Turck CW, Watson SP, Weiss A (2004) SHIP family inositol phosphatases interact with and negatively regulate the Tec tyrosine kinase. J Biol Chem 279: 55089-55096.
  73. Shan X, Czar MJ, Bunnell SC, Liu P, Liu Y, et al. (2000) Deficiency of PTEN in Jurkat T cells causes constitutive localization of Itk to the plasma membrane and hyperresponsiveness to CD3 stimulation. Mol Cell Biol 20: 6945-6957.
  74. Berg LJ, Finkelstein LD, Lucas JA, Schwartzberg PL (2005) Tec family kinases in T lymphocyte development and function. Annu Rev Immunol 23: 549-600.
  75. August A, Gibson S, Kawakami Y, Kawakami T, Mills GB, et al. (1994) CD28 is associated with and induces the immediate tyrosine phosphorylation and activation of the Tec family kinase ITK/EMT in the human Jurkat leukemic T-cell line. Proc Natl Acad Sci U S A 91: 9347-9351.
  76. Yang WC, Ghiotto M, Barbarat B, Olive D (1999) The role of Tec protein-tyrosine kinase in T cell signaling. J Biol Chem 274: 607-617.
  77. Yang, WC, Olive D (1999) Tec kinase is involved in transcriptional regulation of IL-2 and IL-4 in the CD28 pathway. Eur J Immunol 29: 1842-1849.
  78. King PD, Sadra A, Teng JM, Bell GM, Dupont B (1998) CD2-mediated activation of the Tec-family tyrosine kinase ITK is controlled by proline-rich stretch-4 of the CD2 cytoplasmic tail. Int Immunol 10: 1009-1016.
  79. Tanaka N, Abe H, Yagita H, Okumura K, Nakamura M, et al. (1997) Itk, a T cell-specific tyrosine kinase, is required for CD2-mediated interleukin-2 promoter activation in the human T cell line Jurkat. Eur J Immunol 27: 834-841.
  80. Fischer AM, Mercer JC, Iyer A, Ragin MJ, August A (2004) Regulation of CXC chemokine receptor 4-mediated migration by the Tec family tyrosine kinase ITK. J Biol Chem 279: 29816-29820.
  81. Takesono A, Horai R, Mandai M, Dombroski D, Schwartzberg PL (2004) Requirement for Tec kinases in chemokine-induced migration and activation of Cdc42 and Rac. Curr Biol 14: 917-922.
  82. Fowell DJ, Shinkai K, Liao XC, Beebe AM, Coffman RL, et al. (1999) Impaired NFATc translocation and failure of Th2 development in Itk-deficient CD4+ T cells. Immunity 11: 399-409.
  83. Schaeffer EM, Yap GS, Lewis CM, Czar MJ, McVicar DW, et al. (2001) Mutation of Tec family kinases alters T helper cell differentiation. Nat Immunol 2: 1183-1188.
  84. Colgan J, Asmal M, Neagu M, Yu B, Schneidkraut J, et al. (2004) Cyclophilin A regulates TCR signal strength in CD4+ T cells via a proline-directed conformational switch in Itk. Immunity 21: 189-201.
  85. Hu Q, Davidson D, Schwartzberg PL, Macchiarini F, Lenardo MJ, et al. (1995) Identification of Rlk, a novel protein tyrosine kinase with predominant expression in the T cell lineage. J Biol Chem 270: 1928-1934.
  86. Kashiwakura J, Suzuki N, Nagafuchi H, Takeno M, Takeba Y, et al. (1999) Txk, a nonreceptor tyrosine kinase of the Tec family, is expressed in T helper type 1 cells and regulates interferon gamma production in human T lymphocytes. J Exp Med 190: 1147-1154.
  87. Tomlinson MG, Kane LP, Su J, Kadlecek TA, Mollenauer MN, et al. (2004) Expression and function of Tec, Itk, and Btk in lymphocytes: evidence for a unique role for Tec. Mol Cell Biol 24: 2455-2466.
  88. Liu KQ, Bunnell SC, Gurniak CB, Berg LJ (1998) T cell receptor-initiated calcium release is uncoupled from capacitative calcium entry in Itk-deficient T cells. J Exp Med 187: 1721-1727.
  89. Schaeffer EM, Debnath J, Yap G, McVicar D, Liao XC, et al. (1999) Requirement for Tec kinases Rlk and Itk in T cell receptor signaling and immunity. Science 284: 638-641.
  90. Lucas JA, Atherly LO, Berg LJ (2002) The absence of Itk inhibits positive selection without changing lineage commitment. J Immunol 168: 6142-6151.
  91. Mariathasan S, Ho SS, Zakarian A, Ohashi PS (2000) Degree of ERK activation influences both positive and negative thymocyte selection. Eur J Immunol 30: 1060-1068.
  92. McNeil LK, Starr TK, Hogquist KA (2005) A requirement for sustained ERK signaling during thymocyte positive selection in vivo. Proc Natl Acad Sci U S A 102: 13574-13579.
  93. Readinger JA, Mueller KL, Venegas AM, Horai R, Schwartzberg PL (2009) Tec kinases regulate T-lymphocyte development and function: new insights into the roles of Itk and Rlk/Txk. Immunol Rev 228: 93-114.
  94. Qi Q, Kannan AK, August A (2011) Tec family kinases: Itk signaling and the development of NKT alphabeta and gammadelta T cells. FEBS J 278: 1970-1979.
  95. Mueller C, August A (2003) Attenuation of immunological symptoms of allergic asthma in mice lacking the tyrosine kinase ITK. J Immunol 170: 5056-5063.
  96. Miller AT, Wilcox HM, Lai Z, Berg LJ (2004) Signaling through Itk promotes T helper 2 differentiation via negative regulation of T-bet. Immunity 21: 67-80.
  97. Matsumoto Y, Oshida T, Obayashi I, Imai Y, Matsui K, et al. (2002) Identification of highly expressed genes in peripheral blood T cells from patients with atopic dermatitis. International archives of allergy and immunology 129: 327-340.
  98. Gomez-Rodriguez J, ZJ Kraus, Schwartzberg PL (2011) Tec family kinases Itk and Rlk / Txk in T lymphocytes: cross-regulation of cytokine production and T-cell fates. FEBS J 278: 1980-1989.
  99. Sahu, N, Venegas AM, Jankovic D, Mitzner W, Gomez-Rodriguez J, et al. (2008) Selective expression rather than specific function of Txk and Itk regulate Th1 and Th2 responses. J Immunol 181: 6125-6131.
  100. Leitenberg D, Bottomly K (1999) Regulation of naive T cell differentiation by varying the potency of TCR signal transduction. Semin Immunol 11: 283-292.
  101. Ranger AM, Oukka M, Rengarajan J, Glimcher LH (1998) Inhibitory function of two NFAT family members in lymphoid homeostasis and Th2 development. Immunity 9: 627-635.
  102. Xanthoudakis S, Viola JP, Shaw KT, Luo C, Wallace JD, et al. (1996) An enhanced immune response in mice lacking the transcription factor NFAT1. Science 272: 892-895.
  103. Brogdon JL, Leitenberg D, Bottomly K (2002) The potency of TCR signaling differentially regulates NFATc/p activity and early IL-4 transcription in naive CD4+ T cells. J Immunol 168: 3825-3832.
  104. Aguado E, Richelme S, Nuñez-Cruz S, Miazek A, Mura AM, et al. (2002) Induction of T helper type 2 immunity by a point mutation in the LAT adaptor. Science 296: 2036-2040.
  105. Sahu, N, August A (2009) ITK inhibitors in inflammation and immune-mediated disorders. Curr Top Med Chem 9: 690-703.
  106. Snow RJ, Abeywardane A, Campbell S, Lord J, Kashem MA, et al. (2007) Hit-to-lead studies on benzimidazole inhibitors of ITK: discovery of a novel class of kinase inhibitors. Bioorg Med Chem Lett 17: 3660-3665.
  107. von Bonin A, Rausch A, Mengel A, Hitchcock M, Krüger M, et al. (2011) Inhibition of the IL-2-inducible tyrosine kinase (Itk) activity: a new concept for the therapy of inflammatory skin diseases. Exp Dermatol 20: 41-47.
  108. Gomez-Rodriguez J, Sahu N, Handon R, Davidson TS, Anderson SM, et al. (2009) Differential expression of interleukin-17A and -17F is coupled to T cell receptor signaling via inducible T cell kinase. Immunity 31: 587-597.
  109. Schaeffer HJ, Weber MJ (1999) Mitogen-activated protein kinases: specific messages from ubiquitous messengers. Mol Cell Biol 19: 2435-2344.
  110. Han J, Ulevitch RJ (1999) Emerging targets for anti-inflammatory therapy. Nat Cell Biol 1: E39-E40.
  111. Davis RJ (2000) Signal transduction by the JNK group of MAP kinases. Cell 103: 239-252.
  112. Mor A, Philips MR (2006) Compartmentalized Ras/MAPK signaling. Annu Rev Immunol 24: 771-800.
  113. Dalby KN, Morrice N, Caudwell FB, Avruch J, Cohen P (1998) Identification of regulatory phosphorylation sites in mitogen-activated protein kinase (MAPK)-activated protein kinase-1a/p90rsk that are inducible by MAPK. J Biol Chem 273: 1496-1505.
  114. Hauge C, Frödin M (2006) RSK and MSK in MAP kinase signalling. J Cell Sci 119: 3021-3023.
  115. Deak M, Clifton AD, Lucocq LM, Alessi DR (1998) Mitogen- and stress-activated protein kinase-1 (MSK1) is directly activated by MAPK and SAPK2/p38, and may mediate activation of CREB. EMBO J 17: 4426-4441.
  116. Fukunaga R, Hunter T (1997) MNK1, a new MAP kinase-activated protein kinase, isolated by a novel expression screening method for identifying protein kinase substrates. EMBO J 16: 1921-1933.
  117. Waskiewicz AJ, Flynn A, Proud CG, Cooper JA (1997) Mitogen-activated protein kinases activate the serine/threonine kinases Mnk1 and Mnk2. EMBO J 16: 1909-1920.
  118. Gille H, Kortenjann M, Thomae O, Moomaw C, Slaughter C, et al. (1995) ERK phosphorylation potentiates Elk-1-mediated ternary complex formation and transactivation. EMBO J 14: 951-962.
  119. Tan Y, Rouse J, Zhang A, Cariati S, Cohen P, et al. (1996) FGF and stress regulate CREB and ATF-1 via a pathway involving p38 MAP kinase and MAPKAP kinase-2. EMBO J 15: 4629-4642.
  120. Ben-Levy R, Leighton IA, Doza YN, Attwood P, Morrice N, et al. (1995) Identification of novel phosphorylation sites required for activation of MAPKAP kinase-2. EMBO J 14: 5920-5930.
  121. Huot, J, Houle F, Marceau F, Landry J (1997) Oxidative stress-induced actin reorganization mediated by the p38 mitogen-activated protein kinase/heat shock protein 27 pathway in vascular endothelial cells. Circ Res 80 383-392.
  122. McLaughlin MM, Kumar S, McDonnell PC, Van Horn S, Lee JC, et al. (1996) Identification of mitogen-activated protein (MAP) kinase-activated protein kinase-3, a novel substrate of CSBP p38 MAP kinase. J Biol Chem 271: 8488-8492.
  123. Wang XZ, Ron D (1996) Stress-induced phosphorylation and activation of the transcription factor CHOP (GADD153) by p38 MAP Kinase. Science 272: 1347-1349.
  124. Han J, Jiang Y, Li Z, Kravchenko VV, Ulevitch RJ (1997) Activation of the transcription factor MEF2C by the MAP kinase p38 in inflammation. Nature 386: 296-299.
  125. Zhao M, New L, Kravchenko VV, Kato Y, Gram H, et al. (1999) Regulation of the MEF2 family of transcription factors by p38. Mol Cell Biol 19: 21-30.
  126. Janknecht R, Hunter T (1997) Convergence of MAP kinase pathways on the ternary complex factor Sap-1a. EMBO J 16: 1620-1627.
  127. Yang SH, Whitmarsh AJ, Davis RJ, Sharrocks AD (1998) Differential targeting of MAP kinases to the ETS-domain transcription factor Elk-1. EMBO J 17: 1740-1749.
  128. Dérijard B, Hibi M, Wu IH, Barrett T, Su B, et al. (1994) JNK1: a protein kinase stimulated by UV light and Ha-Ras that binds and phosphorylates the c-Jun activation domain. Cell 76: 1025-1037.
  129. Kallunki T, Su B, Tsigelny I, Sluss HK, Dérijard B, et al. (1994) JNK2 contains a specificity-determining region responsible for efficient c-Jun binding and phosphorylation. Genes Dev 8: 2996-3007.
  130. Dai T, Rubie E, Franklin CC, Kraft A, Gillespie DA, et al. (1995) Stress-activated protein kinases bind directly to the delta domain of c-Jun in resting cells: implications for repression of c-Jun function. Oncogene 10: 849-855.
  131. Gupta S, Campbell D, Dérijard B, Davis RJ (1995) Transcription factor ATF2 regulation by the JNK signal transduction pathway. Science 267: 389-393.
  132. Kallunki T, Deng T, Hibi M, Karin M (1996) c-Jun can recruit JNK to phosphorylate dimerization partners via specific docking interactions. Cell 87: 929-939.
  133. Cavigelli M, Dolfi F, Claret FX, Karin M (1995) Induction of c-fos expression through JNK-mediated TCF/Elk-1 phosphorylation. EMBO J 14: 5957-5964.
  134. Chow CW, Rincón M, Cavanagh J, Dickens M, Davis RJ (1997) Nuclear accumulation of NFAT4 opposed by the JNK signal transduction pathway. Science 278: 1638-1641.
  135. Chow CW, Dong C, Flavell RA, Davis RJ (2000) c-Jun NH(2)-terminal kinase inhibits targeting of the protein phosphatase calcineurin to NFATc1. Mol Cell Biol 20: 5227-5234.
  136. Dong C, Yang DD, Wysk M, Whitmarsh AJ, Davis RJ, et al. (1998) Defective T cell differentiation in the absence of Jnk1. Science 282: 2092-2095.
  137. Dong C, Yang DD, Tournier C, Whitmarsh AJ, Xu J, et al. (2000) JNK is required for effector T-cell function but not for T-cell activation. Nature 405: 91-94.
  138. Crompton T, Gilmour KC, Owen MJ (1996) The MAP kinase pathway controls differentiation from double-negative to double-positive thymocyte. Cell 86: 243-251.
  139. Michie AM, Sébastien T, David LW, Juan Zúñiga-Pflücker C (1999) Extracellular signal-regulated kinase (ERK) activation by the pre-T cell receptor in developing thymocytes in vivo. J Exp Med 190: 1647-1656.
  140. Iritani BM, Alberola-Ila J, Forbush KA, Perimutter RM (1999) Distinct signals mediate maturation and allelic exclusion in lymphocyte progenitors. Immunity10: 713-722.
  141. Pagès G, Guérin S, Grall D, Bonino F, Smith A, et al. (1999) Defective thymocyte maturation in p44 MAP kinase (Erk 1) knockout mice. Science 286: 1374-1377.
  142. DelgadoP, Fernández E, Dave V, Kappes D, Alarcón B (2000) CD3delta couples T-cell receptor signalling to ERK activation and thymocyte positive selection. Nature 406: 426-430.
  143. Nel AE, Gupta S, Lee L, Ledbetter JA, Kanner SB (1995) Ligation of the T-cell antigen receptor (TCR) induces association of hSos1, ZAP-70, phospholipase C-gamma 1, and other phosphoproteins with Grb2 and the zeta-chain of the TCR. J Biol Chem 270: 18428-18436.
  144. Yablonski D, Kuhne MR, Kadlecek T, Weiss A (1998) Uncoupling of nonreceptor tyrosine kinases from PLC-gamma1 in an SLP-76-deficient T cell. Science 281: 413-416.
  145. Kane LP, Lin J, Weiss A (2000) Signal transduction by the TCR for antigen. Curr Opin Immunol12: 242-249.
  146. Fields PE, Gajewski TF, Fitch FW (1996) Blocked Ras activation in anergic CD4+ T cells. Science 271: 1276-1278.
  147. Li W, Whaley CD, Mondino A, Mueller DL (1996) Blocked signal transduction to the ERK and JNK protein kinases in anergic CD4+ T cells. Science 271: 1272-1276.
  148. Yamashita M, Kimura M, Kubo M, Shimizu C, Tada T, et al. (1999) T cell antigen receptor-mediated activation of the Ras/mitogen-activated protein kinase pathway controls interleukin 4 receptor function and type-2 helper T cell differentiation. Proc Natl Acad Sci U S A 96: 1024-1029.
  149. Yamashita M, Shinnakasu R, Asou H, Kimura M, Hasegawa A, et al. (2005) Ras-ERK MAPK cascade regulates GATA3 stability and Th2 differentiation through ubiquitin-proteasome pathway. J Biol Chem 280: 29409-29419.
  150. Jorritsma PJ, Brogdon JL, Bottomly K (2003) Role of TCR-induced extracellular signal-regulated kinase activation in the regulation of early IL-4 expression in naive CD4+ T cells. J Immunol 170: 2427-2434.
  151. Tripathi P, Sahoo N, Ullah U, Kallionpää H, Suneja A, et al. (2012) A novel mechanism for ERK-dependent regulation of IL4 transcription during human Th2-cell differentiation. Immunol Cell Biol 90: 676-687.
  152. Yamane H, Zhu J, Paul WE (2005) Independent roles for IL-2 and GATA-3 in stimulating naive CD4+ T cells to generate a Th2-inducing cytokine environment. J Exp Med 202: 793-804.
  153. Miyoshi J, Higashi T, Mukai H, Ohuchi T, Kakunaga T (1991) Structure and transforming potential of the human cot oncogene encoding a putative protein kinase. Mol Cell Biol 11: 4088-4096.
  154. C Patriotis, Markois A, Chernoff J, Tsichlis PN (1994) Tpl-2 acts in concert with Ras and Raf-1 to activate mitogen-activated protein kinase. Proc Natl Acad Sci of the USA 91: 9755-9759.
  155. Patriotis C, Makris A, Bear SE, Tsichlis PN (1993) Tumor progression locus 2 (Tpl-2) encodes a protein kinase involved in the progression of rodent T-cell lymphomas and in T-cell activation. Proc Natl Acad Sci U S A 90: 2251-2255.
  156. Dumitru CD, Ceci JD, Tsatsanis C, Kontoyiannis D, Stamatakis K, et al. (2000) TNF-alpha induction by LPS is regulated posttranscriptionally via a Tpl2/ERK-dependent pathway. Cell 103: 1071-1083.
  157. Rousseau S, Papoutsopoulou M, Symons A, Cook D, Lucocq JM, et al. (2008) TPL2-mediated activation of ERK1 and ERK2 regulates the processing of pre-TNF alpha in LPS-stimulated macrophages. J Cell Sci 121: 149-154.
  158. Watford WT, Hissong BD, Durant LR, Yamane H, Muul LM, et al. (2008) Tpl2 kinase regulates T cell interferon-gamma production and host resistance to Toxoplasma gondii. J Exp Med 205: 2803-2812.
  159. Watford WT, Wang CC, Tsatsanis C, Mielke LA, Eliopoulos AG, et al. (2010) Ablation of tumor progression locus 2 promotes a type 2 Th cell response in Ovalbumin-immunized mice. J Immunol 184: 105-113.
  160. Sugimoto K, Ohata M, Miyoshi J, Ishizaki H, Tsuboi N, et al. (2004) A serine/threonine kinase, Cot/Tpl2, modulates bacterial DNA-induced IL-12 production and Th cell differentiation. J Clin Invest 114: 857-866.
  161. Belich MP, Salmerón A, Johnston LH, Ley SC (1999) TPL-2 kinase regulates the proteolysis of the NF-kappaB-inhibitory protein NF-kappaB1 p105. Nature 397: 363-368.
  162. Diehl NL, Enslen H, Fortner KA, Merritt C, Stetson N, et al. (2000) Activation of the p38 mitogen-activated protein kinase pathway arrests cell cycle progression and differentiation of immature thymocytes in vivo. J Exp Med 191: 321-334.
  163. Sugawara T, Moriguchi T, Nishida E, Takahama Y (1998) Differential roles of ERK and p38 MAP kinase pathways in positive and negative selection of T lymphocytes. Immunity 9: 565-574.
  164. Fernández E (2000) Thymocyte development past the CD4(+)CD8(+) stage requires an active p38 mitogen-activated protein kinase. Blood 95: 1356-1361.
  165. Hsu SC, Wu CC, Han J, Lai MZ (2003) Involvement of p38 mitogen-activated protein kinase in different stages of thymocyte development. Blood 101: 970-976.
  166. Lu HT, Yang DD, Wysk M, Gatti E, Mellman I, et al. (1999) Defective IL-12 production in mitogen-activated protein (MAP) kinase kinase 3 (Mkk3)-deficient mice. EMBO J 18: 1845-1857.
  167. Rincón M, Enslen H, Raingeaud J, Recht M, Zapton T, et al. (1998) Interferon-gamma expression by Th1 effector T cells mediated by the p38 MAP kinase signaling pathway. EMBO J 17: 2817-2829.
  168. Yu JJ, Tripp CS, Russell JH (2003) Regulation and phenotype of an innate Th1 cell: role of cytokines and the p38 kinase pathway. J Immunol 171: 6112-6118.
  169. Yang Z, Zhang X, Darrah PA, Mosser DM (2010) The regulation of Th1 responses by the p38 MAPK. J Immunol 185: 6205-6213.
  170. Rincon M, Pedraza-Alva G (2003) JNK and p38 MAP kinases in CD4+ and CD8+ T cells. Immunol Rev. 192: 131-42.
  171. Berenson LS, Yang J, Sleckman BP, Murphy TL, Murphy KM (2006) Selective requirement of p38alpha MAPK in cytokine-dependent, but not antigen receptor-dependent, Th1 responses. J Immunol 176: 4616-4621.
  172. Maneechotesuwan K, Xin Y, Ito K, Jazrawi E, Lee KY, et al. (2007) Regulation of Th2 cytokine genes by p38 MAPK-mediated phosphorylation of GATA-3. J Immunol 178: 2491-2498.
  173. Chi H, Lu B, Takekawa M, Davis RJ, Flavell RA (2004) GADD45beta/GADD45gamma and MEKK4 comprise a genetic pathway mediating STAT4-independent IFNgamma production in T cells. EMBO J 23: 1576-1586.
  174. Salvador JM, Mittelstadt PR, Guszczynski T, Copeland TD, Yamaguchi H, et al. (2005) Alternative p38 activation pathway mediated by T cell receptor-proximal tyrosine kinases. Nat Immunol 6: 390-395.
  175. Jirmanova L, Sarma DN, Jankovic D, Mittelstadt PR, Ashwell JD (2009) Genetic disruption of p38alpha Tyr323 phosphorylation prevents T-cell receptor-mediated p38alpha activation and impairs interferon-gamma production. Blood 113: 2229-2237.
  176. Rincon, M, Whitmarsh A, Yang DD, Weiss L, Dérijard B, et al. (1998) The JNK pathway regulates the In vivo deletion of immature CD4(+)CD8(+) thymocytes. J Exp Medicine 188: 1817-1830.
  177. Sabapathy K, Kallunki T, David JP, Graef I, Karin M, et al. (2001) c-Jun NH2-terminal kinase (JNK)1 and JNK2 have similar and stage-dependent roles in regulating T cell apoptosis and proliferation. J Exp Med 193: 317-328.
  178. Sabapathy K, Hu Y, Kallunki T, Schreiber M, David JP, et al. (1999) JNK2 is required for efficient T-cell activation and apoptosis but not for normal lymphocyte development. Curr Biol 9: 116-125.
  179. Behrens A, Sabapathy K, Graef I, Cleary M, Crabtree GR, et al. (2001) Jun N-terminal kinase 2 modulates thymocyte apoptosis and T cell activation through c-Jun and nuclear factor of activated T cell (NF-AT). Proc Natl Acad Sci U S A 98: 1769-1774.
  180. Weiss L, Whitmarsh AJ, Yang DD, Rincón M, Davis RJ, et al. (2000) Regulation of c-Jun NH(2)-terminal kinase (Jnk) gene expression during T cell activation. J Exp Med 191: 139-146.
  181. Rincón M, Dérijard B, Chow CW, Davis RJ, Flavell RA (1997) Reprogramming the signalling requirement for AP-1 (activator protein-1) activation during differentiation of precursor CD4+ T-cells into effector Th1 and Th2 cells. Genes Funct 1: 51-68.
  182. Yang DD, Conze D, Whitmarsh AJ, Barrett T, Davis RJ, et al. (1998) Differentiation of CD4+ T cells to Th1 cells requires MAP kinase JNK2. Immunity 9: 575-585.
  183. Constant SL, Dong C, Yang DD, Wysk M, Davis RJ, et al. (2000) JNK1 is required for T cell-mediated immunity against Leishmania major infection. J Immunol 165: 2671-2676.
  184. Gao M, Labuda T, Xia Y, Gallagher E, Fang D, et al. (2004) Jun turnover is controlled through JNK-dependent phosphorylation of the E3 ligase Itch. Science 306: 271-275.
  185. Li B, Tournier C, Davis RJ, Flavell RA (1999) Regulation of IL-4 expression by the transcription factor JunB during T helper cell differentiation. EMBO J 18: 420-432.
  186. Hartenstein B, Teurich S, Hess J, Schenkel J, Schorpp-Kistner M et al. (2002) Th2 cell-specific cytokine expression and allergen induced airway inflammation depend on JunB. EMBO J 21: 6321-6329.
  187. Fang D, Chris E, Baixue G, Nan F, Yoav A, et al. (2002) Dysregulation of T lymphocyte function in itchy mice: a role for Itch in Th2 differentiation. Nat Immunol 3: 281-287.
  188. Gallagher E, Gao M, Liu YC, Karin M (2006) Activation of the E3 ubiquitin ligase Itch through a phosphorylation-induced conformational change. Proc Natl Acad Sci U S A 103: 1717-1722.
  189. Enzler T, Chang X, Facchinetti V, Melino G, Karin M, et al. (2009) MEKK1 binds HECT E3 ligase Itch by its amino-terminal RING motif to regulate Th2 cytokine gene expression. J Immunol 183: 3831-3838.
  190. Venuprasad K, Elly C, Gao M, Salek-Ardakani S, Harada Y, et al. (2006) Convergence of Itch-induced ubiquitination with MEKK1-JNK signaling in Th2 tolerance and airway inflammation. J Clin Invest 116: 1117-1126.
  191. Guo Z, Clydesdale G, Cheng J, Kim K, Gan L, et al. (2002) Disruption of Mekk2 in mice reveals an unexpected role for MEKK2 in modulating T-cell receptor signal transduction. Mol Cell Biol 22: 5761-5768.
  192. Wang X, Chang X, Facchinetti V, Zhuang Y, Su B (2009) MEKK3 is essential for lymphopenia-induced T cell proliferation and survival. J Immunol 182: 3597-3608.
  193. Wang X, Zhang F, Chen F, Liu D, Zheng Y, et al. (2011) MEKK3 regulates IFN-gamma production in T cells through the Rac1/2-dependent MAPK cascades. J Immunol 186: 5791-5800.
  194. Chang X, Chang X, Facchinetti V, Zhuang Y, Su B (2011) The kinases MEKK2 and MEKK3 regulate transforming growth factor-beta-mediated helper T cell differentiation. Immunity 34: 201-212.
  195. Kretzschmar M, Doody J, Timokhina I, Massagué J (1999) A mechanism of repression of TGFbeta/ Smad signaling by oncogenic Ras. Genes Dev 13: 804-816.
  196. Derynck R, Zhang YE (2003) Smad-dependent and Smad-independent pathways in TGF-beta family signalling. Nature 425: 577-584.
  197. Matsuura I, Wang G, He D, Liu F (2005) Identification and characterization of ERK MAP kinase phosphorylation sites in Smad3. Biochemistry 44: 12546-12553.
  198. Wrighton KH, Lin X, Feng XH (2009) Phospho-control of TGF-beta superfamily signaling. Cell Res 19: 8-20.
  199. Lloyd CM, Hessel EM (2010) Functions of T cells in asthma: more than just T(H)2 cells. Nat Rev Immunol 10: 838-848.
  200. Okkenhaug K, Bilancio A, Farjot G, Priddle H, Sancho S, et al. (2002) Impaired B and T cell antigen receptor signaling in p110delta PI 3-kinase mutant mice. Science 297: 1031-1034.
  201. Okkenhaug K, Ali K, Vanhaesebroeck B (2007) Antigen receptor signalling: a distinctive role for the p110delta isoform of PI3K. Trends Immunol 28: 80-87.
  202. Sadhu C, Dick K, Tino WT, Staunton DE (2003) Selective role of PI3K delta in neutrophil inflammatory responses. Biochem Biophys Res Commun 308: 764-769.
  203. Lee KS, Lee HK, Hayflick JS, Lee YC, Puri KD (2006) Inhibition of phosphoinositide 3-kinase delta attenuates allergic airway inflammation and hyperresponsiveness in murine asthma model. FASEB J 20: 455-465.
  204. Holmes D, (2011) PI3K pathway inhibitors approach junction. Nat Rev Drug Discov 10: 563-564.
  205. Maas A, Hendriks RW (2001) Role of Bruton's tyrosine kinase in B cell development. Dev Immunol 8: 171-181.
  206. Setoguchi R, Kinashi T, Sagara H, Hirosawa K, Takatsu K (1998) Defective degranulation and calcium mobilization of bone-marrow derived mast cells from Xid and Btk-deficient mice. Immunol Lett 64: 109-118.
  207. Fiedler K, Sindrilaru A, Terszowski G, Kokai E, Feyerabend TB, et al. (2011) Neutrophil development and function critically depend on Bruton tyrosine kinase in a mouse model of X-linked agammaglobulinemia. Blood 117: 1329-1339.
  208. MacGlashan D Jr, Honigberg LA, Smith A, Buggy J, Schroeder JT (2011) Inhibition of IgE-mediated secretion from human basophils with a highly selective Bruton's tyrosine kinase, Btk, inhibitor. Int Immunopharmacol 11: 475-479.
  209. Chang BY, Huang MM, Francesco M, Chen J, Sokolove J, et al. (2011) The Bruton tyrosine kinase inhibitor PCI-32765 ameliorates autoimmune arthritis by inhibition of multiple effector cells. Arthritis Res Ther 13: R115.
  210. Uckun FM, Zheng Y, Cetkovic-Cvrlje M, Vassilev A, Lisowski E, et al. (2002) In vivo pharmacokinetic features, toxicity profile, and chemosensitizing activity of alpha-cyano-beta-hydroxy-beta- methyl-N-(2,5-dibromophenyl)propenamide (LFM-A13), a novel antileukemic agent targeting Bruton's tyrosine kinase. Clin Cancer Res 8: 1224-1233.
  211. Subramanian T, Namasivayam KM, Shanmugasundaram ER (1982) In vivo and in vitro studies on the binding nature of terreic acid with macromolecules such as protein and nucleic acids. Toxicol Lett 10: 249-253.
  212. Xu D, Kim Y, Postelnek J, Vu MD, Hu DQ, et al. (2012) RN486, a selective Bruton's tyrosine kinase inhibitor, abrogates immune hypersensitivity responses and arthritis in rodents. J Pharmacol Exp Ther 341: 90-103.
  213. Ponader S, Chen SS, Buggy JJ, Balakrishnan K, Gandhi V, et al. (2012) The Bruton tyrosine kinase inhibitor PCI-32765 thwarts chronic lymphocytic leukemia cell survival and tissue homing in vitro and in vivo. Blood 119: 1182-1189.
  214. Iyer AS, Morales JL, Huang W, Ojo F, Ning G, et al. (2011) Absence of Tec family kinases interleukin-2 inducible T cell kinase (Itk) and Bruton's tyrosine kinase (Btk) severely impairs Fc epsilonRI-dependent mast cell responses. J Biol Chem 286: 9503-9513.
  215. Iyer AS, August A (2008) The Tec family kinase, IL-2-inducible T cell kinase, differentially controls mast cell responses. J Immunol 180: 7869-7877.
  216. Forssell J, Sideras P, Eriksson C, Malm-Erjefält M, Rydell-Törmänen K, et al. (2005) Interleukin-2-inducible T cell kinase regulates mast cell degranulation and acute allergic responses. Am J Respir Cell Mol Biol 32: 511-520.
  217. Ferrara TJ, Mueller C, Sahu N, Ben-Jebria A, August A (2006) Reduced airway hyperresponsiveness and tracheal responses during allergic asthma in mice lacking tyrosine kinase inducible T-cell kinase. J Allergy Clin Immunol 117: 780-786.
  218. Sahu, N, Mueller C, Fischer A, August A (2008) Differential sensitivity to Itk kinase signals for T helper 2 cytokine production and chemokine-mediated migration. J Immunol 180: 3833-3838.
  219. Herdemann M, Weber A, Jonveaux J, Schwoebel F, Stoeck M, et al. (2011) Optimisation of ITK inhibitors through successive iterative design cycles. Bioorg Med Chem Lett 21: 1852-1856.
  220. Herdemann M, Heit I, Bosch FU, Quintini G, Scheipers C, et al. (2010) Identification of potent ITK inhibitors through focused compound library design including structural information. Bioorg Med Chem Lett 20: 6998-7003.
  221. Kutach AK, Villaseñor AG, Lam D, Belunis C, Janson C, et al. (2010) Crystal structures of IL-2-inducible T cell kinase complexed with inhibitors: insights into rational drug design and activity regulation. Chem Biol Drug Des 76: 154-163.
  222. Cook BN, Bentzien J, White A, Nemoto PA, Wang J, et al, (2009) Discovery of potent inhibitors of interleukin-2 inducible T cell kinase (ITK) through structure-based drug design. Bioorg Med Chem Lett 19: 773-777.
  223. Lin TA, McIntyre KW, Das J, Liu C, O'Day KD, et al. (2004) Selective Itk inhibitors block T-cell activation and murine lung inflammation. Biochemistry 43: 11056-11062.
  224. Marshall CJ (1994) MAP kinase kinase kinase, MAP kinase kinase and MAP kinase. Curr Opin Genet Dev 4: 82-89.
  225. Dudley DT, Pang L, Decker SJ, Bridges AJ, Saltiel AR (1995) A synthetic inhibitor of the mitogen-activated protein kinase cascade. Proc Natl Acad Sci U S A 92: 7686-7689.
  226. Favata MF, Horiuchi KY, Manos EJ, Daulerio AJ, Stradley DA, et al. (1998) Identification of a novel inhibitor of mitogen-activated protein kinase kinase. J Biol Chem 273: 18623-18632.
  227. Tsang F, Koh AH, Ting WL, Wong PT, Wong WS (1998) Effects of mitogen-activated protein kinase kinase inhibitor PD 098059 on antigen challenge of guinea-pig airways in vitro. Br J Pharmacol 125: 61-68.
  228. Wong WS, Tsang F, Li H, Ma B (1999) Effects of inhibitors of the tyrosine kinase signaling cascade on an in vitro model of allergic airways. Asian Pac J Allergy Immunol 17: 229-237.
  229. Laporte JD, Moore PE, Abraham JH, Maksym GN, Fabry B, et al. (1999) Role of ERK MAP kinases in responses of cultured human airway smooth muscle cells to IL-1beta. Am J Physiol 277: L943-951.
  230. Newton R, Cambridge L, Hart LA, Stevens DA, Lindsay MA, et al. (2000) The MAP kinase inhibitors, PD098059, UO126 and SB203580, inhibit IL-1beta-dependent PGE(2) release via mechanistically distinct processes. Br J Pharmacol 130: 1353-1361.
  231. Pahl A, Zhang M, Kuss H, Szelenyi I, Brune K (2002) Regulation of IL-13 synthesis in human lymphocytes: implications for asthma therapy. Br J Pharmacol 135: 1915-1926.
  232. Duan W, Chan JH, Wong CH, Leung BP, Wong WS (2004) Anti-inflammatory effects of mitogen-activated protein kinase kinase inhibitor U0126 in an asthma mouse model. J Immunol 172: 7053-7059.
  233. Tsitoura, DC, Rothman PB (2004) Enhancement of MEK/ERK signaling promotes glucocorticoid resistance in CD4+ T cells. J Clin Invest 113: 619-627.
  234. Fremin C, Meloche S (2010) From basic research to clinical development of MEK1/2 inhibitors for cancer therapy. J Hematol Oncol 3: 8.
  235. Sebolt-Leopold JS, Dudley DT, Herrera R, Van Becelaere K, Wiland A, et al. (1999) Blockade of the MAP kinase pathway suppresses growth of colon tumors in vivo. Nat Med 5: 810-816.
  236. Lorusso PM, Adjei AA, Varterasian M, Gadgeel S, Reid J, et al. (2005) Phase I and pharmacodynamic study of the oral MEK inhibitor CI-1040 in patients with advanced malignancies. J Clin Oncol 23: 5281-5293.
  237. Rinehart J, Adjei AA, Lorusso PM, Waterhouse D, Hecht JR, et al. (2004) Multicenter phase II study of the oral MEK inhibitor, CI-1040, in patients with advanced non-small-cell lung, breast, colon, and pancreatic cancer. J Clin Oncol 22: 4456-4462.
  238. Banerjee A, Koziol-White C, Panettieri R Jr (2012) p38 MAPK inhibitors, IKK2 inhibitors, and TNFα inhibitors in COPD. Curr Opin Pharmacol 12: 287-292.
  239. Chung KF (2011) p38 mitogen-activated protein kinase pathways in asthma and COPD. Chest 139: 1470-1479.
  240. Chopra P, Kanoje V, Semwal A, Ray A (2008) Therapeutic potential of inhaled p38 mitogen-activated protein kinase inhibitors for inflammatory pulmonary diseases. Expert Opin Investig Drugs 17: 1411-1425.
  241. Lee JC, Kumar S, Griswold DE, Underwood DC, Votta BJ, et al. (2000) Inhibition of p38 MAP kinase as a therapeutic strategy. Immunopharmacology 47: 185-201.
  242. Royce SG, Tang ML (2009) The effects of current therapies on airway remodeling in asthma and new possibilities for treatment and prevention. Curr Mol Pharmacol 2: 169-181.
  243. Kim SR, Lee KS, Park SJ, Jeon MS, Lee YC (2012) Inhibition of p38 MAPK reduces expression of vascular endothelial growth factor in allergic airway disease. J Clin Immunol 32: 574-586.
  244. Nath P, Leung SY, Williams A, Noble A, Chakravarty SD, et al. (2006) Importance of p38 mitogen-activated protein kinase pathway in allergic airway remodelling and bronchial hyperresponsiveness. Eur J Pharmacol 544: 160-167.
  245. Duan W, Chan JH, McKay K, Crosby JR, Choo HH, et al. (2005) Inhaled p38alpha mitogen-activated protein kinase antisense oligonucleotide attenuates asthma in mice. Am J Respir Crit Care Med 171: 571-578.
  246. Underwood DC, Osborn RR, Kotzer CJ, Adams JL, Lee JC, et al. (2000) SB 239063, a potent p38 MAP kinase inhibitor, reduces inflammatory cytokine production, airways eosinophil infiltration, and persistence. J Pharmacol Exp Ther 293: 281-288.
  247. O'Byrne PM, Inman MD, Parameswaran K (2001) The trials and tribulations of IL-5, eosinophils, and allergic asthma. J Allergy Clin Immunol 108: 503-508.
  248. Mori A, Okudaira H, Kobayashi N, Akiyama K (2001) Selective regulation of T cell IL-5 synthesis by OM-01, JTE-711 and p38 MAP kinase inhibitor: independent control of Th2 cytokines, IL-4 and IL-5. Int Arch Allergy Immunol 124: 172-175.
  249. Kankaanranta H, De Souza PM, Barnes PJ, Salmon M, Giembycz MA, et al, (1999) SB 203580, an inhibitor of p38 mitogen-activated protein kinase, enhances constitutive apoptosis of cytokine-deprived human eosinophils. J Pharmacol Exp Ther 290: 621-628.
  250. Bush A, Pedersen S, Hedlin G, Baraldi E, Barbato A, et al. (2011) Pharmacological treatment of severe, therapy-resistant asthma in children: what can we learn from where? Eur Respir J 38: 947-958.
  251. Barnes PJ (2004) Corticosteroid resistance in airway disease. Proc Am Thorac Soc 1: 264-268.
  252. Bhavsar P, Hew M, Khorasani N, Torrego A, Barnes PJ, et al. (2008) Relative corticosteroid insensitivity of alveolar macrophages in severe asthma compared with non-severe asthma. Thorax 63: 784-790.
  253. Bhavsar P, Khorasani N, Hew M, Johnson M, Chung KF (2010) Effect of p38 MAPK inhibition on corticosteroid suppression of cytokine release in severe asthma. Eur Respir J 35: 750-756.
  254. Mercado N, To Y, Kobayashi Y, Adcock IM, Barnes PJ, et al. (2011) p38 mitogen-activated protein kinase-γ inhibition by long-acting β2 adrenergic agonists reversed steroid insensitivity in severe asthma. Mol Pharmacol 80: 1128-1135.
  255. Mercado N, Hakim A, Kobayashi Y, Meah S, Usmani OS, et al. (2012) Restoration of corticosteroid sensitivity by p38 mitogen activated protein kinase inhibition in peripheral blood mononuclear cells from severe asthma. PLoS One 7: e41582.
  256. Robins S, Roussel L, Schachter A, Risse PA, Mogas AK, et al. (2011) Steroid-insensitive ERK1/2 activity drives CXCL8 synthesis and neutrophilia by airway smooth muscle. Am J Respir Cell Mol Biol 45: 984-990.
  257. Sabapathy K (2012) Role of JNK Pathway in Health and Disease. Protein Phosphorylation in Health and Disease, Volume 106, S. Shenolikar edition. Academic Press, Elsevier.
  258. Assi K, Pillai R, Gómez-Muñoz A, Owen D, Salh B (2006) The specific JNK inhibitor SP600125 targets tumour necrosis factor-alpha production and epithelial cell apoptosis in acute murine colitis. Immunology 118: 112-121.
  259. Mitsuyama K, Suzuki A, Tomiyasu N, Tsuruta O, Kitazaki S, et al. (2006) Pro-inflammatory signaling by Jun-N-terminal kinase in inflammatory bowel disease. Int J Mol Med 17: 449-455.
  260. Reinecke K, Eminel S, Dierck F, Roessner W, Kersting S, et al. (2012) The JNK inhibitor XG-102 protects against TNBS1489 induced colitis. PloS one 2012. 7: e30985.
  261. Williams AS, Issa R, Leung SY, Nath P, Ferguson GD, et al. (2007) Attenuation of ozone-induced airway inflammation and hyper-responsiveness by c-Jun NH2 terminal kinase inhibitor SP600125. J Pharmacol Exp Ther 322: 351-359.
  262. Nath P, Eynott P, Leung SY, Adcock IM, Bennett BL, et al. (2005) Potential role of c-Jun NH2-terminal kinase in allergic airway inflammation and remodelling: effects of SP600125. Eur J Pharmacol 506: 273-283.
  263. Eynott PR, Xu L, Bennett BL, Noble A, Leung SY, et al. (2004) Effect of an inhibitor of Jun N-terminal protein kinase, SP600125, in single allergen challenge in sensitized rats. Immunology 112: 446-453.
  264. Murphy KM, Stockinger B (2010) Effector T cell plasticity: flexibility in the face of changing circumstances. Nature immunology 11: 674-680.
  265. Wynn ML, Ventura AC, Sepulchre JA, García HJ, Merajver SD (2011) Kinase inhibitors can produce off-target effects and activate linked pathways by retroactivity. BMC Syst Biol 5: 156.
Citation: Kannan Y, Wilson MS (2012) TEC and MAPK Kinase Signalling Pathways in T helper (TH) cell Development, TH2 Differentiation and Allergic Asthma. J Clin Cell Immunol S12:011.

Copyright: © 2012 Kannan Y, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top