Skip to main content
Howard Hughes Medical Institute Author Manuscripts logoLink to Howard Hughes Medical Institute Author Manuscripts
. Author manuscript; available in PMC: 2016 Dec 16.
Published in final edited form as: Curr Opin Immunol. 2014 May 27;29:69–78. doi: 10.1016/j.coi.2014.04.006

Complementary diversification of dendritic cells and innate lymphoid cells

Carlos G Briseño 1, Theresa L Murphy 1, Kenneth M Murphy 1,2,*
PMCID: PMC5161034  NIHMSID: NIHMS833974  PMID: 24874447

Abstract

Dendritic cells (DCs) are professional antigen presenting cells conventionally thought to mediate cellular adaptive immune responses. Recent studies have led to the recognition of a non-redundant role for DCs in orchestrating innate immune responses, and in particular, for DC subset-specific interactions with innate lymphoid cells (ILCs). Recently recognized as important effectors of early immune responses, ILCs develop into subsets which mirror the transcriptional and cytokine profile of their T cell subset counterparts. DC diversification into functional subsets provides for modules of pathogen sensing and cytokine production that direct pathogen-appropriate ILC and T cell responses. This review focuses on the recent advances in the understanding of DC development, and their function in orchestrating the innate immune modules.

Introduction

Dendritic cells (DCs) arise from the bone marrow (BM) and seed peripheral tissues and lymphoid organs [1]. Initially identified by their robust capacity for activating T cells, they are now recognized as modulators of both innate and adaptive responses [2]. Four different lineages can be classified as DCs: classical DCs (cDCs) [1], plasmacytoid DCs (pDCs) [3;4], monocyte derived DCs (moDCs) [57] and Langerhans cells [8]. cDCs are also heterogeneous, and can be divided into at least two major branches, but are likely more diverse [9]. In the mouse spleen, the two major cDC subpopulations are distinguished by expression of the surface markers CD8α and CD172a (Sirpα), whereas in peripheral tissues, patterns of other markers such as CD103 and CD11b can be used to identify the splenic cDC subset equivalents [10;11*]. This review focuses on recent progress in the development and functional diversity of cDC subsets, and in particular their modes of orchestrating innate immune responses.

DCs develop from distinct BM progenitors

The cellular stages involved in murine DC development, briefly described here, have been recently reviewed [12;13]. DCs originate in the BM from the common myeloid progenitor (CMP) [14]. The macrophage-DC progenitor (MDP) develops from the CMP and is restricted to cDC, pDC, macrophage and monocyte lineages [15;16]. Subsequently, a common DC progenitor (CDP) that expresses the macrophage colony-stimulating factor receptor (M-CSFR) and fms-related tyrosine kinase 3 (Flt3), but lower levels of stem cell factor (c-Kit) as compared to the MDP, has only pDC and cDC potential [17;18]. The pre-cDC, which develops from the CDP, exits from the BM and seeds lymphoid and peripheral tissues, where it differentiates into both cDC subsets [17;19]. Unlike cDCs, pDCs develop fully in the BM. Recently, segregation of DC progenitors based on M-CSFR expression has identified an M-CSFR-negative population that preferentially gives rise to pDCs [20*]. Immature pDCs in the BM can be identified as CD11c+B220+CCR9Ly49Q; upon maturation, pDCs acquire expression of CCR9 and Ly49Q and egress from the BM [21]. However, these pre-cDC and pDC progenitors, as currently defined, do not entirely exclude cDC and pDC potential, respectively [19;20*;22]. The splenic pre-cDC retains approximately 50% of the pDC potential relative to unfractionated BM [19], while the CCR9 pDC progenitors can differentiate into CD11b+ cDCs upon granulocyte macrophage colony-stimulating factor (GM-CSF) stimulation [22]. Thus, further refinements in the definitions of these committed progenitors may be required before a mechanistic pathway of their development can emerge.

Cytokine regulation of DC homeostasis

DC development is dependent on the growth receptor Flt3 and its downstream transcription factor STAT3 [2325]. Flt3 expression is maintained on DC progenitors and mature cDCs [26*], and loss of Flt3 signaling results in significantly reduced cDCs in vivo [27]. Treatment with Flt3 ligand (Flt3L) induces CDP proliferation and cDC expansion [27]. Other cytokines, such as GM-CSF, are also involved in DC homeostasis and loss of both GM-CSF and Flt3L has been shown to exacerbate DC deficiencies in vivo [28]. However, the exact function of GM-CSF in cDC homeostasis has been complicated by discordant observations [29;30]. One study has found that peripheral CD103+ cDCs develop normally in mice lacking GM-CSFR but are characterized by lower levels of CD103 expression relative to control mice [29]. In contrast, a subsequent study has suggested that loss of GM-CSF signaling reduces CD103+ cDCs in lung and CD103+CD11b+ cDCs in small intestine. However, in that study decreased CD103 expression has been interpreted to indicate the developmental failure of particular cDC peripheral populations in the absence of GM-CSF signaling [30].

Zbtb46 identifies cDCs committed progenitors

Zbtb46 is a transcription factor specifically expressed in cDCs within the immune system [31*;32*]. Analysis of BM from mice harboring a Zbtb46-GFP reporter allele found that progenitors expressing Zbtb46 had lost potential for generation of pDCs and only develop into cDCs [31*]. However, the function of Zbtb46 in cDC development is not fully understood. Overexpression of Zbtb46 by retrovirus into unfractionated BM cells causes a bias in favor of the development of cDCs and against neutrophils in vitro; nonetheless, loss of Zbtb46, as assessed using homozygous Zbtb46gfp/gfp mice, does not adversely impact neutrophil or augment cDC population frequencies [31*]. A role for Zbtb46 as a transcriptional repressor in cDCs at steady state has been reported using an independent Zbtb46−/− mouse strain [33]. Microarray analysis of Zbtb46-deficient cDCs shows an upregulation of more than 2,000 genes, based on a historical comparison to microarrays of wild type cDCs [33]. Further, loss of Zbtb46 results in a partial activation phenotype of cDCs at steady state, in particular higher expression of MHC class II and co-stimulatory molecules [33]. However, similar results have not been reported with homozygous Zbtb46gfp/gfp mice. Zbtb46 may prove useful in allowing the further characterization of cDC progenitors, and understanding the pathways governing its expression could help identify the developmental stages of cDCs.

L-myc mediates cDC homeostasis in vivo

A recent study reported that L-myc (Mycl1), a c-Myc (Myc) and N-Myc paralogue, is specifically expressed in pDCs and cDCs [34]. c-Myc is highly expressed in progenitor cells and is required for proliferation and homeostasis [35], but in the murine DC lineage, Mycl1 and not Myc is expressed from the CDP stage onwards and is maintained in cDCs outside the BM [34]. Treatment of cDCs with GM-CSF induces L-myc expression and loss of the latter results in a significant reduction in cDC proliferative capacity and cDC numbers in the lung [34], a tissue rich in GM-CSF [36]. Since loss of GM-CSFR also impairs cDC expansion [33], these findings suggest L-myc activity is required for GM-CSF induced proliferation. Functionally, Mycl1−/− mice have a reduction in the extent of T cell priming during infection with the bacterial pathogen Listeria monocytogenes and vesicular stomatitis virus [34]. The reason for the switch from c-Myc to L-Myc expression during the development of DCs is still obscure, as it would seem that c-Myc should also be capable of supporting mitogenic activity downstream of GM-CSF. Possible reasons include constraints on c-Myc expression that might impair c-Myc levels during inflammation, or alternatively, different transcriptional targets of c-Myc and L-myc, issues that require further investigation.

Innate lymphoid cells mediate defense in response to distinct cDC subsets

Early studies on DC subsets have identified CD8α+ cDCs as specialized cross-presenting cells that induce CD8+ T cell responses in response to viral infections [3742], while CD8α-negative cDCs induce CD4+ T cell responses [43]. More recent studies, however, have highlighted the important roles cDC subsets play in modulating a vast array of immune responses, and in particular, their critical role in activating innate lymphoid cells (ILCs) by secretion of pro-inflammatory cytokines[4446**;47**].

ILCs are lymphocytes lacking recombined antigen-recognition receptors that are dependent on the transcription factors ID2 and PLZF for their development [4851**]. These populations of cells are involved in lymphoid organogenesis and tissue repair, as well as pathogen clearance [5254]. Over several years, it has come to be recognized that ILCs develop into subsets, now classified as groups ILC1, ILC2, and ILC3 [55], that share a substantial similarity in cytokine profiles and transcription factor dependence with CD4+ T cell subsets TH1, TH2 and TH17, respectively [56]. ILC1s secrete high levels of interferon (IFN)-γ upon IL-12 stimulation and are dependent on Nfil3 (E4BP4), T-bet, Ets-1 and IL-15 for their development and function [5764]. ILC2s secrete high levels of IL-4, IL-5 and IL-13 in response to parasitic helminths and allergen challenges [50;6567] and are dependent on the transcription factors Gata3, Rorα and Gfi1 [6871]. ILC3s, which include lymphotoxin inducer (LTi) cells, CD4+ LTi-like cells, and NKp46+ cells, secrete high levels of IL-22 upon IL-23 stimulation and are dependent on RORγt and AHR [7275]. Recent studies have shown a critical role for CD8α+ and CD11b+ cDCs as the obligate sources of inflammatory cytokines required for the activation of ILC1s and ILC3s, respectively [44;46**;47**].

Batf3-dependent CD8α+ cDCs support Type I immunity

The transcriptional regulation of CD8α+ cDCs has been studied extensively. CD8α+ cDC development is dependent on Irf8, ID2, Nfil3, Batf3 and Bcl-6 [7681]. However, the developmental stages at which these transcription factors are required are still unclear. Irf8 is expressed in early progenitors, and loss of Irf8 impairs cDC and pDC development [76;77;82]. ID2 deficiency only affects CD8α+ cDCs and not pDCs, suggesting ID2 is involved in cDC development downstream of Irf8 [78]. Nfil3−/− mice also have impaired CD8α+ cDC development, and Nfil3-deficient pre-cDCs have a two-fold decrease in Batf3 expression, suggesting Nfil3 acts upstream of Batf3 [79]; however, it is unclear if Batf3 is indeed a direct target of Nfil3. Finally, expression of Batf3 and Bcl-6 is not induced until later stages of CD8α+ cDC development [26;81]. CD8α+ cDC development, which is impaired by Batf3 deficiency, can be compensated by Irf8 interactions with Batf and Batf2 [83], which have the ability to interact with Irf8 in the same manner as Batf3 [84]. Importantly, the specific function for Batf3 in the development of CD8α+ cDCs is unclear; nonetheless, analysis of Bat/3−/− mice has provided ample insight into the critical role CD8α+ cDCs play in modulating innate responses.

The Batf3-dependent CD8α+ cDC is necessary for induction of type I responses to Toxoplasma gondii infection, based on their non-redundant secretion of IL-12, which is necessary for NK cell activation and concomitant production of IFN-γ (Fig. 1A) [44;45;8588]. Recognition of T. gondii infection by the CD8α+ cDC requires appropriate TLR signaling, since loss of Myd88 in cDCs during parasite infection caused a significant reduction in IL-12 production resulting in impaired parasite clearance [45]. Recent studies have shown that cDC sensing of T. gondii is dependent on TLR11/TLR12, which recognize profilin, resulting in IL-12 production and parasite clearance [89*;90*]. The susceptibility of Batf3-deficient mice to this pathogen [44], therefore, demonstrates that a DC subset, previously considered non-redundant strictly in adaptive immune functions, plays a critical role in coordinating innate immune responses to a lethal pathogen.

Figure 1. Dendritic cells orchestrate pathogen-appropriate innate and adaptive immune responses.

Figure 1

(a and b) A pre-cDC progenitor develops into both the Batf3-dependent CD8α+ cDC and the Notch2-dependent CD11b+ cDC subsets. (a) CD8α+ cDCs respond to Toxoplasma gondii infection by sensing the antigen profilin via TLRs 11/12. Activated CD8α+ cDCs secrete IL-12 which induces IFN-γ secretion by ILC1/NK cells. IFN-γ activates infected cells, such as MΦ’s, to produce nitric oxide and reactive oxygen species, which are necessary for parasite control. In parallel, activated CD8α+ cDCs migrate to lymph nodes and promote differentiation of TH1 cells, which enhance parasite clearance. (b) Infection with Citrobacter rodentium activates CD11b+ cDCs in the small intestine by an unknown mechanism. Upon activation, tissue resident CD11b+ cDCs produce high levels of IL-23 which induces IL-22 production in Rorγt dependent ILC3s and possibly acts on other cells as well. IL-22 activates the epithelial layer to secrete the bactericidal lectin RegIIIγ, which is necessary for C. rodentium clearance. CD11b+ cDCs also migrate to the mesenteric lymph node where they induce T cell differentiation into TH17 and TH22 cells. These effector T cells migrate back to the site of infection and are the adaptive source of IL-22. cDC: classical dendritic cell. TLR: Toll-like Receptor. ILC3: Innate lymphoid cell type 3. IFN-γ: Interferon-γ. MΦ: Macrophage

Recently, examination of Batf3−/− mice has revealed a role for CD8α+ cDCs in modulating invariant natural killer T cell (iNKT) responses. iNKTs express an invariant Vα14 chain and respond to glycolipid antigens, such as α-galactosylceramide (α-GalCer) [91] presented on CD1d molecules [92;93]. cDCs express CD1d and mediate iNKT negative selection in the thymus [94]. Furthermore, cDCs upregulate co-stimulatory molecules and secrete IL-12 in response to treatment with α-GalCer [95;96] and DCs pulsed with such glycolipid prolong IFN-γ production by iNKT cells [97]. Further, iNKT responses are reduced in Batf3−/− mice, suggesting that CD8α+ cDCs are the main regulator of iNKT activation [98*]. Nonetheless, the mechanism mediating this finding requires further analysis, as it is unclear if antigen presentation on CD1d molecules by CD8α+ cDCs is necessary for iNKT induction, or whether IL-12 secretion is sufficient for this process to occur.

Development and ontogeny of CD11b+ cDCs

The transcriptional networks regulating CD11b+ cDC development remain incompletely understood. Several transcription factors including Irf4, Notch2, and RelB, have been implicated in CD11b+ cDC development and homeostasis, however, a transcription factor that selectively controls development of CD11b+ cDCs remains unidentified. For some time, it was thought that Irf4 was required for this branch of cDCs [77]. Nonetheless, it is now it is recognized that Irf4−/− mice retain CD11b+ cDCs, which fail to induce CD4 expression and to migrate to lymph nodes during immune responses [99*]. In peripheral tissues, Irf4−/− mice have a two-fold decrease in CD103+ CD11b+ DCs in the small intestine lamina propria (SI-LP), and are completely absent in the mesenteric lymph node (MLN) [100*;101*]. Notch2 signaling is necessary for the development of ESAM+ CD11b+ splenic cDCs and CD103+ CD11b+ cDCs in the SI-LP, but this reduction is not observed in other tissues [102]. Non-canonical NF-κB signaling is also involved in CD11b+ cDC homeostasis; loss of Relb [103;104], Ltbr[105], Nik or Nfkb1[47**] results in a partial reduction of splenic CD11b+ cDCs. However, the basis of these transcriptional requirements is still unclear. For example, the involvement of Relb in CD11b+ cDC development has been inferred from the decrease in the number of CD8α-negative cDCs rather than loss of CD11b+ cDCs [104]. Conceivably, Irf4-expressing cDCs develop in the absence of RelB but fail to express a limited set of surface markers such as CD4.

The origins of SI-LP DCs have been elegantly shown in studies testing the developmental potential of DC progenitors by adoptively transferring MDPs, CDPs and pre-cDCs in vivo. All three progenitor populations are able to develop into CD103+ and CD103+CD11b+ cDCs in the SI-LP. Surprisingly, only the MDP gives rise to CD103CD11b+ cDCs, suggesting that this population may originate from monocytes and not a DC committed progenitor [11*;106*]. Analysis of Zbtb46-GFP mice has identified heterogeneous expression of Zbtb46 in the CD11b+ CD103F4/80 DC population in the SI-LP [47**]. Partial deletion of that population is also observed in the diphtheria toxin (DT)-based cDC depletion model using Zbtb46-DT receptor (DTR) mice, suggesting a mixed contribution from multiple myeloid lineages to this population [107*]. It is possible that Zbtb46-expressing CD103CD11b+ and CD103+ CD11b+ cDCs are developmentally related, and that expression of CD103 defines fully mature SI-LP CD11b+ cDCs. Alternatively, CD103CD11b+ cDCs could be a separate lineage. In a murine model of DSS-induced colitis, Ly6Chi monocytes can develop into moDCs in a model of DSS-induced colitis [108*]. During inflammation, colonic Ly6Chi monocytes give rise to a CX3CRintLyC6lo population which has a transcriptional profile similar to cDCs. Specifically, these cells express Zbtb46 and Flt3, genes restricted in the periphery to cDCs. Functionally, these moDCs were able to migrate to lymphatic tissue, process antigen and induce T cell proliferation [108*]. These findings provide evidence for the differentiation of monocytes into DCs during inflammation; however this phenomenon has not been analyzed at steady state. Analysis of reporter mice combined with lineage tracing could help identify monocyte contributions to CD11b+ DCs.

Notch2-dependent CD11b+ cDCs mediate pathogen-selective defenses in the intestine

The dependence of splenic CD8α+ and peripheral CD103+ cDCs on Batf3 for their development has helped to establish the functional restriction of this branch of cDCs in defense against intracellular pathogens [39;40;44;80;109]. In contrast, for Batf3-independent cDCs, some analysis using other genetic models indicates distinct functional specialization of CD11b+ cDC subsets. It was known that IL-23 and IL-22 production by innate cells is crucial during early stages of infection with Citrobacter rodentium, a mouse model for enteropathogenic Escherichia coli [110*112]. Recent studies have identified CD11b+ cDCs as the non-redundant source of IL-23 in response to administration of bacterial flagellin [46**] and conditional deletion of Notch2 in cDCs has revealed that CD11b+ cDCs are required for clearance of C. rodentium by modulating the IL-23/IL-22 cytokine axis (Fig. 1B) [47**;107*]. Mixed chimera analysis indicated that the Notch2-dependent CD11b+ cDCs are the non-redundant critical source of IL-23 for resistance to C. rodentium; specifically, the susceptibility of Notch2-deficient mice to this pathogen appears to be due to failure of innate defense [47**]. Importantly Irf4−/− and Ccr7−/− mice, whose DCs exhibit a defect in migration [99*;113], do not show the same dramatic susceptibility to C. rodentium infection as mice lacking Notch2 in cDCs [47**]. This result suggests that IL-23 secretion by tissue resident CD11b+ cDCs is sufficient for inducing IL-22 production by ILC3s and, hence, effective local control of the pathogen.

IL-23 has also been implicated in the stabilization of Th17 commitment [114]. However, the non-redundant production of IL-23 by CD11b+ cDCs during infection with C. rodentium does not necessarily indicate that these cells are also involved in TH17 polarization. Nevertheless, recent studies have shown a role for CD11b+ cDCs in TH17 differentiation and function. Deletion of Notch2 in cDCs leads to a decrease in TH17 numbers in the MLN [102], and loss of Irf4 hinders the induction of TH17 differentiation by small intestine CD103+CD11b+ cDCs [100*;101*]. These findings have been further corroborated by specific deletion of SI-LP CD103+CD11b+ cDCs using a human Langerin-DTA transgenic mouse model which also resulted in a decrease in TH17 cell numbers [115]. However, further analysis is needed to resolve the role DCs play in TH17 induction both at steady state and during infection. It has been recognized that other cytokines such as IL-6 and TGF-β, but not IL-23, are required for polarization of TH17 cells from naïve T cells [116;117]. Further, secretion of IL-22 by CD4+ T cells can occur independently of IL-23 [118]. These results suggest that CD11b+ cDCs may be dispensable for initial differentiation of TH17, but instead may enhance TH17 commitment and cytokine production via secretion of IL-23 [119; 120].

Although significant progress has been made in understanding the immune responses necessary for clearance of enteropathogenic bacteria, several aspects are still unclear. The physiological stimuli required for IL-23 secretion by cDCs are not understood. For example, it is unknown whether C. rodentium directly activates cDCs or if secondary signaling induces IL-23 production in cDCs. Although flagellin can induce IL-23 secretion by CD11b+ DCs, C. rodentium is an aflagellate organism and should not stimulate cDCs via TLR5 [121]. Infection with a flagellate organism such as Salmonella typhimurium could provide insight on the mechanisms regulating IL-23 secretion by CD11b+ DCs. Further, it is unclear if the CD103+ or the CD103-negative CD11b+ cDC populations in the SI-LP are sufficient for clearance of C. rodentium. Although loss of Notch2 severely diminishes SI-LP CD103+ CD11b+ DCs, this deficiency could also impair CD103CD11b+ cDC function [115]. Finally, other aspects of IL-23 signaling in addition to IL-22 induction have not been studied. Mortality and severity of disease during infection by C. rodentium is more severe in Il23a−/− mice when compared to Il22−/− mice [110*,118], further examination of the effects of IL-23 on other cells could provide insight on the pleiotropic functions of this cytokine.

CD11b+ cDCs mediate type II immunity

The specific secretion of IL-12 and IL-23 by CD8α- and CD11b–expressing cDCs, respectively, suggests that a specific DC subset that regulates type II immunity may exist. Depletion of CD11c+ expressing cells in Itgax-DTR mice significantly impairs TH2 responses against the parasitic helminth Schistosoma mansoni [122]. Recent studies suggest that the CD11b+ cDC subsets may modulate type II responses. A CD11b+ cDC subset expressing FCεR1 initiates in vivo type II immune responses to immunization with house dust mite (HDM) antigen [123]. In addition, both CD11b+ cDCs and moDCs are rapidly recruited to lung airways upon HDM challenge, yet only CD11b+ cDCs are able to migrate to the lymph node and trigger TH2 responses [124*]. Nevertheless, moDCs are sufficient to induce TH2 responses in Flt3l−/− mice immunized with a high dose of HDM antigen [125*]. Recently, it has been shown that loss of Irf4 in CD11c+ cells impairs TH2 responses in a model of allergy based on HDM antigen immunization [126]. Further, using Irf4−/− mice, two recent studies have identified a subset of dermal cDCs expressing CD301b and PDL2 that mediate TH2 responses during allergic responses or infection with Nippostrongylus brasiliensis [127;128]. However, these studies did not address a possible defect in DC migration from tissues to lymph nodes due to Irf4 deficiency [99*]. It is possible that reduced TH2 responses in these studies [127;128] result from impaired cDC migration to lymph nodes rather than a selective Irf4 involvement in TH2 priming. Also, it is unclear if Irf4-expressing CD11b+ cDCs are a uniform group [9]. Finally, the involvement of CD11b+ cDCs in modulating innate type II responses is unstudied.

Concluding remarks

At present, two cDC subsets have been examined with respect to their specialization in providing defense against different types of pathogens. The Batf3-dependent subset of cDCs, which is characterized by high levels of Irf8, appear specialized in priming CTL responses against many viruses and for promoting type I defense against intracellular pathogens, such as T. gondii. This latter function is not restricted to priming TH1 responses, but also enhancing innate defenses that involve early production of IFN-γ produced by innate lymphocytes such as ILC1 or NK cells [44;45]. Notch2-dependent subsets of CD11b+ cDCs have been shown to specialize in defense against one pathogen, C. rodentium, but are dispensable for defense against pathogens such as T. gondii and L. monocytogenes. Thus, in both cases, specialized DC subsets appear to engage innate immune cells in a manner that is not simply an adjuvant to defense, but critical for resistance. An unresolved issue is how the different subsets manage to acquire their specialized functions. To some degree, the particular expression patterns of innate molecular sensors and the capacity for specific cytokine production may be involved, although the molecular basis for these have not been well worked out.

Different sensing and effector capacities exhibited by distinct cDC subsets can provide a system to drive activation of pathogen-appropriate ILC subsets, as is the development of pathogen-appropriate CD4+ T cell subsets. The similarity in transcriptional features underlying ILC and T cell diversity [56] might suggest that the latter was adapted during evolution from the former, and that both effectors arms rest on the same platform of DC diversity. If ILCs evolved before T cells, then cDC subsets may have first developed to link particular pathogen sensors with the production of cytokines needed to activate pathogen-appropriate ILC subsets. Later, with the introduction of RAG-dependent adaptive immunity, this established DC subset program would acquire the capacity for antigen presentation in addition to the pre-existing cytokine circuitry. CD4+ T cells would adapt to the use of these same cytokines to trigger their development down the transcriptional pathways first established in ILCs. Testing this speculation may be difficult, since the ancestral species representative of this transition may no longer exist.

Highlights.

  • Transcriptional requirements of DC subset development.

  • Batf3-dependent CD8α+ DCs mediate type I immune responses.

  • CD11b+ DCs orchestrate ILC3 and TH17 responses.

Acknowledgments

The authors thank Vivek Durai, Ansuman T. Satpathy and Xiaodi Wu for their critical reading of the manuscript. Supported by the Howard Hughes Medical Institute (K.M.M.).

Reference List

  • 1.Steinman RM, Cohn ZA. Identification of a novel cell type in peripheral lymphoid organs of mice. I. Morphology, quantitation, tissue distribution. J Exp.Med. 1973;137:1142–1162. doi: 10.1084/jem.137.5.1142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Fernandez NC, Lozier A, Flament C, Ricciardi-Castagnoli P, Bellet D, Suter M, Perricaudet M, Tursz T, Maraskovsky E, Zitvogel L. Dendritic cells directly trigger NK cell functions: crosstalk relevant in innate anti-tumor immune responses in vivo. Nat.Med. 1999;5:405–411. doi: 10.1038/7403. [DOI] [PubMed] [Google Scholar]
  • 3.Siegal FP, Kadowaki N, Shodell M, Fitzgerald-Bocarsly PA, Shah K, Ho, Antonenko S, Liu YJ. The nature of the principal type 1 interferon-producing cells in human blood. Science. 1999;284:1835–1837. doi: 10.1126/science.284.5421.1835. [DOI] [PubMed] [Google Scholar]
  • 4.Cella M, Jarrossay D, Facchetti F, Alebardi O, Nakajima H, Lanzavecchia A, Colonna M. Plasmacytoid monocytes migrate to inflamed lymph nodes and produce large amounts of type I interferon. Nature Medicine. 1999;5:919–923. doi: 10.1038/11360. [DOI] [PubMed] [Google Scholar]
  • 5.Randolph GJ, Inaba K, Robbiani DF, Steinman RM, Muller WA. Differentiation of phagocytic monocytes into lymph node dendritic cells in vivo. Immunity. 1999;11:753–761. doi: 10.1016/s1074-7613(00)80149-1. [DOI] [PubMed] [Google Scholar]
  • 6.Serbina NV, Salazar-Mather TP, Biron CA, Kuziel WA, Pamer EG. TNF/iNOS-producing dendritic cells mediate innate immune defense against bacterial infection. Immunity. 2003;19:59–70. doi: 10.1016/s1074-7613(03)00171-7. [DOI] [PubMed] [Google Scholar]
  • 7.Geissmann F, Jung S, Littman DR. Blood monocytes consist of two principal subsets with distinct migratory properties. Immunity. 2003;19:71–82. doi: 10.1016/s1074-7613(03)00174-2. [DOI] [PubMed] [Google Scholar]
  • 8.Schuler G, Steinman RM. Murine epidermal Langerhans cells mature into potent immunostimulatory dendritic cells in vitro. J.Exp.Med. 1985;161:526–546. doi: 10.1084/jem.161.3.526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Jaitin DA, Kenigsberg E, Keren-Shaul H, Elefant N, Paul F, Zaretsky I, Mildner A, Cohen N, Jung S, Tanay A, Amit I. Massively parallel single-cell RNA-seq for marker-free decomposition of tissues into cell types. Science. 2014;343:776–779. doi: 10.1126/science.1247651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Ginhoux F, Liu K, Helft J, Bogunovic M, Greter M, Hashimoto D, Price J, Yin N, Bromberg J, Lira SA, Stanley ER, Nussenzweig M, Merad M. The origin and development of nonlymphoid tissue CD103+ DCs. J Exp.Med. 2009;206:3115–3130. doi: 10.1084/jem.20091756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Bogunovic M, Ginhoux F, Helft J, Shang L, Hashimoto D, Greter M, Liu K, Jakubzick C, Ingersoll MA, Leboeuf M, Stanley ER, Nussenzweig M, Lira SA, Randolph GJ, Merad M. Origin of the lamina propria dendritic cell network. Immunity. 2009;31:513–525. doi: 10.1016/j.immuni.2009.08.010. These two studies (ref. 11 and 106) have shown DC progenitors and monocytes give rise to CD11c+ cells in the small intestine lamina propria.
  • 12.Satpathy AT, Wu X, Albring JC, Murphy KM. Re(de)fining the dendritic cell lineage. Nat.Immunol. 2012;13:1145–1154. doi: 10.1038/ni.2467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Merad M, Sathe P, Helft J, Miller J, Mortha A. The dendritic cell lineage: ontogeny and function of dendritic cells and their subsets in the steady state and the inflamed setting. Annu.Rev.Immunol. 2013;31:563–604. doi: 10.1146/annurev-immunol-020711-074950. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Traver D, Akashi K, Manz M, Merad M, Miyamoto T, Engleman EG, Weissman IL. Development of CD8alpha-positive dendritic cells from a common myeloid progenitor. Science. 2000;290:2152–2154. doi: 10.1126/science.290.5499.2152. [DOI] [PubMed] [Google Scholar]
  • 15.Fogg DK, Sibon C, Miled C, Jung S, Aucouturier P, Littman DR, Cumano A, Geissmann F. A clonogenic bone marrow progenitor specific for macrophages and dendritic cells. Science. 2006;311:83–87. doi: 10.1126/science.1117729. [DOI] [PubMed] [Google Scholar]
  • 16.Auffray C, Fogg DK, Narni-Mancinelli E, Senechal B, Trouillet C, Saederup N, Leemput J, Bigot K, Campisi L, Abitbol M, Molina T, Charo I, Hume DA, Cumano A, Lauvau G, Geissmann F. CX3CR1+ CD115+ CD135+ common macrophage/DC precursors and the role of CX3CR1 in their response to inflammation. J Exp.Med. 2009;206:595–606. doi: 10.1084/jem.20081385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Naik SH, Sathe P, Park HY, Metcalf D, Proietto AI, Dakic A, Carotta S, O’Keeffe M, Bahlo M, Papenfuss A, Kwak JY, Wu L, Shortman K. Development of plasmacytoid and conventional dendritic cell subtypes from single precursor cells derived in vitro and in vivo. Nat Immunol. 2007;8:1217–1226. doi: 10.1038/ni1522. [DOI] [PubMed] [Google Scholar]
  • 18.Onai N, Obata-Onai A, Schmid MA, Ohteki T, Jarrossay D, Manz MG. Identification of clonogenic common Flt3(+) M-CSFR+ plasmacytoid and conventional dendritic cell progenitors in mouse bone marrow. Nat. 2007;8:1207–1216. doi: 10.1038/ni1518. [DOI] [PubMed] [Google Scholar]
  • 19.Liu K, Waskow C, Liu X, Yao K, Hoh J, Nussenzweig M. Origin of dendritic cells in peripheral lymphoid organs of mice. Nat Immunol. 2007;8:578–583. doi: 10.1038/ni1462. [DOI] [PubMed] [Google Scholar]
  • 20. Onai N, Kurabayashi K, Hosoi-Amaike M, Toyama-Sorimachi N, Matsushima K, Inaba K, Ohteki T. A clonogenic progenitor with prominent plasmacytoid dendritic cell developmental potential. Immunity. 2013;38:943–957. doi: 10.1016/j.immuni.2013.04.006. This study identifies an M-CSFR-negative DC progenitor with strong bias towards pDC development.
  • 21.Omatsu Y, Iyoda T, Kimura Y, Maki A, Ishimori M, Toyama-Sorimachi N, Inaba K. Development of murine plasmacytoid dendritic cells defined by increased expression of an inhibitory NK receptor, Ly49Q. J.Immunol. 2005;174:6657–6662. doi: 10.4049/jimmunol.174.11.6657. [DOI] [PubMed] [Google Scholar]
  • 22.Schlitzer A, Loschko J, Mair K, Vogelmann R, Henkel L, Einwachter H, Schiemann M, Niess JH, Reindl W, Krug A. Identification of CCR9- murine plasmacytoid DC precursors with plasticity to differentiate into conventional DCs. Blood. 2011;117:6562–6570. doi: 10.1182/blood-2010-12-326678. [DOI] [PubMed] [Google Scholar]
  • 23.McKenna HJ, Stocking KL, Miller RE, Brasel K, De Smedt T, Maraskovsky E, Maliszewski CR, Lynch DH, Smith J, Pulendran B, Roux ER, Teepe M, Lyman SD, Peschon JJ. Mice lacking flt3 ligand have deficient hematopoiesis affecting hematopoietic progenitor cells, dendritic cells, and natural killer cells. Blood. 2000;95:3489–3497. [PubMed] [Google Scholar]
  • 24.Karsunky H, Merad M, Cozzio A, Weissman IL, Manz MG. Flt3 ligand regulates dendritic cell development from Flt3+ lymphoid and myeloid-committed progenitors to Flt3+ dendritic cells in vivo. J.Exp.Med. 2003;198:305–313. doi: 10.1084/jem.20030323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Laouar Y, Welte T, Fu XY, Flavell RA. STAT3 is required for Flt3L–dependent dendritic cell differentiation. Immunity. 2003;19:903–912. doi: 10.1016/s1074-7613(03)00332-7. [DOI] [PubMed] [Google Scholar]
  • 26. Miller JC, Brown BD, Shay T, Gautier EL, Jojic V, Cohain A, Pandey G, Leboeuf M, Elpek KG, Helft J, Hashimoto D, Chow A, Price J, Greter M, Bogunovic M, Bellemare-Pelletier A, Frenette PS, Randolph GJ, Turley SJ, Merad M, Gautier EL, Jakubzick C, Randolph GJ, Best AJ, Knell J, Goldrath A, Miller J, Brown B, Merad M, Jojic V, Koller D, Cohen N, Brennan P, Brenner M, Shay T, Regev A, Fletcher A, Elpek K, Bellemare-Pelletier A, Malhotra D, Turley S, Jianu R, Laidlaw D, Collins J, Narayan K, Sylvia K, Kang J, Gazit R, Rossi DJ, Kim F, Rao TN, Wagers A, Shinton SA, Hardy RR, Monach P, Bezman NA, Sun JC, Kim CC, Lanier LL, Heng T, Kreslavsky T, Painter M, Ericson J, Davis S, Mathis D, Benoist C. Deciphering the transcriptional network of the dendritic cell lineage. Nat.Immunol. 2012;13:888–899. doi: 10.1038/ni.2370. This study extensively analyzes the transcriptional profile of the dendritic cell lineages.
  • 27.Waskow C, Liu K, Darrasse-Jeze G, Guermonprez P, Ginhoux F, Merad M, Shengelia T, Yao K, Nussenzweig M. The receptor tyrosine kinase Flt3 is required for dendritic cell development in peripheral lymphoid tissues. Nat Immunol. 2008;9:676–683. doi: 10.1038/ni.1615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Kingston D, Schmid MA, Onai N, Obata-Onai A, Baumjohann D, Manz MG. The concerted action of GM-CSF and Flt3-ligand on in vivo dendritic cell homeostasis. Blood. 2009;114:835–843. doi: 10.1182/blood-2009-02-206318. [DOI] [PubMed] [Google Scholar]
  • 29.Edelson BT, Bradstreet TR, KC W, Hildner K, Herzog JW, Sim J, Russell JH, Murphy TL, Unanue ER, Murphy KM. Batf3-dependent CD11b(low/−) peripheral dendritic cells are GM-CSF-independent and are not required for Th cell priming after subcutaneous immunization. PLoS One. 2011;6:e25660. doi: 10.1371/journal.pone.0025660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Greter M, Helft J, Chow A, Hashimoto D, Mortha A, Agudo-Cantero J, Bogunovic M, Gautier EL, Miller J, Leboeuf M, Lu G, Aloman C, Brown BD, Pollard JW, Xiong H, Randolph GJ, Chipuk JE, Frenette PS, Merad M. GM-CSF controls nonlymphoid tissue dendritic cell homeostasis but is dispensable for the differentiation of inflammatory dendritic cells. Immunity. 2012;36:1031–1046. doi: 10.1016/j.immuni.2012.03.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Satpathy AT, KC W, Albring JC, Edelson BT, Kretzer NM, Bhattacharya D, Murphy TL, Murphy KM. Zbtb46 expression distinguishes classical dendritic cells and their committed progenitors from other immune lineages. J.Exp.Med. 2012;209:1135–1152. doi: 10.1084/jem.20120030. These two studies (Ref. 31 and 32) identify Zbtb46 to be selectively expressed in dendritic cells within the immune system.
  • 32. Meredith MM, Liu K, Darrasse-Jeze G, Kamphorst AO, Schreiber HA, Guermonprez P, Idoyaga J, Cheong C, Yao KH, Niec RE, Nussenzweig MC. Expression of the zinc finger transcription factor zDC (Zbtb46, Btbd4) defines the classical dendritic cell lineage. J.Exp.Med. 2012;209:1153–1165. doi: 10.1084/jem.20112675. These two studies (Ref. 31 and 32) identify Zbtb46 to be selectively expressed in dendritic cells within the immune system.
  • 33.Meredith MM, Liu K, Kamphorst AO, Idoyaga J, Yamane A, Guermonprez P, Rihn S, Yao KH, Silva IT, Oliveira TY, Skokos D, Casellas R, Nussenzweig MC. Zinc finger transcription factor zDC is a negative regulator required to prevent activation of classical dendritic cells in the steady state. J.Exp.Med. 2012;209:1583–1593. doi: 10.1084/jem.20121003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.KC W, Satpathy AT, Rapaport AS, Briseno CG, Wu X, Albring JC, Russler-Germain EV, Kretzer NM, Durai V, Persaud SP, Edelson BT, Loschko J, Cella M, Allen PM, Nussenzweig MC, Colonna M, Sleckman BP, Murphy TL, Murphy KM. L-Myc expression by dendritic cells is required for optimal T-cell priming. Nature. 2014 doi: 10.1038/nature12967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Laurenti E, Varnum-Finney B, Wilson A, Ferrero I, Blanco-Bose WE, Ehninger A, Knoepfler PS, Cheng PF, MacDonald HR, Eisenman RN, Bernstein ID, Trumpp A. Hematopoietic stem cell function and survival depend on c-Myc and N-Myc activity. Cell Stem Cell. 2008;3:611–624. doi: 10.1016/j.stem.2008.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Scott CL, Hughes DA, Cary D, Nicola NA, Begley CG, Robb L. Functional analysis of mature hematopoietic cells from mice lacking the beta c chain of the granulocyte-macrophage colony-stimulating factor receptor. Blood. 1998;92:4119–4127. [PubMed] [Google Scholar]
  • 37.den Haan JM, Lehar SM, Bevan MJ. CD8(+) but not CD8(−) dendritic cells cross-prime cytotoxic T cells in vivo. J Exp.Med. 2000;192:1685–1696. doi: 10.1084/jem.192.12.1685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.den Haan JM, Bevan MJ. Antigen presentation to CD8+ T cells: cross-priming in infectious diseases. Curr.Opin.Immunol. 2001;13:437–441. doi: 10.1016/s0952-7915(00)00238-7. [DOI] [PubMed] [Google Scholar]
  • 39.Torti N, Walton SM, Murphy KM, Oxenius A. Batf3 transcription factor-dependent DC subsets in murine CMV infection: differential impact on T-cell priming and memory inflation. Eur.J.Immunol. 2011;41:2612–2618. doi: 10.1002/eji.201041075. [DOI] [PubMed] [Google Scholar]
  • 40.Pinto AK, Daffis S, Brien JD, Gainey MD, Yokoyama WM, Sheehan KC, Murphy KM, Schreiber RD, Diamond MS. A temporal role of type I interferon signaling in CD8+ T cell maturation during acute West Nile virus infection. PLoS Pathog. 2011;7:e1002407. doi: 10.1371/journal.ppat.1002407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Nopora K, Bernhard CA, Ried C, Castello AA, Murphy KM, Marconi P, Koszinowski U, Brocker T. MHC class I cross-presentation by dendritic cells counteracts viral immune evasion. Front Immunol. 2012;3:348. doi: 10.3389/fimmu.2012.00348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Helft J, Manicassamy B, Guermonprez P, Hashimoto D, Silvin A, Agudo J, Brown BD, Schmolke M, Miller JC, Leboeuf M, Murphy KM, Garcia-Sastre A, Merad M. Cross-presenting CD103+ dendritic cells are protected from influenza virus infection. J.Clin.Invest. 2012;122:4037–4047. doi: 10.1172/JCI60659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Dudziak D, Kamphorst AO, Heidkamp GF, Buchholz VR, Trumpfheller C, Yamazaki S, Cheong C, Liu K, Lee HW, Park CG, Steinman RM, Nussenzweig MC. Differential antigen processing by dendritic cell subsets in vivo. Science. 2007;315:107–111. doi: 10.1126/science.1136080. [DOI] [PubMed] [Google Scholar]
  • 44.Mashayekhi M, Sandau MM, Dunay IR, Frickel EM, Khan A, Goldszmid RS, Sher A, Ploegh HL, Murphy TL, Sibley LD, Murphy KM. CD8a+ Dendritic Cells Are the Critical Source of Interleukin-12 that Controls Acute Infection by Toxoplasma gondii Tachyzoites. Immunity. 2011;35:249–259. doi: 10.1016/j.immuni.2011.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Hou B, Benson A, Kuzmich L, DeFranco AL, Yarovinsky F. Critical coordination of innate immune defense against Toxoplasma gondii by dendritic cells responding via their Toll-like receptors. Proc.Natl.Acad.Sci U.S A. 2011;108:278–283. doi: 10.1073/pnas.1011549108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Kinnebrew MA, Buffie CG, Diehl GE, Zenewicz LA, Leiner I, Hohl TM, Flavell RA, Littman DR, Pamer EG. Interleukin 23 production by intestinal CD103(+)CD11b(+) dendritic cells in response to bacterial flagellin enhances mucosal innate immune defense. Immunity. 2012;36:276–287. doi: 10.1016/j.immuni.2011.12.011. First study to demonstrate that the CD103+CD11b+ is the non-redundant source of IL-23 upon bacterial sensing.
  • 47. Satpathy AT, Briseno CG, Lee JS, Ng D, Manieri NA, KC W, Wu X, Thomas SR, Lee WL, Turkoz M, McDonald KG, Meredith MM, Song C, Guidos CJ, Newberry RD, Ouyang W, Murphy TL, Stappenbeck TS, Gommerman JL, Nussenzweig MC, Colonna M, Kopan R, Murphy KM. Notch2-dependent classical dendritic cells orchestrate intestinal immunity to attaching-and-effacing bacterial pathogens. Nat.Immunol. 2013 doi: 10.1038/ni.2679. This study shows CD11b+ dendritic cells are the obligate innate source of IL-23 during C. rodentium infection.
  • 48.Yokota Y, Mansouri A, Mori S, Sugawara S, Adachi S, Nishikawa S, Gruss P. Development of peripheral lymphoid organs and natural killer cells depends on the helix-loop-helix inhibitor Id2. Nature. 1999;397:702–706. doi: 10.1038/17812. [DOI] [PubMed] [Google Scholar]
  • 49.Satoh-Takayama N, Lesjean-Pottier S, Vieira P, Sawa S, Eberl G, Vosshenrich CA, Di Santo JP. IL-7 and IL-15 independently program the differentiation of intestinal CD3-NKp46+ cell subsets from Id2-dependent precursors. J.Exp.Med. 2010;207:273–280. doi: 10.1084/jem.20092029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Moro K, Yamada T, Tanabe M, Takeuchi T, Ikawa T, Kawamoto H, Furusawa J, Ohtani M, Fujii H, Koyasu S. Innate production of T(H)2 cytokines by adipose tissue-associated c-Kit(+)Sca-1(+) lymphoid cells. Nature. 2010;463:540–544. doi: 10.1038/nature08636. [DOI] [PubMed] [Google Scholar]
  • 51. Constantinides MG, McDonald BD, Verhoef PA, Bendelac A. A committed precursor to innate lymphoid cells. Nature. 2014 doi: 10.1038/nature13047. This study uses lineage tracing of PLZF expressing cells to identify a committed innate lymphoid cell progenitor.
  • 52.Kelly KA, Scollay R. Seeding of neonatal lymph nodes by T cells and identification of a novel population of CD3-CD4+ cells. Eur.J.Immunol. 1992;22:329–334. doi: 10.1002/eji.1830220207. [DOI] [PubMed] [Google Scholar]
  • 53.Fallon PG, Ballantyne SJ, Mangan NE, Barlow JL, Dasvarma A, Hewett DR, McIlgorm A, Jolin HE, McKenzie AN. Identification of an interleukin (IL)-25-dependent cell population that provides IL-4, IL-5, and IL-13 at the onset of helminth expulsion. J.Exp.Med. 2006;203:1105–1116. doi: 10.1084/jem.20051615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Scandella E, Bolinger B, Lattmann E, Miller S, Favre S, Littman DR, Finke D, Luther SA, Junt T, Ludewig B. Restoration of lymphoid organ integrity through the interaction of lymphoid tissue-inducer cells with stroma of the T cell zone. Nat.Immunol. 2008;9:667–675. doi: 10.1038/ni.1605. [DOI] [PubMed] [Google Scholar]
  • 55.Spits H, Artis D, Colonna M, Diefenbach A, Di Santo JP, Eberl G, Koyasu S, Locksley RM, McKenzie AN, Mebius RE, Powrie F, Vivier E. Innate lymphoid cells--a proposal for uniform nomenclature. Nat.Rev.Immunol. 2013;13:145–149. doi: 10.1038/nri3365. [DOI] [PubMed] [Google Scholar]
  • 56.Spits H, Cupedo T. Innate lymphoid cells: emerging insights in development, lineage relationships, and function. Annu.Rev.Immunol. 2012;30:647–675. doi: 10.1146/annurev-immunol-020711-075053. [DOI] [PubMed] [Google Scholar]
  • 57.Gascoyne DM, Long E, Veiga-Fernandes H, de Boer J, Williams O, Seddon B, Coles M, Kioussis D, Brady HJ. The basic leucine zipper transcription factor E4BP4 is essential for natural killer cell development. Nat.Immunol. 2009;10:1118–1124. doi: 10.1038/ni.1787. [DOI] [PubMed] [Google Scholar]
  • 58.Barton K, Muthusamy N, Fischer C, Ting CN, Walunas TL, Lanier LL, Leiden JM. The Ets-1 transcription factor is required for the development of natural killer cells in mice. Immunity. 1998;9:555–563. doi: 10.1016/s1074-7613(00)80638-x. [DOI] [PubMed] [Google Scholar]
  • 59.Kennedy MK, Glaccum M, Brown SN, Butz EA, Viney JL, Embers M, Matsuki N, Charrier K, Sedger L, Willis CR, Brasel K, Morrissey PJ, Stocking K, Schuh JC, Joyce S, Peschon JJ. Reversible defects in natural killer and memory CD8 T cell lineages in interleukin 15-deficient mice. J.Exp.Med. 2000;191:771–780. doi: 10.1084/jem.191.5.771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Suzuki H, Duncan GS, Takimoto H, Mak TW. Abnormal development of intestinal intraepithelial lymphocytes and peripheral natural killer cells in mice lacking the IL-2 receptor beta chain. J.Exp.Med. 1997;185:499–505. doi: 10.1084/jem.185.3.499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.DiSanto JP, Muller W, Guy-Grand D, Fischer A, Rajewsky K. Lymphoid development in mice with a targeted deletion of the interleukin 2 receptor gamma chain. Proc.Natl.Acad.Sci.U.S A. 1995;92:377–381. doi: 10.1073/pnas.92.2.377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Gordon SM, Chaix J, Rupp LJ, Wu J, Madera S, Sun JC, Lindsten T, Reiner SL. The transcription factors T-bet and Eomes control key checkpoints of natural killer cell maturation. Immunity. 2012;36:55–67. doi: 10.1016/j.immuni.2011.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Sciume G, Hirahara K, Takahashi H, Laurence A, Villarino AV, Singleton KL, Spencer SP, Wilhelm C, Poholek AC, Vahedi G, Kanno Y, Belkaid Y, O’Shea JJ. Distinct requirements for T-bet in gut innate lymphoid cells. J.Exp.Med. 2012;209:2331–2338. doi: 10.1084/jem.20122097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Rankin LC, Groom JR, Chopin M, Herold MJ, Walker JA, Mielke LA, McKenzie AN, Carotta S, Nutt SL, Belz GT. The transcription factor T-bet is essential for the development of NKp46+ innate lymphocytes via the Notch pathway. Nat.Immunol. 2013;14:389–395. doi: 10.1038/ni.2545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Neill DR, Wong SH, Bellosi A, Flynn RJ, Daly M, Langford TK, Bucks C, Kane CM, Fallon PG, Pannell R, Jolin HE, McKenzie AN. Nuocytes represent a new innate effector leukocyte that mediates type-2 immunity. Nature. 2010;464:1367–1370. doi: 10.1038/nature08900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Price AE, Liang HE, Sullivan BM, Reinhardt RL, Eisley CJ, Erle DJ, Locksley RM. Systemically dispersed innate IL-13-expressing cells in type 2 immunity. Proc.Natl.Acad.Sci.U.S A. 2010;107:11489–11494. doi: 10.1073/pnas.1003988107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Saenz SA, Siracusa MC, Perrigoue JG, Spencer SP, Urban JF, Jr, Tocker JE, Budelsky AL, Kleinschek MA, Kastelein RA, Kambayashi T, Bhandoola A, Artis D. IL25 elicits a multipotent progenitor cell population that promotes T(H)2 cytokine responses. Nature. 2010;464:1362–1366. doi: 10.1038/nature08901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Klein Wolterink RG, Serafini N, van Nimwegen M, Vosshenrich CA, de Bruijn MJ, Fonseca PD, Veiga FH, Hendriks RW, Di Santo JP. Essential, dose-dependent role for the transcription factor Gata3 in the development of IL-5+ and IL-13+ type 2 innate lymphoid cells. Proc.Natl.Acad.Sci.U.S A. 2013 doi: 10.1073/pnas.1217158110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Wong SH, Walker JA, Jolin HE, Drynan LF, Hams E, Camelo A, Barlow JL, Neill DR, Panova V, Koch U, Radtke F, Hardman CS, Hwang YY, Fallon PG, McKenzie AN. Transcription factor RORalpha is critical for nuocyte development. Nat.Immunol. 2012;13:229–236. doi: 10.1038/ni.2208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Halim TY, MacLaren A, Romanish MT, Gold MJ, McNagny KM, Takei F. Retinoic-acid-receptor-related orphan nuclear receptor alpha is required for natural helper cell development and allergic inflammation. Immunity. 2012;37:463–474. doi: 10.1016/j.immuni.2012.06.012. [DOI] [PubMed] [Google Scholar]
  • 71.Spooner CJ, Lesch J, Yan D, Khan AA, Abbas A, Ramirez-Carrozzi V, Zhou M, Soriano R, Eastham-Anderson J, Diehl L, Lee WP, Modrusan Z, Pappu R, Xu M, DeVoss J, Singh H. Specification of type 2 innate lymphocytes by the transcriptional determinant Gfi1. Nat.Immunol. 2013;14:1229–1236. doi: 10.1038/ni.2743. [DOI] [PubMed] [Google Scholar]
  • 72.Satoh-Takayama N, Vosshenrich CA, Lesjean-Pottier S, Sawa S, Lochner M, Rattis F, Mention JJ, Thiam K, Cerf-Bensussan N, Mandelboim O, Eberl G, Di Santo JP. Microbial flora drives interleukin 22 production in intestinal NKp46+ cells that provide innate mucosal immune defense. Immunity. 2008;29:958–970. doi: 10.1016/j.immuni.2008.11.001. [DOI] [PubMed] [Google Scholar]
  • 73.Luci C, Reynders A, Ivanov II, Cognet C, Chiche L, Chasson L, Hardwigsen J, Anguiano E, Banchereau J, Chaussabel D, Dalod M, Littman DR, Vivier E, Tomasello E. Influence of the transcription factor RORgammat on the development of NKp46+ cell populations in gut and skin. Nat.Immunol. 2009;10:75–82. doi: 10.1038/ni.1681. [DOI] [PubMed] [Google Scholar]
  • 74.Sanos SL, Bui VL, Mortha A, Oberle K, Heners C, Johner C, Diefenbach A. RORgammat and commensal microflora are required for the differentiation of mucosal interleukin 22-producing NKp46+ cells. Nat.Immunol. 2009;10:83–91. doi: 10.1038/ni.1684. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Lee JS, Cella M, McDonald KG, Garlanda C, Kennedy GD, Nukaya M, Mantovani A, Kopan R, Bradfield CA, Newberry RD, Colonna M. AHR drives the development of gut ILC22 cells and postnatal lymphoid tissues via pathways dependent on and independent of Notch. Nat.Immunol. 2012;13:144–151. doi: 10.1038/ni.2187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Schiavoni G, Mattei F, Sestili P, Borghi P, Venditti M, Morse HC, III, Belardelli F, Gabriele L. ICSBP is essential for the development of mouse type I interferon-producing cells and for the generation and activation of CD8alpha(+) dendritic cells. J Exp.Med. 2002;196:1415–1425. doi: 10.1084/jem.20021263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Tamura T, Tailor P, Yamaoka K, Kong HJ, Tsujimura H, O’Shea JJ, Singh H, Ozato K. IFN regulatory factor-4 and -8 govern dendritic cell subset development and their functional diversity. J Immunol. 2005;174:2573–2581. doi: 10.4049/jimmunol.174.5.2573. [DOI] [PubMed] [Google Scholar]
  • 78.Hacker C, Kirsch RD, Ju XS, Hieronymus T, Gust TC, Kuhl C, Jorgas T, Kurz SM, Rose-John S, Yokota Y, Zenke M. Transcriptional profiling identifies Id2 function in dendritic cell development. Nat Immunol. 2003;4:380–386. doi: 10.1038/ni903. [DOI] [PubMed] [Google Scholar]
  • 79.Kashiwada M, Pham NL, Pewe LL, Harty JT, Rothman PB. NFIL3/E4BP4 is a key transcription factor for CD8{alpha}+ dendritic cell development. Blood. 2011;117:6193–6197. doi: 10.1182/blood-2010-07-295873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Hildner K, Edelson BT, Purtha WE, Diamond M, Matsushita H, Kohyama M, Calderon B, Schraml BU, Unanue ER, Diamond MS, Schreiber RD, Murphy TL, Murphy KM. Batf3 deficiency reveals a critical role for CD8alpha+ dendritic cells in cytotoxic T cell immunity. Science. 2008;322:1097–1100. doi: 10.1126/science.1164206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Watchmaker PB, Lahl K, Lee M, Baumjohann D, Morton J, Kim SJ, Zeng R, Dent A, Ansel KM, Diamond B, Hadeiba H, Butcher EC. Comparative transcriptional and functional profiling defines conserved programs of intestinal DC differentiation in humans and mice. Nat.Immunol. 2014;15:98–108. doi: 10.1038/ni.2768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Tsujimura H, Tamura T, Ozato K. Cutting edge: IFN consensus sequence binding protein/IFN regulatory factor 8 drives the development of type I IFN-producing plasmacytoid dendritic cells. J Immunol. 2003;170:1131–1135. doi: 10.4049/jimmunol.170.3.1131. [DOI] [PubMed] [Google Scholar]
  • 83.Tussiwand R, Lee WL, Murphy TL, Mashayekhi M, Wumesh KC, Albring JC, Satpathy AT, Rotondo JA, Edelson BT, Kretzer NM, Wu X, Weiss LA, Glasmacher E, Li P, Liao W, Behnke M, Lam SS, Aurthur CT, Leonard WJ, Singh H, Stallings CL, Sibley LD, Schreiber RD, Murphy KM. Compensatory dendritic cell development mediated by BATF-IRF interactions. Nature. 2012;490:502–507. doi: 10.1038/nature11531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Murphy TL, Tussiwand R, Murphy KM. Specificity through cooperation: BATF-IRF interactions control immune-regulatory networks. Nat.Rev.Immunol. 2013;13:499–509. doi: 10.1038/nri3470. [DOI] [PubMed] [Google Scholar]
  • 85.Gazzinelli RT, Wysocka M, Hayashi S, Denkers EY, Hieny S, Caspar P, Trinchieri G, Sher A. Parasite-induced IL-12 stimulates early IFN-gamma synthesis and resistance during acute infection with Toxoplasma gondii. J Immunol. 1994;153:2533–2543. [PubMed] [Google Scholar]
  • 86.Gazzinelli RT, Hieny S, Wynn TA, Wolf S, Sher A. Interleukin 12 is required for the T-lymphocyte-independent induction of interferon gamma by an intracellular parasite and induces resistance in T-cell-deficient hosts. Proc.Natl.Acad.Sci U.S A. 1993;90:6115–6119. doi: 10.1073/pnas.90.13.6115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Sher A, Oswald IP, Hieny S, Gazzinelli RT. Toxoplasma gondii induces a T-independent IFN-gamma response in natural killer cells that requires both adherent accessory cells and tumor necrosis factor-alpha. J.Immunol. 1993;150:3982–3989. [PubMed] [Google Scholar]
  • 88.Reis e Sousa C, Hieny S, Scharton-Kersten T, Jankovic D, Charest H, Germain RN, Sher A. In vivo microbial stimulation induces rapid CD40 ligand-independent production of interleukin 12 by dendritic cells and their redistribution to T cell areas. J.Exp.Med. 1997;186:1819–1829. doi: 10.1084/jem.186.11.1819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Koblansky AA, Jankovic D, Oh H, Hieny S, Sungnak W, Mathur R, Hayden MS, Akira S, Sher A, Ghosh S. Recognition of Profilin by Toll-like Receptor 12 Is Critical for Host Resistance to Toxoplasma gondii. Immunity. 2013;38:119–130. doi: 10.1016/j.immuni.2012.09.016. These studies (Ref. 89 and 90) show TLR11/12 recognize the antigen profilin on Toxoplasma gondii .
  • 90. Raetz M, Kibardin A, Sturge CR, Pifer R, Li H, Burstein E, Ozato K, Larin S, Yarovinsky F. Cooperation of TLR12 and TLR11 in the IRF8-dependent IL-12 response to Toxoplasma gondii profilin. J.Immunol. 2013;191:4818–4827. doi: 10.4049/jimmunol.1301301. These studies (Ref. 89 and 90) show TLR11/12 recognize the antigen profilin on Toxoplasma gondii .
  • 91.Kawano T, Cui J, Koezuka Y, Toura I, Kaneko Y, Motoki K, Ueno H, Nakagawa R, Sato H, Kondo E, Koseki H, Taniguchi M. CD1d–restricted and TCR-mediated activation of valpha14 NKT cells by glycosylceramides. Science. 1997;278:1626–1629. doi: 10.1126/science.278.5343.1626. [DOI] [PubMed] [Google Scholar]
  • 92.Lantz O, Bendelac A. An invariant T cell receptor alpha chain is used by a unique subset of major histocompatibility complex class I-specific CD4+ and CD4-8- T cells in mice and humans. J.Exp.Med. 1994;180:1097–1106. doi: 10.1084/jem.180.3.1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Bendelac A, Lantz O, Quimby ME, Yewdell JW, Bennink JR, Brutkiewicz RR. CD1 recognition by mouse NK1+ T lymphocytes. Science. 1995;268:863–865. doi: 10.1126/science.7538697. [DOI] [PubMed] [Google Scholar]
  • 94.Chun T, Page MJ, Gapin L, Matsuda JL, Xu H, Nguyen H, Kang HS, Stanic AK, Joyce S, Koltun WA, Chorney MJ, Kronenberg M, Wang CR. CD1d–expressing dendritic cells but not thymic epithelial cells can mediate negative selection of NKT cells. J.Exp.Med. 2003;197:907–918. doi: 10.1084/jem.20021366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Fujii S, Shimizu K, Smith C, Bonifaz L, Steinman RM. Activation of natural killer T cells by alpha-galactosylceramide rapidly induces the full maturation of dendritic cells in vivo and thereby acts as an adjuvant for combined CD4 and CD8 T cell immunity to a coadministered protein. J.Exp.Med. 2003;198:267–279. doi: 10.1084/jem.20030324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Fujii S, Liu K, Smith C, Bonito AJ, Steinman RM. The linkage of innate to adaptive immunity via maturing dendritic cells in vivo requires CD40 ligation in addition to antigen presentation and CD80/86 costimulation. J.Exp.Med. 2004;199:1607–1618. doi: 10.1084/jem.20040317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Fujii S, Shimizu K, Kronenberg M, Steinman RM. Prolonged IFN-gamma-producing NKT response induced with alpha-galactosylceramide-loaded DCs. Nat.Immunol. 2002;3:867–874. doi: 10.1038/ni827. [DOI] [PubMed] [Google Scholar]
  • 98. Arora P, Baena A, Yu KO, Saini NK, Kharkwal SS, Goldberg MF, Kunnath-Velayudhan S, Carreno LJ, Venkataswamy MM, Kim J, Lazar-Molnar E, Lauvau G, Chang YT, Liu Z, Bittman R, Al Shamkhani A, Cox LR, Jervis PJ, Veerapen N, Besra GS, Porcelli SA. A single subset of dendritic cells controls the cytokine bias of natural killer T cell responses to diverse glycolipid antigens. Immunity. 2014;40:105–116. doi: 10.1016/j.immuni.2013.12.004. This study shows that Batf3-dependent dendritic cells mediate iNKT responses.
  • 99. Bajana S, Roach K, Turner S, Paul J, Kovats S. IRF4 promotes cutaneous dendritic cell migration to lymph nodes during homeostasis and inflammation. J.Immunol. 2012;189:3368–3377. doi: 10.4049/jimmunol.1102613. Irf4 deficiency impairs migration of peripheral CD11b+ dendritic cells to lymph nodes.
  • 100. Schlitzer A, McGovern N, Teo P, Zelante T, Atarashi K, Low D, Ho AW, See P, Shin A, Wasan PS, Hoeffel G, Malleret B, Heiseke A, Chew S, Jardine L, Purvis HA, Hilkens CM, Tam J, Poidinger M, Stanley ER, Krug AB, Renia L, Sivasankar B, Ng LG, Collin M, Ricciardi-Castagnoli P, Honda K, Haniffa M, Ginhoux F. IRF4 Transcription Factor-Dependent CD11b(+) Dendritic Cells in Human and Mouse Control Mucosal IL-17 Cytokine Responses. Immunity. 2013;38:970–983. doi: 10.1016/j.immuni.2013.04.011. These two studies (Ref. 100 and 101) show that CD11b+ DCs require Irf4 to drive TH17 polarization.
  • 101. Persson EK, Uronen-Hansson H, Semmrich M, Rivollier A, Hagerbrand K, Marsal J, Gudjonsson S, Hakansson U, Reizis B, Kotarsky K, Agace WW. IRF4 Transcription-Factor-Dependent CD103(+)CD11b(+) Dendritic Cells Drive Mucosal T Helper 17 Cell Differentiation. Immunity. 2013;38:958–969. doi: 10.1016/j.immuni.2013.03.009. These two studies (Ref. 100 and 101) show that CD11b+ DCs require Irf4 to drive TH17 polarization.
  • 102.Lewis KL, Caton ML, Bogunovic M, Greter M, Grajkowska LT, Ng D, Klinakis A, Charo IF, Jung S, Gommerman JL, Ivanov II, Liu K, Merad M, Reizis B. Notch2 receptor signaling controls functional differentiation of dendritic cells in the spleen and intestine. Immunity. 2011;35:780–791. doi: 10.1016/j.immuni.2011.08.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Wu L, D’Amico A, Winkel KD, Suter M, Lo D, Shortman K. RelB is essential for the development of myeloid-related CD8alpha- dendritic cells but not of lymphoid-related CD8alpha+ dendritic cells. Immunity. 1998;9:839–847. doi: 10.1016/s1074-7613(00)80649-4. [DOI] [PubMed] [Google Scholar]
  • 104.Kobayashi T, Walsh PT, Walsh MC, Speirs KM, Chiffoleau E, King CG, Hancock WW, Caamano JH, Hunter CA, Scott P, Turka LA, Choi Y. TRAF6 is a critical factor for dendritic cell maturation and development. Immunity. 2003;19:353–363. doi: 10.1016/s1074-7613(03)00230-9. [DOI] [PubMed] [Google Scholar]
  • 105.Kabashima K, Banks TA, Ansel KM, Lu TT, Ware CF, Cyster JG. Intrinsic lymphotoxin-beta receptor requirement for homeostasis of lymphoid tissue dendritic cells. Immunity. 2005;22:439–450. doi: 10.1016/j.immuni.2005.02.007. [DOI] [PubMed] [Google Scholar]
  • 106. Varol C, Vallon-Eberhard A, Elinav E, Aychek T, Shapira Y, Luche H, Fehling HJ, Hardt WD, Shakhar G, Jung S. Intestinal lamina propria dendritic cell subsets have different origin and functions. Immunity. 2009;31:502–512. doi: 10.1016/j.immuni.2009.06.025. These studies (ref. 11 and 106) show DC progenitors and monocytes give rise to CD11c+ cells in the small intestine lamina propria.
  • 107. Schreiber HA, Loschko J, Karssemeijer RA, Escolano A, Meredith MM, Mucida D, Guermonprez P, Nussenzweig MC. Intestinal monocytes and macrophages are required for T cell polarization in response to Citrobacter rodentium. J.Exp.Med. 2013;210:2025–2039. doi: 10.1084/jem.20130903. This study uses a novel transgenic mouse model in which monocytes and macrophages are selectively depleted to elucidate the origin of CD11c+ cells in the small intestine.
  • 108. Zigmond E, Varol C, Farache J, Elmaliah E, Satpathy AT, Friedlander G, Mack M, Shpigel N, Boneca IG, Murphy KM, Shakhar G, Halpern Z, Jung S. Ly6C hi monocytes in the inflamed colon give rise to proinflammatory effector cells and migratory antigen-presenting cells. Immunity. 2012;37:1076–1090. doi: 10.1016/j.immuni.2012.08.026. This study provides evidence for the differentiation of monocytes into antigen presenting cells.
  • 109.Zelenay S, Keller AM, Whitney PG, Schraml BU, Deddouche S, Rogers NC, Schulz O, Sancho D, Reis Sousa E. The dendritic cell receptor DNGR-1 controls endocytic handling of necrotic cell antigens to favor cross-priming of CTLs in virus-infected mice. J.Clin.Invest. 2012;122:1615–1627. doi: 10.1172/JCI60644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Zheng Y, Valdez PA, Danilenko DM, Hu Y, Sa SM, Gong Q, Abbas AR, Modrusan Z, Ghilardi N, de Sauvage FJ, Ouyang W. Interleukin-22 mediates early host defense against attaching and effacing bacterial pathogens. Nat.Med. 2008;14:282–289. doi: 10.1038/nm1720. This study shows IL-22 is necessary for protection against C. rodentium infection.
  • 111.Simmons CP, Clare S, Ghaem-Maghami M, Uren TK, Rankin J, Huett A, Goldin R, Lewis DJ, Macdonald TT, Strugnell RA, Frankel G, Dougan G. Central role for B lymphocytes and CD4+ T cells in immunity to infection by the attaching and effacing pathogen Citrobacter rodentium. Infect.Immun. 2003;71:5077–5086. doi: 10.1128/IAI.71.9.5077-5086.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Mundy R, Macdonald TT, Dougan G, Frankel G, Wiles S. Citrobacter rodentium of mice and man. Cell Microbiol. 2005;7:1697–1706. doi: 10.1111/j.1462-5822.2005.00625.x. [DOI] [PubMed] [Google Scholar]
  • 113.Ohl L, Mohaupt M, Czeloth N, Hintzen G, Kiafard Z, Zwirner J, Blankenstein T, Henning G, Forster R. CCR7 governs skin dendritic cell migration under inflammatory and steady-state conditions. Immunity. 2004;21:279–288. doi: 10.1016/j.immuni.2004.06.014. [DOI] [PubMed] [Google Scholar]
  • 114.Ghoreschi K, Laurence A, Yang XP, Tato CM, McGeachy MJ, Konkel JE, Ramos HL, Wei L, Davidson TS, Bouladoux N, Grainger JR, Chen Q, Kanno Y, Watford WT, Sun HW, Eberl G, Shevach EM, Belkaid Y, Cua DJ, Chen W, O’Shea JJ. Generation of pathogenic T(H)17 cells in the absence of TGF-beta signalling. Nature. 2010;467:967–971. doi: 10.1038/nature09447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Welty NE, Staley C, Ghilardi N, Sadowsky MJ, Igyarto BZ, Kaplan DH. Intestinal lamina propria dendritic cells maintain T cell homeostasis but do not affect commensalism. J.Exp.Med. 2013 doi: 10.1084/jem.20130728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Bettelli E, Carrier Y, Gao W, Korn T, Strom TB, Oukka M, Weiner HL, Kuchroo VK. Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature. 2006;441:235–238. doi: 10.1038/nature04753. [DOI] [PubMed] [Google Scholar]
  • 117.Mangan PR, Harrington LE, O’Quinn DB, Helms WS, Bullard DC, Elson CO, Hatton RD, Wahl SM, Schoeb TR, Weaver CT. Transforming growth factor-beta induces development of the T(H)17 lineage. Nature. 2006;441:231–234. doi: 10.1038/nature04754. [DOI] [PubMed] [Google Scholar]
  • 118. Basu R, O’Quinn DB, Silberger DJ, Schoeb TR, Fouser L, Ouyang W, Hatton RD, Weaver CT. Th22 cells are an important source of IL-22 for host protection against enteropathogenic bacteria. Immunity. 2012;37:1061–1075. doi: 10.1016/j.immuni.2012.08.024. This study shows that IL-22 production by T cells during C. rodentium is dependent on IL-6 and not IL-23.
  • 119.Zheng Y, Danilenko DM, Valdez P, Kasman I, Eastham-Anderson J, Wu J, Ouyang W. Interleukin-22, a T(H)17 cytokine, mediates IL-23-induced dermal inflammation and acanthosis. Nature. 2007;445:648–651. doi: 10.1038/nature05505. [DOI] [PubMed] [Google Scholar]
  • 120.Liang SC, Tan XY, Luxenberg DP, Karim R, Dunussi-Joannopoulos K, Collins M, Fouser LA. Interleukin (IL)-22 and IL-17 are coexpressed by Th17 cells and cooperatively enhance expression of antimicrobial peptides. J.Exp.Med. 2006;203:2271–2279. doi: 10.1084/jem.20061308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Khan MA, Bouzari S, Ma C, Rosenberger CM, Bergstrom KS, Gibson DL, Steiner TS, Vallance BA. Flagellin-dependent and -independent inflammatory responses following infection by enteropathogenic Escherichia coli and Citrobacter rodentium. Infect.Immun. 2008;76:1410–1422. doi: 10.1128/IAI.01141-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Phythian-Adams AT, Cook PC, Lundie RJ, Jones LH, Smith KA, Barr TA, Hochweller K, Anderton SM, Hammerling GJ, Maizels RM, MacDonald AS. CD11c depletion severely disrupts Th2 induction and development in vivo. J.Exp.Med. 2010;207:2089–2096. doi: 10.1084/jem.20100734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Hammad H, Plantinga M, Deswarte K, Pouliot P, Willart MA, Kool M, Muskens F, Lambrecht BN. Inflammatory dendritic cells--not basophils--are necessary and sufficient for induction of Th2 immunity to inhaled house dust mite allergen. J Exp.Med. 2010;207:2097–2111. doi: 10.1084/jem.20101563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Mesnil C, Sabatel CM, Marichal T, Toussaint M, Cataldo D, Drion PV, Lekeux P, Bureau F, Desmet CJ. Resident CD11b(+)Ly6C(−) lung dendritic cells are responsible for allergic airway sensitization to house dust mite in mice. PLoS One. 2012;7:e53242. doi: 10.1371/journal.pone.0053242. These studies (Ref. 124 and 125) identify the role of CD11b+ DCs in mediating allergic responses in the lung.
  • 125. Plantinga M, Guilliams M, Vanheerswynghels M, Deswarte K, Branco-Madeira F, Toussaint W, Vanhoutte L, Neyt K, Killeen N, Malissen B, Hammad H, Lambrecht BN. Conventional and Monocyte-Derived CD11b(+) Dendritic Cells Initiate and Maintain T Helper 2 Cell-Mediated Immunity to House Dust Mite Allergen. Immunity. 2013 doi: 10.1016/j.immuni.2012.10.016. These studies (Ref. 124 and 125) identify the role of CD11b+ DCs in mediating allergic responses in the lung.
  • 126.Williams JW, Tjota MY, Clay BS, Vander LB, Bandukwala HS, Hrusch CL, Decker DC, Blaine KM, Fixsen BR, Singh H, Sciammas R, Sperling AI. Transcription factor IRF4 drives dendritic cells to promote Th2 differentiation. Nat.Commun. 2013;4:2990. doi: 10.1038/ncomms3990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Kumamoto Y, Linehan M, Weinstein JS, Laidlaw BJ, Craft JE, Iwasaki A. CD301b(+) dermal dendritic cells drive T helper 2 cell-mediated immunity. Immunity. 2013;39:733–743. doi: 10.1016/j.immuni.2013.08.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Gao Y, Nish SA, Jiang R, Hou L, Licona-Limon P, Weinstein JS, Zhao H, Medzhitov R. Control of T helper 2 responses by transcription factor IRF4-dependent dendritic cells. Immunity. 2013;39:722–732. doi: 10.1016/j.immuni.2013.08.028. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES