Skip to main content
Mediators of Inflammation logoLink to Mediators of Inflammation
. 2013 Nov 28;2013:731023. doi: 10.1155/2013/731023

Interferon-Regulatory Factors Determine Macrophage Phenotype Polarization

Roman Günthner 1, Hans-Joachim Anders 1,*
PMCID: PMC3863528  PMID: 24379524

Abstract

The mononuclear phagocyte system regulates tissue homeostasis as well as all phases of tissue injury and repair. To do so changing tissue environments alter the phenotype of tissue macrophages to assure their support for sustaining and amplifying their respective surrounding environment. Interferon-regulatory factors are intracellular signaling elements that determine the maturation and gene transcription of leukocytes. Here we discuss how several among the 9 interferon-regulatory factors contribute to macrophage polarization.

1. Introduction

During development mononuclear phagocyte progenitors populate most tissues where they differentiate into transcriptionally and functionally diverse phenotypes [13]; for example, bone marrow, liver, and lung harbor macrophages with an enormous capacity to clear airborne particles, gut-derived pathogens, or cell nuclei expelled from erythroblasts, respectively [4]. In contrast, skin, kidney, and brain host a dense network of dendritic cells [4, 5]. Upon tissue injury M-CSF drives resident mononuclear phagocyte to proliferate [6] or circulating monocytes recruit to the site of injury. It is the local microenvironment that then determines mononuclear phagocyte polarization to distinct phenotypes, which can vary between disorders or between the different stages of a disease process [7]. Several factors mediate mononuclear phagocyte polarization, as being mostly described by in vitro experiments [7, 8]. However, attempts to translate this simplistic model to disease states in vivo often failed to cover all aspects of heterogeneous and changing tissue environments. For example, ischemia-reperfusion injury induces transient sterile inflammation because dying tissue cells release damage-associated molecular patterns (DAMPs) that polarize macrophages toward a classically activated M1-like phenotype [9, 10]. This process is associated with NF-κB and STAT1 pathway activation [2]. Macrophages apoptosis or their phenotype switches towards alternatively activated, M2-like macrophages that produce IL-10 and TGF-β, induce resolution of inflammation, and enforce tissue regeneration [1115]. Failure of this phenotype switch leads to persistent tissue inflammation, atrophy, and fibrosis [16]. The uptake of neutrophils, epithelium-derived alarmins, and Th2 cytokines IL-4 and IL-13 supports this phenotype switch [11]. As disease processes do not always occur in a serial manner, concomitant proinflammatory and anti-inflammatory macrophages infiltrates often populate organs affected by persistent injury, for example, in slowly progressive lesions of organ transplants [17, 18].

Current data suggest that the family of the interferon-regulatory factors (IRFs) plays an important role in regulating macrophage polarization. IRFs are intracellular proteins that regulate immune cell maturation [19]. Here we provide a summary on IRF biology that is focused on the IRF's role in macrophage phenotype control and the associated contributions to tissue inflammation and remodeling.

2. The Family of Interferon-Regulatory Factors

The IRFs were discovered in search of transcription factors that bind to the conserved virus response elements within the promoters of type I IFN genes [19]. It was found that both NF-κB and IRF-3 activate IFN-β gene transcription while IFN-α gene expression is entirely based on IRFs [19]. The generation of Irf-deficient mice led to the discovery of additional regulatory roles of the IRFs for cell growth, for immune cell maturation and activation, and for apoptosis. In mammals the IRF gene family consists of nine members: IRF-1, IRF-2, IRF-3, IRF-4, IRF-5, IRF-6, IRF-7, IRF-8/ICSBP, and IRF-9. Their respective IRF proteins share significant homologies at the N-terminal 115 amino acids where they share a conserved tryptophan pentad repeat DNA-binding domain [20]. These include a DNA-binding domain of five tryptophan repeats of which three recognize the GAAA and AANNNGAA sequence motifs, that is, the IFN-stimulated response elements [20]. However, the variable domains at the C-terminus determine the functional specificity of the nine IRFs, their potential to interact with each other via IRF-association domains, and their cell type-specific actions [21] (Figure 1). Accordingly, the IRFs have been subdivided into the “interferonic” IRFs (IRF-2, -3, -7, and -9), the “stress-responsive” IRFs (IRF-1 and -5), the “hematopoietic” IRFs (IRF-4 and -8), and the “morphogenic” IRF-6 [22]. The genetic and biological characteristics of the IRF family members are listed in Table 1.

Figure 1.

Figure 1

Structural domain organization and important posttranslational modifications of IRFs. Proteins are illustrated by N-terminus on the left and C-terminus on the right. Each of the nine IRFs consists of a conserved pentad repeat DNA-binding domain. Regulatory and repression domains are mostly located in the C-terminal domain. IRF-association domains 1/2(IADs) mediate the interaction with other IRF-family members. Yellow arrows indicate the phosphorylation site within the domain. Posttranslational modifications are illustrated in the right column. Numbers of amino acids for each IRF are given next to structural scheme.

Table 1.

Interferon-regulator factors and macrophage polarization.

Chromosome Expression Effect on macrophages Favoured macrophage phenotype
IRF-1 5q31 Ubiquitous Induced by IFN-γ, mediates TLR/MyD88 signaling, interacts with NF-κB (“enhanceosome”), and suppresses IL-4 promotor M1

IRF-2 4q34 Ubiquitous Suppression of IRF-1-mediated IFN and Cox-2 induction, complex role in LPS-induced cytokine release, and suppression of STAT1/3 signalling Context-dependent

IRF-3 19q13 Ubiquitous Promotes TRIF-signalling, drives IFN-γ/IL-1-mediated IL-10 secretion M2

IRF-4 6p25 Hematopoietic cells Induced by IL-4 via Jmjd3, inhibits MyD88 signalling by blocking IRF-5/MyD88 interaction, promotes IL-4 and IL-10 secretion M2

IRF-5 7q32 Ubiquitous Interacts with MyD88 needed for MyD88 signalling, drives IL-12p35 and IL-23p19 secretion M1

IRF-6 1q32 Keratinocytes

IRF-7 11p15 Ubiquitous Type I interferon induction

IRF-8 16q24 Hematopoietic cells Induced by IFN-γ, mediates TLR-mediated induction of IFN-β, IL-12p40, IL-12p35, and iNOS, and mediates Notch and TLR signalling for M1 polarization M1

IRF-9 14q11 Ubiquitous Regulates type I interferon signalling

TLR: Toll-like receptor, IL: interleukin, IFN: interferon, Cox: cyclooxygenase, LPS: lipopolysaccharide, iNOS: inducible NO synthase.

3. IRFs in Macrophage Polarization

3.1. IRF-1

IRF-1 was first described in 1980s as a 325-amino acid-long nonredundant transcription factor for type I IFNs upon TLR3 ligation [2325]. IRF-1 is only weakly expressed in resting DCs and macrophages but is induced by IFN-γ up to 8-fold in M1 polarized macrophages [26]. IRF-1 interacts with MyD88 to migrate into the nucleus where it triggers TLR-mediated expression of proinflammatory genes [27, 28]. Casein kinase II activates IRF-1 by phosphorylation [29]. The protein complex formed by IRF-1, NF-κB, and Jun that bind to the IFN-β promotor was named “enhanceosome” [28, 3032]. Sumoylation represses the transcriptional activity of IRF-1 [33]. LPS challenge requires IRF-1 to induce TLR3, TLR6, and TLR9 in macrophages [34]. In fact, Irf1-deficient macrophages almost entirely lack inducible nitric oxide synthase (iNOS) production upon LPS and IFN-γ stimulation [35]. This way, IRF-1 contributes to the priming of classically activated, M1-like macrophage polarization in inflammatory tissue environments that involve IFN-γ-producing NKT cells or Th1 T cells [36]. At the same time, IRF-1 suppresses the binding of other transcription factors to the IL-4 promoter, which inhibits alternative macrophage activation [37]. This process supports host defense against intracellular pathogens but also accounts for M1 macrophage-related immunopathology [35, 36, 38]. The latter is particularly evident in sterile inflammation, for example, in ischemia-reperfusion injury [39, 40].

3.2. IRF-2

IRF-2 is 349-amino acid-long and displays considerable sequence homology with IRF-1 [23]. IRF-2 competes with IRF-1 for the same cis-acting recognition sequences in gene promoters [41]. Hence, IRF-2 is a negative regulator of IRF-1-mediated type I IFN and Cox-2 induction [23, 31]. IRF-2 has a more complex role in cytokine regulation as it suppresses LPS-induced TNF expression while augmenting LPS-induced IL-1, IL-6, IL-12, and IFN-γ secretion [42]. Sumoylation increases IRF-2's ability to inhibit IRF-1 transcriptional activity [43]. LPS challenge regulates TLR3, TLR4, and TLR5 via IRF-2 in macrophages [34]. IRF-2 suppresses caspase-1-mediated programmed cell death by interfering with the transcriptional regulation of caspase-1 and by suppressing STAT1/3 signaling [44]. Irf-2-deficient mice are highly susceptible to Listeria monocytogenes infection, which seems to be related to IRF-2's role in mediating the IFN-γ-induced oxidative burst that kills the pathogen inside intracellular compartments of macrophages [45]. However, this was iNOS transcription independent. IRF-2 rather regulates iNOS in a posttranscriptional manner [46]. The net effect of IRF-2 on sterile inflammation seems to be immunosuppressive as Irf-2-deficient mice are more susceptible to lymphocytic choriomeningitis virus infection as well as to ischemia-reperfusion injury-related tissue inflammation while that latter was suppressed in mice that overexpress IRF-2 [47]. IRF-2's negative regulatory effect on type I IFN expression also suppresses inflammatory skin disease involving CD8 T cells [48]. In addition, IRF-2 is needed for the development of splenic and epidermal CD4+ dendritic cells [49].

3.3. IRF-3

IRF-3 was discovered by searching genes with homology sequences with IRF-1 and IRF-2 [50]. This 427-amino acid protein shares a number of characteristics with IRF-7 [51]. Unlike IRF-7, that confers MyD88 signaling, IRF-3 is involved in TRIF-dependent signaling pathways. After binding pathogens, pattern-recognition receptors like TLR-3, TLR-4, or RIG-I recruit TRIF to trigger an IRF-3-mediated induction of type I IFNs [5255]. Additional cytoplasmic DNA recognition receptors use the STING pathway to activate IRF-3 [56]. The transcriptional activation of the IFN-β gene requires an enhanceosome of 7 additional proteins that create a continuous surface that recognizes the DNA-binding element [57]. Phosphorylation of TLR3's specific tyrosine residues can initiate two distinct signaling pathways. One activates TBK-1 and the other activates PI3 kinase and Akt for full phosphorylation and activation of IRF-3 [58, 59]. Cytoplasmic IRF-3 is inactive unless phosphoactivation of IRF-3 triggers unfolding of the autoinhibitory elements and exposes the hydrophobic surface to interaction with CREBBP to translocate to the nucleus [60]. By contrast, ubiquitination inactivates IRF-3 [61]. GM-CSF-primed M1-like macrophages display a diminished IRF-3 axis and enhanced activation of MyD88. In contrast, M-CSF stimulated macrophages that develop an M2-like phenotype show defective NF-κB activation and enhanced TRIF-mediated IRF-3 induction upon LPS stimulation [62, 63]. Hence, the IRF-3 axis is rather enabled in M2-like macrophages than in M1-like macrophages. But does IRF-3 also contribute to the development of an alternatively activated macrophage phenotype? One study transduced IRF-3 into primary human microglia. Stimulation with IFN-γ/IL-1 suppressed proinflammatory mediators like IL-6, TNF-α, or IL-1β, whereas anti-inflammatory mediators including IL-10 were enhanced [64]. Altogether the data suggest that IRF-3 is associated with anti-inflammatory microenvironments and contributes to the polarization toward a M2 macrophage phenotype. However, IRF-3 also induces a number of inflammatory cytokines such as CCL5 and IFN-β [65].

3.4. IRF-4

IRF-4, first described in 1995, is a 450-amino acid-long “hematopoietic” protein with considerable homology with IRF-1 and IRF-2 [66]. IRF-4 contributes to the maturation of multiple myeloid and lymphoid cell types from their lineage-specific progenitors [19, 67]. IRF-4 competes with IRF-5 for binding to the adaptor MyD88 that transmits TLR outside-in signaling to NF-κB and other proinflammatory transcription factors [27]. As IRF-5 is needed for signal transduction the competitive action of IRF-4 for MyD88 binding renders IRF-4 an endogenous TLR signaling antagonist that can suppress M1 macrophage polarization [68]. IL-10 induction needs IRF-4 and IRF-4 overexpression enhances IL-4 and IL-10 secretion [69]. On the contrary, IRF-4−/− mice are more sensitive to LPS-induced sepsis and exhibit higher production of proinflammatory cytokines like TNF and IL-6 [70]. IL-4 induces macrophages to upregulate IRF-4 and contributes to their M2 polarization [71]. Accordingly, IRF-4 deficiency leads to decreased expression of M2 marker genes like Arg1, Ym1, and Fizz1 [72]. In fact, Jumonji domain-containing-3 (Jmjd3), a histone 3 Lys27 (3K27) demethylase, regulates the trimethylation at H3K27 of a selected number of genes including IRF-4. This mechanism controls IRF-4 induction and is needed for M2 macrophage polarization, for example, in the host defense during helminth infection [72]. Interestingly, IL-4-induced STAT6 signaling regulates Jmjd3 [73]. Hence, polarization of alternatively activated macrophages through IL-4 seems to be mediated via STAT6-Jmjd3-IRF-4 signaling and reveals an essential role of IRF-4 in macrophage polarization for helminth control.

3.5. IRF-5

IRF-5 is a 504-amino acid-long stress-responsive IRF [22]. IRF-5 is required for TLR-mediated induction of IL-6, TNF, IL-12, and other proinflammatory cytokines [74]. IRF-5 competes with IRF-4 for binding to the signaling adapter MyD88 and the downstream subsequent activation of proinflammatory transcription factors [27]. Its capacity to induce inflammatory cytokines and B cell transcription factors implies its role in host defense and autoimmune disorders [75, 76]. This competitive interaction involves IRF-5 in the polarization into M1 macrophages [68]. In fact, the balance between IRF-4 and IRF-5 seems to be a major determinant of M1 versus M2 macrophage polarization. For example, M-CSF induces IRF-4 in human monocyte-derived macrophages while GM-CSF induces IRF-5, which results in two phenotypically different macrophage phenotypes [77]. M1 macrophages express high levels of IRF-5 where it not only mediates the expression of proinflammatory cytokines but also suppresses the immunoregulatory cytokine IL-10 [68]. IRF-5 itself is regulated by the transcriptional corepressor KAP1/TRIM28 to avoid overshooting secretion of TNF and other mediators that induce immunopathology [78]. KAP1/TRIM28 regulates IRF-5 by recruiting histone deacetylases and methyltransferases that can silence IRF-5-related gene expression [78]. IRF-5-mediated polarization of monocytic phagocytes involves the secretion of various IL-12 family members including IL-12p35 and IL-23p19, which support Th17 T cell immunity, an element of adaptive immunity that contributes to autoimmune disorders [79]. In fact, gain of function mutations in the IRF-5 gene exists that increases TLR- or NOD-mediated secretion of proinflammatory cytokines [80]. Such variants also predispose to autoimmune diseases like systemic lupus erythematous [8183], which may be related to these phenomena.

3.6. IRF-6

IRF-6 is a so-called “morphogenic” IRF of 467 amino acid length. IRF-6 has a large structural homology with IRF-5 but does not seem to share its functional properties or contribute to macrophage biology, which is related to the tissue-specific expression of IRF-6. IRF-6 mutations rather predispose to cleft lip or palate and other abnormalities of limb, skin, and craniofacial morphogenesis [84, 85].

3.7. IRF-7

IRF-7, together with IRF-3, is a 503-amino acid central and nonredundant mediator of viral nucleic acid-induced induction of IFN-α [19, 86, 87]. IRF-7 drives the differentiation of monocytes to macrophages but a direct role in macrophage polarization has not been reported.

3.8. IRF-8

IRF-8, also known as interferon consensus sequence-binding protein (ICSBP), is a 393-amino acid-long “hematopoietic” IRF [22]. IRF-8 (like IRF-4) has a dominant role in the maturation and differentiation of monocytes and macrophages from their immature progenitors, while it represses neutrophil production [8890]. IFN-γand LPS slow down the intrinsic mobility of IRF8 inside the nucleus to enforce its chromatin interaction for the initiation of transcription [91]. IFN-γinduces IRF-8 and IRF-8 drives to induction of IFN-β, IL-12p40, IL-12p35, and iNOS upon TLR stimulation, that is, M1 macrophage gene profile [92]. In addition, IRF-8 integrates outside-in signaling of Notch receptors and TLRs for the induction of genes that define an M1 macrophage phenotype [93]. IRF-8 selectively modulates TLR4 signaling via IRAK2-dependent activation of MNK1 and eIF4E-regulated translation. IRF-8 itself is regulated by small ubiquitin-like modifiers (SUMO) 2/3 at the lysine residue 310. SUMO3-conjugated IRF8 cannot induce IRF-8 target genes [94]. Upon macrophage activation, SUMOylation of IRF-8 is reduced as the deSUMOylating enzyme, sentrin-specific peptidase 1 (SENP1), inactivates SUMOylation-related IRF-8 repression. As such IRF-8 SUMO conjugation/deconjugation represents a previously unrecognized mechanism of macrophage phenotype control.

3.9. IRF-9

IRF-9 is a 424-amino acid-long regulator of type I IFN signaling. It forms a DNA-binding complex with the STAT1 homodimer, for example, for the induction of CXCL10 [95]. A specific role in macrophage polarization has not been reported.

4. Summary and Perspective

Macrophages contribute to tissue homeostasis and all phases of tissue injury and repair. Tissue environments prime macrophages to distinct phenotypes to assure that their functional properties enforce the surrounding environment, whether this may be inflammation, the resolution of inflammation, tissue repair (fibrosis), or the resolution of extracellular matrix. Members of the IRF family are an integral component of the macrophage polarization process and, hence, regulate the phenotypic plasticity and heterogeneity of tissue macrophages. Research in this area is still in progress, but our present working model refers to IRF-1, IRF-5, and IRF8 as factors driving the proinflammatory, classically activated (M1) macrophage phenotype, while IRF-3 and IRF-4 promote anti-inflammatory, alternatively activated (M2) macrophages (Figure 2). Future work in this area will certainly refine this concept and define additional functions of the IRF's in this context and elucidate additional mechanisms of how changing tissue environments shape immune effector cells to meet the tissue needs in homeostasis and disease.

Figure 2.

Figure 2

Working model of the role of interferon-regulatory factors in macrophage polarization. Circulating monocytes reach tissues by rolling and adhesion at luminal surfaces of activated endothelia, which is followed by transmigration into the interstitial tissue compartment. The local environment will prime M0 macrophage polarization, a process to which interferon-regulatory factors (IRFs) contribute in a phenotype-specific manner. See text for details.

Conflict of Interests

All authors do not have competing interests and do not have a financial relation to profiting companies or commercial products.

Acknowledgments

Hans-Joachim Anders was supported by the Deutsche Forschungsgemeinschaft AN371/15-1 and GRK1202. Roman Günthner was supported by the GRK1202.

Abbreviations

CCL:

Chemokine ligand

CCR:

Chemokine receptor

Cox-2:

Cyclooxygenase 2

CREBBP:

CREB-binding protein

DAMP:

Danger-associated molecular pattern

eIF4E:

Eukaryotic translation initiation factor 4E

GM-CSF:

Granulocyte-macrophage colony-stimulating factor

IFN:

Interferon

iNOS:

Inducible nitric oxide synthase

IRF:

Interferon-regulatory factor

Jmjd3:

Jumonji domain-containing-3

KAP1:

KRAB-associated protein-1

M-CSF:

Macrophage colony-stimulating factor

MNK1:

MAPK signal-integrating kinase 1

NO:

Nitric oxide

NOD:

Nucleotide-binding oligomerization domain-containing protein

PAMP:

Pathogen-associated molecular pattern

RIG-I:

Retinoic acid-inducible gene 1

SENP1:

Sentrin-specific protease 1

STAT:

Signal transducer and activator of transcription

SUMO:

Small ubiquitin-like modifier

STING:

Stimulator of interferon genes

TGF-β:

Transforming growth factor-β

TLR:

Toll-like receptor

TNF:

Tumor necrosis factor

TRIF:

TIR-domain-containing adapter-inducing interferon-β

TRIM28:

Tripartite motif-containing 28.

References

  • 1.Geissmann F, Manz MG, Jung S, Sieweke MH, Merad M, Ley K. Development of monocytes, macrophages, and dendritic cells. Science. 2010;327(5966):656–661. doi: 10.1126/science.1178331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Gordon S, Taylor PR. Monocyte and macrophage heterogeneity. Nature Reviews Immunology. 2005;5(12):953–964. doi: 10.1038/nri1733. [DOI] [PubMed] [Google Scholar]
  • 3.Gautier EL, Shay T, Miller J, et al. Gene-expression profiles and transcriptional regulatory pathways that underlie the identity and diversity of mouse tissue macrophages. Nature Immunology. 2012;13(11):1118–1128. doi: 10.1038/ni.2419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Wynn TA, Chawla A, Pollard JW. Macrophage biology in development, homeostasis and disease. Nature. 2013;496(7446):445–455. doi: 10.1038/nature12034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Soos TJ, Sims TN, Barisoni L, et al. CX3CR1+ interstitial dendritic cells form a contiguous network throughout the entire kidney. Kidney International. 2006;70(3):591–596. doi: 10.1038/sj.ki.5001567. [DOI] [PubMed] [Google Scholar]
  • 6.Jenkins SJ, Ruckerl D, Cook PC, et al. Local macrophage proliferation, rather than recruitment from the blood, is a signature of TH2 inflammation. Science. 2011;332(6035):1284–1288. doi: 10.1126/science.1204351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Murray PJ, Wynn TA. Protective and pathogenic functions of macrophage subsets. Nature Reviews Immunology. 2011;11(11):723–737. doi: 10.1038/nri3073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Sica A, Mantovani A. Macrophage plasticity and polarization: in vivo veritas. The Journal of Clinical Investigation. 2012;122(3):787–795. doi: 10.1172/JCI59643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Rock KL, Latz E, Ontiveros F, Kono H. The sterile inflammatory response. Annual Review of Immunology. 2010;28:321–342. doi: 10.1146/annurev-immunol-030409-101311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Swaminathan S, Griffin MD. First responders: understanding monocyte-lineage traffic in the acutely injured kidney. Kidney International. 2008;74(12):1509–1511. doi: 10.1038/ki.2008.555. [DOI] [PubMed] [Google Scholar]
  • 11.Gordon S. Alternative activation of macrophages. Nature Reviews Immunology. 2003;3(1):23–35. doi: 10.1038/nri978. [DOI] [PubMed] [Google Scholar]
  • 12.Anders H-J, Ryu M. Renal microenvironments and macrophage phenotypes determine progression or resolution of renal inflammation and fibrosis. Kidney International. 2011;80(9):915–925. doi: 10.1038/ki.2011.217. [DOI] [PubMed] [Google Scholar]
  • 13.Kim M-G, Su Boo C, Sook Ko Y, et al. Depletion of kidney CD11c+ F4/80+ cells impairs the recovery process in ischaemia/reperfusion-induced acute kidney injury. Nephrology Dialysis Transplantation. 2010;25(9):2908–2921. doi: 10.1093/ndt/gfq183. [DOI] [PubMed] [Google Scholar]
  • 14.Lee S, Huen S, Nishio H, et al. Distinct macrophage phenotypes contribute to kidney injury and repair. Journal of the American Society of Nephrology. 2011;22(2):317–326. doi: 10.1681/ASN.2009060615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Zhang MZ, Yao B, Yang S, et al. CSF-1 signaling mediates recovery from acute kidney injury. The Journal of Clinical Investigation. 2012;122(12):4519–4532. doi: 10.1172/JCI60363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Lech M, Gröbmayr R, Ryu M, et al. Macrophage phenotype controls longterm AKI outcomes-kidney regeneration versus atrophy. doi: 10.1681/ASN.2013020152. Journal of the American Society of Nephrology. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Dehmel S, Wang S, Schmidt C, et al. Chemokine receptor Ccr5 deficiency induces alternative macrophage activation and improves long-term renal allograft outcome. European Journal of Immunology. 2010;40(1):267–278. doi: 10.1002/eji.200939652. [DOI] [PubMed] [Google Scholar]
  • 18.Duffield JS, Forbes SJ, Constandinou CM, et al. Selective depletion of macrophages reveals distinct, opposing roles during liver injury and repair. The Journal of Clinical Investigation. 2005;115(1):56–65. doi: 10.1172/JCI22675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Tamura T, Yanai H, Savitsky D, Taniguchi T. The IRF family transcription factors in immunity and oncogenesis. Annual Review of Immunology. 2008;26:535–584. doi: 10.1146/annurev.immunol.26.021607.090400. [DOI] [PubMed] [Google Scholar]
  • 20.Escalante CR, Yie J, Thanos D, Aggarwal AK. Structure of IRF-1 with bound DNA reveals determinants of interferon regulation. Nature. 1998;391(6662):103–106. doi: 10.1038/34224. [DOI] [PubMed] [Google Scholar]
  • 21.Lohoff M, Mak TW. Roles of interferon-regulatory factors in T-helper-cell differentiation. Nature Reviews Immunology. 2005;5(2):125–135. doi: 10.1038/nri1552. [DOI] [PubMed] [Google Scholar]
  • 22.Takaoka A, Tamura T, Taniguchi T. Interferon regulatory factor family of transcription factors and regulation of oncogenesis. Cancer Science. 2008;29(3):467–478. doi: 10.1111/j.1349-7006.2007.00720.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Matsuyama T, Kimura T, Kitagawa M, et al. Targeted disruption of IRF-1 or IRF-2 results in abnormal type I IFN gene induction and aberrant lymphocyte development. Cell. 1993;75(1):83–97. [PubMed] [Google Scholar]
  • 24.Miyamoto M, Fujita T, Kimura Y, et al. Regulated expression of a gene encoding a nuclear factor, IRF-1, that specifically binds to IFN-β gene regulatory elements. Cell. 1988;54(6):903–913. doi: 10.1016/s0092-8674(88)91307-4. [DOI] [PubMed] [Google Scholar]
  • 25.Fujita T, Kimura Y, Miyamoto M, Barsoumian EL, Taniguchi T. Induction of endogenous IFN-α and IFN-β genes by a regulatory transcription factor, IRF-1. Nature. 1989;337(6204):270–272. doi: 10.1038/337270a0. [DOI] [PubMed] [Google Scholar]
  • 26.Martinez FO, Gordon S, Locati M, Mantovani A. Transcriptional profiling of the human monocyte-to-macrophage differentiation and polarization: new molecules and patterns of gene expression. Journal of Immunology. 2006;177(10):7303–7311. doi: 10.4049/jimmunol.177.10.7303. [DOI] [PubMed] [Google Scholar]
  • 27.Negishi H, Fujita Y, Yanai H, et al. Evidence for licensing of IFN-γ-induced IFN regulatory factor 1 transcription factor by MyD88 in Toll-like receptor-dependent gene induction program. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(41):15136–15141. doi: 10.1073/pnas.0607181103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Liu J, Cao S, Herman LM, Ma X. Differential regulation of interleukin (IL)-12 p35 and p40 gene expression and interferon (IFN)-gamma-primed IL-12 production by IFN regulatory factor 1. The Journal of Experimental Medicine. 2003;198(8):1265–1276. doi: 10.1084/jem.20030026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Lin R, Hiscott J. A role for casein kinase II phosphorylation in the regulation of IRF-1 transcriptional activity. Molecular and Cellular Biochemistry. 1999;191(1-2):169–180. [PubMed] [Google Scholar]
  • 30.Carey M. The enhanceosome and transcriptional synergy. Cell. 1998;92(1):5–8. doi: 10.1016/s0092-8674(00)80893-4. [DOI] [PubMed] [Google Scholar]
  • 31.Blanco JCG, Contursi C, Salkowski CA, DeWitt DL, Ozato K, Vogel SN. Interferon regulatory factor (IRF)-1 and IRF-2 regulate interferon γ- dependent cyclooxygenase 2 expression. The Journal of Experimental Medicine. 2000;191(12):2131–2144. doi: 10.1084/jem.191.12.2131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Sanceau J, Kaisho T, Hirano T, Wietzerbin J. Triggering of the human interleukin-6 gene by interferon-γ and tumor necrosis factor-α in monocytic cells involves cooperation between interferon regulatory factor-1, NFκB, and Sp1 transcription factors. The Journal of Biological Chemistry. 1995;270(46):27920–27931. doi: 10.1074/jbc.270.46.27920. [DOI] [PubMed] [Google Scholar]
  • 33.Nakagawa K, Yokosawa H. PIAS3 induces SUMO-1 modification and transcriptional repression of IRF-1. FEBS Letters. 2002;530(1–3):204–208. doi: 10.1016/s0014-5793(02)03486-5. [DOI] [PubMed] [Google Scholar]
  • 34.Nhu MQ, Cuesta N, Vogel SN. Transcriptional regulation of lipopolysaccharide (LPS)-induced Toll-like receptor (TLR) expression in murine macrophages: role of interferon regulatory factors 1 (IRF-1) and 2 (IRF-2) Journal of Endotoxin Research. 2006;12(5):285–295. doi: 10.1179/096805106X118834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Kamijo R, Harada H, Matsuyama T, et al. Requirement for transcription factor IRF-1 in NO synthase induction in macrophages. Science. 1994;263(5153):1612–1615. doi: 10.1126/science.7510419. [DOI] [PubMed] [Google Scholar]
  • 36.Taki S, Sato T, Ogasawara K, et al. Multistage regulation of Th1-type immune responses by the transcription factor IRF-1. Immunity. 1997;6(6):673–679. doi: 10.1016/s1074-7613(00)80443-4. [DOI] [PubMed] [Google Scholar]
  • 37.Elser B, Lohoff M, Kock S, et al. IFN-γ represses IL-4 expression via IRF-1 and IRF-2. Immunity. 2002;17(6):703–712. doi: 10.1016/s1074-7613(02)00471-5. [DOI] [PubMed] [Google Scholar]
  • 38.Khan IA, Matsuura T, Fonseka S, Kasper LH. Production of nitric oxide (NO) is not essential for protection against acute Toxoplasma gondii infection in IRF-1-/- Mice. Journal of Immunology. 1996;156(2):636–643. [PubMed] [Google Scholar]
  • 39.Iadecola C, Salkowski CA, Zhang F, et al. The transcription factor interferon regulatory factor 1 is expressed after cerebral ischemia and contributes to ischemic brain injury. The Journal of Experimental Medicine. 1999;189(4):719–727. doi: 10.1084/jem.189.4.719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Wang Y, John R, Chen J, et al. IRF-1 promotes inflammation early after ischemic acute kidney injury. Journal of the American Society of Nephrology. 2009;20(7):1544–1555. doi: 10.1681/ASN.2008080843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Harada H, Fujita T, Miyamoto M, et al. Structurally similar but functionally distinct factors, IRF-1 and IRF-2, bind to the same regulatory elements of IFN and IFN-inducible genes. Cell. 1989;58(4):729–739. doi: 10.1016/0092-8674(89)90107-4. [DOI] [PubMed] [Google Scholar]
  • 42.Cuesta N, Salkowski CA, Thomas KE, Vogel SN. Regulation of lipopolysaccharide sensitivity by IFN regulatory factor-21. Journal of Immunology. 2003;170(11):5739–5747. doi: 10.4049/jimmunol.170.11.5739. [DOI] [PubMed] [Google Scholar]
  • 43.Han K-J, Jiang L, Shu H-B. Regulation of IRF2 transcriptional activity by its sumoylation. Biochemical and Biophysical Research Communications. 2008;372(4):772–778. doi: 10.1016/j.bbrc.2008.05.103. [DOI] [PubMed] [Google Scholar]
  • 44.Cuesta N, Nhu QM, Zudaire E, Polumuri S, Cuttitta F, Vogel SN. IFN regulatory factor-2 regulates macrophage apoptosis through a STAT1/3- and caspase-1-dependent mechanism. Journal of Immunology. 2007;178(6):3602–3611. doi: 10.4049/jimmunol.178.6.3602. [DOI] [PubMed] [Google Scholar]
  • 45.Fehr T, Schoedon G, Odermatt B, et al. Crucial role of interferon consensus sequence binding protein, but neither of interferon regulatory factor 1 nor of nitric oxide synthesis for protection against murine listeriosis. The Journal of Experimental Medicine. 1997;185(5):921–931. doi: 10.1084/jem.185.5.921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Salkowski CA, Barber SA, Detore GR, Vogel SN. Differential dysregulation of nitric oxide production in macrophages with targeted disruptions in IFN regulatory factor-1 and -2 genes. Journal of Immunology. 1996;156(9):3107–3110. [PubMed] [Google Scholar]
  • 47.Klune JR, Dhupar R, Kimura S, et al. Interferon regulatory factor-2 is protective against hepatic ischemia-reperfusion injury. American Journal of Physiology. Gastrointestinal and Liver Physiology. 2012;303(5):G666–G673. doi: 10.1152/ajpgi.00050.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Hida S, Ogasawara K, Sato K, et al. CD8+ T cell-mediated skin disease in mice lacking IRF-2, the transcriptional attenuator of interferon-α/β signaling. Immunity. 2000;13(5):643–655. doi: 10.1016/s1074-7613(00)00064-9. [DOI] [PubMed] [Google Scholar]
  • 49.Ichikawa E, Hida S, Omatsu Y, et al. Defective development of splenic and epidermal CD4+ dendritic cells in mice deficient for IFN regulatory factor-2. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(11):3909–3914. doi: 10.1073/pnas.0400610101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Au W-C, Moore PA, Lowther W, Juang Y-T, Pitha PM. Identification of a member of the interferon regulatory factor family that binds to the interferon-stimulated response element and activates expression of interferon-induced genes. Proceedings of the National Academy of Sciences of the United States of America. 1995;92(25):11657–11661. doi: 10.1073/pnas.92.25.11657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Nguyen H, Hiscott J, Pitha PM. The growing family of interferon regulatory factors. Cytokine and Growth Factor Reviews. 1997;8(4):293–312. doi: 10.1016/s1359-6101(97)00019-1. [DOI] [PubMed] [Google Scholar]
  • 52.Yamamoto M, Sato S, Hemmi H, et al. Role of adaptor TRIF in the MyD88-independent toll-like receptor signaling pathway. Science. 2003;301(5633):640–643. doi: 10.1126/science.1087262. [DOI] [PubMed] [Google Scholar]
  • 53.Fitzgerald KA, Rowe DC, Barnes BJ, et al. LPS-TLR4 signaling to IRF-3/7 and NF-κB involves the toll adapters TRAM and TRIF. The Journal of Experimental Medicine. 2003;198(7):1043–1055. doi: 10.1084/jem.20031023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Honda K, Taniguchi T. IRFs: master regulators of signalling by Toll-like receptors and cytosolic pattern-recognition receptors. Nature Reviews Immunology. 2006;6(9):644–658. doi: 10.1038/nri1900. [DOI] [PubMed] [Google Scholar]
  • 55.Akira S, Uematsu S, Takeuchi O. Pathogen recognition and innate immunity. Cell. 2006;124(4):783–801. doi: 10.1016/j.cell.2006.02.015. [DOI] [PubMed] [Google Scholar]
  • 56.Ishikawa H, Ma Z, Barber GN. STING regulates intracellular DNA-mediated, type i interferon-dependent innate immunity. Nature. 2009;461(7265):788–792. doi: 10.1038/nature08476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Panne D, Maniatis T, Harrison SC. An atomic model of the interferon-beta enhanceosome. Cell. 2007;129(6):1111–1123. doi: 10.1016/j.cell.2007.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Takahasi K, Suzuki NN, Horiuchi M, et al. X-ray crystal structure of IRF-3 and its functional implications. Nature Structural Biology. 2003;10(11):922–927. doi: 10.1038/nsb1001. [DOI] [PubMed] [Google Scholar]
  • 59.Panne D, McWhirter SM, Maniatis T, Harrison SC. Interferon regulatory factor 3 is regulated by a dual phosphorylation- dependent switch. The Journal of Biological Chemistry. 2007;282(31):22816–22822. doi: 10.1074/jbc.M703019200. [DOI] [PubMed] [Google Scholar]
  • 60.Qin BY, Liu C, Srinath H, et al. Crystal structure of IRF-3 in complex with CBP. Structure. 2005;13(9):1269–1277. doi: 10.1016/j.str.2005.06.011. [DOI] [PubMed] [Google Scholar]
  • 61.Zhang M, Tian Y, Wang R-P, et al. Negative feedback regulation of cellular antiviral signaling by RBCK1-mediated degradation of IRF3. Cell Research. 2008;18(11):1096–1104. doi: 10.1038/cr.2008.277. [DOI] [PubMed] [Google Scholar]
  • 62.Fleetwood AJ, Dinh H, Cook AD, Hertzog PJ, Hamilton JA. GM-CSF- and M-CSF-dependent macrophage phenotypes display differential dependence on Type I interferon signaling. Journal of Leukocyte Biology. 2009;86(2):411–421. doi: 10.1189/jlb.1108702. [DOI] [PubMed] [Google Scholar]
  • 63.Biswas SK, Gangi L, Paul S, et al. A distinct and unique transcriptional program expressed by tumor-associated macrophages (defective NF-κB and enhanced IRF-3/STAT1 activation) Blood. 2006;107(5):2112–2122. doi: 10.1182/blood-2005-01-0428. [DOI] [PubMed] [Google Scholar]
  • 64.Tarassishin L, Suh H-S, Lee SC. Interferon regulatory factor 3 plays an anti-inflammatory role in microglia by activating the PI3K/Akt pathway. Journal of Neuroinflammation. 2011;8, article 187 doi: 10.1186/1742-2094-8-187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Lin R, Heylbroeck C, Genin P, Pitha PM, Hiscott J. Essential role of interferon regulatory factor 3 in direct activation of RANTES chemokine transcription. Molecular and Cellular Biology. 1999;19(2):959–966. doi: 10.1128/mcb.19.2.959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Eisenbeis CF, Singh H, Storb U. Pip, a novel IRF family member, is a lymphoid-specific, PU.1-dependent transcriptional activator. Genes and Development. 1995;9(11):1377–1387. doi: 10.1101/gad.9.11.1377. [DOI] [PubMed] [Google Scholar]
  • 67.Mittrücker H-W, Matsuyama T, Grossman A, et al. Requirement for the transcription factor LSIRF/IRF4 for mature B and T lymphocyte function. Science. 1997;275(5299):540–543. doi: 10.1126/science.275.5299.540. [DOI] [PubMed] [Google Scholar]
  • 68.Lawrence T, Natoli G. Transcriptional regulation of macrophage polarization: enabling diversity with identity. Nature Reviews Immunology. 2011;11(11):750–761. doi: 10.1038/nri3088. [DOI] [PubMed] [Google Scholar]
  • 69.Ahyi A-NN, Chang H-C, Dent AL, Nutt SL, Kaplan MH. IFN regulatory factor 4 regulates the expression of a subset of Th2 cytokines. Journal of Immunology. 2009;183(3):1598–1606. doi: 10.4049/jimmunol.0803302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Honma K, Udono H, Kohno T, et al. Interferon regulatory factor 4 negatively regulates the production of proinflammatory cytokines by macropages in response to LPS. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(44):16001–16006. doi: 10.1073/pnas.0504226102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.El Chartouni C, Schwarzfischer L, Rehli M. Interleukin-4 induced interferon regulatory factor (Irf) 4 participates in the regulation of alternative macrophage priming. Immunobiology. 2010;215(9-10):821–825. doi: 10.1016/j.imbio.2010.05.031. [DOI] [PubMed] [Google Scholar]
  • 72.Satoh T, Takeuchi O, Vandenbon A, et al. The Jmjd3-Irf4 axis regulates M2 macrophage polarization and host responses against helminth infection. Nature Immunology. 2010;11(10):936–944. doi: 10.1038/ni.1920. [DOI] [PubMed] [Google Scholar]
  • 73.Ishii M, Wen H, Corsa CAS, et al. Epigenetic regulation of the alternatively activated macrophage phenotype. Blood. 2009;114(15):3244–3254. doi: 10.1182/blood-2009-04-217620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Takaoka A, Yanai H, Kondo S, et al. Integral role of IRF-5 in the gene induction programme activated by Toll-like receptors. Nature. 2005;434(7030):243–249. doi: 10.1038/nature03308. [DOI] [PubMed] [Google Scholar]
  • 75.Paun A, Bankoti R, Joshi T, Pitha PM, Stäger S. Critical role of IRF-5 in the development of t helper 1 responses to Leishmania donovani infection. PLoS Pathogens. 2011;7(1) doi: 10.1371/journal.ppat.1001246.e1001246 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Savitsky DA, Yanai H, Tamura T, Taniguchi T, Honda K. Contribution of IRF5 in B cells to the development of murine SLE-like disease through its transcriptional control of the IgG2a locus. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(22):10154–10159. doi: 10.1073/pnas.1005599107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Lacey DC, Achuthan A, Fleetwood AJ, et al. Defining GM-CSF- and macrophage-CSF-dependent macrophage responses by in vitro models. Journal of Immunology. 2012;188(11):5752–5765. doi: 10.4049/jimmunol.1103426. [DOI] [PubMed] [Google Scholar]
  • 78.Eames HL, Saliba DG, Krausgruber T, Lanfrancotti A, Ryzhakov G, Udalova IA. KAP1/TRIM28: an inhibitor of IRF5 function in inflammatory macrophages. Immunobiology. 2012;217(12):1315–1324. doi: 10.1016/j.imbio.2012.07.026. [DOI] [PubMed] [Google Scholar]
  • 79.Krausgruber T, Blazek K, Smallie T, et al. IRF5 promotes inflammatory macrophage polarization and TH1-TH17 responses. Nature Immunology. 2011;12(3):231–238. doi: 10.1038/ni.1990. [DOI] [PubMed] [Google Scholar]
  • 80.Hedl M, Abraham C. IRF5 risk polymorphisms contribute to interindividual variance in pattern recognition receptor-mediated cytokine secretion in human monocyte-derived cells. Journal of Immunology. 2012;188(11):5348–5356. doi: 10.4049/jimmunol.1103319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Feng D, Stone RC, Eloranta M-L, et al. Genetic variants and disease-associated factors contribute to enhanced interferon regulatory factor 5 expression in blood cells of patients with systemic lupus erythematosus. Arthritis and Rheumatism. 2010;62(2):562–573. doi: 10.1002/art.27223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Graham RR, Kyogoku C, Sigurdsson S, et al. Three functional variants of IFN regulatory factor 5 (IRF5) define risk and protective haplotypes for human lupus. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(16):6758–6763. doi: 10.1073/pnas.0701266104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Stone RC, Feng D, Deng J, et al. Interferon regulatory factor 5 activation in monocytes of systemic lupus erythematosus patients is triggered by circulating autoantigens independent of type i interferons. Arthritis and Rheumatism. 2012;64(3):788–798. doi: 10.1002/art.33395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Zucchero TM, Cooper ME, Maher BS, et al. Interferon regulatory factor 6 (IRF6) gene variants and the risk of isolated cleft lip or palate. The New England Journal of Medicine. 2004;351(8):769–780. doi: 10.1056/NEJMoa032909. [DOI] [PubMed] [Google Scholar]
  • 85.Ingraham CR, Kinoshita A, Kondo S, et al. Abnormal skin, limb and craniofacial morphogenesis in mice deficient for interferon regulatory factor 6 (Irf6) Nature Genetics. 2006;38(11):1335–1340. doi: 10.1083/ng1903. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Sato M, Suemori H, Hata N, et al. Distinct and essential roles of transcription factors IRF-3 and IRF-7 in response to viruses for IFN-α/β gene induction. Immunity. 2000;13(4):539–548. doi: 10.1016/s1074-7613(00)00053-4. [DOI] [PubMed] [Google Scholar]
  • 87.Sato M, Hata N, Asagiri M, Nakaya T, Taniguchi T, Tanaka N. Positive feedback regulation of type I IFN genes by the IFN-inducible transcription factor IRF-7. FEBS Letters. 1998;441(1):106–110. doi: 10.1016/s0014-5793(98)01514-2. [DOI] [PubMed] [Google Scholar]
  • 88.Qi C-F, Li Z, Raffeld M, Wang H, Kovalchuk AL, Morse HC., III Differential expression of IRF8 in subsets of macrophages and dendritic cells and effects of IRF8 deficiency on splenic B cell and macrophage compartments. Immunologic Research. 2009;45(1):62–74. doi: 10.1007/s12026-008-8032-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Wang H, Morse HC., III IRF8 regulates myeloid and B lymphoid lineage diversification. Immunologic Research. 2009;43(1–3):109–117. doi: 10.1007/s12026-008-8055-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Becker AM, Michael DG, Satpathy AT, Sciammas R, Singh H, Bhattacharya D. IRF-8 extinguishes neutrophil production and promotes dendritic cell lineage commitment in both myeloid and lymphoid mouse progenitors. Blood. 2012;119(9):2003–2012. doi: 10.1182/blood-2011-06-364976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Laricchia-Robbio L, Tamura T, Karpova T, Sprague BL, McNally JG, Ozato K. Partner-regulated interaction of IFN regulatory factor 8 with chromatin visualized in live macrophages. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(40):14368–14373. doi: 10.1073/pnas.0504014102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Holtschke T, Löhler J, Kanno Y, et al. Immunodeficiency and chronic myelogenous leukemia-like syndrome in mice with a targeted mutation of the ICSBP gene. Cell. 1996;87(2):307–317. doi: 10.1016/s0092-8674(00)81348-3. [DOI] [PubMed] [Google Scholar]
  • 93.Xu H, Zhu J, Smith S, et al. Notch-RBP-J signaling regulates the transcription factor IRF8 to promote inflammatory macrophage polarization. Nature Immunology. 2012;13(7):642–650. doi: 10.1038/ni.2304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Chang TH, Xu S, Tailor P, Kanno T, Ozato K. The small ubiquitin-like modifier-deconjugating enzyme sentrin-specific peptidase 1 switches IFN regulatory factor 8 from a repressor to an activator during macrophage activation. Journal of Immunology. 2012;189(7):3548–3556. doi: 10.4049/jimmunol.1201104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Reich NC, Liu L. Tracking STAT nuclear traffic. Nature Reviews Immunology. 2006;6(8):602–612. doi: 10.1038/nri1885. [DOI] [PubMed] [Google Scholar]

Articles from Mediators of Inflammation are provided here courtesy of Wiley

RESOURCES