Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Jun 1.
Published in final edited form as: Semin Immunol. 2009 Apr 21;21(3):164–171. doi: 10.1016/j.smim.2009.03.001

Barrier immunity and IL-17

Benjamin R Marks 1, Joe Craft 1,2
PMCID: PMC2692766  NIHMSID: NIHMS104805  PMID: 19386512

Abstract

CD4+ TH17 cells display a featured role in barrier immunity. This effector population of T cells is important for clearance of microorganisms but can also promote autoimmunity at barrier sites. Recent work has indicated that these effector cells share a pathway with CD4+ regulatory T cells (TR cells) that also have a critical function in barrier protection and immune regulation. The development and function of TH17 cells, and their relationship with TR cells are discussed.

Keywords: effector T cells, IL-17, mucosa, regulatory T cells, TH17 cells

Introduction

Mucosal and epithelial surfaces harbor a significant number of immune cells that are necessary to provide host protection at these susceptible sites for pathogen entry. These surfaces, including the gastrointestinal tract, lungs, skin and reproductive tract, present a particularly difficult and precarious scenario for the cells of the immune system. Unlike the sterile environment of the systemic circulation, the blood and the lymphatics, the epithelial layer is bathed in a sea of microorganisms. The evolution of the immune system has occurred so that a symbiotic relationship exists between the commensal bacteria of the gut and human hosts. Teleologically one may predict that the epithelial barriers are constructed so that microorganisms are sequestered outside of a tight barrier so that immune response to resident bacteria is rare. However, more recent data presents the opposite story and indicates that the gut microflora are a critical component to proper immune function and in their absence, the immune system has inadequate development often leading to immune dysregulation. Thus, the immune system that has co-evolved in the setting of diverse microorganisms that are beneficial and potentially harmful to the host has developed a specialized approach to addressing the complex nature of the barrier layer between host and the outside world.

The phenotype of T cells at the barrier surfaces

Given the unique conditions existing at the body's interface with the external environment, the cells of the immune system at these sites operate differently than those in the lymphoid organs. Within the secondary lymphoid organs (SLO), the majority of T cells are naive, identified by the expression of L-selectin (CD62L) and CD45RA (in humans) that upon activation produce IL-2[1]. Naïve cells are primarily recruited to the secondary lymphoid organs such as the lymph node and spleen mediated by L-selectin[2] and the chemokine (C-C motif) receptor 7 (CCR7)[3]. In these locations, naïve cells are continuously surveying available antigen from resident and migrating dendritic cells[4] and awaiting the correct combination of T cell receptor (TCR) signaling in the context of co-stimulation to initiate activation[5]. Following stimulation, T cells differentiate into distinct effector lineages dictated by the activating environment so that the proper immune response occurs[6]. This process requires several days for naïve T cell activation to produce a functional effector T cell exported from the lymph node[7].

In contrast, conventional αβ T cells residing at barrier sites have an effector or memory cell phenotype[8]. These T cells express CD44, a molecule important for their non-specific exit from the systemic circulation and residence within peripheral organs[9, 10]. They also express distinct chemokine receptors and integrins that allow recruitment to specific sites. Gastrointestinal tract tropism is determined by CCR9 mediated recruitment via the small intestine's expression of CCL25 (TECK) in combination with α4β7 expressed on effector T cells promoting adhesion to mucosal addressin cell adhesion molecule-1 (MAdCAM-1) expressed by the postcapillary endothelial cells in the small intestine[11, 12]. For efficient dermal and epidermal homing, T cells express CCR4 and CCR10 that bind to CCL17 and CCL27, respectively, expressed by the skin during resting and inflammatory conditions[13-16]. In addition, T cells necessary for the protection against inhaled pathogens are directed to the lung by the expression of CCR3 and CCR5. While unique adhesion molecules direct effector T cells to specific barrier locations, other molecules such as α4β1 and CCR6 play a more general but important role in recruitment to mucosa and the skin[17, 18].

Another important distinction between the αβ T cells that reside in the epithelial surfaces compared to the compartment within the lymph node or spleen is the reduced threshold for activation and rapid response to pathogens[8]. At barrier surfaces, effector T cell populations are primed for cytokine secretion. These cells are activated much faster at least partially due to altered co-stimulatory molecule expression. Naïve T cells rely on CD28 as their second signal while memory cells located at epithelial layers make use of ICOS as well as others. The cells at these sites are primed and ready to respond rapidly to any sign of infection to help remove the source before it results in infection[19].

The effector T cell paradigm

For the past 20 years, the TH differentiation paradigm consisted of two mutually exclusive pathways, TH1 and TH2, defined by distinct cytokine production and immune function[20]. The first, TH1, is characterized by the production of IFNγ and directs cell-mediated immunity. This subset requires the transcription factor T-bet[21], is induced by IFN-γ and IL-12 from macrophages or DCs[22] and requires STAT1 and STAT4 signaling[23, 24]. TH1 responses are necessary for intracellular pathogen clearance[25] and in mice induce B cells to produce IgG2a[26]. The other subset, TH2, is characterized by the production of IL-4, IL-5 and IL-13[27] and in mice, directs B cell secretion of IgE and IgG1[26]. This subset requires the transcription factor GATA3[28], is induced by IL-4 or thymic stromal lymphopoietin from basophils[29], and requires STAT6 signaling[30]. TH2 responses are necessary for clearance of extracellular parasitic infections and cause allergic disease[25]. The two subsets are distinct lineages in that GATA3 and T-bet negatively regulate each other and the presence of IFNγ prevents TH2 and IL-4 prevents TH1[31] differentiation.

More recent work indicates that another subset of effector TH cells exists. This lineage is defined by the production of IL-17 and has been given the name TH17. These cells are critical for protection against extracellular bacteria and fungi and are responsible for several autoimmune conditions[32].

The TH17 subset

Interleukin 17: structure and function

The TH17 subset of helper T cells is defined by the production of the IL-17 cytokine. This cytokine was first described 15 years ago and was originally given the name cytotoxic T-lymphocyte-associated antigen 8 (CTLA-8)[33, 34], and later renamed IL-17[35]. Subsequent work revealed that it was the first identified in a family of six cytokines, now referred to as IL-17A through F, with IL-17F showing the highest degree of homology with IL-17A followed by B, D, C, and E[36]. The cellular source of IL-17 was originally identified in activated T cells[33, 34] but more recently been expanded to include γδ T cells[37], CD8+ memory T cells[38], neutrophils[38] and monocytes[39]. CD4+ T cells are considered the significant producers of this cytokine.

IL-17 is pro-inflammatory and important for the clearance of extracellular pathogens and multiple autoimmune disorders. Experimental models using mice with defective IL-17 signaling or treated with depleting antibodies show increased susceptibility to lung infection by Klebsiella pneumonia and Mycoplamsa pneumoniae and a defect in clearance of Candida albicans and Escherichia coli [40-43]. This effect has been linked to IL-17-mediated neutrophil recruitment as well as induction of anti-microbial proteins from resident cells. IL-17 stimulates a host of inflammatory cytokines and chemokines, including granulocyte colony-stimulating factor (G-CSF), macrophage inflammatory protein-2 (MIP-2), IL-8, monocyte chemotactic protein-1 (MCP-1), CXCL-8, CXCL-1 and CXCL-10[36, 44-47] along with other inflammatory mediators such as prostaglandin E2, nitric oxide, matrix metalloproteases, acute phase proteins and IL-6[45, 46, 48, 49]. Along the same lines, IL-17 can promote unfavorable immune responses indicated by this cytokine's role in multiple autoimmune disorders such as rheumatoid arthritis[50], psoriasis[51], inflammatory bowl disease[52], asthma[53], and multiple sclerosis[54, 55].

Given the robust immune response mediated by IL-17, it is not surprising that targets for this cytokine are highly diverse. Studies of mRNA expression indicate that the receptor is present on hematopoetic cells, osteoblasts, fibroblasts, endothelial cells and epithelial cells in the lung, liver, spleen and kidney[35, 56]. In fact, this family of receptors is as complex as their ligands. Sequences homology searches have revealed that there are five members, IL-17RA to IL-17RE[36, 57]. This group represents a unique family containing domains not observed previously and is structured as a single-pass transmembrane proteins with an extracellular domain and a long intracellular tail[35]. Further analysis indicated that all receptors except IL-17RA have alternative splicing variants that introduce early stop codons allowing for the receptor to be secreted[58, 59] and potentially act as a decoy to help reduce IL-17 signaling during an immune response.

To date, functional studies of the IL-17 receptor family are still lacking, with most analyses limited to IL-17RA and more recently IL-17RC. While both of these receptors can bind to IL-17 and IL-17F, IL-17RA has a log fold decreased affinity for IL-17F[60] while IL-17RC binds equally to both[61]. IL-17RA exists as a preformed homodimer[62] or can function as a heterodimer pairing with IL-17RC[63]. In fact, IL-17RA may be a generic receptor for all IL-17 family members given the recent report that both IL-17RA and IL-17RB are necessary for IL-17E (IL-25) signaling, a TH2 inducing pathway, very different from IL-17[64].

Upon ligand binding, the IL-17 receptor undergoes a conformational change facilitating dissociation of the intracellular region. The IL-17RA has a cytoplasmic tail with motifs similar to the TLR-IL-1 receptor (TIR)[65] superfamily, now termed SEFIR domain (similar expression to FGF receptor, IL-17 receptor, Toll-IL-1R)[66] but does not require the myeloid differentiation factor 88 (MyD88) for signaling. Upon IL-17RA engagement, signaling via Act1[67] promotes TRAF6 ubiquitination of the receptor[68] and activates the NF-κB transcription factor pathway[67, 69-71]. However, even with multiple family members and overlap of IL-17RA and IL-17RC in binding, there is minimal redundancy in the function of IL-17RA given that targeted deletion of this receptor causes profound defects in host defense[47].

TH17: a new helper subset

As mentioned earlier, CD4+ helper T cells have been divided into two distinct effector lineages, TH1 and TH2[20]. These two subsets develop during the course of an infection and are selected by the inflammatory signals provided by the innate immune system so that a particular type of immune response can be carried out. These two effector lineages have somewhat opposing functions[72-74] with TH1 cells driving cell-mediated immune responses that can cause tissue damage and experimentally characterized as the pathway necessary for delayed type hypersensitivity (DTH)[75], while TH2 cells promoting antibody-mediated responses and are associated with allergy and the IgE isotype[76].

Although the TH1/TH2 paradigm explained many experimental systems and disease models, there were some inconsistencies that provided a framework for the introduction of a new subset[32, 77]. These experimental observations related to the IL-12 cytokine family and the discovery of another member, IL-23. IL-12 is a heterodimeric cytokine that is composed of p40 and p35 subunits to make a complete cytokine p70[78, 79] that signals via the IL-12 receptor consisting of the IL-12Rβ1 and IL-12Rβ2[80, 81]. This signaling pathway is necessary for TH1 development and genetic deficiency of this cytokine prevents IFNγ production from T cells and mice normally resistant to Leishmania major die from the infection[82-84].

However, other experimental models, particular experimental autoimmune encephalitis (EAE) raised questions regarding the simplicity of IL-12 and the IFNγ inducing effect. Initial work using antibodies to block p40 or mice deficient in this subunit showed resistance to EAE, indicating that IL-12 and presumably IFNγ were necessary for disease development[85-87], except that, in the absence of IFNγ, mice were still susceptible to the disease[88]. The data indicated a divergent function between the p40 subunit of IL-12 and IFNγ induction. Support for this notion occurred when p35 deficiency had the same effect on EAE as the IFNγ knockout and opposite effect when p40 was lacking[89]. These observations indicated that the p40 subunit had functions apart from pairing with p35 to induce a TH1 response.

This issue was resolved when another IL-12 family member, IL-23 was discovered[90] and shown to be critical for the induction of IL-17 from CD4+ T cells[91]. The IL-23 cytokine shares the p40 subunit but pairs with a unique p19 protein that together bind to a receptor composed of the shared IL-12Rβ1 and a unique IL-23 receptor[92] (Table 1). Similar to the IL-12 requirement for IFNγ production, it was shown that the related cytokine IL-23 induces T cells to produce IL-17[91]. Soon after, EAE disease induction was proven to be dependent upon IL-23 derived TH17 cells and actually protected by the IL-12/IFNγ pathway[55, 89]. Thus, very similar IL-12 family member cytokines with a common p40 subunit and IL-12Rβ1 induce distinct effector pathways consisting of TH1 and TH17.

Table 1. IL-12 and IL-23 cytokines and receptors.

cytokine structure receptor structure T helper pathway
unique common unique common
IL-12 p35 p40 IL-12Rβ2 IL-12Rβ1 TH1
(STAT4)
IL-23 p19 p40 IL-23R IL-12Rβ1 TH17
(STAT3)

TH17 development

The discovery of IL-23 and the identification of its role in IL-17 mediated disease set the stage for characterization of the new subset, TH17[55, 89], and understanding how these effector cells develop. Similar to TH1 and TH2, the TH17 lineage has a distinct in vitro differentiation pathway. Sorting of naïve T cells and culturing in the presence of TGFβ induces FoxP3 transcription factor and converts the majority of T cells to regulatory T cells (TR); however, addition of IL-6 to the culture conditions changed the phenotype to IL-17 production[93, 94]. In fact, activating T cells with TR, producers of TGFβ, and DCs stimulated with a Toll like receptor agonist, a source of IL-6, also converted the naïve population to TH17[95]. In these in vitro skewing experiments, the absence of IL-23 had no effect on TH17 development leading to the conclusion that TGFβ and IL-6 mediate the initial TH17 differentiation while IL-23 is important for survival and expansion. In vivo experiments using genetically altered mice with non-functional TGFβ receptor signaling or impaired T cell production confirmed the role of this cytokine in IL-17 differentiation and subsequent EAE disease development[96, 97].

TH17 transcription factors and signaling pathways

Analogous to T-bet, GATA3 and FoxP3 for TH1, TH2 and TREG, the transcription factor retinoic acid receptor-related orphan receptor (ROR)γt directs the differentiation of TH17. Mice that lack RORγt cannot make IL-17 producing T cells and retroviral transduction into naïve cells promotes TH17 development[98]. In addition, another transcription factor of the same family, RORα, plays a synergistic role with RORγt in TH17 differentiation[99]. It has recently been shown that Runx1 is an important transcription factor in binding to RORγt and FoxP3 to promote efficient TH17 development[100]. Interferon-regulatory factor (IRF) 4, a mediator of TH2 development is also required but not specific for TH17 induction[101].

The specific sequence of cell signaling events involved in TH17 development and function has been partially elucidated. IL-23 and IL-6 activate STAT3 signaling, now considered necessary and unique for TH17 differentiation[102]. As part of activation, the TH17 cells make IL-21 that provides autocrine signaling and can replace the need for IL-6[103-105]. As part of this process, the suppressor of cytokine signaling (Socs) 3 is turned off as it functions as a negative regulator[106].

TH17 regulation

Analogous to the inhibitory effects that TH1 cytokines have on TH2 development and the reverse, TH17 function is also influenced by TH1 and TH2 cytokines. In vitro activation and differentiation of naïve T cells to the TH17 lineage is enhanced with blocking of IFNγ and IL-4[94, 107]. Mice lacking the TH1 transcription factor T-bet develop exaggerated TH17 levels in the setting of autoimmune disease such as myocarditis or during Mycobacterium bacterial infection[108, 109]. However, this data does not determine if the presence of these cytokines prevents development of TH17 cells or regulates the secretion of IL-17 follow lineage commitment. In addition, IL-27, another member of the IL-12 cytokine family, with TH1 inducing properties, can inhibit TH17 independent of its TH1 promoting function[110]. Along the same lines, IL-17E also known as IL-25, has suppressive function and its absence promotes enhanced IL-17 levels that exacerbates EAE[110] and allows for increased TH17 cell development in the gut[111]. Thus, the IL-17 cytokine contributes to a specific type of inflammatory response and as appropriate is carefully regulated by other cytokines to promote swift resolution of toxic inflammatory conditions to minimize injury to the host.

TH17: An effector lineage sharing a regulatory T cell pathway

The TH17 subset is often considered a parallel effector lineage to TH1 and TH2[112] with a distinct role in the pathogenesis of specific autoimmune conditions and a mediator of microbial clearance (Table 2). Early work indicating a dependency on IL-23 presented an analogous developmental pathway to TH1 induction by IL-12[32] and presented the initial idea that the secretion of related factors IL-12 or IL-23 determined the fate of the developing immune response. However, subsequent data established IL-23 as a survival factor and identified TGFβ in conjunction with IL-6 as the lineage determining cytokines[93-95]. Thus, the comparison to TH1 development was diminished and a link to TR cells was introduced.

Table 2. Description of Helper T cell subsets.

cytokines TF* signaling in vitro skewing pathogen clearance disease
TH1 IFNγ T-bet eomes STAT1 STAT4 IL-12 anti-IL4 intracellular delayed type hypersensitivity
TH2 IL-4, IL-5, IL-13 GATA3 STAT6 IL-4 anti-IFNγ parasites asthma, allergy
TH17 IL-17, IL-17F, IL-22 RORγt RORα STAT3 TGFβ/IL-6/IL-23/IL-21 anti-IL-2 extracellular EAE, RA, IBD, psoriasis
Treg TGFβ, IL-10 FoxP3 STAT5 TGFβ/IL-2 tumor
*

transcription factors

Activating naïve T cells in vitro in the presence of TGFβ alone promotes development of TR with the addition of IL-6 diverting differentiation to TH17. This was the first indication that this inflammatory subset shared a common lineage with TR[93, 94]. Further support comes from IL-2, a cytokine necessary for TR survival[113]. This cytokine constrains the development of TH17 cells so that TGFβ/IL-6 in the presence of IL-2 had significantly reduced TH17 development and expanded FoxP3+ TR cells[114]. However, inflammatory conditions, such as provision of IL-1 with TGFβ/IL-6 in the presence of IL-2 rescued the IL-2 inhibitory effect and restored TH17 differentiation[115]. Additional reports find that FoxP3+ TR cells can be converted directly to TH17 producing cells with the correct inflammatory conditions[116]. The vitamin A metabolite, retinoic acid, produced by DCs within the gut, is responsible for preventing inflammation by diverting TH17 cells into TR[117].

A convincing piece of work proving a common lineage between TH17 and TR cells comes from a study using reporter mice to track the expression of FoxP3 and RORγt in T cells. The authors showed that TGFβ signaled in a concentration dependent manner to promote the expression of both FoxP3 and RORγt. FoxP3 directly bound to RORγt preventing TH17 differentiation an effect relieved by IL-6, IL-21 and IL-23[118]. An additional report confirms the suppressive function of FoxP3 on RORγt and adds that Runx1 is critical in binding both transcriptions factors to promote TH17 development[100].

Additional support for a TH17/TR shared developmental pathway was provided by identification of T cells fated to become TR but unable to express FoxP3 due to an insertion of GFP in place of this gene. In so doing, the TR fated cells in the absence of FoxP3 converted to RORγt expressing cells and produced IL-17[119]. Thus, in the absence of FoxP3, natural mechanisms selecting for TR development default to TH17, suggesting that altering thymic conditions such as IL-6 or IL-1 may select for TH17 cells from the TR compartment.

TH17 associated cytokines

The TH17 subset is associated with several other cytokines that contribute to this subset's unique function. The IL-17 family member, IL-17F, the closest related cytokine to IL-17 within this family, is also secreted by this lineage[55]. IL-17F can function similar to IL-17 by inducing production of IL-6, IL-8 and CXCL1 from in vitro cultured cells and administration of exogenous IL-17F during asthma induction promotes neutrophil recruitment[120, 121]. However, despite similarities in protein sequence and function, IL-17F does not have complete redundancy with IL-17. For example, IL-17 knock out mice exhibit reduced arithritis[122], EAE[123] and allergic responses[124]. In fact, recent data indicates a distinct role during gut inflammation given that IL-17 knockout mice have reduced survival during DSS-induced colitis while IL-17F deficient mice are protected[125].

The IL-10 family member, IL-22, is also an established TH17 associated cytokine. In vitro activation of naïve T cells in the presence of TH17 skewing conditions, i.e. TGFβ and IL-6, promotes IL-22 production[126, 127]. The source of IL-22 is limited to T cells, NK cells and NK T cells[128]. The receptor for IL-22 consists of the IL-10 receptor (IL-10R) β and IL-22 receptor (IL-22R)[129]. While the IL-10Rβ has broad expression, the IL-22R is limited to the skin, liver, lung and pancreas but not detected in T cells[128, 130]. Thus, IL-22 promotes signaling to peripheral organs and does not directly influence T cell responses.

IL-22 is an important factor during inflammation. This cytokine in cooperation with IL-17 induces anti-microbial peptide activation to enhance clearance of bacterial infections[126]. IL-22 knockout mice indicate that this cytokine has an important role in psoriasis and hepatitis. In the setting of dermal inflammation, IL-22 plays a pathologic role in promoting acanthosis[127]. Conversely, during acute inflammation of the liver, IL-22 is protective and reduces liver enzyme elevation[131].

IL-17: an important cytokine in immune barrier function

The IL-17 cytokine is a major player in the immune responses at epithelial surfaces. This factor is important for efficient clearance of pathogenic infections and responsible for significant autoimmune pathology.

Lung

IL-17 is critical for protecting the host from lung-associated pathogens. Studies using IL-17RA and IL-23 knockout mice highlight the importance of this cytokine in the clearance of the pathogen Klebsiella pneumonia[41, 132]. Other bacterial infections such as Mycobacterium tuberculosis[133] and Mycobacterium bovis[134] can induce an IL-17 response that is important for preventing lethal disease. A role for IL17 has been suggested in viral infections such as in synergistic recruitment of neutrophils in human rhinovirus infection[135]. This cytokine has also been linked to opportunistic fungal infections such as the HIV related Pneumocystis carinii[136] and Candida albicans[40].

While IL-17 is considered beneficial for protecting the lungs from the constant exposure to potential pathogens, this cytokine is responsible for directing inflammation during allergic asthma. This cytokine is increased in the airways of people with asthma consistent with its inflammatory role in promoting inflammation[137]. However, its function in allergic lung inflammation is not clear. While IL-17 contributes to the recruitment of neutrophils and eosinophils to the lungs, IL-17RA knockout mice have worse TH2 disease indicating an inhibitory/protective role in mediating TH2 type disease[138]. Additional studies in mice indicate that the IL-17 cytokine promotes a distinct type of inflammatory lung disease. Mice that receive TH2 skewed T cells were responsive to treatment with dexamethasone while TH17 skewed cells induced significant airway inflammation but unresponsive to steroid treatement[139]. Thus, this subset can direct unwanted lung inflammation and may help to explain why some people are resistant to conventional asthma therapy.

Gastrointestinal tract

At this mucosal surface of the gastrointestinal tract, the body is exposed to an abundance of microorganisms, most of which are important for preventing overgrowth of pathogenic bacteria and necessary for immune homeostasis. As such, the immune system has developed mechanisms to distinguish between the harmful and the helpful residents of our gut. One indication that the TH17 subset plays an important role at this site comes from the study that initially identified RORγt as the lineage specific transcription factor. In this study, the authors found the highest concentration of TH17 cells were within the lamina propria of the small intestine, almost 10% of αβ T cells[98]. This finding is quite striking when considering that it was later shown that mucosal DCs were poor inducers of TH17 cells, secondarily to the production of retinoic acid, compared to their lymph node counterparts that were much more TH17 permissive[117]. These findings appear to be contradictory unless one considers that the TH17 cells located in the lamina propria originate somewhere else and the retinoic acid from the mucosal DCs control these imported potentially pathogenic T cells.

There is a dynamic interaction between the commensal bacteria and the immune cells of the gut. Toll-like receptors are responsible for mediating this cross talk and instructing the immune system appropriately[140]. For example, toll-like receptor 9 detects gut flora DNA to regulate the balance between TR and TH17 cells of the gut. In the absence of this signaling pathway, the T cells of the gut are overwhelmed with regulatory cells and prevent productive immune function[141]. Similarly, the native gut flora provides a balance between related IL-17 cytokines, IL-17 and IL-17E (IL-25). Pathogen-free conditions promote TH17 overgrowth and elevated IL-23 while restoration of microorganisms signals for IL-17E to re-establish the proper balance and promote healthy intestinal homeostasis[111].

The IL-17/IL-23 axis is a critical player in the promotion of inflammatory bowl disease. Mouse models of gut inflammation were originally attributed to TH1 effector subsets based on the observation that antibodies directed at the p40 subunit of IL-12 proved to be an effective treatment[142]. However, the identification of p40 as a shared subunit between IL-12 and IL-23 has prompted a re-examination of gut inflammation. It is now established that the IL-17 pathway is an important cytokine involved in autoimmune disease of the gut[143, 144]. In human studies, it was observed that IL-23 and IL-17 are elevated in patients with IBD[145] and that treatment with anti-p40 antibodies is very effective in preventing the disease symptoms[146], most likely due to reduction in IL-23 and subsequently IL-17. The IL-23/IL-17 pathway has also been implicated in the promotion of unwanted gut inflammation with the identification of genetic variations of IL-23, STAT3 and other TH17 associated genes linked to Crohn's disease and ulcerative colitis[147].

Skin

The dermis/epidermis is another very large barrier organ housing a distinct immune cell population. At this site, similar to the gut, microorganisms are ubiquitous along the outside border. Here too, TH17 cells are present and help to protect this potential danger zone from pathogen entry. People with an inability to clear the opportunistic fungal infection Candida suffer from mucocutaneous candidiasis. A recent report indicates that peripheral blood mononuclear cells from these patients have reduced IL-17 and IL-22 mRNA and that their CCR6+ IL-17+ T cells are significantly reduced[148]. Mouse studies confirm the role of IL-17 in preventing this cutaneous yeast infection[40].

In addition, the TH17 associated cytokine IL-22 has been at the forefront of autoimmune pathology of the dermis. Mouse models of psoriasis indicate that IL-22 production promotes keratinocyte survival and drives acanthosis[127]. Furthermore, it has been reported that human psoriatic lesions have increased IL-23 mRNA compared to healthy skin[149]. As one might expect, IL-17 expression from psoriatic plaques correlated with disease severity and cytokine levels normalized following treatment with cyclosporine[150].

γδ T cells as a source of IL-17 at epithelial surfaces

The γδ T cell subset makes up a small fraction of the total T cell compartment but serves a distinct conserved function. These T cells develop in the thymus, require random recombination events similar to αβ T cells but are produced in waves of subsets defined by the individual γ and δ receptors that they express. The different subsets of γδ T cells vary in function and location[151], [152]. Thus, this group of immune cells has the potential to generate great diversity in antigen recognition but somehow targets specific T cell subsets to reside in distinct locations to provide a protective function unique to the individual site. γδ T cells are not restricted to classical MHC class I or II molecules like their CD8+ or CD4+ αβ counterparts. A small population of γδ T cells has been shown to recognize MHC class IB antigens T10 and T22 inmice[153] and other “stress markers” such as MICA in humans through their TCR[154] or by expression of NKG2D[155, 156]. Thus, γδ T cells do not require foreign antigen to induce activation and promote inflammation but rather respond to endogenous signals that indicate pathogen entry into protected sites. It may be this unconventional activation mechanism that has selected for their niche within the immune system.

γδ T cells have several mechanisms in which they contribute to immune responses. These T cells have been noted to provide common T cell cytokines such as IFNγ and IL-17 during the innate phase of inflammation. γδ T cells produce IL-17 and depending on the timing, can represent a majority of cells producing this cytokine[157]. This subset is unique in that these T cells do not require priming to allow effector function causing a delay in IL-17 production but rather can secrete this cytokine immediately upon activation. This ability to produce IL-17 is part of the thymic developmental pathway that selects the individual γδ subsets given the observation that Vγ4 thymocytes in mice can produce IL-17 while the Vγ1 cells have minimal production[158] which may be related to the affinity for ligand binding taking place within the thymus[159]. As such, γδ T cells may be an important source of the IL-17 cytokine. Given the role of IL-17 in neutrophil recruitment and other early inflammatory responses, this population of T cells being instructed in the thymus to populate host organs, especially the barrier surfaces, and having an ability to produce IL-17 immediately upon activation presents a unique model for T cell production of IL-17 and helps fine tune the immune response.

Conclusions

T cells are critical in the complex regulation of barrier immunity. These unique sites require dynamic interactions between the cells of the immune system and the surrounding environment. The immune system has evolved not only to prevent unwanted activation in response to microorganisms residing at these sites but makes use of these species to shape the mature compartment guarding the epithelial lining. A newly described helper T cell subset, TH17, has proven to be a major player in protecting the host barrier surfaces. IL-17 has been implicated in clearance of bacterial, viral and fungal infections occurring in the lung, gut and skin as well as the pathogenic mediator of multiple mucosal and cutaneous autoimmune diseases.

The TH17 subset, while often described as a third effector lineage, parallel to the TH1 and TH2 subsets, has certain characteristics that may place these cells in a class of their own. One of the most striking findings is the direct lineage relationship with TR cells. These two subsets require a common cytokine TFGβ for lineage commitment and recent work has observed direct interactions between the lineage specific transcription factor FoxP3 and RORγt. While TH1 and TH2 cells function as late contributors to pathogen clearance, the TH17 subset is implicated in neutrophil recruitment, a function that is necessary during early inflammation, before the conventional adaptive phase of the immune response. Thus, one may speculate that this subset is already pre-formed waiting for the correct signal to promote inflammation and direct efficient immune clearance and host protection. We are still adding pieces to the puzzle defining the TH17 subset to reveal the true role of this unique and critical helper T cell subset.

Acknowledgments

This work was supported in part by NIH Grants AR40072, AR44076, and P30 AR053495, and by support from Rheuminations, Inc., the Arthritis Foundation and the Connecticut Chapter of the Lupus Foundation of America.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Truneh A, Albert F, Golstein P, Schmitt-Verhulst AM. Early steps of lymphocyte activation bypassed by synergy between calcium ionophores and phorbol ester. Nature. 1985;313:318–20. doi: 10.1038/313318a0. [DOI] [PubMed] [Google Scholar]
  • 2.Gallatin WM, Weissman IL, Butcher EC. A cell-surface molecule involved in organ-specific homing of lymphocytes. Nature. 1983;304:30–4. doi: 10.1038/304030a0. [DOI] [PubMed] [Google Scholar]
  • 3.Gunn MD, Kyuwa S, Tam C, Kukiuchi T, Matsuzawa A, Williams LT, et al. Mice lacking expression of secondary lymphoid organ chemokine have defects in lymphocyte homing and dendritic cell localization. J Exp Med. 1999;189:451–60. doi: 10.1084/jem.189.3.451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Mempel TR, Henrickson SE, Von Andrian UH. T-cell priming by dendritic cells in lymph nodes occurs in three distinct phases. Nature. 2004;427:154–9. doi: 10.1038/nature02238. [DOI] [PubMed] [Google Scholar]
  • 5.June CH, Ledbetter JA, Linsley PS, Thompson CB. Role of the CD28 receptor in T-cell activation. Immunol Today. 1990;11:211–6. doi: 10.1016/0167-5699(90)90085-n. [DOI] [PubMed] [Google Scholar]
  • 6.Dong C, Juedes AE, Temann UA, Shresta S, Allison JP, Ruddle NH, et al. ICOS co-stimulatory receptor is essential for T-cell activation and function. Nature. 2001;409:97–101. doi: 10.1038/35051100. [DOI] [PubMed] [Google Scholar]
  • 7.Campbell DJ, Butcher EC. Rapid acquisition of tissue-specific homing phenotypes by CD4(+) T cells activated in cutaneous or mucosal lymphoid tissues. J Exp Med. 2002;195:135–41. doi: 10.1084/jem.20011502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Masopust D, Vezys V, Marzo AL, Lefrancois L. Preferential localization of effector memory cells in nonlymphoid tissue. Science. 2001;291:2413–7. doi: 10.1126/science.1058867. [DOI] [PubMed] [Google Scholar]
  • 9.St John T, Meyer J, Idzerda R, Gallatin WM. Expression of CD44 confers a new adhesive phenotype on transfected cells. Cell. 1990;60:45–52. doi: 10.1016/0092-8674(90)90714-p. [DOI] [PubMed] [Google Scholar]
  • 10.Jalkanen S, Reichert RA, Gallatin WM, Bargatze RF, Weissman IL, Butcher EC. Homing receptors and the control of lymphocyte migration. Immunol Rev. 1986;91:39–60. doi: 10.1111/j.1600-065x.1986.tb01483.x. [DOI] [PubMed] [Google Scholar]
  • 11.Berlin C, Berg EL, Briskin MJ, Andrew DP, Kilshaw PJ, Holzmann B, et al. Alpha 4 beta 7 integrin mediates lymphocyte binding to the mucosal vascular addressin MAdCAM-1. Cell. 1993;74:185–95. doi: 10.1016/0092-8674(93)90305-a. [DOI] [PubMed] [Google Scholar]
  • 12.Hamann A, Andrew DP, Jablonski-Westrich D, Holzmann B, Butcher EC. Role of alpha 4-integrins in lymphocyte homing to mucosal tissues in vivo. J Immunol. 1994;152:3282–93. [PubMed] [Google Scholar]
  • 13.Campbell JJ, Haraldsen G, Pan J, Rottman J, Qin S, Ponath P, et al. The chemokine receptor CCR4 in vascular recognition by cutaneous but not intestinal memory T cells. Nature. 1999;400:776–80. doi: 10.1038/23495. [DOI] [PubMed] [Google Scholar]
  • 14.Homey B, Alenius H, Muller A, Soto H, Bowman EP, Yuan W, et al. CCL27-CCR10 interactions regulate T cell-mediated skin inflammation. Nat Med. 2002;8:157–65. doi: 10.1038/nm0202-157. [DOI] [PubMed] [Google Scholar]
  • 15.Morales J, Homey B, Vicari AP, Hudak S, Oldham E, Hedrick J, et al. CTACK, a skin-associated chemokine that preferentially attracts skin-homing memory T cells. Proc Natl Acad Sci U S A. 1999;96:14470–5. doi: 10.1073/pnas.96.25.14470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Reiss Y, Proudfoot AE, Power CA, Campbell JJ, Butcher EC. CC chemokine receptor (CCR)4 and the CCR10 ligand cutaneous T cell-attracting chemokine (CTACK) in lymphocyte trafficking to inflamed skin. J Exp Med. 2001;194:1541–7. doi: 10.1084/jem.194.10.1541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Sixt M, Bauer M, Lammermann T, Fassler R. Beta1 integrins: zip codes and signaling relay for blood cells. Curr Opin Cell Biol. 2006;18:482–90. doi: 10.1016/j.ceb.2006.08.007. [DOI] [PubMed] [Google Scholar]
  • 18.Williams IR. Chemokine receptors and leukocyte trafficking in the mucosal immune system. Immunol Res. 2004;29:283–92. doi: 10.1385/IR:29:1-3:283. [DOI] [PubMed] [Google Scholar]
  • 19.Sallusto F, Geginat J, Lanzavecchia A. Central memory and effector memory T cell subsets: function, generation, and maintenance. Annu Rev Immunol. 2004;22:745–63. doi: 10.1146/annurev.immunol.22.012703.104702. [DOI] [PubMed] [Google Scholar]
  • 20.Mosmann TR, Cherwinski H, Bond MW, Giedlin MA, Coffman RL. Two types of murine helper T cell clone. I. Definition according to profiles of lymphokine activities and secreted proteins. J Immunol. 1986;136:2348–57. [PubMed] [Google Scholar]
  • 21.Szabo SJ, Kim ST, Costa GL, Zhang X, Fathman CG, Glimcher LH. A novel transcription factor, T-bet, directs Th1 lineage commitment. Cell. 2000;100:655–69. doi: 10.1016/s0092-8674(00)80702-3. [DOI] [PubMed] [Google Scholar]
  • 22.Hsieh CS, Macatonia SE, Tripp CS, Wolf SF, O'Garra A, Murphy KM. Development of TH1 CD4+ T cells through IL-12 produced by Listeria-induced macrophages. Science. 1993;260:547–9. doi: 10.1126/science.8097338. [DOI] [PubMed] [Google Scholar]
  • 23.Afkarian M, Sedy JR, Yang J, Jacobson NG, Cereb N, Yang SY, et al. T-bet is a STAT1-induced regulator of IL-12R expression in naive CD4+ T cells. Nat Immunol. 2002;3:549–57. doi: 10.1038/ni794. [DOI] [PubMed] [Google Scholar]
  • 24.Jacobson NG, Szabo SJ, Weber-Nordt RM, Zhong Z, Schreiber RD, Darnell JE, et al. Interleukin 12 signaling in T helper type 1 (Th1) cells involves tyrosine phosphorylation of signal transducer and activator of transcription (Stat)3 and Stat4. J Exp Med. 1995;181:1755–62. doi: 10.1084/jem.181.5.1755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Abbas AK, Murphy KM, Sher A. Functional diversity of helper T lymphocytes. Nature. 1996;383:787–93. doi: 10.1038/383787a0. [DOI] [PubMed] [Google Scholar]
  • 26.Snapper CM, Paul WE. Interferon-gamma and B cell stimulatory factor-1 reciprocally regulate Ig isotype production. Science. 1987;236:944–7. doi: 10.1126/science.3107127. [DOI] [PubMed] [Google Scholar]
  • 27.Mosmann TR, Coffman RL. TH1 and TH2 cells: different patterns of lymphokine secretion lead to different functional properties. Annu Rev Immunol. 1989;7:145–73. doi: 10.1146/annurev.iy.07.040189.001045. [DOI] [PubMed] [Google Scholar]
  • 28.Zheng W, Flavell RA. The transcription factor GATA-3 is necessary and sufficient for Th2 cytokine gene expression in CD4 T cells. Cell. 1997;89:587–96. doi: 10.1016/s0092-8674(00)80240-8. [DOI] [PubMed] [Google Scholar]
  • 29.Sokol CL, Barton GM, Farr AG, Medzhitov R. A mechanism for the initiation of allergen-induced T helper type 2 responses. Nat Immunol. 2007;9:310–18. doi: 10.1038/ni1558. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Kaplan MH, Grusby MJ. Regulation of T helper cell differentiation by STAT molecules. J Leukoc Biol. 1998;64:2–5. doi: 10.1002/jlb.64.1.2. [DOI] [PubMed] [Google Scholar]
  • 31.Ouyang W, Ranganath SH, Weindel K, et al. Inhibition of Th1 development mediated by GATA-3 through an IL-4-independent mechanism. Immunity. 1998;9:745–55. doi: 10.1016/s1074-7613(00)80671-8. [DOI] [PubMed] [Google Scholar]
  • 32.Harrington LE, Mangan PR, Weaver CT. Expanding the effector CD4 T-cell repertoire: the Th17 lineage. Curr Opin Immunol. 2006;18:349–56. doi: 10.1016/j.coi.2006.03.017. [DOI] [PubMed] [Google Scholar]
  • 33.Rouvier E, Luciani MF, Mattei MG, Denizot F, Golstein P. CTLA-8, cloned from an activated T cell, bearing AU-rich messenger RNA instability sequences, and homologous to a herpesvirus saimiri gene. J Immunol. 1993;150:5445–56. [PubMed] [Google Scholar]
  • 34.Yao Z, Timour M, Painter S, Fanslow W, Spriggs M. Complete nucleotide sequence of the mouse CTLA8 gene. Gene. 1996;168:223–5. doi: 10.1016/0378-1119(95)00778-4. [DOI] [PubMed] [Google Scholar]
  • 35.Yao Z, Fanslow WC, Seldin MF, Rousseau AM, Painter SL, Comeau MR, et al. Herpesvirus Saimiri encodes a new cytokine, IL-17, which binds to a novel cytokine receptor. Immunity. 1995;3:811–21. doi: 10.1016/1074-7613(95)90070-5. [DOI] [PubMed] [Google Scholar]
  • 36.Kolls JK, Linden A. Interleukin-17 family members and inflammation. Immunity. 2004;21:467–76. doi: 10.1016/j.immuni.2004.08.018. [DOI] [PubMed] [Google Scholar]
  • 37.Lockhart E, Green AM, Flynn JL. IL-17 production is dominated by gammadelta T cells rather than CD4 T cells during Mycobacterium tuberculosis infection. J Immunol. 2006;177:4662–9. doi: 10.4049/jimmunol.177.7.4662. [DOI] [PubMed] [Google Scholar]
  • 38.Ferretti S, Bonneau O, Dubois GR, Jones CE, Trifilieff A. IL-17, produced by lymphocytes and neutrophils, is necessary for lipopolysaccharide-induced airway neutrophilia: IL-15 as a possible trigger. J Immunol. 2003;170:2106–12. doi: 10.4049/jimmunol.170.4.2106. [DOI] [PubMed] [Google Scholar]
  • 39.Zhou Q, Desta T, Fenton M, Graves DT, Amar S. Cytokine profiling of macrophages exposed to Porphyromonas gingivalis, its lipopolysaccharide, or its FimA protein. Infect Immun. 2005;73:935–43. doi: 10.1128/IAI.73.2.935-943.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Huang W, Na L, Fidel PL, Schwarzenberger P. Requirement of interleukin-17A for systemic anti-Candida albicans host defense in mice. J Infect Dis. 2004;190:624–31. doi: 10.1086/422329. [DOI] [PubMed] [Google Scholar]
  • 41.Ye P, Garvey PB, Zhang P, et al. Interleukin-17 and lung host defense against Klebsiella pneumoniae infection. Am J Respir Cell Mol Biol. 2001;25:335–40. doi: 10.1165/ajrcmb.25.3.4424. [DOI] [PubMed] [Google Scholar]
  • 42.Shibata K, Yamada H, Hara H, Kishihara K, Yoshikai Y. Resident Vdelta1+ gammadelta T cells control early infiltration of neutrophils after Escherichia coli infection via IL-17 production. J Immunol. 2007;178:4466–72. doi: 10.4049/jimmunol.178.7.4466. [DOI] [PubMed] [Google Scholar]
  • 43.Wu Q, Martin RJ, Rino JG, Breed R, Torres RM, Chu HW. IL-23-dependent IL-17 production is essential in neutrophil recruitment and activity in mouse lung defense against respiratory Mycoplasma pneumoniae infection. Microbes Infect. 2007;9:78–86. doi: 10.1016/j.micinf.2006.10.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Awane M, Andres PG, Li DJ, Reinecker HC. NF-kappa B-inducing kinase is a common mediator of IL-17-, TNF-alpha-, and IL-1 beta-induced chemokine promoter activation in intestinal epithelial cells. J Immunol. 1999;162:5337–44. [PubMed] [Google Scholar]
  • 45.Moseley TA, Haudenschild DR, Rose L, Reddi AH. Interleukin-17 family and IL-17 receptors. Cytokine Growth Factor Rev. 2003;14:155–74. doi: 10.1016/s1359-6101(03)00002-9. [DOI] [PubMed] [Google Scholar]
  • 46.Shen F, Ruddy MJ, Plamondon P, Gaffen SL. Cytokines link osteoblasts and inflammation: microarray analysis of interleukin-17- and TNF-alpha-induced genes in bone cells. J Leukoc Biol. 2005;77:388–99. doi: 10.1189/jlb.0904490. [DOI] [PubMed] [Google Scholar]
  • 47.Ye P, Rodriguez FH, Kanaly S, Stocking KL, Schurr J, Schwarzenberger P, et al. Requirement of interleukin 17 receptor signaling for lung CXC chemokine and granulocyte colony-stimulating factor expression, neutrophil recruitment, and host defense. J Exp Med. 2001;194:519–27. doi: 10.1084/jem.194.4.519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Fossiez F, Djossou O, Chomarat P, Flores-Romo L, Ait-Yahia S, Maat C, et al. T cell interleukin-17 induces stromal cells to produce proinflammatory and hematopoietic cytokines. J Exp Med. 1996;183:2593–603. doi: 10.1084/jem.183.6.2593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.LeGrand A, Fermor B, Fink C, Pisetsky DS, Weinberg JB, Vail TP, et al. Interleukin-1, tumor necrosis factor alpha, and interleukin-17 synergistically up-regulate nitric oxide and prostaglandin E2 production in explants of human osteoarthritic knee menisci. Arthritis Rheum. 2001;44:2078–83. doi: 10.1002/1529-0131(200109)44:9<2078::AID-ART358>3.0.CO;2-J. [DOI] [PubMed] [Google Scholar]
  • 50.Koenders MI, Lubberts E, Oppers-Walgreen B, van den Bersselaar L, Helsen MM, Di Padova FE, et al. Blocking of interleukin-17 during reactivation of experimental arthritis prevents joint inflammation and bone erosion by decreasing RANKL and interleukin-1. Am J Pathol. 2005;167:141–9. doi: 10.1016/S0002-9440(10)62961-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.van Beelen AJ, Teunissen MB, Kapsenberg ML, de Jong EC. Interleukin-17 in inflammatory skin disorders. Curr Opin Allergy Clin Immunol. 2007;7:374–81. doi: 10.1097/ACI.0b013e3282ef869e. [DOI] [PubMed] [Google Scholar]
  • 52.Fujino S, Andoh A, Bamba S, Ogawa A, Hata K, Araki Y, et al. Increased expression of interleukin 17 in inflammatory bowel disease. Gut. 2003;52:65–70. doi: 10.1136/gut.52.1.65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Hellings PW, Kasran A, Liu Z, Vandekerckhove P, Wuyts A, Overbergh L, et al. Interleukin-17 orchestrates the granulocyte influx into airways after allergen inhalation in a mouse model of allergic asthma. Am J Respir Cell Mol Biol. 2003;28:42–50. doi: 10.1165/rcmb.4832. [DOI] [PubMed] [Google Scholar]
  • 54.Lock C, Hermans G, Pedotti R, Pedotti R, Brendolan A, Schadt E, et al. Gene-microarray analysis of multiple sclerosis lesions yields new targets validated in autoimmune encephalomyelitis. Nat Med. 2002;8:500–8. doi: 10.1038/nm0502-500. [DOI] [PubMed] [Google Scholar]
  • 55.Langrish CL, Chen Y, Blumenschein WM, Mattson J, Basham B, Sedgwick JD, et al. IL-23 drives a pathogenic T cell population that induces autoimmune inflammation. J Exp Med. 2005;201:233–40. doi: 10.1084/jem.20041257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Yao Z, Spriggs MK, Derry JM, Strockbine L, Park LS, VandenBos T, et al. Molecular characterization of the human interleukin (IL)-17 receptor. Cytokine. 1997;9:794–800. doi: 10.1006/cyto.1997.0240. [DOI] [PubMed] [Google Scholar]
  • 57.Weaver CT, Hatton RD, Mangan PR, Harrington LE. IL-17 family cytokines and the expanding diversity of effector T cell lineages. Annu Rev Immunol. 2007;25:821–52. doi: 10.1146/annurev.immunol.25.022106.141557. [DOI] [PubMed] [Google Scholar]
  • 58.Haudenschild D, Moseley T, Rose L, Reddi AH. Soluble and transmembrane isoforms of novel interleukin-17 receptor-like protein by RNA splicing and expression in prostate cancer. J Biol Chem. 2002;277:4309–16. doi: 10.1074/jbc.M109372200. [DOI] [PubMed] [Google Scholar]
  • 59.Tian E, Sawyer JR, Largaespada DA, Jenkins NA, Copeland NG, Shaughnessy JD., Jr Evi27 encodes a novel membrane protein with homology to the IL17 receptor. Oncogene. 2000;19:2098–109. doi: 10.1038/sj.onc.1203577. [DOI] [PubMed] [Google Scholar]
  • 60.Hymowitz SG, Filvaroff EH, Yin JP, et al. IL-17s adopt a cystine knot fold: structure and activity of a novel cytokine, IL-17F, and implications for receptor binding. Embo J. 2001;20:5332–41. doi: 10.1093/emboj/20.19.5332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Kuestner RE, Taft DW, Haran A, Brandt CS, Brender T, Lum K, et al. Identification of the IL-17 receptor related molecule IL-17RC as the receptor for IL-17F. J Immunol. 2007;179:5462–73. doi: 10.4049/jimmunol.179.8.5462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Kramer JM, Yi L, Shen F, Maitra A, Jiao X, Jin T, et al. Evidence for ligand-independent multimerization of the IL-17 receptor. J Immunol. 2006;176:711–5. doi: 10.4049/jimmunol.176.2.711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Toy D, Kugler D, Wolfson M, Vanden Bos T, Gurgel J, Derry J, et al. Cutting edge: interleukin 17 signals through a heteromeric receptor complex. J Immunol. 2006;177:36–9. doi: 10.4049/jimmunol.177.1.36. [DOI] [PubMed] [Google Scholar]
  • 64.Rickel EA, Siegel LA, Yoon BR, Rottman JB, Kugler DG, Swart DA, et al. Identification of functional roles for both IL-17RB and IL-17RA in mediating IL-25-induced activities. J Immunol. 2008;181:4299–310. doi: 10.4049/jimmunol.181.6.4299. [DOI] [PubMed] [Google Scholar]
  • 65.Rock FL, Hardiman G, Timans JC, Kastelein RA, Bazan JF. A family of human receptors structurally related to Drosophila Toll. Proc Natl Acad Sci U S A. 1998;95:588–93. doi: 10.1073/pnas.95.2.588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Novatchkova M, Leibbrandt A, Werzowa J, Neubuser A, Eisenhaber F. The STIR-domain superfamily in signal transduction, development and immunity. Trends Biochem Sci. 2003;28:226–9. doi: 10.1016/S0968-0004(03)00067-7. [DOI] [PubMed] [Google Scholar]
  • 67.Qian Y, Liu C, Hartupee J, Altuntas CZ, Gulen MF, Jane-Wit D, et al. The adaptor Act1 is required for interleukin 17-dependent signaling associated with autoimmune and inflammatory disease. Nat Immunol. 2007;8:247–56. doi: 10.1038/ni1439. [DOI] [PubMed] [Google Scholar]
  • 68.Rong Z, Cheng L, Ren Y, Li Z, Li Y, Li H, et al. Interleukin-17F signaling requires ubiquitination of interleukin-17 receptor via TRAF6. Cell Signal. 2007;19:1514–20. doi: 10.1016/j.cellsig.2007.01.025. [DOI] [PubMed] [Google Scholar]
  • 69.Chang SH, Park H, Dong C. Act1 adaptor protein is an immediate and essential signaling component of interleukin-17 receptor. J Biol Chem. 2006;281:35603–7. doi: 10.1074/jbc.C600256200. [DOI] [PubMed] [Google Scholar]
  • 70.Maitra A, Shen F, Hanel W, Mossman K, Tocker J, Swart D, et al. Distinct functional motifs within the IL-17 receptor regulate signal transduction and target gene expression. Proc Natl Acad Sci U S A. 2007;104:7506–11. doi: 10.1073/pnas.0611589104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Schwandner R, Yamaguchi K, Cao Z. Requirement of tumor necrosis factor receptor-associated factor (TRAF)6 in interleukin 17 signal transduction. J Exp Med. 2000;191:1233–40. doi: 10.1084/jem.191.7.1233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Fernandez-Botran R, Sanders VM, Mosmann TR, Vitetta ES. Lymphokine-mediated regulation of the proliferative response of clones of T helper 1 and T helper 2 cells. J Exp Med. 1988;168:543–58. doi: 10.1084/jem.168.2.543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Gajewski TF, Fitch FW. Anti-proliferative effect of IFN-gamma in immune regulation. I. IFNgamma inhibits the proliferation of Th2 but not Th1 murine helper T lymphocyte clones. J Immunol. 1988;140:4245–52. [PubMed] [Google Scholar]
  • 74.Gajewski TF, Goldwasser E, Fitch FW. Anti-proliferative effect of IFN-gamma in immune regulation. II. IFN-gamma inhibits the proliferation of murine bone marrow cells stimulated with IL-3, IL-4, or granulocyte-macrophage colony-stimulating factor. J Immunol. 1988;141:2635–42. [PubMed] [Google Scholar]
  • 75.Cher DJ, Mosmann TR. Two types of murine helper T cell clone. II. Delayed-type hypersensitivity is mediated by TH1 clones. J Immunol. 1987;138:3688–94. [PubMed] [Google Scholar]
  • 76.Coffman RL, Carty J. A T cell activity that enhances polyclonal IgE production and its inhibition by interferon-gamma. J Immunol. 1986;136:949–54. [PubMed] [Google Scholar]
  • 77.McKenzie BS, Kastelein RA, Cua DJ. Understanding the IL-23-IL-17 immune pathway. Trends Immunol. 2006;27:17–23. doi: 10.1016/j.it.2005.10.003. [DOI] [PubMed] [Google Scholar]
  • 78.Kobayashi M, Fitz L, Ryan M, Ryan M, Hewick RM, Clark SC, et al. Identification and purification of natural killer cell stimulatory factor (NKSF), a cytokine with multiple biologic effects on human lymphocytes. J Exp Med. 1989;170:827–45. doi: 10.1084/jem.170.3.827. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Stern AS, Podlaski FJ, Hulmes JD, Pan YC, Quinn PM, Wolitzky AG, et al. Purification to homogeneity and partial characterization of cytotoxic lymphocyte maturation factor from human B-lymphoblastoid cells. Proc Natl Acad Sci U S A. 1990;87:6808–12. doi: 10.1073/pnas.87.17.6808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Presky DH, Yang H, Minetti LJ, Chua AO, Nabavi N, Wu CY, et al. A functional interleukin 12 receptor complex is composed of two beta-type cytokine receptor subunits. Proc Natl Acad Sci U S A. 1996;93:14002–7. doi: 10.1073/pnas.93.24.14002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Wu C, Wang X, Gadina M, O'Shea JJ, Presky DH, Magram J. IL-12 receptor beta 2 (IL-12R beta 2)-deficient mice are defective in IL-12-mediated signaling despite the presence of high affinity IL-12 binding sites. J Immunol. 2000;165:6221–8. doi: 10.4049/jimmunol.165.11.6221. [DOI] [PubMed] [Google Scholar]
  • 82.Magram J, Connaughton SE, Warrier RR, Carvajal DM, Wu CY, Ferrante J, et al. IL-12-deficient mice are defective in IFN gamma production and type 1 cytokine responses. Immunity. 1996;4:471–81. doi: 10.1016/s1074-7613(00)80413-6. [DOI] [PubMed] [Google Scholar]
  • 83.Mattner F, Magram J, Ferrante J, Launois P, Di Padova K, Behin R, et al. Genetically resistant mice lacking interleukin-12 are susceptible to infection with Leishmania major and mount a polarized Th2 cell response. Eur J Immunol. 1996;26:1553–9. doi: 10.1002/eji.1830260722. [DOI] [PubMed] [Google Scholar]
  • 84.Park AY, Hondowicz BD, Scott P. IL-12 is required to maintain a Th1 response during Leishmania major infection. J Immunol. 2000;165:896–902. doi: 10.4049/jimmunol.165.2.896. [DOI] [PubMed] [Google Scholar]
  • 85.Leonard JP, Waldburger KE, Goldman SJ. Prevention of experimental autoimmune encephalomyelitis by antibodies against interleukin 12. J Exp Med. 1995;181:381–6. doi: 10.1084/jem.181.1.381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Constantinescu CS, Wysocka M, Hilliard B, Ventura ES, Lavi E, Trinchieri G, et al. Antibodies against IL-12 prevent superantigen-induced and spontaneous relapses of experimental autoimmune encephalomyelitis. J Immunol. 1998;161:5097–104. [PubMed] [Google Scholar]
  • 87.Segal BM, Dwyer BK, Shevach EM. An interleukin (IL)-10/IL-12 immunoregulatory circuit controls susceptibility to autoimmune disease. J Exp Med. 1998;187:537–46. doi: 10.1084/jem.187.4.537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Ferber IA, Brocke S, Taylor-Edwards C, et al. Mice with a disrupted IFN-gamma gene are susceptible to the induction of experimental autoimmune encephalomyelitis (EAE) J Immunol. 1996;156:5–7. [PubMed] [Google Scholar]
  • 89.Cua DJ, Sherlock J, Chen Y, Murphy CA, Joyce B, Seymour B, et al. Interleukin-23 rather than interleukin-12 is the critical cytokine for autoimmune inflammation of the brain. Nature. 2003;421:744–8. doi: 10.1038/nature01355. [DOI] [PubMed] [Google Scholar]
  • 90.Oppmann B, Lesley R, Blom B, Timans JC, Xu Y, Hunte B, et al. Novel p19 protein engages IL-12p40 to form a cytokine, IL-23, with biological activities similar as well as distinct from IL-12. Immunity. 2000;13:715–25. doi: 10.1016/s1074-7613(00)00070-4. [DOI] [PubMed] [Google Scholar]
  • 91.Aggarwal S, Ghilardi N, Xie MH, de Sauvage FJ, Gurney AL. Interleukin-23 promotes a distinct CD4 T cell activation state characterized by the production of interleukin-17. J Biol Chem. 2003;278:1910–4. doi: 10.1074/jbc.M207577200. [DOI] [PubMed] [Google Scholar]
  • 92.Parham C, Chirica M, Timans J, Vaisberg E, Travis M, Cheung J, et al. A receptor for the heterodimeric cytokine IL-23 is composed of IL-12Rbeta1 and a novel cytokine receptor subunit, IL-23R. J Immunol. 2002;168:5699–708. doi: 10.4049/jimmunol.168.11.5699. [DOI] [PubMed] [Google Scholar]
  • 93.Bettelli E, Carrier Y, Gao W, Korn T, Strom TB, Oukka M, et al. Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature. 2006;441:235–8. doi: 10.1038/nature04753. [DOI] [PubMed] [Google Scholar]
  • 94.Mangan PR, Harrington LE, O'Quinn DB, Helms WS, Bullard DC, Elson CO, et al. Transforming growth factor-beta induces development of the T(H)17 lineage. Nature. 2006;441:231–4. doi: 10.1038/nature04754. [DOI] [PubMed] [Google Scholar]
  • 95.Veldhoen M, Hocking RJ, Atkins CJ, Locksley RM, Stockinger B. TGFbeta in the context of an inflammatory cytokine milieu supports de novo differentiation of IL-17-producing T cells. Immunity. 2006;24:179–89. doi: 10.1016/j.immuni.2006.01.001. [DOI] [PubMed] [Google Scholar]
  • 96.Li MO, Wan YY, Flavell RA. T cell-produced transforming growth factor-beta1 controls T cell tolerance and regulates Th1- and Th17-cell differentiation. Immunity. 2007;26:579–91. doi: 10.1016/j.immuni.2007.03.014. [DOI] [PubMed] [Google Scholar]
  • 97.Veldhoen M, Hocking RJ, Flavell RA, Stockinger B. Signals mediated by transforming growth factor-beta initiate autoimmune encephalomyelitis, but chronic inflammation is needed to sustain disease. Nat Immunol. 2006;7:1151–6. doi: 10.1038/ni1391. [DOI] [PubMed] [Google Scholar]
  • 98.Ivanov II, McKenzie BS, Zhou L, Tadokoro CE, Lepelley A, Lafaille JJ, et al. The orphan nuclear receptor RORgammat directs the differentiation program of proinflammatory IL-17+ T helper cells. Cell. 2006;126:1121–33. doi: 10.1016/j.cell.2006.07.035. [DOI] [PubMed] [Google Scholar]
  • 99.Yang XO, Pappu BP, Nurieva R, Akimzhanov A, Kang HS, Chung Y, et al. T Helper 17 Lineage Differentiation Is Programmed by Orphan Nuclear Receptors RORalpha and RORgamma. Immunity. 2007 doi: 10.1016/j.immuni.2007.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Zhang F, Meng G, Strober W. Interactions among the transcription factors Runx1, RORgammat and Foxp3 regulate the differentiation of interleukin 17-producing T cells. Nat Immunol. 2008 doi: 10.1038/ni.1663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Brustle A, Heink S, Huber M, Rosenplanter C, Stadelmann C, Yu P, et al. The development of inflammatory T(H)-17 cells requires interferon-regulatory factor 4. Nat Immunol. 2007;8:958–66. doi: 10.1038/ni1500. [DOI] [PubMed] [Google Scholar]
  • 102.Yang XO, Panopoulos AD, Nurieva R, Chang SH, Wang D, Watowich SS, et al. STAT3 regulates cytokine-mediated generation of inflammatory helper T cells. J Biol Chem. 2007;282:9358–63. doi: 10.1074/jbc.C600321200. [DOI] [PubMed] [Google Scholar]
  • 103.Korn T, Bettelli E, Gao W, Awasthi A, Jager A, Strom TB, et al. IL-21 initiates an alternative pathway to induce proinflammatory T(H)17 cells. Nature. 2007;448:484–7. doi: 10.1038/nature05970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Nurieva R, Yang XO, Martinez G, Zhang Y, Panopoulos AD, Ma L, et al. Essential autocrine regulation by IL-21 in the generation of inflammatory T cells. Nature. 2007;448:480–3. doi: 10.1038/nature05969. [DOI] [PubMed] [Google Scholar]
  • 105.Zhou L, Ivanov II, Spolski R, Min R, Shenderov K, Egawa T, et al. IL-6 programs T(H)-17 cell differentiation by promoting sequential engagement of the IL-21 and IL-23 pathways. Nat Immunol. 2007;8:967–74. doi: 10.1038/ni1488. [DOI] [PubMed] [Google Scholar]
  • 106.Chen Z, Laurence A, Kanno Y, Pacher-Zavisin M, Zhu BM, Tato C, et al. Selective regulatory function of Socs3 in the formation of IL-17-secreting T cells. Proc Natl Acad Sci U S A. 2006;103:8137–42. doi: 10.1073/pnas.0600666103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Park H, Li Z, Yang XO, Chang SH, Nurieva R, Wang YH, et al. A distinct lineage of CD4 T cells regulates tissue inflammation by producing interleukin 17. Nat Immunol. 2005;6:1133–41. doi: 10.1038/ni1261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Cruz A, Khader SA, Torrado E, Fraga A, Pearl JE, Pedrosa J, et al. Cutting edge: IFN-gamma regulates the induction and expansion of IL-17-producing CD4 T cells during mycobacterial infection. J Immunol. 2006;177:1416–20. doi: 10.4049/jimmunol.177.3.1416. [DOI] [PubMed] [Google Scholar]
  • 109.Rangachari M, Mauermann N, Marty RR, Dirnhofer S, Kurrer MO, Komnenovic V, et al. T-bet negatively regulates autoimmune myocarditis by suppressing local production of interleukin 17. J Exp Med. 2006;203:2009–19. doi: 10.1084/jem.20052222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Batten M, Li J, Yi S, Kljavin NM, Danilenko DM, Lucas S, et al. Interleukin 27 limits autoimmune encephalomyelitis by suppressing the development of interleukin 17-producing T cells. Nat Immunol. 2006;7:929–36. doi: 10.1038/ni1375. [DOI] [PubMed] [Google Scholar]
  • 111.Zaph C, Du Y, Saenz SA, Nair MG, Perrigoue JG, Taylor BC, et al. Commensal-dependent expression of IL-25 regulates the IL-23-IL-17 axis in the intestine. J Exp Med. 2008;205:2191–8. doi: 10.1084/jem.20080720. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Weaver CT, Harrington LE, Mangan PR, Gavrieli M, Murphy KM. Th17: an effector CD4 T cell lineage with regulatory T cell ties. Immunity. 2006;24:677–88. doi: 10.1016/j.immuni.2006.06.002. [DOI] [PubMed] [Google Scholar]
  • 113.Setoguchi R, Hori S, Takahashi T, Sakaguchi S. Homeostatic maintenance of natural Foxp3(+) CD25(+) CD4(+) regulatory T cells by interleukin (IL)-2 and induction of autoimmune disease by IL-2 neutralization. J Exp Med. 2005;201:723–35. doi: 10.1084/jem.20041982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Laurence A, Tato CM, Davidson TS, Kanno Y, Chen Z, Yao Z, et al. Interleukin-2 signaling via STAT5 constrains T helper 17 cell generation. Immunity. 2007;26:371–81. doi: 10.1016/j.immuni.2007.02.009. [DOI] [PubMed] [Google Scholar]
  • 115.Kryczek I, Wei S, Vatan L, Escara-Wilke J, Szeliga W, Keller ET, et al. Cutting edge: opposite effects of IL-1 and IL-2 on the regulation of IL-17+ T cell pool IL-1 subverts IL-2-mediated suppression. J Immunol. 2007;179:1423–6. doi: 10.4049/jimmunol.179.3.1423. [DOI] [PubMed] [Google Scholar]
  • 116.Xu L, Kitani A, Fuss I, Strober W. Cutting edge: regulatory T cells induce CD4+CD25-Foxp3- T cells or are self-induced to become Th17 cells in the absence of exogenous TGF-beta. J Immunol. 2007;178:6725–9. doi: 10.4049/jimmunol.178.11.6725. [DOI] [PubMed] [Google Scholar]
  • 117.Mucida D, Park Y, Kim G, Turovskaya O, Scott I, Kronenberg M, et al. Reciprocal TH17 and regulatory T cell differentiation mediated by retinoic acid. Science. 2007;317:256–60. doi: 10.1126/science.1145697. [DOI] [PubMed] [Google Scholar]
  • 118.Zhou L, Lopes JE, Chong MM, Ivanov II, Min R, Victora GD, et al. TGF-beta-induced Foxp3 inhibits T(H)17 cell differentiation by antagonizing RORgammat function. Nature. 2008;453:236–40. doi: 10.1038/nature06878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Gavin MA, Rasmussen JP, Fontenot JD, Vasta V, Manganiello VC, Beavo JA, et al. Foxp3-dependent programme of regulatory T-cell differentiation. Nature. 2007;445:771–5. doi: 10.1038/nature05543. [DOI] [PubMed] [Google Scholar]
  • 120.Hizawa N, Kawaguchi M, Huang SK, Nishimura M. Role of interleukin-17F in chronic inflammatory and allergic lung disease. Clin Exp Allergy. 2006;36:1109–14. doi: 10.1111/j.1365-2222.2006.02550.x. [DOI] [PubMed] [Google Scholar]
  • 121.Oda N, Canelos PB, Essayan DM, Plunkett BA, Myers AC, Huang SK. Interleukin-17F induces pulmonary neutrophilia and amplifies antigen-induced allergic response. Am J Respir Crit Care Med. 2005;171:12–8. doi: 10.1164/rccm.200406-778OC. [DOI] [PubMed] [Google Scholar]
  • 122.Nakae S, Nambu A, Sudo K, Iwakura Y. Suppression of immune induction of collagen-induced arthritis in IL-17-deficient mice. J Immunol. 2003;171:6173–7. doi: 10.4049/jimmunol.171.11.6173. [DOI] [PubMed] [Google Scholar]
  • 123.Komiyama Y, Nakae S, Matsuki T, Nambu A, Ishigame H, Kakuta S, et al. IL-17 plays an important role in the development of experimental autoimmune encephalomyelitis. J Immunol. 2006;177:566–73. doi: 10.4049/jimmunol.177.1.566. [DOI] [PubMed] [Google Scholar]
  • 124.Nakae S, Komiyama Y, Nambu A, Sudo K, Iwase M, Homma I, et al. Antigen-specific T cell sensitization is impaired in IL-17-deficient mice, causing suppression of allergic cellular and humoral responses. Immunity. 2002;17:375–87. doi: 10.1016/s1074-7613(02)00391-6. [DOI] [PubMed] [Google Scholar]
  • 125.Yang XO, Chang SH, Park H, Nurieva R, Shah B, Acero L, et al. Regulation of inflammatory responses by IL-17F. J Exp Med. 2008;205:1063–75. doi: 10.1084/jem.20071978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Liang SC, Tan XY, Luxenberg DP, Karim R, Dunussi-Joannopoulos K, Collins M, et al. Interleukin (IL)-22 and IL-17 are coexpressed by Th17 cells and cooperatively enhance expression of antimicrobial peptides. J Exp Med. 2006;203:2271–9. doi: 10.1084/jem.20061308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Zheng Y, Danilenko DM, Valdez P, Kasman I, Eastham-Anderson J, Wu J, et al. Interleukin-22, a T(H)17 cytokine, mediates IL-23-induced dermal inflammation and acanthosis. Nature. 2007;445:648–51. doi: 10.1038/nature05505. [DOI] [PubMed] [Google Scholar]
  • 128.Wolk K, Kunz S, Witte E, Friedrich M, Asadullah K, Sabat R. IL-22 increases the innate immunity of tissues. Immunity. 2004;21:241–54. doi: 10.1016/j.immuni.2004.07.007. [DOI] [PubMed] [Google Scholar]
  • 129.Kotenko SV, Izotova LS, Mirochnitchenko OV, Esterova E, Dickensheets H, Donnelly RP, et al. Identification, cloning, and characterization of a novel soluble receptor that binds IL-22 and neutralizes its activity. J Immunol. 2001;166:7096–103. doi: 10.4049/jimmunol.166.12.7096. [DOI] [PubMed] [Google Scholar]
  • 130.Aggarwal S, Xie MH, Maruoka M, Foster J, Gurney AL. Acinar cells of the pancreas are a target of interleukin-22. J Interferon Cytokine Res. 2001;21:1047–53. doi: 10.1089/107999001317205178. [DOI] [PubMed] [Google Scholar]
  • 131.Zenewicz LA, Yancopoulos GD, Valenzuela DM, Murphy AJ, Karow M, Flavell RA. Interleukin-22 but not interleukin-17 provides protection to hepatocytes during acute liver inflammation. Immunity. 2007;27:647–59. doi: 10.1016/j.immuni.2007.07.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Happel KI, Dubin PJ, Zheng M, Ghilardi N, Lockhart C, Quinton LJ, et al. Divergent roles of IL-23 and IL-12 in host defense against Klebsiella pneumoniae. J Exp Med. 2005;202:761–9. doi: 10.1084/jem.20050193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Khader SA, Bell GK, Pearl JE, Fountain JJ, Rangel-Moreno J, Cilley GE, et al. IL-23 and IL-17 in the establishment of protective pulmonary CD4+ T cell responses after vaccination and during Mycobacterium tuberculosis challenge. Nat Immunol. 2007;8:369–77. doi: 10.1038/ni1449. [DOI] [PubMed] [Google Scholar]
  • 134.Umemura M, Yahagi A, Hamada S, Begum MD, Watanabe H, Kawakami K, et al. IL-17-mediated regulation of innate and acquired immune response against pulmonary Mycobacterium bovis bacille Calmette-Guerin infection. J Immunol. 2007;178:3786–96. doi: 10.4049/jimmunol.178.6.3786. [DOI] [PubMed] [Google Scholar]
  • 135.Wiehler S, Proud D. Interleukin-17A modulates human airway epithelial responses to human rhinovirus infection. Am J Physiol Lung Cell Mol Physiol. 2007;293:L505–15. doi: 10.1152/ajplung.00066.2007. [DOI] [PubMed] [Google Scholar]
  • 136.Rudner XL, Happel KI, Young EA, Shellito JE. Interleukin-23 (IL-23)-IL-17 cytokine axis in murine Pneumocystis carinii infection. Infect Immun. 2007;75:3055–61. doi: 10.1128/IAI.01329-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Molet S, Hamid Q, Davoine F, Nutku E, Taha R, Page N, et al. IL-17 is increased in asthmatic airways and induces human bronchial fibroblasts to produce cytokines. J Allergy Clin Immunol. 2001;108:430–8. doi: 10.1067/mai.2001.117929. [DOI] [PubMed] [Google Scholar]
  • 138.Schnyder-Candrian S, Togbe D, Couillin I, Mercier I, Brombacher F, Quesniaux V, et al. Interleukin-17 is a negative regulator of established allergic asthma. J Exp Med. 2006;203:2715–25. doi: 10.1084/jem.20061401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.McKinley L, Alcorn JF, Peterson A, Dupont RB, Kapadia S, Logar A, et al. TH17 cells mediate steroid-resistant airway inflammation and airway hyperresponsiveness in mice. J Immunol. 2008;181:4089–97. doi: 10.4049/jimmunol.181.6.4089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Rakoff-Nahoum S, Paglino J, Eslami-Varzaneh F, Edberg S, Medzhitov R. Recognition of commensal microflora by toll-like receptors is required for intestinal homeostasis. Cell. 2004;118:229–41. doi: 10.1016/j.cell.2004.07.002. [DOI] [PubMed] [Google Scholar]
  • 141.Hall JA, Bouladoux N, Sun CM, Wohlfert EA, Blank RB, Zhu Q, et al. Commensal DNA Limits Regulatory T Cell Conversion and Is a Natural Adjuvant of Intestinal Immune Responses. Immunity. 2008 doi: 10.1016/j.immuni.2008.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Neurath MF, Fuss I, Kelsall BL, Stuber E, Strober W. Antibodies to interleukin 12 abrogate established experimental colitis in mice. J Exp Med. 1995;182:1281–90. doi: 10.1084/jem.182.5.1281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Hue S, Ahern P, Buonocore S, Kullberg MC, Cua DJ, McKenzie BS, et al. Interleukin-23 drives innate and T cell-mediated intestinal inflammation. J Exp Med. 2006;203:2473–83. doi: 10.1084/jem.20061099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Kullberg MC, Jankovic D, Feng CG, Hue S, Gorelick PL, McKenzie BS, et al. IL-23 plays a key role in Helicobacter hepaticus-induced T cell-dependent colitis. J Exp Med. 2006;203:2485–94. doi: 10.1084/jem.20061082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Schmidt C, Giese T, Ludwig B, Mueller-Molaian I, Marth T, Zeuzem S, et al. Expression of interleukin-12-related cytokine transcripts in inflammatory bowel disease: elevated interleukin-23p19 and interleukin-27p28 in Crohn's disease but not in ulcerative colitis. Inflamm Bowel Dis. 2005;11:16–23. doi: 10.1097/00054725-200501000-00003. [DOI] [PubMed] [Google Scholar]
  • 146.Mannon PJ, Fuss IJ, Mayer L, Elson CO, Sandborn WJ, Present D, et al. Anti-interleukin-12 antibody for active Crohn's disease. N Engl J Med. 2004;351:2069–79. doi: 10.1056/NEJMoa033402. [DOI] [PubMed] [Google Scholar]
  • 147.Cho JH. The genetics and immunopathogenesis of inflammatory bowel disease. Nat Rev Immunol. 2008;8:458–66. doi: 10.1038/nri2340. [DOI] [PubMed] [Google Scholar]
  • 148.Eyerich K, Foerster S, Rombold S, Seidl HP, Behrendt H, Hofmann H, et al. Patients with chronic mucocutaneous candidiasis exhibit reduced production of Th17-associated cytokines IL-17 and IL-22. J Invest Dermatol. 2008;128:2640–5. doi: 10.1038/jid.2008.139. [DOI] [PubMed] [Google Scholar]
  • 149.Lee E, Trepicchio WL, Oestreicher JL, Pittman D, Wang F, Chamian F, et al. Increased expression of interleukin 23 p19 and p40 in lesional skin of patients with psoriasis vulgaris. J Exp Med. 2004;199:125–30. doi: 10.1084/jem.20030451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Lowes MA, Kikuchi T, Fuentes-Duculan J, Cardinale I, Zaba LC, Haider AS, et al. Psoriasis vulgaris lesions contain discrete populations of Th1 and Th17 T cells. J Invest Dermatol. 2008;128:1207–11. doi: 10.1038/sj.jid.5701213. [DOI] [PubMed] [Google Scholar]
  • 151.Itohara S, Farr AG, Lafaille JJ, Bonneville M, Takagaki Y, Haas W, et al. Homing of a gamma delta thymocyte subset with homogeneous T-cell receptors to mucosal epithelia. Nature. 1990;343:754–7. doi: 10.1038/343754a0. [DOI] [PubMed] [Google Scholar]
  • 152.Scotet E, Nedellec S, Devilder MC, Allain S, Bonneville M. Bridging innate and adaptive immunity through gammadelta T-dendritic cell crosstalk. Front Biosci. 2008;13:6872–85. doi: 10.2741/3195. [DOI] [PubMed] [Google Scholar]
  • 153.Crowley MP, Fahrer AM, Baumgarth N, Hampl J, Gutgemann I, Teyton L, et al. A population of murine gammadelta T cells that recognize an inducible MHC class Ib molecule. Science. 2000;287:314–6. doi: 10.1126/science.287.5451.314. [DOI] [PubMed] [Google Scholar]
  • 154.Wu J, Groh V, Spies T. T cell antigen receptor engagement and specificity in the recognition of stress-inducible MHC class I-related chains by human epithelial gamma delta T cells. J Immunol. 2002;169:1236–40. doi: 10.4049/jimmunol.169.3.1236. [DOI] [PubMed] [Google Scholar]
  • 155.Bauer S, Groh V, Wu J, Steinle A, Phillips JH, Lanier LL, et al. Activation of NK cells and T cells by NKG2D, a receptor for stress-inducible MICA. Science. 1999;285:727–9. doi: 10.1126/science.285.5428.727. [DOI] [PubMed] [Google Scholar]
  • 156.Pennington DJ, Vermijlen D, Wise EL, Clarke SL, Tigelaar RE, Hayday AC. The integration of conventional and unconventional T cells that characterizes cell-mediated responses. Adv Immunol. 2005;87:27–59. doi: 10.1016/S0065-2776(05)87002-6. [DOI] [PubMed] [Google Scholar]
  • 157.Stark MA, Huo Y, Burcin TL, Morris MA, Olson TS, Ley K. Phagocytosis of apoptotic neutrophils regulates granulopoiesis via IL-23 and IL-17. Immunity. 2005;22:285–94. doi: 10.1016/j.immuni.2005.01.011. [DOI] [PubMed] [Google Scholar]
  • 158.Roark CL, Simonian PL, Fontenot AP, Born WK, O'Brien RL. gammadelta T cells: an important source of IL-17. Curr Opin Immunol. 2008;20:353–7. doi: 10.1016/j.coi.2008.03.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Jensen KD, Su X, Shin S, Li L, Youssef S, Yamasaki S, et al. Thymic selection determines gammadelta T cell effector fate: antigen-naive cells make interleukin-17 and antigen-experienced cells make interferon gamma. Immunity. 2008;29:90–100. doi: 10.1016/j.immuni.2008.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES