Sunteți pe pagina 1din 27

Interferon-: an overview of signals, mechanisms and functions

Kate Schroder,*, Paul J. Hertzog,, Timothy Ravasi,*, and David A. Hume*,,1 *Institute for Molecular Bioscience, University of Queensland, St. Lucia, Brisbane, Australia; CRC for Chronic Inammatory Diseases, Parkville, Victoria, Australia; and Molecular Genetics Group, Institute of Reproduction and Development, Monash Medical Centre, Monash University, Clayton, Victoria, Australia

Abstract: Interferon- (IFN-) coordinates a diverse array of cellular programs through transcriptional regulation of immunologically relevant genes. This article reviews the current understanding of IFN- ligand, receptor, signal transduction, and cellular effects with a focus on macrophage responses and to a lesser extent, responses from other cell types that inuence macrophage function during infection. The current model for IFN- signal transduction is discussed, as well as signal regulation and factors conferring signal specicity. Cellular effects of IFN- are described, including up-regulation of pathogen recognition, antigen processing and presentation, the antiviral state, inhibition of cellular proliferation and effects on apoptosis, activation of microbicidal effector functions, immunomodulation, and leukocyte trafcking. In addition, integration of signaling and response with other cytokines and pathogen-associated molecular patterns, such as tumor necrosis factor-, interleukin-4, type I IFNs, and lipopolysaccharide are discussed. J. Leukoc. Biol. 75: 163189; 2004.
Key Words: macrophage cytokine lipopolysaccharide Toll-like receptor inammation

Macrophages respond to a range of different cell products during the innate and acquired immune response. Of these, IFN- (originally called macrophage-activating factor) is among the most important. Macrophage stimulation with IFN- induces direct antimicrobial and antitumor mechanisms as well as up-regulating antigen processing and presentation pathways. IFN- orchestrates leukocyte attraction and directs growth, maturation, and differentiation of many cell types [2 4], in addition to enhancing natural killer (NK) cell activity [5] and regulating B cell functions such as immunoglobulin (Ig) production and class switching [4, 6]. This review provides a brief overview of IFN- biology with respect to the wellcharacterized responses that alter macrophage function during infectious challenge.

THE IFNs
The IFNs were originally discovered as agents that interfere with viral replication [7]. Initially, they were classied by the secreting cell type but are now classied into type I and type II according to receptor specicity and sequence homology. The type I IFNs are comprised of multiple IFN- subtypes (14 20, depending on species), IFN-, IFN-, and IFN-, all of which are structurally related and bind to a common heterodimeric receptor (IFNAR, comprised of IFNAR1 and IFNAR2 chains). Although type I IFNs are secreted at low levels by almost all cell types, hematopoietic cells are the major producers of IFN- and IFN-, whereas broblasts are a major cellular source of IFN- [8]. IFN- is also produced by macrophages under appropriate stimulus (discussed later). Viral infection is the classic stimulus for IFN- and IFN- expression [8, 9]. Secretion of IFN- has only been reported in ruminants [10]. IFN- is the sole type II IFN. It is structurally unrelated to type I IFNs, binds to a different receptor, and is encoded by a separate chromosomal locus. Initially, it was believed that CD4 T helper cell type 1 (Th1) lymphocytes, CD8 cytotoxic lymphocytes, and NK cells exclusively produced IFN- [8, 11]. However, there is now evidence that other cells, such

INTRODUCTION
This review focuses on the interplay between interferon- (IFN-) and macrophages in inammation and acquired immunity during infection. Macrophages are extremely versatile cells involved in a number of complex functions in disease and health. A pathogen encounters macrophages soon after host entry, and the result of these encounters is fundamental to the hosts ability to mount an effective immune response. Macrophage activation, broadly dened as acquisition of competence to execute a complex function [1], is among the rst actions to occur in innate immunity to potential pathogens. In most situations, macrophages are activated to acquire microbicidal effector functions and secrete proinammatory cytokines, resulting in inammation and recruitment of immune cells and subsequent elimination of the microbe by phagocytosis or release of toxic metabolites.

1 Correspondence: Institute for Molecular Bioscience, University of Queensland, St. Lucia, Brisbane 4072, Australia. E-mail: D.Hume@imb.uq.edu.au Received June 2, 2003; revised July 25, 2003; accepted July 27, 2003; doi: 10.1189/jlb.0603252.

Journal of Leukocyte Biology Volume 75, February 2004 163

as B cells, NKT cells, and professional antigen-presenting cells (APCs) secrete IFN- (reviewed in refs. [5, 1216]). IFN- production by professional APCs [monocyte/macrophage, dendritic cells (DCs)] acting locally may be important in cell self-activation and activation of nearby cells [12, 13]. IFN- secretion by NK cells and possibly professional APCs is likely to be important in early host defense against infection, whereas T lymphocytes become the major source of IFN- in the adaptive immune response [12, 17]. IFN- production is controlled by cytokines secreted by APCs, most notably interleukin (IL)-12 and IL-18. These cytokines serve as a bridge to link infection with IFN- production in the innate immune response [18 24]. Macrophage recognition of many pathogens induces secretion of IL-12 and chemokines [e.g., macrophage-inammatory protein-1 (MIP1); ref. 25]. These chemokines attract NK cells to the site of inammation, and IL-12 promotes IFN- synthesis in these cells [25, 26]. In macrophages, NK and T cells, the combination of IL-12 and IL-18 stimulation further increases IFN- production [20, 23, 24, 27, 28]. Negative regulators of IFN- production include IL-4, IL10, transforming growth factor-, and glucocorticoids [17, 21, 2729]. Given the complexity of IFN- regulation, it is not surprising that inbred mouse strains vary in their ability to secrete this cytokine; for example, T lymphocytes of C57BL/6 and C3H mice secrete signicantly higher amounts of IFN- compared with the T lymphocytes of BALB/c and B10.D2 mice. Increased IFN- production in these strains is associated with greater resistance to bacteria and viruses [30 32]. Many excellent reviews on the regulation of IFN- production have been published recently and the reader is referred to these for further information [1113].

THE IFN- RECEPTOR


Functional IFN- receptor (IFNGR) is comprised of two ligand-binding IFNGR1 chains associated with two signal-transducing IFNGR2 chains and associated signaling machinery. IFNGR1 and IFNGR2 chains belong to the class II cytokine receptor family, a class of receptors that bind ligand in the small angle of a V formed by the two Ig-like folds that constitute the extracellular domain. It is likely that the IFNGR chains and other family members (type I IFN receptor and tissue factor) evolved from primitive adhesive molecules [33, 34]. The IFNGR2 chain is generally the limiting factor in IFN- responsiveness, as the IFNGR1 chain is usually in surplus [8, 35]. The IFNGR2 chain is constitutively expressed, but its expression level may be tightly regulated according to the state of cellular differentiation or activation [8]. For example, some CD4 Th1 populations have very low levels of cell-surface expression of the IFNGR2 chain and thus low expression of active IFNGR, leading to a functional blockade of some aspects of IFN- signaling [36, 37]. As the growth-inhibitory effects of IFN- are blocked, T cells expressing low levels of the IFNGR2 chain continue to proliferate during IFN- treatment. Conversely, IFN- exposure to CD4 Th2 populations displaying high levels of IFNGR2 inhibits proliferation and
164 Journal of Leukocyte Biology Volume 75, February 2004

may induce the apoptotic program [35, 38, 39]. This mechanism may aid in the Th2-to-Th1 phenotype switch apparent when CD4 cells are treated with IFN-: the Th2 population decreases as a result of growth-inhibitory and proapoptotic effects of IFN-, and the Th1 population continues to proliferate as a result of blockade of IFN- function. As a consequence, IFN- signaling causes IFNGR2 down-regulation and switching to a Th1 phenotype as a T cell population but not at a single-cell level [37]. Both IFNGR chains lack intrinsic kinase/phosphatase activity and so must associate with signaling machinery for signal transduction. The IFNGR1 intracellular domain contains binding motifs for the Janus tyrosine kinase (Jak)1 and the latent cytosolic factor, signal transducer and activator of transcription (Stat)1. In humans, the Jak1-binding motif LPKS is a membrane-proximal sequence located at residues 266 269 [40 42]. The human Stat1-binding site YDKPH is positioned at residues 440 444 [43]. This motif contains an essential Y440 phosphorylation site that is phosphorylated during signal transduction to allow Stat1 recruitment to the receptor [42 44]. Residues 441DKPH444 are responsible for the binding specicity of IFNGR1 and Stat1 [41, 43, 44]. The Jak1- and Stat1binding motifs are required for receptor phosphorylation, signal transduction, and induction of biological response. Receptor-ligand internalization is mediated by an isoleucine-leucine sequence at residues 270 and 271 of the mature IFNGR1 chain in humans. Mutant receptors with deletion or substitution of this sequence cannot internalize ligand, although signal transduction is not affected [42, 45]. The intracellular region of IFNGR2 contains a noncontiguous binding motif for recruitment of Jak2 kinase for participation in signal transduction. In humans, these sites are 263 PPSIP267 and 270IEEYL274 [46, 47]. The IFNGR2 chain is not tyrosine phosphorylated during signal transduction [46]. Cross-linking experiments with labeled human IFN- demonstrated that IFN- only associates with IFNGR2 when the IFNGR1 chain is present, indicating that the primary interaction between the IFNGR2 chain and the IFN-:IFNGR1 complex is found between IFNGR1 and IFNGR2, although it is likely that IFNGR2 also interacts weakly with the ligand [46, 48]. IFN-:IFNGR1 and IFNGR1:IFNGR2 interactions are species-specic. A number of studies with chimeric receptors in which the intracellular, transmembrane, or extracellular domains were swapped between the human and murine IFNGR chains showed that species specicity in interactions of the ligand:receptor complex is restricted to the receptor extracellular domains [49 53].

IFN- SYSTEM DYSFUNCTION


IFN-/ and IFNGR1/ mice showed no overt developmental defects, and their immune system appeared to develop normally [54]. However, these mice show deciencies in natural resistance to bacterial, parasitic, and viral infections such as vaccinia virus, Theilers murine encephalomyelitis virus, Leishmania major, Toxoplasma gondii, Listeria monocytogenes, and several poorly virulent mycobacteria species [54 59].
http://www.jleukbio.org

Some viruses (e.g., vaccinia virus, Theilers murine encephalomyelitis virus, and lymphocytic choriomeningitis virus) appear to require both type I and II IFN pathways, whereas viruses such as Semliki forest virus and vesicular stomatitis virus require a predominantly type I IFN response for efcient clearing [60]. These observations have prompted the suggestion that the two IFN systems may have evolved to complement each other in overlapping but nonredundant activities to defend against a broad spectrum of pathogens [60, 61]. Historically, the type I IFNs were considered to be primarily antiviral agents with limited immunomodulatory activity, whereas IFN- was considered to be primarily an immunomodulator with limited antiviral activity [8]. It is now apparent that both types of IFN possess antiviral and immunomodulatory activities to some degree [62]. The antiviral activity of both types of IFNs is often important in the early stages of viral infection, but their immunomodulatory activities become important in later stages of infection when the adaptive immune response becomes critical [60]. Patients with inactivating mutations of the human IFNGR1 or IFNGR2 chains show clinical presentation similar to the mouse models. Human loss-of-function mutations in the IFNGR1 or IFNGR2 chain are closely associated with severe susceptibility to poorly virulent mycobacteria, often presenting as early onset bacillus Calmette-Guerin infection and fatality in childhood [6371]. Partial deciency of either chain gives a milder phenotype and a much better prognosis [63 65]. Loss of functional IFNGR1 appears to be associated only with increased susceptibility to infection with some viruses and intracellular bacteria; increased susceptibility to other common bacterial and fungi pathogens have not been reported [72]. In addition to recurrent infection, infants with decient production of IFN- exhibited decreased neutrophil mobility and NK cell activity, highlighting the importance of IFN- in the inammatory response and immunoregulation [73]. It is interesting that natural IFN- polymorphisms have been correlated with increased longevity [74]. It has been proposed that a slightly dampened inammatory status caused by an IFN- polymorphism, while not enough to signicantly impact on the individuals ability to clear infection, may prevent or defer inammation-related diseases such as cardiovascular disease, neurodegeneration, osteoarthritis, osteoporosis, and diabetes [74]. Aside from functions in host defense, IFN- may also contribute to autoimmune pathology. Although IFN- production was shown to be disease-limiting in autoimmune models such as murine experimental allergic encephalomyelitis (EAE) [75, 76], it may contribute to autoimmune nephritis [77, 78]. In humans, IFN- is implicated in pathology of diseases such as systemic lupus erythematosus [79, 80], multiple sclerosis [81], and insulin-dependent diabetes mellitus [82 84]. IFN- function is signicant in tumor surveillance. IFNGR1 knockout (KO) mice, cells with dominant-negative IFNGR1 mutations, and cells treated with IFN--neutralizing antibodies display compromised tumor rejection [85 87]. IFN- protects against tumor development and directs the immunogenic phenotype of tumors that arise in an immunocompetent host (the cancer immunoediting concept; reviewed in refs. [88, 89]).

SIGNAL TRANSDUCTION
IFN- primarily signals through the Jak-Stat pathway, a pathway used by over 50 cytokines, growth factors, and hormones to affect gene regulation [90]. Jak-Stat signaling involves sequential receptor recruitment and activation of members of the Janus family of kinases (Jaks: Jaks 13 and Tyk2) and the Stats (Stats 1 6, including Stat5a and Stat5b) to control transcription of target genes via specic response elements. As this signaling mechanism is a recurring theme amongst members of the cytokine receptor superfamily, IFN--induced Jak-Stat signaling has emerged as the global paradigm for class II cytokine receptor signal transduction. The IFNGR1 and IFNGR2 subunits of the IFN- receptor were thought not to strongly associate with each other in the absence of ligand [13, 40, 47, 91, 92], but new techniques allowing the study of receptor chain interactions in intact cells have shown the receptor complex is assembled before ligand binding [93]. Upon ligand binding, the intracellular domains of the receptor chains open out to allow association of downstream signaling components. Biologically active IFN- is a noncovalent homodimer formed by the self-association of two mature polypeptides in an antiparallel orientation [94, 95] and thus binds IFNGR1 in 2:2 binding stoichiometry [47, 9597]. Ligand binding induces Jak2 autophosphorylation and activation, which allows Jak1 transphosphorylation by Jak2 (Fig. 1) [98]. The activated Jak1 phosphorylates functionally critical tyrosines on residue 440 of each IFNGR1 chain to form two adjacent docking sites for the SH2 domains of latent Stat1 [41, 99 101]. The receptor-recruited Stat1 pair is phosphorylated near the C terminus at Y701, probably by Jak2 [98]. Phosphorylation induces dissociation of a Stat1 homodimer (also known as -IFN activation factor) from the receptor [100]. The four critical tyrosines (contained by Jak1, Jak2, IFNGR1, and Stat1) are phosphorylated within 1 min of IFN- treatment [41, 99]. The dissociated Stat1 homodimer enters the nucleus and binds to promoter elements to initiate or suppress transcription of IFN--regulated genes (reviewed in refs. [102104]). Stat1 homodimers bind DNA at GAS elements of consensus sequence TTCN(2-4)GAA [105]. The key function of Stat1 in mediating IFN- signal transduction is indicated by the phenotype of Stat1/ mice, which largely phenocopy IFNGR1/ mice during IFN- stimulation [106]. The rst wave of IFN--induced transcription occurs within 1530 min of IFN- treatment [107]. Many of the induced genes are in fact transcription factors (for example, IRF-1), which are activated by IFN- and are able to further drive regulation of the next wave of transcription. Transcription of many IFN--responsive genes is controlled by a GAS element or an ISRE [108].

Stat1
The crystal structure for tyrosine phosphorylated Stat1 bound to DNA has been determined [109]. The structure veried that the SH2 domain necessary for dimerization is also involved in

Schroder et al. Signals, mechanisms, and functions of IFN- 165

IFN

JAK2

JAK1

JAK1

JAK2

JAK2

JAK1

JAK1

ST AT STA 1 T1

JAK2

Cytosol

IRF-1

STAT1

STAT2

STAT1

STAT1

STA T1

IRF-9

IRF-9

STAT1

eg. ICAM-1 MIG

GAS

IFN-regulated gene

ISRE

STAT1

IRF-E/GAS/IRF-E IRF-1

Stat1

GAS

Fig. 1. The current paradigm for IFN- signal transduction. Ligand binding causes a conformational change in the IFN-R (IFNGR1, yellow; IFNGR2, green), such that the inactive Jak2 kinase undergoes autophosphorylation and activation, which in turn allows Jak1 transphosphorylation by Jak2. The activated Jak1 phosphorylates functionally critical tyrosines on residue 440 of each IFNGR1 chain to form two adjacent docking sites for the Src homology (SH)2 domains of latent Stat1. The receptor-recruited Stat1 pair is phosphorylated near the C terminus at Y701. Phosphorylation induces dissociation of a Stat1 homodimer from the receptor. To a lesser extent, IFN- signaling also produces Stat1:Stat1:IFN regulatory factor (IRF)-9 and Stat1:Stat2:IRF-9 [IFN-stimulated gene factor 3 (ISGF3)] complexes. Stat1 homodimers travel to the nucleus and bind to promoter IFN--activation site (GAS) elements to initiate/suppress transcription of IFN--regulated genes. Many of IFN--regulated genes are in fact transcription factors (e.g., IRF-1), which are activated by IFN- and are able to drive regulation of the next wave of transcription (e.g., induction of IFN-). Stat1:Stat1:IRF-9 heterodimers, ISGF3, and IRF-1 are able to bind to IFN-stimulated response element (ISRE) promoter regions in target genes to regulate transcription. IRF-1 is also able to promote transcription of Stat1 through an unusual ISRE site (IRF-E/GAS/IRF-E). Signaling molecules activated by IFN- are depicted in red text. ICAM-1, Intercellular adhesion molecule-1; MIG, monokine induced by IFN-; iNOS, inducible nitric oxide synthase.

DNA binding. The C-terminal transactivation domain contains the Y701 phosphorylation site and also a S727 phosphorylation site, both of which are functionally important for efcient signaling [110 113]. All Stat proteins possess a SH2 domain that when tyrosine phosphorylated, mediates homodimerization or heterodimerization with other Stat molecules [114, 115]. Although the simplistic model described above suggests that IFN- signaling produces active Stat1 homodimers alone, other active complexes such as Stat1 heterodimers (e.g., Stat1:Stat2) and heterotrimers (e.g., Stat1:Stat1:IRF-9, Stat1:Stat2:IRF-9) form during signaling [102, 116 118]. Stat2 is the only Stat that does not contain a DNA-binding domain [119]. The Stat1:
166 Journal of Leukocyte Biology Volume 75, February 2004

Stat2:IRF-9 complex (known as ISGF3) was thought to be the typical type I IFN transcription factor. It is now evident that type I IFN signals primarily through ISGF3 but also uses Stat1 homodimers and conversely, type II IFN signals primarily through Stat1 homodimers but also uses ISGF3 to activate transcription [116]. This may explain many of the overlapping effects of the two types of IFNs. Phosphorylation of Stat1 at S727 is essential for maximal ability to activate transcription of target genes [111113, 120 122]. A number of different stimuli induce Stat1 serine phosphorylation, including types I and II IFN, lipopolysaccharide (LPS), IL-2, IL-12, tumor necrosis factor (TNF-), and
http://www.jleukbio.org

ST AT 1
IRF-1

IRF-1 STAT1
P

STA T1

IRF-1

Nucleus
eg. IRF-2 iNOS IFN
STAT1

IFN-regulated gene

platelet-derived growth factor [110 112, 123, 124]. This may be a mechanism whereby S727 serves as an avenue for modulation of IFN- signaling by independent extracellular cues. For example, LPS signaling increases Stat1 S727 phosphorylation independently of Y701 phosphorylation in macrophages, thereby augmenting cellular responses to IFN- [110]. The ability of Stat1 to activate or repress gene transcription depends on the presence of other transcription factors binding to the promoter element and Stat1 interaction with other factors. Stat1 activation is necessary but not sufcient for transcription of a number of genes [125, 126]. Likewise, factors such as oncostatin M, which participate in Jak-Stat signaling to produce activated Stat1, do not induce transcription of IFN-inducible genes [127]. Stat1 interaction with transcription factors such as IRF-9, upstream stimulatory factor-1, specicity protein-1, and heat shock factor-1 have been reported (reviewed in ref. [119]) and are likely to inuence specicity of DNA binding and transactivator ability. Phosphorylation of S727 is necessary for interaction with some proteins, such as BRCA1 and MCM-5, and this interaction potentiates Stat1mediated transcription [128, 129]. In some promoters, transcription is maximal when more than one GAS-binding transcriptional activator binds to tandem GAS sites in the target gene promoter [130, 131]. In the absence of IFN-induced tyrosine phosphorylation, Stat1 drives constitutive transcription of genes such as caspases [132, 133]. The Stat1 KO mouse showed increased lymphocyte survival and proliferation, which was attributed to insufcient caspase levels to carry out the apoptotic program [133]. Stat1 also constitutively regulates expression of the low molecular protein 2 (LMP2) gene [134]. A complex of unphosphorylated Stat1 and IRF-1 may mediate this. The mechanism of Stat1 entry into the nucleus is still controversial, but the involvement of importin--1 (NPI-1) is implied [135]. Two major models have been suggested, one being the classic nuclear importin mechanism and the other involving a requirement for intracellular IFN-. Nuclear entry of Stat1 is apparent at 15 min and almost complete after 30 min exposure to IFN- [136]. Stat1 nuclear translocation does not appear to be a result of liberation from a cytoplasmic anchor or transport by the cellular cytoskeleton or microtubules [136]. Experiments with green uorescent protein-tagged Stat1 implied that it travels through the cytoplasm to the nucleus by random walk movement [136]. A model was proposed whereby Stat1 dimerization allows interaction with importin NPI-1, which then mediates translocation through the nuclear pore [135]. The nuclear membrane forms an efcient barrier to inactivated Stat1, and entry requires a gain of nuclear localization sequence (NLS) function [136 138]. To date, no classic NLS (single stretch of basic residues or bipartite motif; ref. [139]) has been identied for Stat1, indicating an as-yet unidentied, novel NLS is unmasked during Stat1 homodimerization, or the NLS of a Stat1 homodimer-binding partner mediates nuclear translocation. The second model for Stat1 nuclear entry involves a requirement for intracellular IFN-. IFN enters the cell using a mechanism that is unclear at present. Nuclear accumulation of IFN- is apparent shortly after cellular exposure to IFN- [140, 141]. Nuclear translocation is mediated by a C-terminal

sequence very similar to the Simian virus-40 T antigen NLS, RKRKRSR132 in mice and 128KTGKRKR134 in humans [139, 142145]. This sequence is a functional NLS and is essential for the full biological activity of IFN- [144 148]. A deletion mutant of the putative NLS in human IFN- is still able to bind with high afnity to the IFNGR1 chain, but induction of biological response is absent [149]. The addition of a heterologous NLS is able to rescue biological responsiveness in this system. Another study tracked the cellular localization of components of the IFNGR complex following ligand exposure [150]. IFN-, Stat1, and IFNGR1 were found to be quickly internalized, colocalized, and accumulated in the nucleus, while the majority of the IFNGR2 subunit remained at the cell surface throughout signaling. Stat1 nuclear translocation is mediated by NPI-1 [135], and immunoprecipitation of Stat1 following IFN- treatment showed that IFN-, IFNGR1, Stat1, and NPI-1 form a complex at the nuclear pore during translocation [142, 143]. Subramaniam et al. [90] proposed a model suggesting a function for intracellular ligand in Stat1 nuclear localization. In this model, extracellular ligand binds to the IFNGR1 chain, causing endocytosis of the complex and recycling of the IFNGR2 chain to the cell surface. The intracellular domain of IFNGR1 associates with activated Stat1 and intracellular IFN-, and the IFN:IFNGR1:Stat1 complex is actively transported through the nuclear pore via IFN-NLS interaction with the NPI-1 importin. Activated Stat1 may then bind to promoter elements to activate transcription. Functional Stat1 is crucial to host response to infection. KO of Stat1 renders mice extremely sensitive to infection by viral and microbial pathogens (e.g., vesicular stomatitis viruses, mouse herpes virus, L. monocytogenes) [106, 151]. This susceptibility has been attributed to defective immune system function as a result of the combined loss of types I and II IFN responses [151]. NO production in Stat1/ macrophages primed with types I or II IFN and treated with LPS is greatly impaired compared with the wild type (WT), resulting in reduced macrophage microbicidal function [106]. Impaired macrophage microbicidal function may contribute to the increased susceptibility to infection apparent in Stat1/ mice. Stat1 KO mice also demonstrated the existence of Stat1independent mechanisms in IFN- signaling. IFN- may exert pro- and antiproliferative signals through Stat1-independent and Stat1-dependent pathways, respectively [152154]. Although IFN- treatment of lymphocytes and broblasts inhibited cell growth, the same treatment in Stat1/ cells promoted proliferation and decreased apoptotic signals [133, 153, 154]. Microarray analysis revealed that many genes, including the growth-promoting c-myc and c-jun genes, were induced in response to IFN- in Stat1/ cells [152]. This signaling was not dependent on the Stat1-docking site of IFNGR1. Although much more susceptible than their WT counterparts, Stat1/ mice are signicantly more resistant to viral infection than IFNGR1/IFNAR1 double-KO mice, indicating the existence of IFN-dependent, Stat1-independent, antiviral mechanisms [155]. A loss-of-function, heterozygous mutation in the Stat1 gene was recently described in humans, resulting in a dominantnegative effect over the WT allele for Stat1 activation [156]. Although these patients demonstrated increased susceptibility
126

Schroder et al. Signals, mechanisms, and functions of IFN- 167

to bacterial infection, viral infections took the normal clinical course. This suggests that IFN--induced, physiologically signicant, Stat1-independent control of antiviral responses also occurs in humans. Mechanisms of Stat1-independent IFN- signaling are currently unclear. One possibility is the existence of an unknown IFN- receptor or IFN- receptor complex linked to an alternative signaling pathway. This hypothesis was suggested when it was found that anti-IFN- antibodies affected the clinical course of EAE in IFNGR1 KO mice [75].

IRF-1
The IRF gene family is intimately involved in mediating the type I and II IFN signal cascades. Members of the IRF family share similar structure, including an amino-terminal DNAbinding domain containing multiple tryptophan residues and a carboxy-terminal transcriptional activation or repression domain (reviewed in ref. [157]). IRF-1, IRF-2, and IRF-9 (p48, ISGF3-) all participate in IFN- signaling. Basal IRF-1 expression has functions in constitutive gene expression [134], but Stat1 and nuclear factor (NF)-B interaction with promoter elements dramatically increases IRF-1 transcription [158 160]. IRF-1 expression is up-regulated in response to types I and II IFN, virus, or cytokines. IRF-1 transcriptional regulatory activity is regulated independently in response to IFN-, virus, and dsRNA [161]. The physiological relevance of IRF-1 function in transcriptional control of IFN--regulated genes is highlighted by IRF-1 KO mice, in which many genes are submaximally induced by IFN- when compared with their WT littermates [162, 163]. IRF-1 functions are not limited to IFN- signaling; IFN--independent activities in viral/bacterial infection, lymphocyte development, regulation of cell cycle, growth inhibition, apoptosis and tumor suppression have been reported [107, 123, 135, 159, 164 168]. IRF-1 drives inducible expression of many target genes through interaction with the IRF-E site, consensus G(A)AAA G T /C /CGAAAG/CT/C [102, 164]. This specicity overlaps with the ISRE consensus site A/GNGAAANNGAAACT, recognized by ISGF3, which is induced by type I IFN and to a lesser extent, type II IFN [102]. In this way, IRF-1 is able to induce a subset of the full spectrum of IFN-inducible genes. IRF-2 also binds to the IRF-E site and largely functions to repress transcription of IRF-1-inducible genes through its transcriptional repressor domain [102, 159].

Negative regulation of IFN- signaling


Stat1 activation is inhibited within 1 h of IFN- treatment, despite the continued presence of extracellular IFN-, and so mechanisms must exist to control the extent of ligand stimulation of IFN- signaling (Fig. 2) [117, 169]. These mechanisms involve every level of the pathway. Following signal transduction, the IFN-:IFNGR1 complex internalizes and enters the endosomal pathway, where the complex dissociates [170]. In many cell types, the IFNGR1 chain eventually recycles to the cell surface in its uncoupled, dephosphorylated form, and the ligand is degraded [45, 171,
168 Journal of Leukocyte Biology Volume 75, February 2004

172]. In some cell types, IFN- signaling may induce degradation of the internalized receptor, thereby down-regulating IFNGR1 surface expression [45]. This may be a mechanism by which these cells can prevent overstimulation of the pathway through cellular desensitization. Such a mechanism may also allow differential regulation of IFN--induced responses in different cell types [37]. One of the most inducible targets of IFN- is a specic feedback inhibitor, SOCS-1, which associates with Jak1/2, interfering with tyrosine kinase activity and inhibiting downstream IFN- signaling [173177]. SOCS-1 may also promote degradation of signaling machinery by binding to and targeting signaling molecules for the ubiquitin-proteasome pathway [178]. Overexpression of SOCS-1 in cells in vitro or in transgenic mice causes loss of responsiveness to IFN- [176, 179]. Conversely, SOCS-1/ mice are hyper-responsive to microbial infection, exhibit enhanced macrophage cytocidal activity, and die of IFN--dependent, multisystem inammatory tissue destruction in the absence of infection [180, 181]. This indicates that IFN- is normally present and performs important functions in vivo, even in the absence of infection. Another SOCS protein, SOCS-3, is also induced by IFN- and negatively regulates IFN- signaling, although perhaps less effectively than SOCS-1 [182]. Aside from SOCS protein function, IFN- signal downregulation can occur by receptor and Jak dephosphorylation by the PTPs Shp1 and Shp2 [183187]. Shp2 null cells exhibit elevated tyrosine phosphorylation and DNA-binding ability of Stat1 and show enhanced suppression of cell growth with IFN- treatment [184]. Shp1 KO mice (motheaten mice, me/ me) die from a condition caused by dysregulation of a large number of cytokines, possibly including IFN- [188]. Stat1 phosphorylation may also be controlled by nuclear dephosphorylation. Stat1 homodimers enter the nucleus and are actively retained by a mechanism dependent on the kinase function of Jak1 [189]. McBride et al. [137] proposed a model in which nuclear export is mediated by a leucine-rich nuclear export signal (NES) at residues 197205, which is hidden when Stat1 is bound to DNA. Stat1 dephosphorylation by a nuclear protein tyrosine phosphate (PTP) allows the release of DNA and subsequent recognition of the NES by chromosome region maintenance/exportin 1 (CRM1), an export receptor for NES-containing proteins [190, 191]. CRM1 then mediates Stat1 translocation through the nuclear pore in a Ran-GTPasedependent manner. This model is supported by work from other groups, demonstrating Stat1 tyrosine dephosphorylation in the nucleus and subsequent recycling of inactive Stat1 to the cytoplasm, and is consistent with the crystal structure of Stat1 bound to DNA [137, 189, 192, 193].

Signaling specicity
The many cytokines that use Jak-Stat signaling direct transcription through specic but not necessarily unique combinations of Jaks and Stats. This raises the question of how signal specicity is maintained when many cytokines use the same signaling machinery. Evidence for a devoted role of Jaks in signal transduction of specic cytokines is controversial. Mutant cell lines lacking Jak1 or Jak2 have demonstrated deciency in types I and II
http://www.jleukbio.org

Receptor expression level

IFN

JAK2

JAK1

JAK1

JAK2

JAK2

JAK1

JAK1

JAK2

SHP-2 SOCS-3 SOCS-1

Cytosol
PTPs/SOCS limit receptor and Jak phosphorylation

ST AT STA 1 T1

IRF-1

STAT1

STAT2

STAT1

STAT1

STA T1

STA T1

IRF-9

IRF-9

PTP

STA T1

STAT1

GAS

IFN-regulated gene

ISRE

Nuclear PTPs dephosphorylate Stat1 causing inactivation and nuclear export

eg.

IRF-1

IRF-2

Fig. 2. Negative regulation of IFN- signaling. A number of factors limit the extent and duration of IFN- signal transduction. Receptor IFNGR2 (green) expression is tightly regulated in some cell types, whereas IFNGR1 (yellow) surface concentration may be regulated by the extent of recycling following receptor:ligand internalization. Protein tyrosine phosphatases (PTPs), such as SH2-containing tyrosine phosphatase-2 (Shp2), dephosphorylate Jak1, Jak2, and IFNGR1, and suppressors of cytokine signaling-1 (SOCS-1) and SOCS-3 interfere with Jak activity to turn off signaling after ligand binding. Stat1 activation is down-regulated by Stat1 dephosphorylation in the nucleus. IRF-2 antagonizes transcriptional activation of many (IRF-1-inducible) genes containing ISRE or IRF-E promoter elements by competing for binding sites without promoting gene expression. Signaling molecules activated by IFN- are depicted in red text.

IFN signaling (Jak1) or IFN- signaling (Jak2) [194, 195]. These deciencies are reected in the Jak1 and Jak2 KO mice, suggesting that these Jaks are functionally specic to their cytokine systems, and other Jaks cannot substitute for them in vivo. Jak1 gene KO mice exhibit a number of abnormalities, all attributable to signaling decits in three cytokine receptor families: type II cytokine receptors (i.e., types I and II IFN and IL-10 receptors); all cytokines that use the c subunit for signaling, thereby promoting lymphopoiesis (e.g., IL-2, IL-4, IL-7 receptors); and gp130 receptor family members (e.g., IL-6, IL-11, leukemia inhibitory factor receptors) [196]. These results showed that in vivo, the role of Jak1 is essential and nonredundant in signal transduction and induction of biological response for a specic range of cytokine receptors. Jak2 gene KO in mice is embryonic lethal as a result of deciencies

in erythropoiesis [197, 198]. Other than suppression of IFN- signaling, the cytokine response decit of Jak2 KO mice is largely nonoverlapping compared with Jak1 KO mice [197, 198]. This indicates that other Jaks cannot functionally substitute for Jak2 in vivo, although Jak3 can substitute for Jak2 in vitro to produce Stat1 homodimers and activate transcription of class I major histomcompatibility complex (MHC) [199]. Although specic Stats appear to be dedicated to specic subsets of cytokine signaling, the same Stat or combination of Stats can be activated in different cytokine systems to give completely different biological effects (e.g., IFN- and oncostatin M as described earlier). Possible mechanisms giving rise to specicity of response with respect to Stats include: different thresholds/timeframes of Stat activation may be required for specic biological responses, and specic Stat com-

Schroder et al. Signals, mechanisms, and functions of IFN- 169

STA T1

IRF-1

STA T1
STAT1

Nucleus

STA T1

IFN-regulated gene

Transcriptional repression by IRF-2 competition for ISRE/IRF-E sites

plexes induced by cytokines may give different biological effects. It is widely agreed that signal specicity is largely conferred by the particular Stats activated in cytokine signaling [106, 151, 200 205]. This is supported by the Stat1 KO mouse, which displays loss of only IFN signaling, despite multiple reports of Stat1 activation in response to a range of other cytokines in vitro [106, 151]. Similarly, KO mice of the other Stat proteins showed a loss of biological responsiveness to a single cytokine, indicating that Stat activation may be specic for a single cytokine in vivo if not in vitro [169]. In addition to the Stats themselves, other factors (e.g., intracellular IFN- as described above, or cell-specic transcription factors), which associate with Stats may contribute to the specicity of the signal.

Cross-talk between IFN-/ and IFN- pathways


The IFN- and IFN-/ signal pathways cross-talk at multiple levels. The signal pathways and target genes used by types I and II IFN are partially overlapping, which gives the opportunity for cross-talk to synergize or antagonize particular functions within the cell. This cross-talk is biologically relevant, as cells in vivo are not stimulated with one cytokine in isolation, rather a cytokine cocktail instructs gene expression through the integration of multiple signaling pathways. Although IFN- primarily signals through Stat1 homodimers in the initial signal cascade, additional signaling molecules are activated. As noted above, the archetypal type I IFN signaling molecule ISGF3 is activated by IFN-, thus providing a mechanism for cross-talk between the types I and II IFN pathways [116]. ISGF3 is able to induce expression of type I IFN, thereby further amplifying the response of IFN-/-induced genes [206, 207]. Conversely, type I IFN can elicit classic type II IFN signaling molecules such as active Stat1 homodimers, which are able to bind to GAS sites to activate transcription of target genes [17, 208]. Takaoka et al. [209] published data indicating that IFN- signaling in mouse embryonic broblasts requires a constitutive subthreshold IFN-/ signal. They have suggested that IFN-/ may promote the interaction of the IFNGR2 and IFNAR1 chains in the caveolar membrane domains of the cell membrane. In this model, low level IFN-/ signal is necessary for maintaining its receptor in the phosphorylated form, thus providing a functional aid for efcient assembly of IFN-activated Stat1 homodimers.

IFN- stimulation induces a replacement of the constitutive proteasome subunits with immunoproteasome subunits. In unstimulated cells, the proteasome enzymatic subunits are 1, 2, and 5, encoded outside of the MHC locus. IFN- induces expression of new subunits, LMP2, MECL-1, and LMP7, which competitively replace 1, 2, and 5 subunits of the proteasome, respectively [210 216]. In this way, new species of proteasome are formed, consisting of LMP2:MECL-1:LMP7 (the immunoproteasome), LMP2:MECL-1:5, or 1:2:LMP7, as well as low levels of the 1:2:5 constitutive proteasome [216]. Inducible proteasome replacement is thought to be a mechanism by which IFN- can increase the quantity, quality, and repertoire of peptides for class I MHC loading. The quantity is increased, as overall expression levels of proteasome are increased. The cleavage specicity of the immunoproteasome may allow production of peptides better able to bind class I MHC and thereby increase efciency in the system. Peptide diversity is thought to increase as a result of differences in cleavage specicities between different species of proteasome. As a whole, this serves to increase levels and diversity of epitopes presented for CD8 T cell recognition in the context of class I MHC and thus increase immune surveillance (reviewed in ref. [216]). This mechanism may have evolved to ensure LMP2/LMP7/MECL-1-dependent epitopes are only produced in sites of inammation and thus avoid autoimmunity without compromising appropriate T cell stimulation [216]. The efciency of peptide generation is further increased with the IFN--induced PA28, which is composed of PA28 and PA29 subunits and associates with the proteasome and alters proteasome proteolytic cleavage preference. It is thought to gear antigen processing toward more efcient generation of TAP- and class I MHC-compatible peptides to increase overall efciency of class I MHC peptide delivery [217, 218]. The IFN--inducible TAP transporter is vital in peptide transport from the cytosol to the ER lumen [307]. TAP transiently associates with class I MHC to aid in efcient peptide loading [307, 308]. Genes encoding the TAP subunits are located within the class II MHC locus and are coordinately expressed with MHC class I [220, 221]. The class I MHC complex is composed of a heavy chain (consisting of 1, 2, and 3 domains) and a light chain (2 microglobulin) and is up-regulated by IFN- stimulation [4, 222, 224 232]. Chaperones such as tapasin and GP96 implicated in aiding in the efcient assembly of peptide:MHC I complexes are also up-regulated by IFN- [4, 309, 310].

Class II antigen presentation pathway CELLULAR EFFECTS OF IFN- Class I antigen presentation pathway
Types I and II IFN up-regulate multiple functions within the class I antigen presentation pathway to increase the quantity and diversity of peptides presented on the cell surface in the context of class I MHC. Up-regulation of cell-surface class I MHC by IFN- (Table 1) is important for host response to intracellular pathogens, as it increases the potential for cytotoxic T cell recognition of foreign peptides and thus promotes the induction of cell-mediated immunity.
170 Journal of Leukocyte Biology Volume 75, February 2004

Of the IFNs, IFN- alone can efciently up-regulate the class II antigen presenting pathway and thus promote peptide-specic activation of CD4 T cells [4, 234]. IFN- treatment further up-regulates class II MHC molecules in cells constitutively expressing class II MHC, such as B cells, DCs, and cells of the monocyte-macrophage lineage (professional APCs) [234]. IFN- is also able to induce class II MHC expression in cells that do not constitutively express these genes (nonprofessional APCs) [76]. IFN- up-regulates the quantity of peptide: MHC II complexes on the cell surface by promoting expression of several key molecules (Table 1): the Ii chain and the components of the class II MHC complex [234 239]; cathephttp://www.jleukbio.org

TABLE 1. Antigen Processing and Presentation Class I antigen presentation pathway Gene/protein up-regulated by IFN- LMP-2, LMP-7, MECL-1

IFN--Regulated Genes and Their Function in Producing IFN- Effects

Function IFN--inducible, enzymatic proteasome subunits, which replace the constitutive proteasome 1, 5, and 2 subunits. May function to optimize peptide diversity and loading to class I MHC and thus alter epitopes presented to the immune system. LMP-2 and LMP-7 map to class II MHC, and multicatalytic endopeptidase complex-like-1 (MECL-1) is encoded outside the MHC Proteasome activator (PA)28:PA28 dimer is a nonenzymatic proteasome subunit, which alters the specicity of peptides generated to increase efciency of class I MHC peptide delivery. The transporter associated with antigen processing (TAP) is a heterodimer consisting of TAP-1 and TAP-2 subunits and belongs to the adenosine 5triphosphate-binding cassette transporter family. TAP functions as a transmembrane pump to transfer peptides from the proteasome into the endoplasmic reticulum (ER) lumen. It also aids in peptide delivery to class I MHC. TAP-1 and TAP-2 map to class II MHC. The heavy chain associates with 2-microglobulin to form the MHC class I complex (MHC I). MHC I displays foreign and self-peptides on the cell surface for immune surveillance by cytotoxic T cells. The class I heavy chain is encoded by the class I MHC locus. Light chain that associates with the class I MHC heavy chain to form the MHC I. MHC I displays foreign and self-peptides on the cell surface for immune surveillance by cytotoxic T cells. The 2-microglobulin light chain is not MHC-encoded. Tapasin is a chaperone that aids in the retention of empty MHC I in the ER and peptide loading into MHC I peptide-binding cleft.

Refs. [210216]

PA28, PA28 TAP-1, TAP-2

[4, 217, 218, 219] [4, 215, 220223]

Class I MHC heavy chain

[4, 224231]

2-microglobulin

[222, 230, 232]

Tapasin

[4, 233]

Class II antigen presentation pathway Gene/protein up-regulated by IFN- 1, 2, 1, 2 MHC II chains Ii chain Function Constituents of the heterodimeric MHC II. MHC II displays foreign and selfpeptides on the cell surface for immune surveillance by CD4 T cells. The MHC II and chains are encoded by the class II MHC locus. The invariant (Ii) chain is a transmembrane chaperone, which trafcs MHC II from the ER to the MHC II endosomal compartment (MIIC). An Ii-derived peptide, CLIP (class II-associated invariant chain peptide), binds to the MHC II peptide-binding groove to prevent inappropriate peptide binding. The DMA:DMB dimer forms DM, a class II-like heterodimeric protein resident in the MIIC. DM functions to remove CLIP from the peptide-binding cleft of MHC II so that it is accessible for peptide loading. DM is encoded by the class II MHC locus. Lysosomal proteases implicated in peptide production for class II MHC loading. Class II transactivator (CIITA) is a transactivator with no DNA-binding motif. It is the limiting component of a complex that induces transcription of key genes involved in class II MHC induction, including constituents of the class II MHC complex, the Ii chain, and DM. Refs. [234237] [237239]

DMA, DMB

[235, 237]

Cathepsins B, H, L CIITA

[240, 241] [234]

Antiviral Effect Gene/protein up-regulated by IFN- PKR Function Protein kinase dsRNA-regulated (PKR) is a serine/threonine kinase activated by dsRNA, which inhibits viral protein synthesis by phosphorylating the subunit of eukaryotic translation initiation factor (eIF-2). PKR is also implicated in NF-B activation, TNF- mRNA splice regulation, induction of apoptosis, and regulation of Stat1 and Stat3 activity. The dsRNA-specic adenosine deaminase (ADAR) catalyzes the deamination of adenosine to form inosine on dsRNA substrates and thus may be responsible for the generation of edited viral mRNA. The cellular translational machinery treats inosine as guanosine, and thus, A 3 I editing of viral mRNA may cause mistranslation into nonfunctional viral proteins to inhibit viral replication. The guanylate-binding proteins (GBP) are GTPases with antiviral properties that function by an unknown mechanism. Refs. [154, 242247]

ADAR

[248]

GBP1, GBP2

[61, 249, 250]

Schroder et al. Signals, mechanisms, and functions of IFN-

171

TABLE 1. Antiproliferative Effect Gene/protein up-regulated by IFN- PKR p21, p27 Rb Function

(Continued)

Refs. [154, 242247] [251] [252] [253256] [257]

p202 Mad1

PKR is an antiviral enzyme, which functions as a serine/threonine kinase when activated by dsRNA. PKR inhibits cellular proliferation by phosphorylating the subunit of eIF-2, thereby halting protein synthesis. May also suppress c-myc function. p21 and p27 are cyclin-dependent kinase (CDK) inhibitors of the Cip/Kip family. p21 and p27 inhibit the activity of CDK2 and CDK4, respectively, causing cell cycle arrest at the G1/S checkpoint. IFN- inhibits the activity of the G1 cyclin:CDK complexes [via up-regulation of CDK inhibitors (CKIs), up-regulating CDK-activating kinase activity, and down-regulating CDC25A activity]. Decreased G1 cyclin:CDK activity results in suppression of retinoblastoma (Rb) phosphorylation, thereby increasing levels of the active (E2F-repressing) form of Rb and preventing transcription of E2F-dependent genes required for S phase. p202 is a strong cell cycle repressor that can bind to E2F and inactivate its DNA-binding activity, thereby preventing transcription of E2F-dependent genes required for S phase. Mad1 antagonizes c-myc function, thereby inhibiting proliferation. Mad1 competes with myc for max binding, forming the mad1:max heterodimer, which has similar DNA sequence specicity to myc:max but suppresses (instead of activating) transcription of genes required for S-phase progression. High levels of mad1 are often associated with differentiation. Function c-myc controls G1/S transition by activating cyclin:CDK complexes and inducing transcription of genes required for S phase. Expression of c-myc is induced by a number of growth factors and cytokines and is down-regulated by antiproliferative agents such as IFN-. IFN- regulates c-myc expression by Rb-dependent down-regulation of E2F activity, as well as Rb-independent pathways that may be a result of PKR action. c-myc is active when associated with max. c-myc activity is antagonized by mad1.

[155, 258] [259262]

Gene/protein downregulated by IFN- c-myc

Refs. [263] [264] [154] [251] [265] [259]

Apoptotic Effect Gene/protein up-regulated by IFN- IRF-1 Caspase 1 (IL-1-converting enzyme) PKR DAPs Cathepsin D Fas/Fas ligand TNF- receptor Function IRF-1 is a tumor-suppressor gene required for the induction of apoptosis by signals such as DNA damage. Many of the proapoptotic effects of IRF-1 are mediated by the IRF-1-induced caspase 1. Caspase-1 is a cysteine protease involved in the generation of bioactive IL-1 and IL-18 and implicated in mediating macrophage apoptosis. PKR is an antiviral enzyme that appears to mediate TNF--induced apoptosis. Mechanisms by which PKR affects apoptosis are still unclear but may involve induction of Fas. The death-associated proteins (DAP15) are mediators of IFN--induced apoptosis by poorly dened mechanisms. Mediates IFN-, Fas/apolipoprotein, and IFN--induced cell death by poorly dened mechanisms. IFN- may increase cellular sensitivity to apoptosis by up-regulating expression of Fas and Fas ligand. IFN- may promote cellular sensitivity to the proapoptotic effects of TNF- by promoting surface expression of a TNF- receptor on tumor cells. Refs. [165, 266, 267] [267269] [267269] [270272] [273276] [277] [278, 279] [280]

Activation of Microbial Effector Functions Production of NO intermediates Gene/protein up-regulated by IFN- iNOS/NOS2 Function The NOS enzymes (NOS1, iNOS, NOS3) catalyze the reduced nictonimide adenine dinucleotide phosphate (NADPH)-dependent conversion of L-arginine to L-citrulline, forming NO as a by-product. Of these, iNOS is the only isoform inducible by cytokine and/or microbial stimulus. Argininosuccinate synthetase produces the L-arginine substrate required for the NO-liberating reaction catalyzed by iNOS. Guanosine 5-triphosphate (GTP)-cyclohydroxylase I supplies the tetrahydrobiopterin cofactor required for NO production by iNOS. Refs. [281]

Argininosuccinate synthetase GTP-cyclohydroxylase I

[282284] [285, 286]

172 Journal of Leukocyte Biology Volume 75, February 2004

http://www.jleukbio.org

TABLE 1. Production of reactive oxygen species (ROS) Gene/protein up-regulated by IFN- gp91
phox

(Continued)

Function gp91 is a subunit of the NADPH oxidase, which associates with gp22 , gp47phox, and gp67phox to form the active complex capable of the generation of ROS during the respiratory burst. gp67phox is a subunit of the NADPH oxidase found in the cytosol in resting cells. Upon cell activation, it translocates to the phagosome membrane and associates with gp22phox, gp47phox, and gp91phox to form the active complex capable of the generation of ROS during the respiratory burst.
phox phox

Refs. [287, 288] [289]

gp67phox

Other antimicrobial mechanisms Gene/protein up-regulated by IFN- NRAMP1 FcRI C2, C4, Factor B Complement receptor CR3 (Mac-1) Function The natural resistance-associated macrophage protein (NRAMP1) confers resistance to macrophage intracellular pathogens by largely unknown mechanisms. Expression of the high-afnity Fc receptor (FcRI) is increased in myeloid cells by IFN- stimulation. FcRI binds extracellular pathogens via IgG in the adaptive phase of the immune response. Complement proteins are secreted by macrophages and broblasts in response to IFN-. Complement functions to opsonize extracellular pathogen for receptor-mediated phagocytosis by mononuclear phagocytes. Complement receptors of mononuclear phagocytes are up-regulated by IFN- to promote receptor-mediated phagocytosis of opsonized extracellular pathogens. Refs. [4, 290] [291] [292294] [295]

Immunomodulation, th Development, and Leukocyte Trafcking Gene/protein up-regulated by IFN- IL-12 IFN-inducible protein 10 (IP-10; CXCL10) Monocyte chemoattractant protein-1 (MCP)-1/JE (CCL2) MIG (CXCL9) MIP-1, MIP-1 (CCL3, CCL4) Regulated on activation, normal T expressed and secreted (RANTES; CCL5) ICAM-1 Vascular cell adhesion molecule-1 (VCAM-1) B7.2 Function NK cell activator and differentiation factor driving CD4 cell development to a Th1 phenotype. Chemoattractant for monocytes and T cells. Chemoattractant for monocyte/macrophages. Chemoattractant for T cells. Chemoattractant for CD4, CD8, and memory T cells. Chemoattractant for memory CD4 T cells and monocyte/macrophages. Adhesion molecule-binding to lymphocyte function-associated antigen-1 and Mac1. Adhesion molecule-binding to very late antigen-4. Surface molecule on APCs that provide costimulus for antigen-specic T cell activation. Refs. [296, 297] [155, 298] [299] [155] [155, 300] [301] [155] [155, 302] [303] [304] [305, 306]

sins B, H, and L, lysosomal proteases implicated in production of antigenic peptides for class II MHC loading [240, 241]; and DM, a regulator of peptide accessibility to the peptide-binding cleft of class II MHC [235, 237]. The CIITA mediates coordinate transcriptional control over these genes. Targeted disruption of CIITA in mice causes complete loss of constitutive or inducible display of class II MHC molecules [311]. Likewise, loss of constitutive/inducible class II MHC display is found in humans with CIITA mutation (Bare Lymphocyte Syndrome) [312314]. As CIITA is the limiting factor in a complex that directs transcriptional regulation, it acts as a switch for rapid up-regulation of class II MHC-related genes by IFN- [315].

IFN- and development of Th1 response


IFN- is a major product of Th1 cells and further skews the immune response toward a Th1 phenotype. IFN- achieves this by promoting characteristic Th1 effector mechanisms: innate cell-mediated immunity (via activation of NK cell effector functions), specic cytotoxic immunity (via T cell:APC interactions), and macrophage activation (discussed below) [4]. IFN--induced, specic cytotoxic immunity is promoted by direct and indirect mechanisms. IFN- promotes specic cytotoxic immunity by indirect mechanisms, such as growth inhibition of Th2 populations and up-regulation of antigen processing, presentation, and APC costimulatory molecules,
173

Schroder et al. Signals, mechanisms, and functions of IFN-

thereby increasing CD4 differentiation. IFN- also inuences na ve CD4 cell differentiation toward a Th1 phenotype more directly. The phenotype adopted by a naive T cell during T cell activation is heavily inuenced by the cytokine milieu present at the time of T cell receptor engagement. IFN- and IL-12 are the prototypic cytokines directing Th1 differentiation during the primary response to antigen, and IL-4 directs differentiation of Th2 populations. IFN- induces IL-12 production in phagocytes [296] and inhibits IL-4 secretion by Th2 populations [39], which may further drive Th1 differentiation in vivo. IL-12 and IFN- coordinate the link between pathogen recognition by innate immune cells and the induction of specic immunity, by mediating a positive feedback loop to amplify the Th1 response [4]. LPS and other pathogen-associated molecular patterns directly trigger IL-12 production upon recognition by macrophages, DCs, and neutrophils [316 319], which in turn induces IFN- secretion in antigen-stimulated, naive CD4 T cells and NK cells [320, 321]. IL-12-induced IFN- participates in positive feedback by further promoting IL-12 production in macrophages [296, 297]. This amplication may be important in initiation or stabilization of the Th1 response (reviewed in refs. [4, 322]). IL-18 is highly synergistic with IL-12 for IFN- production and has important functions in the in vivo Th response [23, 24]. Secretion of large amounts of IFN- or IL-4 is the dening feature of Th1 or Th2 cells, respectively [323]. These cytokines function to directly promote cell-mediated immunity (IFN-) or humoral immunity (IL-4) and to reciprocally antagonize each others actions, further polarizing the response (reviewed in ref. [108]). For example, IL-4 inhibits IFN--dependent activation of macrophage-effector functions [324, 325]. Conversely, IFN- antagonizes IL-4-dependent induction of the IgE receptor, FcRII [326].

chinery. Phosphorylation by PKR prevents recycling of eIF-2 GTP from its guanosine 5-diphosphate-bound form, thereby inhibiting viral and cellular protein synthesis [242]. PKR is implicated in numerous other functions, including activation of NF-B [243, 334], TNF- transcript splice regulation [244], induction of fas-mediated apoptosis [245, 246], and regulation of Stat1 activity [154, 247]. It is likely that IFNs induce many as-yet undiscovered antiviral proteins, as the enzymes above and described in Table 1 do not account for all IFN-induced antiviral effects [335, 336]. Also, the host of factors mediating the profound growth-inhibitory and proapoptotic effects of IFN- would have no small part in limiting viral replication within the cell and spread to other cells.

Cell cycle, growth, and apoptosis


Macrophages are produced in large amounts by the bone marrow, and many die shortly after production by a process of programmed cell death (apoptosis) [337]. After release from the bone marrow, macrophages proliferate, differentiate, acquire specialized functions (e.g., microbicidal activity), or die by apoptosis depending on the presence or absence of extracellular cues. Many reports indicate that IFN- arrests the macrophage cell cycle and provides a survival signal, and others suggest that it serves as a proapoptotic signal. Although recent advances in cell cycle control and apoptosis have greatly increased our understanding in these functions in isolation, the inter-relationship between these intimately related processes is still somewhat confusing. For this reason, the effects of IFN- on cellular proliferation and apoptosis are described here separately. One of the most easily observed effects of IFN- is cell growth inhibition. Types I and II IFN are able to protect against pathogen-induced apoptosis and suppress colony stimulating factor type 1 (CSF-1)-dependent growth of bone marrow-derived macrophages (BMM) [252]. Granulocyte macrophageCSF, which is usually released by T cells at the same time as IFN-, is able to protect against the growth-inhibitory effects of IFN- on macrophages, and this may be physiologically important in vivo [338]. The IFNs most commonly arrest the cell cycle at the G1/S checkpoint, although blockages of other cell cycle stages have been reported [339 341]. It has been suggested that IFN-induced G1 arrest may actually reect exit from the cell cycle into G0 [342]. IFNs inhibit proliferation primarily by increasing protein levels of Ink4 and Cip/Kip CKIs [252255, 343, 344]. IFN- transcriptionally induces p21 and p27 CKIs (Table 1) [256, 345, 346]. p21 and p27 inhibit the activity of cyclin E:CDK2 and cyclin D:CDK4 complexes, respectively, thereby arresting the cell cycle at the G1/S boundary. Decreased cyclin:CDK activity during G1 results in hypophosphorylation of the tumor suppressor Rb. In its hypophosphorylated state, Rb sequesters the E2F family of transcription factors from activation of genes (e.g., c-myc) required for cell cycle progression from G1 to S phase [292]. c-myc controls G1/S transition by activating cyclin:CDK complexes and inducing transcription of genes required for S phase (reviewed in ref. [263]). Expression of c-myc is induced rapidly by a number of growth factors and cytokines and is down-regulated by IFN- [154, 264]. IFN- inuences
http://www.jleukbio.org

The IFN-induced antiviral state


The IFN system regulates innate and adaptive immunity to viral infection. Viral invasion directly triggers induction of type I IFNs, through a mechanism involving IRF-3 and IRF-7 (reviewed in ref. [327]). Although both types of IFN are crucial in the immediate cellular response to viral infection, the immunomodulatory activities of IFN- (discussed elsewhere) become important later in the response in coordinating the immune response and establishing an antiviral state for longer term control [54, 60, 328, 329]. IFN--induced antiviral mechanisms include induction key antiviral enzymes (Table 1), most notably PKR, which is a serine/threonine kinase greatly induced by types I and II IFN stimulation [242, 330]. PKR is inactive in its constitutive form and requires an activating signal for autophosphorylation. dsRNA, a necessary intermediate in replication of RNA viruses, is the best characterized activator of PKR, although other agents such as heparin are able to activate PKR [331]. Association of PKR with dsRNA is likely to cause a conformational change that unmasks the catalytic domain responsible for PKR autophosphorylation [242, 332]. PKR is then activated for dsRNA-independent phosphorylation of specic cellular substrates [333]. One of these substrates is the eIF-2 subunit, a rate-limiting factor in the normal cellular translational ma174 Journal of Leukocyte Biology Volume 75, February 2004

c-myc expression through multiple pathways, including suppression of Rb phosphorylation, resulting in decreased E2F activity as well as Rb-independent pathways, which may be a result of PKR action [251]. In its active form, c-myc is associated with max, and the myc:max heterodimer activates transcription of genes required for cell cycle progression [265]. The level of active myc is negatively regulated by mad1, which sequesters the max coactivator away from myc and suppresses transcription of myc-inducible genes [259]. In this way, the mad1-to-c-myc ratio determines whether a cell will proliferate (as a result of high levels of myc:max) or arrest in G1 (as a result of high levels of mad1:max) [259, 260]. In addition to decreasing the abundance of myc, IFN- increases mad1 levels, thereby further antagonizing myc activity and inhibiting CSF-1-dependent proliferation in BMM [347]. Apoptosis is a process in which the cell directs its own demise by activating intrinsic suicide machinery (reviewed in refs. [337, 348]). In macrophages, apoptosis may be desirable to prevent colonization by intracellular pathogens. Conversely, many pathogens actually induce host-cell death through secretion of macrophage-specic toxins. Induction of apoptosis by signals such as DNA damage requires the IRF-1 tumor-suppressor gene [165, 266, 267]. Levels of IRF-1 may be a deciding factor in whether IFN- induces or protects from apoptosis on treated cells [35, 349, 350]. It is proposed that IFN- treatment of cells with high levels of functional IFNGR very rapidly activates Stat1, thereby producing high levels of IRF-1 that are able to induce apoptosis. In contrast, IFN- treatment of cells with low levels of functional IFNGR may activate Stat1 more slowly, thereby producing lower levels of IRF-1 that are not sufcient to induce apoptosis [35]. Experiments in which overexpression of functional IFNGR on normally low-level IFNGR-expressing cells changed the IFN- response of these cells from an antiapoptotic/proliferative phenotype to a proapoptotic phenotype are consistent with this hypothesis [35]. This may also explain why myeloid cells are more sensitive to the proapoptotic actions of IFN- than other cells such as T cells, as myeloid cells express relatively high numbers of functional IFNGR on the cell surface [35]. Many of the proapoptotic effects of IRF-1 are mediated by the IRF-1-induced caspase 1 (IL-1-converting enzyme) [267 269]. Caspase 1 is a cysteine protease implicated in mediating macrophage apoptosis by various stimuli including LPS [351 353] and is involved in generation of bioactive IL-1 and IL-18 [354]. Caspase 1 expression and resulting apoptosis can be IFN--inducible [269] or can occur through IFN--independent IRF-1 activity [267]. IFN- also induces a number of other proapoptotic molecules. These include PKR, the DAPs, cathepsin D, and surface expression of Fas and the TNF- receptor (Table 1).

Activation of microbicidal effector functions


One of the most important effects of IFN- on macrophages is the activation of microbicidal effector functions. Macrophages activated by IFN- display increased pinocytosis and receptormediated phagocytosis as well as enhanced microbial killing ability. IFN--activated microbicidal ability includes induction of the NADPH-dependent phagocyte oxidase (NADPH oxidase) system (respiratory burst), priming for NO produc-

tion, tryptophan depletion, and up-regulation of lysosomal enzymes promoting microbe destruction (Table 1) [222]. Similar microbicidal mechanisms are activated by IFN- in neutrophils [4]. Macrophages kill bacteria, viruses, protozoa, helminths, fungi, and tumor cells primarily by production of ROS and reactive nitrogen intermediates (RNI) via induction of the NADPH oxidase system and iNOS, respectively (reviewed in refs. [281, 355, 356]). ROS and RNI use the advantages of low molecular weight, reactivity, and lipophilicity to easily penetrate the microbial cell wall/coat to inict injury. The importance of these molecules in host defense is highlighted by the highly pathogen-susceptible phenotype of mice doubly decient in NADPH oxidase and iNOS enzymes [357]. The timeframes and situations in which macrophage use these cytocidal activities are generally different. ROS production becomes apparent within 1 h after stimulus, whereas RNI generation requires 24 h after encountering a pathogen product. In general, ROS are used to target extracellular pathogens during phagocytosis or that are too large for phagocytosis, whereas RNI target intracellular pathogens and upon appropriate stimulus, extracellular pathogens and tumor cells. NO is produced in the NADPH-dependent conversion of L-arginine to L-citrulline by the NOS enzymes (NOS13). The NOS2/iNOS isoform alone is inducible by cytokine and/or microbial stimulus [281]. The NO generated by this enzyme exists in equilibrium among a number of different redox states (called RNI in this review) that have differing biological activities. It should be noted that much of the work presented below applies to mice rather than humans, and there is at present much controversy of the physiological relevance of NO production in human host defense (reviewed in ref. [281]). A wealth of evidence supports a crucial function of iNOS and RNI in host defense (reviewed in ref. [281]). IFN-dependent RNI production is associated with increased ability of phagocytic cells to kill ingested pathogens, and this ability is inhibited by the addition of a substrate analog inhibitor of iNOS such as N-methyl-L-arginine [358 361]. Furthermore, mice in which the iNOS gene has been mutated show greater susceptibility to parasitic infection and a dampened, nonspecic inammatory response [362]. Production of RNI also inhibits replication of a range of viruses, through ill-dened mechanisms [358, 363, 364]. IFN- induces RNI production by up-regulating expression of substrate, cofactor, and catalyst required for NO generation. IFN- up-regulates argininosuccinate synthetase (which produces the L-arginine substrate), GTP-cyclohydroxylase I (which supplies the tetrahydrobiopterin cofactor required for NO production), and the iNOS enzyme [285, 286, 365367]. Maximal induction of iNOS transcription requires priming and triggering stimuli such as priming with IFN- and subsequent triggering with LPS or TNF- [368, 369]. The type I IFNs cannot fully substitute for IFN- in triggering NO production [368]. Superoxide and its reactive products are also important toxic effector molecules of the nonspecic cytotoxic response. The superoxide anion, O2, is generated by the multicomponent avocytochrome enzyme phagocyte oxidase during a process called respiratory burst. This enzyme is composed of two

Schroder et al. Signals, mechanisms, and functions of IFN- 175

cytochrome b558 subunits, gp91phox and gp22phox, concentrated in the phagosome membrane, and two cytosolic components, p47phox and p67phox. Upon appropriate stimulus (e.g., phagocytosis), the cytosolic components translocate to the membrane to form the active complex, which generates superoxide in the phagosome through transfer of a transported electron to molecular oxygen. The superoxide anion generated by the respiratory burst spontaneously reacts to form hydrogen peroxide (H2O2), hydroxyl radicals (OH), and hypochlorous acid (HOCl) [281]. The toxic oxidants produced by the respiratory burst are also able to react with those produced by iNOS, thereby forming a large number of different toxic species (e.g., peroxynitrite) to mediate cytotoxicity by a wide variety of mechanisms [281, 370, 371]. The primary mechanism of IFN--induced up-regulation of ROS production in phagocytes is transcriptional induction of the gp91phox and p67phox subunits of the NADPH oxidase complex [287289, 372]. The biological importance of this pathway in host defense is emphasized by the human genetic disease, chronic granulomatous disease, caused by defects in the NADPH oxidase system and characterized by recurrent and often fatal infections. IFN- also promotes microbe destruction by augmenting surface expression of the high-afnity FcRI on mononuclear phagocytes, thereby promoting antibody-dependent, cell-mediated cytotoxicity [291]. Complement-mediated phagocytosis is also up-regulated by IFN- through increased complement secretion and complement receptor surface expression on mononuclear phagocytes [292]. Tryptophan depletion from cells in response to IFN- may be an antiparasitic mechanism in humans [373]. IFN- upregulates the expression of indoeamine 2,3 dioxygenase, an enzyme responsible for the generation of N-formylkynurenine from tryptophan [374 377]. It is interesting that IFN- also induces expression of tryptophanyl-tRNA synthetase, a housekeeping gene that catalyzes tRNA Trp aminoacetylation during protein synthesis [378, 379]. It is proposed that this enzyme may also aid in tryptophan depletion by incorporating free tryptophan into translating protein and may ensure synthesis of Trp-rich proteins with immunologically important functions (e.g., the IRF family and many MHC molecules) during cellular tryptophan depletion [380].

chemokine milieu to extravasate into the tissue via interactions between adhesion molecules presented on leukocyte and endothelial surfaces (diapedesis) [4]. IFN- regulates this process by up-regulating expression of chemokines (e.g., IP-10, MCP-1, MIG, MIP-1/, RANTES; see Table 1) and adhesion molecules (e.g., ICAM-1, VCAM-1; see Table 1). TNF- and IL-1 synergistically regulate many of these molecules [4].

IFN- priming of the macrophage LPS response


LPS/endotoxin is a cell wall constituent of gram-negative bacteria that activates macrophage microbicidal effector functions and the production of proinammatory cytokines (e.g., TNF-, IL-1, IL-6). LPS:macrophage interaction can be benecial or deleterious to the host. Macrophages sense and are activated to ght and clear bacterial infection through LPS recognition, a mechanism obviously advantageous to the host; however, high doses of LPS that trigger excessive production of inammatory mediators can result in the potentially lethal systemic disorder, septic shock. Macrophages recognition of lipid A (the active moiety of LPS), requires Toll-like receptor (TLR) family member TLR4. The TLR family is a homolog of the Drosophila antifungal protein Toll. To date, 10 mammalian TLRs have been identied (TLR110), and many have proposed or demonstrated functions in innate immunity. LPS:TLR4 interaction triggers a signaling cascade (Fig. 3), which ultimately results in activation of NF-B and activated protein-1 (AP-1) and transcriptional control over genes directing immune function in macrophages (reviewed in refs. [383386]). MyD88 functions as an adaptor to connect the intracellular portion of TLR4 to downstream signaling components and is required for many, but not all, LPS-induced effects [387]. For this reason, a TLR4 signaling paradigm has emerged in which LPS signals through MyD88-dependent and MyD88-independent pathways. MyD88 gene KO mice demonstrated that the MyD88-dependent pathway is necessary for rapid activation of MAPK, AP-1, and NF-B, the production of inammatory cytokines, and septic shock [387]. The MyD88-independent pathway results in slower activation of MAPK, AP-1, and NF-B and does not contribute to the same extent to the inammatory response. IFN- primes macrophages for more rapid and heightened responses to LPS [58, 388, 389] as well as other TLR agonists, such as unmethylated CpG motifs present in bacterial DNA (CpG DNA) [390]. Pretreatment with IFN- is necessary for the induction of some genes in response to LPS. iNOS is the archetypal example of such a gene, although the need for exogenous priming by IFN- can be overcome by adequate production of endogenous type I IFN in response to high doses of LPS [391]. For other genes, IFN- is able to shift the dose-response curve such that transcription is induced at lower concentrations of LPS, thereby superinducing transcription in IFN--primed cells treated with LPS compared with the same concentration of LPS alone [392]. IFN- priming does not occur for all LPS-induced responses; IFN- appears to prime for a select subset of LPS-induced genes. For example, IFN- pretreatment promotes LPS-induced TNF- but not procoagulant activity enzyme production [393]. IFN- receptor KO mice are highly resistant to LPS-induced toxicity [394]. This highlights the physiological signicance of
http://www.jleukbio.org

Immunomodulation and leukocyte trafcking


The ability of IFN- to coordinate the transition from innate immunity to adaptive immunity distinguishes it from the other IFNs. Mechanisms by which IFN- coordinates this transition include aiding in the development of a Th1-type response (described earlier), directly promoting B cell isotype switching to IgG2a [54, 381, 382], and regulation of local leukocyteendothelial interactions (Table 1). IFN- orchestrates the trafcking of specic immune cells to sites of inammation through up-regulating expression of adhesion molecules and chemokines. Unstimulated leukocytes cycle continuously between the blood and the lymph. IFN- and NO produced at the site of inammation cause local dilation of the blood vessels, thereby decreasing the local blood ow rate and causing gathering of blood in leaky vessels. Specic leukocyte subsets are instructed by the cytokine/
176 Journal of Leukocyte Biology Volume 75, February 2004

LBP

LPS CD14

MD-2

TLR4


Cytosol

MyD88

MyD88-dependent pathway

IRAK TRAF6 NIK IKK

TRIF?

MyD88-independent pathway

!
p38 MAPK JNK/SAPK ERK

IRF-3

Rapid activation of NFB and AP-1

Slow activation of NFB and AP-1

"
NF-B AP-1

#
Transcription of IFN/

Nucleus

Transcription of target genes


eg. IL-1, IL-6, TNF

Transcription of target genes


eg. GARG16, IRG-1, IP-10

Fig. 3. The current paradigm for optimal LPS signal transduction through TLR4. The LPS:LPS-binding protein (LBP) complex binds to CD14. The LPS:LBP:CD14 complex probably activates TLR4, and optimal LPS-induced signaling requires the association of the MD-2 accessory component with the TLR4 extracellular domain. Downstream signaling is dependent on sequential recruitment and activation of signaling molecules. In the MyD88-dependent pathway, the signaling adaptor MyD88 is recruited to the receptor, and the signal is relayed via sequential recruitment and activation of the serine/threonine kinase IL-1R-associated kinase (IRAK), TNF receptor-associated factor (TRAF)6, NF-B-inducing kinase (NIK), and IB kinase for the rapid activation of NF-B and AP-1 (IKK). MyD88 adaptor-like/Toll/IL-1 receptor/resistance (TIR) domain-containing adaptor protein also contributes to the MyD88-dependent pathway, through ill-dened mechanisms. The MyD88-independent pathway is less well characterized but involves activation of IRF-3 and more slowly activates NF-B and AP-1. It has recently been suggested that TIR domain-containing adaptor-inducing IFN- (TRIF) participates in the MyD88-independent pathway upstream of IRF-3. Activation of the MyD88-dependent and -independent pathways also activates the mitogen-activated protein kinase (MAPK) pathways, c-Jun N-terminal kinase (JNK)/stress-activated protein kinase (SAPK), p38 MAPK, and extracellular-regulated kinase (ERK) pathways. The MyD88-dependent pathway ultimately controls transcription of target genes, including many of the proinammatory cytokines implicated in the pathophysiology of septic shock. Genes regulated by the MyD88-independent pathway, however, do not contribute to septic shock to the same extent. Possible mechanisms giving rise to the priming effect apparent when LPS-stimulated macrophages are pretreated with IFN- include (numbered on the gure): (1) IFN- up-regulates expression of the TLR4 and MD-2 components of the LPS recognition complex; (2) IFN- upregulates expression of MyD88 and IRAK; (3) LPS activates Stat1, possibly through the MAPK pathway, thus augmenting the IFN- response; (4) IFN- promotes LPS-induced NF-B activation; (5) LPS induces type I IFN, which may participate in cross-talk with the IFN- signaling pathway to augment the macrophage IFN- response; and (6) LPS/TNF-- and IFN--activated/induced signaling molecules can often bind to promoter regions of target genes to synergistically regulate transcription. GARG, Glucocorticoid attenuated response gene; IRG, immune-responsive gene.

IFN- priming in vivo, as it demonstrates that IFN- is normally produced during the response to LPS and acts to amplify LPS-induced cellular responses. The priming phenomenon relies on IFN- and LPS reciprocal cross-regulation of signaling molecules of the other pathway. IFN- inuences LPS-dependent signaling capabilities by promoting ligand-receptor interactions as well as downstream signaling machinery. Efcient LPS:TLR4 interaction requires the MD-2 and CD14 accessory molecules, and subsequent maximal signaling requires MyD88 signaling adaptor. In a number of experiments, IFN- was shown to promote transcription of TLR4, subsequent TLR4 surface expression, and LPSbinding ability in macrophages [395397]. Furthermore, IFN-

inhibits the LPS-dependent down-regulation of TLR4 surface expression apparent in unprimed cells [395]. IFN- stimulation may also promote LPS-dependent signaling by promoting expression of the MD-2 accessory molecule, the MyD88 adaptor, and the IRAK signaling molecule [395, 398]. LPS mediates many of its effects through the NF-B transcription factor. In the macrophage-like cell line RAW264.7, it was found that IFN- pretreatment promoted NF-B activation upon LPS exposure, as well as more rapid DNA-binding kinetics and faster degradation of the NF-B inhibitor, IB- [399]. In human monocytes also, IFN- pretreatment causes superinduction of active NF-B upon LPS treatment [400]. When these cells were unstimulated, they exhibited high levels

Schroder et al. Signals, mechanisms, and functions of IFN- 177

of the p50 subunit of NF-B but only low levels of p65. IFN- priming caused an increase in p65 mRNA as a result of increased transcript stability. Many IFN--inducible genes are also TNF--inducible, and these genes are often superinduced by the combination of these factors [366, 372, 401, 402]. TNF- is a macrophage-derived cytokine secreted in response to LPS, which can act in an autocrine manner to mediate many LPS-induced effects via NF-B [403]. IFN- priming of TNF- responses is thus responsible for the priming of a subset of LPS-induced genes. In many cases, synergistic induction by the combination of these factors may be a result of the combined presence of Stat1 and NF-B-binding sites in the promoter elements of responsive genes [160, 404]. Synergy may be also be a result of cross-talk between the IFN- and TNF- signaling pathways: TNF- stimulation is able to induce transcription of IRF-1 and promotes IFN--induced Stat1 activation [405, 406]; IFN-induced PKR is able to signal to NF-B [407]; and IFN- increases surface expression of the TNF- receptor [280, 408, 409], although the signicance of this is uncertain, as an increased receptor expression level has not been linked with increased response to TNF- [410, 411]. Although probably not complete, these mechanisms of increasing cellular sensitivity to LPS may start to explain the shift in the LPS dose-response curve seen with IFN- priming. Genes that need IFN- and LPS treatment for induction may require a certain threshold of LPS signaling for transactivation, and so IFN- treatment shifts the LPS dose-response curve by increasing cellular sensitivity. Alternatively, these genes may require specic cross-talk between pathways before transcriptional activation can occur. Examples of such cross-talk may include LPS-induced Stat1 serine phosphorylation or crosstalk between IFN-- and LPS-induced type I IFN pathways. LPS is able to promote the IFN- signaling pathway through Stat1. As noted earlier, phosphorylation of S727 of Stat1 occurs in response to IFN- or LPS exposure and is necessary for maximal Stat1 activation. LPS treatment following IFN- exposure increases the percentage of molecules phosphorylated on Y701 and S727 and subsequent Stat1 DNA-binding activity relative to IFN- treatment alone, thereby augmenting IFN-dependent, Stat1-mediated gene expression [123, 223]. IFN- and LPS may use different pathways for Stat1 serine phosphorylation, as LPS-induced Stat1 serine phosphorylation is sensitive to a p38 MAPK inhibitor, whereas the serine phosphorylation induced by IFN- is not [110]. Another study suggested that although p38 MAPK is necessary, it is not sufcient for Stat1 serine phosphorylation by LPS, and protein kinase C- is also required to mediate phosphorylation [412]. In addition to promoting serine phosphorylation, LPS treatment following IFN- exposure augments Stat1 tyrosine phosphorylation of residue 701 [399]. Macrophage exposure to LPS triggers production of endogenous type I IFN that can act on the macrophage population in an autocrine/paracrine manner. This has been identied as the causal agent of LPS-induced Stat1 tyrosine phosphorylation [413]. It has long been recognized that LPS induces transcription of type I IFN [414], but only in recent years has the signicance of autocrine/paracrine functions of type I IFN in mediating LPS-induced gene expression been acknowledged [391, 415,
178 Journal of Leukocyte Biology Volume 75, February 2004

416]. Stat1 is tyrosine-phosphorylated as a result of autocrine/ paracrine type I IFN and is required for the full LPS response in IFN--unprimed macrophages [416]. LPS-induced Stat1 Y701 phosphorylation was found to be essential for iNOS expression [413], and macrophages from mice unable to respond to type I IFN through targeted disruption of the IFNAR1 receptor subunit exhibited limited capacity to induce iNOS in response to LPS [391]. iNOS-producing capacity in response to LPS was able to be rescued by exposure to IFN-, which presumably restored Stat1 activation [391]. LPS induces transcription of type I IFN through an ill-dened, MyD88-independent pathway involving IRF-3 [417]. The contribution of type I IFN to IFN- priming of the LPS response has not been formally studied; however, as IFN- up-regulates expression of IFNAR in other cell types (P. J. Hertzog, unpublished observations) and promotes IFN-/ expression, it is possible that IFN- sensitizes the cells to LPS-induced IFN-/. When the many levels of cross-talk between the IFN- and IFN-/ pathway are considered, it is likely that in vivo IFN- potentiates LPS-induced IFN-/ effects and conversely, LPS augments IFN--mediated effects (e.g., by Stat1 Y701 and S727 phosphorylation). Synergy between LPS- and IFN--induced transcription factors in expression of target genes also contributes to the priming phenotype. Genes such as IRF-1, IP-10, ICAM-1, and iNOS contain Stat1 and NF-B-binding sites in their promoter, and maximal transcription requires both signals [158, 160, 402, 404, 418 423]. This is a demonstrated mechanism in IFN--dependent gene superinduction in response to LPS for a number of genes and is likely to be a global mechanism for synergistic, coordinate regulation of large synexpression groups. Genes that require IFN- and LPS treatment for induction (e.g., iNOS) may also be explained by this mechanism, as it is likely that the IFN-- and LPS-induced transcription factors are limiting for gene induction at a single-cell level. This theory is consistent with the all-or-nothing transcription of iNOS and many other IFN-/LPS-regulated genes observed at a single-cell level (ref. [424] and references therein).

CONCLUDING REMARKS
IFN- is a remarkable cytokine that orchestrates many distinct cellular programs through transcriptional control over large numbers of genes. Many IFN--induced effects resulting in heightened immune surveillance and immune system function during infection have been discussed in this review. As pathogen products such as LPS and CpG DNA augment local IFN- production, and IFN- augments the immune system response to these agonists, an important function of IFN- during in vivo infection is suggested, whereby IFN- participates in an amplication loop to increase immune system sensitivity and response to pathogens. This model is supported by the in vivo and in vitro priming phenotype apparent when LPS-stimulated macrophages are pretreated with IFN-, and also by the high resistance of IFNGR/ mice to LPS-induced septic shock. The ability of IFN- to synergize or antagonize the effects of cytokines, growth factors, and pathogen-associated molecular pattern (PAMP)-signaling pathways (e.g., TNF-, IL-4, CSF-1,
http://www.jleukbio.org

IFN-/, LPS, and CpG DNA) is particularly important in macrophage biology, as macrophages constantly receive multiple signals and need to integrate them to give a response appropriate to the extracellular milieu. The reductionist approach used by many scientists has greatly furthered our understanding of mechanisms by which IFN- coordinates its pleiotropic effects. However, a cell is not exposed to one stimulus in isolation in vivo, and so, one of the great challenges to IFN biologists today is to understand the mechanisms by which IFN- signaling and cellular effects integrate with other growth factor, cytokine, and PAMP signaling pathways. This understanding will enable a more realistic view of in vivo macrophage function during naturally occurring human infection.

17. 18.

19.

20.

21.

22.

23.

ACKNOWLEDGMENT
The authors thank Katharine Irvine for critical reading of the manuscript and helpful suggestions.

24.

25.

26.

REFERENCES
27. 1. Adams, D. O. (1989) Molecular interactions in macrophage activation. Immunol. Today 10, 3335. 2. Perussia, B., Dayton, E. T., Fanning, V., Thiagarajan, P., Hoxie, J., Trinchieri, G. (1983) Immune interferon and leukocyte-conditioned medium induce normal and leukemic myeloid cells to differentiate along the monocytic pathway. J. Exp. Med. 158, 2058 2080. 3. Young, H. A., Hardy, K. J. (1995) Role of interferon-gamma in immune cell regulation. J. Leukoc. Biol. 58, 373381. 4. Boehm, U., Klamp, T., Groot, M., Howard, J. C. (1997) Cellular responses to interferon-gamma. Annu. Rev. Immunol. 15, 749 795. 5. Carnaud, C., Lee, D., Donnars, O., Park, S. H., Beavis, A., Koezuka, Y., Bendelac, A. (1999) Cutting edge: cross-talk between cells of the innate immune system: NKT cells rapidly activate NK cells. J. Immunol. 163, 4647 4650. 6. Finkelman, F. D., Katona, I. M., Mosmann, T. R., Coffman, R. L. (1988) IFN-gamma regulates the isotypes of Ig secreted during in vivo humoral immune responses. J. Immunol. 140, 10221027. 7. Isaacs, A., Lindermann, J. (1957) Virus interference. I. The interferon. Proc. R. Soc. Lond. B Biol. Sci. 147, 258 267. 8. Bach, E. A., Aguet, M., Schreiber, R. D. (1997) The IFN gamma receptor: a paradigm for cytokine receptor signaling. Annu. Rev. Immunol. 15, 563591. 9. Jonasch, E., Haluska, F. G. (2001) Interferon in oncological practice: review of interferon biology, clinical applications, and toxicities. Oncologist 6, 34 55. 10. Bazer, F. W., Spencer, T. E., Ott, T. L. (1997) Interferon tau: a novel pregnancy recognition signal. Am. J. Reprod. Immunol. 37, 412 420. 11. Young, H. A. (1996) Regulation of interferon-gamma gene expression. J. Interferon Cytokine Res. 16, 563568. 12. Frucht, D. M., Fukao, T., Bogdan, C., Schindler, H., OShea, J. J., Koyasu, S. (2001) IFN-gamma production by antigen-presenting cells: mechanisms emerge. Trends Immunol. 22, 556 560. 13. Gessani, S., Belardelli, F. (1998) IFN-gamma expression in macrophages and its possible biological signicance. Cytokine Growth Factor Rev. 9, 117123. 14. Yoshimoto, T., Takeda, K., Tanaka, T., Ohkusu, K., Kashiwamura, S., Okamura, H., Akira, S., Nakanishi, K. (1998) IL-12 up-regulates IL-18 receptor expression on T cells, Th1 cells, and B cells: synergism with IL-18 for IFN-gamma production. J. Immunol. 161, 3400 3407. 15. Flaishon, L., Hershkoviz, R., Lantner, F., Lider, O., Alon, R., Levo, Y., Flavell, R. A., Shachar, I. (2000) Autocrine secretion of interferon gamma negatively regulates homing of immature B cells. J. Exp. Med. 192, 13811388. 16. Harris, D. P., Haynes, L., Sayles, P. C., Duso, D. K., Eaton, S. M., Lepak, N. M., Johnson, L. L., Swain, S. L., Lund, F. E. (2000) Reciprocal

28.

29.

30.

31.

32.

33.

34.

35.

36.

37.

38.

regulation of polarized cytokine production by effector B and T cells. Nat. Immunol. 1, 475 482. Sen, G. C. (2001) Viruses and interferons. Annu. Rev. Microbiol. 55, 255281. Golab, J., Zagozdzon, Stoklosal, T., Kaminski, R., Kozar, K., Jakobisiak, M. (2000) Direct stimulation of macrophages by IL-12 and IL-18 a bridge too far? Immunol. Lett. 72, 153157. Munder, M., Mallo, M., Eichmann, K., Modolell, M. (2001) Direct stimulation of macrophages by IL-12 and IL-18 a bridge built on solid ground. Immunol. Lett. 75, 159 160. Munder, M., Mallo, M., Eichmann, K., Modolell, M. (1998) Murine macrophages secrete interferon gamma upon combined stimulation with interleukin (IL)-12 and IL-18: a novel pathway of autocrine macrophage activation. J. Exp. Med. 187, 21032108. Fukao, T., Matsuda, S., Koyasu, S. (2000) Synergistic effects of IL-4 and IL-18 on IL-12-dependent IFN-gamma production by dendritic cells. J. Immunol. 164, 64 71. Otani, T., Nakamura, S., Toki, M., Motoda, R., Kurimoto, M., Orita, K. (1999) Identication of IFN-gamma-producing cells in IL-12/IL-18treated mice. Cell. Immunol. 198, 111119. Akira, S. (2000) The role of IL-18 in innate immunity. Curr. Opin. Immunol. 12, 59 63. Dinarello, C. A. (1999) IL-18: a TH1-inducing, proinammatory cytokine and new member of the IL-1 family. J. Allergy Clin. Immunol. 103, 1124. Salazar-Mather, T. P., Hamilton, T. A., Biron, C. A. (2000) A chemokineto-cytokine-to-chemokine cascade critical in antiviral defense. J. Clin. Invest. 105, 985993. Pien, G. C., Satoskar, A. R., Takeda, K., Akira, S., Biron, C. A. (2000) Cutting edge: selective IL-18 requirements for induction of compartmental IFN-gamma responses during viral infection. J. Immunol. 165, 4787 4791. Schindler, H., Lutz, M. B., Rollinghoff, M., Bogdan, C. (2001) The production of IFN-gamma by IL-12/IL-18-activated macrophages requires STAT4 signaling and is inhibited by IL-4. J. Immunol. 166, 30753082. Fukao, T., Frucht, D. M., Yap, G., Gadina, M., OShea, J. J., Koyasu, S. (2001) Inducible expression of Stat4 in dendritic cells and macrophages and its critical role in innate and adaptive immune responses. J. Immunol. 166, 4446 4455. Hochrein, H., Shortman, K., Vremec, D., Scott, B., Hertzog, P., OKeeffe, M. (2001) Differential production of IL-12, IFN-alpha, and IFN-gamma by mouse dendritic cell subsets. J. Immunol. 166, 5448 5455. Major, A. S., Cuff, C. F. (1996) Effects of the route of infection on immunoglobulin G subclasses and specicity of the reovirus-specic humoral immune response. J. Virol. 70, 5968 5974. Autenrieth, I. B., Reissbrodt, R., Saken, E., Berner, R., Vogel, U., Rabsch, W., Heesemann, J. (1994) Desferrioxamine-promoted virulence of Yersinia enterocolitica in mice depends on both desferrioxamine type and mouse strain. J. Infect. Dis. 169, 562567. Autenrieth, I. B., Beer, M., Bohn, E., Kaufmann, S. H., Heesemann, J. (1994) Immune responses to Yersinia enterocolitica in susceptible BALB/c and resistant C57BL/6 mice: an essential role for gamma interferon. Infect. Immun. 62, 2590 2599. Bazan, J. F. (1990) Structural design and molecular evolution of a cytokine receptor superfamily. Proc. Natl. Acad. Sci. USA 87, 6934 6938. Thoreau, E., Petridou, B., Kelly, P. A., Djiane, J., Mornon, J. P. (1991) Structural symmetry of the extracellular domain of the cytokine/growth hormone/prolactin receptor family and interferon receptors revealed by hydrophobic cluster analysis. FEBS Lett. 282, 26 31. Bernabei, P., Coccia, E. M., Rigamonti, L., Bosticardo, M., Forni, G., Pestka, S., Krause, C. D., Battistini, A., Novelli, F. (2001) Interferongamma receptor 2 expression as the deciding factor in human T, B, and myeloid cell proliferation or death. J. Leukoc. Biol. 70, 950 960. Pernis, A., Gupta, S., Gollob, K. J., Garfein, E., Coffman, R. L., Schindler, C., Rothman, P. (1995) Lack of interferon gamma receptor beta chain and the prevention of interferon gamma signaling in TH1 cells. Science 269, 245247. Bach, E. A., Szabo, S. J., Dighe, A. S., Ashkenazi, A., Aguet, M., Murphy, K. M., Schreiber, R. D. (1995) Ligand-induced autoregulation of IFNgamma receptor beta chain expression in T helper cell subsets. Science 270, 12151218. Maggi, E., Parronchi, P., Manetti, R., Simonelli, C., Piccinni, M. P., Rugiu, F. S., De Carli, M., Ricci, M., Romagnani, S. (1992) Reciprocal regulatory effects of IFN-gamma and IL-4 on the in vitro development of human Th1 and Th2 clones. J. Immunol. 148, 21422147.

Schroder et al. Signals, mechanisms, and functions of IFN- 179

39. Gajewski, T. F., Fitch, F. W. (1988) Anti-proliferative effect of IFNgamma in immune regulation. I. IFN-gamma inhibits the proliferation of Th2 but not Th1 murine helper T lymphocyte clones. J. Immunol. 140, 4245 4252. 40. Kaplan, D. H., Greenlund, A. C., Tanner, J. W., Shaw, A. S., Schreiber, R. D. (1996) Identication of an interferon-gamma receptor alpha chain sequence required for JAK-1 binding. J. Biol. Chem. 271, 9 12. 41. Greenlund, A. C., Farrar, M. A., Viviano, B. L., Schreiber, R. D. (1994) Ligand-induced IFN gamma receptor tyrosine phosphorylation couples the receptor to its signal transduction system (p91). EMBO J. 13, 15911600. 42. Farrar, M. A., Fernandez-Luna, J., Schreiber, R. D. (1991) Identication of two regions within the cytoplasmic domain of the human interferongamma receptor required for function. J. Biol. Chem. 266, 19626 19635. 43. Farrar, M. A., Campbell, J. D., Schreiber, R. D. (1992) Identication of a functionally important sequence in the C terminus of the interferongamma receptor. Proc. Natl. Acad. Sci. USA 89, 11706 11710. 44. Cook, J. R., Jung, V., Schwartz, B., Wang, P., Pestka, S. (1992) Structural analysis of the human interferon gamma receptor: a small segment of the intracellular domain is specically required for class I major histocompatibility complex antigen induction and antiviral activity. Proc. Natl. Acad. Sci. USA 89, 1131711321. 45. Farrar, M. A., Schreiber, R. D. (1993) The molecular cell biology of interferon-gamma and its receptor. Annu. Rev. Immunol. 11, 571 611. 46. Kotenko, S. V., Izotova, L. S., Pollack, B. P., Mariano, T. M., Donnelly, R. J., Muthukumaran, G., Cook, J. R., Garotta, G., Silvennoinen, O., Ihle, J. N., et al. (1995) Interaction between the components of the interferon gamma receptor complex. J. Biol. Chem. 270, 2091520921. 47. Bach, E. A., Tanner, J. W., Marsters, S., Ashkenazi, A., Aguet, M., Shaw, A. S., Schreiber, R. D. (1996) Ligand-induced assembly and activation of the gamma interferon receptor in intact cells. Mol. Cell. Biol. 16, 3214 3221. 48. Pestka, S., Kotenko, S. V., Muthukumaran, G., Izotova, L. S., Cook, J. R., Garotta, G. (1997) The interferon gamma (IFN-gamma) receptor: a paradigm for the multichain cytokine receptor. Cytokine Growth Factor Rev. 8, 189 206. 49. Muthukumaran, G., Donnelly, R. J., Ebensperger, C., Mariano, T. M., Garotta, G., Dembic, Z., Poast, J., Baron, S., Pestka, S. (1996) The intracellular domain of the second chain of the interferon-gamma receptor is interchangeable between species. J. Interferon Cytokine Res. 16, 1039 1045. 50. Gibbs, V. C., Williams, S. R., Gray, P. W., Schreiber, R. D., Pennica, D., Rice, G., Goeddel, D. V. (1991) The extracellular domain of the human interferon gamma receptor interacts with a species-specic signal transducer. Mol. Cell. Biol. 11, 5860 5866. 51. Hibino, Y., Kumar, C. S., Mariano, T. M., Lai, D. H., Pestka, S. (1992) Chimeric interferon-gamma receptors demonstrate that an accessory factor required for activity interacts with the extracellular domain. J. Biol. Chem. 267, 37413749. 52. Hemmi, S., Merlin, G., Aguet, M. (1992) Functional characterization of a hybrid human-mouse interferon gamma receptor: evidence for speciesspecic interaction of the extracellular receptor domain with a putative signal transducer. Proc. Natl. Acad. Sci. USA 89, 27372741. 53. Kalina, U., Ozmen, L., Di Padova, K., Gentz, R., Garotta, G. (1993) The human gamma interferon receptor accessory factor encoded by chromosome 21 transduces the signal for the induction of 2,5-oligoadenylatesynthetase, resistance to virus cytopathic effect, and major histocompatibility complex class I antigens. J. Virol. 67, 17021706. 54. Huang, S., Hendriks, W., Althage, A., Hemmi, S., Bluethmann, H., Kamijo, R., Vilcek, J., Zinkernagel, R. M., Aguet, M. (1993) Immune response in mice that lack the interferon-gamma receptor. Science 259, 17421745. 55. Pearl, J. E., Saunders, B., Ehlers, S., Orme, I. M., Cooper, A. M. (2001) Inammation and lymphocyte activation during mycobacterial infection in the interferon-gamma-decient mouse. Cell. Immunol. 211, 4350. 56. van den Broek, M. F., Muller, U., Huang, S., Zinkernagel, R. M., Aguet, M. (1995) Immune defence in mice lacking type I and/or type II interferon receptors. Immunol. Rev. 148, 518. 57. Buchmeier, N. A., Schreiber, R. D. (1985) Requirement of endogenous interferon-gamma production for resolution of Listeria monocytogenes infection. Proc. Natl. Acad. Sci. USA 82, 7404 7408. 58. Kamijo, R., Le, J., Shapiro, D., Havell, E. A., Huang, S., Aguet, M., Bosland, M., Vilcek, J. (1993) Mice that lack the interferon-gamma receptor have profoundly altered responses to infection with Bacillus Calmette-Guerin and subsequent challenge with lipopolysaccharide. J. Exp. Med. 178, 14351440.

59. Suzuki, Y., Orellana, M. A., Schreiber, R. D., Remington, J. S. (1988) Interferon-gamma: the major mediator of resistance against Toxoplasma gondii. Science 240, 516 518. 60. Muller, U., Steinhoff, U., Reis, L. F., Hemmi, S., Pavlovic, J., Zinkernagel, R. M., Aguet, M. (1994) Functional role of type I and type II interferons in antiviral defense. Science 264, 1918 1921. 61. Kimura, T., Kadokawa, Y., Harada, H., Matsumoto, M., Sato, M., Kashiwazaki, Y., Tarutani, M., Tan, R. S., Takasugi, T., Matsuyama, T., Mak, T. W., Noguchi, S., Taniguchi, T. (1996) Essential and non-redundant roles of p48 (ISGF3 gamma) and IRF-1 in both type I and type II interferon responses, as revealed by gene targeting studies. Genes Cells 1, 115124. 62. Biron, C. A. (2001) Interferons alpha and beta as immune regulatorsa new look. Immunity 14, 661 664. 63. Dofnger, R., Jouanguy, E., Dupuis, S., Fondaneche, M. C., Stephan, J. L., Emile, J. F., Lamhamedi-Cherradi, S., Altare, F., Pallier, A., Barcenas-Morales, G., Meinl, E., Krause, C., Pestka, S., Schreiber, R. D., Novelli, F., Casanova, J. L. (2000) Partial interferon-gamma receptor signaling chain deciency in a patient with bacille Calmette-Guerin and Mycobacterium abscessus infection. J. Infect. Dis. 181, 379 384. 64. Jouanguy, E., Lamhamedi-Cherradi, S., Lammas, D., Dorman, S. E., Fondaneche, M. C., Dupuis, S., Dofnger, R., Altare, F., Girdlestone, J., Emile, J. F., Ducoulombier, H., Edgar, D., Clarke, J., Oxelius, V. A., Brai, M., Novelli, V., Heyne, K., Fischer, A., Holland, S. M., Kumararatne, D. S., Schreiber, R. D., Casanova, J. L. (1999) A human IFNGR1 small deletion hotspot associated with dominant susceptibility to mycobacterial infection. Nat. Genet. 21, 370 378. 65. Jouanguy, E., Lamhamedi-Cherradi, S., Altare, F., Fondaneche, M. C., Tuerlinckx, D., Blanche, S., Emile, J. F., Gaillard, J. L., Schreiber, R., Levin, M., Fischer, A., Hivroz, C., Casanova, J. L. (1997) Partial interferon-gamma receptor 1 deciency in a child with tuberculoid bacillus Calmette-Guerin infection and a sibling with clinical tuberculosis. J. Clin. Invest. 100, 2658 2664. 66. Jouanguy, E., Altare, F., Lamhamedi-Cherradi, S., Casanova, J. L. (1997) Infections in IFNGR-1-decient children. J. Interferon Cytokine Res. 17, 583587. 67. Newport, M. J., Huxley, C. M., Huston, S., Hawrylowicz, C. M., Oostra, B. A., Williamson, R., Levin, M. (1996) A mutation in the interferongamma-receptor gene and susceptibility to mycobacterial infection. N. Engl. J. Med. 335, 19411949. 68. Pierre-Audigier, C., Jouanguy, E., Lamhamedi, S., Altare, F., Rauzier, J., Vincent, V., Canioni, D., Emile, J. F., Fischer, A., Blanche, S., Gaillard, J. L., Casanova, J. L. (1997) Fatal disseminated Mycobacterium smegmatis infection in a child with inherited interferon gamma receptor deciency. Clin. Infect. Dis. 24, 982984. 69. Roesler, J., Konk, B., Wendisch, J., Heyden, S., Paul, D., Friedrich, W., Casanova, J. L., Leupold, W., Gahr, M., Rosen-Wolff, A. (1999) Listeria monocytogenes and recurrent mycobacterial infections in a child with complete interferon-gamma-receptor (IFNgammaR1) deciency: mutational analysis and evaluation of therapeutic options. Exp. Hematol. 27, 1368 1374. 70. Jouanguy, E., Altare, F., Lamhamedi, S., Revy, P., Emile, J. F., Newport, M., Levin, M., Blanche, S., Seboun, E., Fischer, A., Casanova, J. L. (1996) Interferon-gamma-receptor deciency in an infant with fatal bacille Calmette-Guerin infection. N. Engl. J. Med. 335, 1956 1961. 71. Dorman, S. E., Holland, S. M. (1998) Mutation in the signal-transducing chain of the interferon-gamma receptor and susceptibility to mycobacterial infection. J. Clin. Invest. 101, 2364 2369. 72. Dorman, S. E., Uzel, G., Roesler, J., Bradley, J. S., Bastian, J., Billman, G., King, S., Filie, A., Schermerhorn, J., Holland, S. M. (1999) Viral infections in interferon-gamma receptor deciency. J. Pediatr. 135, 640 643. 73. Davies, E. G., Isaacs, D., Levinsky, R. J. (1982) Defective immune interferon production and natural killer activity associated with poor neutrophil mobility and delayed umbilical cord separation. Clin. Exp. Immunol. 50, 454 460. 74. Lio, D., Scola, L., Crivello, A., Bonafe, M., Franceschi, C., Olivieri, F., Colonna-Romano, G., Candore, G., Caruso, C. (2002) Allele frequencies of 874T3 A single nucleotide polymorphism at the rst intron of interferon-gamma gene in a group of Italian centenarians. Exp. Gerontol. 37, 315319. 75. Espejo, C., Penkowa, M., Saez-Torres, I., Xaus, J., Celada, A., Montalban, X., Martinez-Caceres, E. M. (2001) Treatment with anti-interferongamma monoclonal antibodies modies experimental autoimmune encephalomyelitis in interferon-gamma receptor knockout mice. Exp. Neurol. 172, 460 468. 76. Billiau, A., Heremans, H., Vandekerckhove, F., Dijkmans, R., Sobis, H., Meulepas, E., Carton, H. (1988) Enhancement of experimental allergic

180

Journal of Leukocyte Biology Volume 75, February 2004

http://www.jleukbio.org

77. 78. 79. 80.

81. 82.

83. 84.

85.

86. 87. 88. 89. 90. 91.

92.

93.

94. 95.

96. 97. 98.

encephalomyelitis in mice by antibodies against IFN-gamma. J. Immunol. 140, 1506 1510. Heremans, H., Billiau, A., Colombatti, A., Hilgers, J., de Somer, P. (1978) Interferon treatment of NZB mice: accelerated progression of autoimmune disease. Infect. Immun. 21, 925930. Jacob, C. O., McDevitt, H. O. (1988) Tumour necrosis factor-alpha in murine autoimmune lupus nephritis. Nature 331, 356 358. Lee, J. Y., Goldman, D., Piliero, L. M., Petri, M., Sullivan, K. E. (2001) Interferon-gamma polymorphisms in systemic lupus erythematosus. Genes Immun. 2, 254 257. Baechler, E. C., Batliwalla, F. M., Karypis, G., Gaffney, P. M., Ortmann, W. A., Espe, K. J., Shark, K. B., Grande, W. J., Hughes, K. M., Kapur, V., Gregersen, P. K., Behrens, T. W. (2003) Interferon-inducible gene expression signature in peripheral blood cells of patients with severe lupus. Proc. Natl. Acad. Sci. USA 100, 2610 2615. Panitch, H. S., Hirsch, R. L., Haley, A. S., Johnson, K. P. (1987) Exacerbations of multiple sclerosis in patients treated with gamma interferon. Lancet 1, 893 895. Sarvetnick, N., Liggitt, D., Pitts, S. L., Hansen, S. E., Stewart, T. A. (1988) Insulin-dependent diabetes mellitus induced in transgenic mice by ectopic expression of class II MHC and interferon-gamma. Cell 52, 773782. Awata, T., Matsumoto, C., Urakami, T., Hagura, R., Amemiya, S., Kanazawa, Y. (1994) Association of polymorphism in the interferon gamma gene with IDDM. Diabetologia 37, 1159 1162. Wang, B., Andre, I., Gonzalez, A., Katz, J. D., Aguet, M., Benoist, C., Mathis, D. (1997) Interferon-gamma impacts at multiple points during the progression of autoimmune diabetes. Proc. Natl. Acad. Sci. USA 94, 13844 13849. Kaplan, D. H., Shankaran, V., Dighe, A. S., Stockert, E., Aguet, M., Old, L. J., Schreiber, R. D. (1998) Demonstration of an interferon gammadependent tumor surveillance system in immunocompetent mice. Proc. Natl. Acad. Sci. USA 95, 7556 7561. Dighe, A. S., Richards, E., Old, L. J., Schreiber, R. D. (1994) Enhanced in vivo growth and resistance to rejection of tumor cells expressing dominant negative IFN gamma receptors. Immunity 1, 447 456. Tannenbaum, C. S., Hamilton, T. A. (2000) Immune-inammatory mechanisms in IFNgamma-mediated anti-tumor activity. Semin. Cancer Biol. 10, 113123. Dunn, G. P., Bruce, A. T., Ikeda, H., Old, L. J., Schreiber, R. D. (2002) Cancer immunoediting: from immunosurveillance to tumor escape. Nat. Immunol. 3, 991998. Ikeda, H., Old, L. J., Schreiber, R. D. (2002) The roles of IFN gamma in protection against tumor development and cancer immunoediting. Cytokine Growth Factor Rev. 13, 95109. Subramaniam, P. S., Torres, B. A., Johnson, H. M. (2001) So many ligands, so few transcription factors: a new paradigm for signaling through the STAT transcription factors. Cytokine 15, 175187. Sakatsume, M., Igarashi, K., Winestock, K. D., Garotta, G., Larner, A. C., Finbloom, D. S. (1995) The Jak kinases differentially associate with the alpha and beta (accessory factor) chains of the interferon gamma receptor to form a functional receptor unit capable of activating STAT transcription factors. J. Biol. Chem. 270, 17528 17534. Marsters, S. A., Pennica, D., Bach, E., Schreiber, R. D., Ashkenazi, A. (1995) Interferon gamma signals via a high-afnity multisubunit receptor complex that contains two types of polypeptide chain. Proc. Natl. Acad. Sci. USA 92, 54015405. Krause, C. D., Mei, E., Xie, J., Jia, Y., Bopp, M. A., Hochstrasser, R. M., Pestka, S. (2002) Seeing the light: preassembly and ligand-induced changes of the interferon gamma receptor complex in cells. Mol. Cell. Proteomics 1, 805 815. Ealick, S. E., Cook, W. J., Vijay-Kumar, S., Carson, M., Nagabhushan, T. L., Trotta, P. P., Bugg, C. E. (1991) Three-dimensional structure of recombinant human interferon-gamma. Science 252, 698 702. Walter, M. R., Windsor, W. T., Nagabhushan, T. L., Lundell, D. J., Lunn, C. A., Zauodny, P. J., Narula, S. K. (1995) Crystal structure of a complex between interferon-gamma and its soluble high-afnity receptor. Nature 376, 230 235. Fountoulakis, M., Zulauf, M., Lustig, A., Garotta, G. (1992) Stoichiometry of interaction between interferon gamma and its receptor. Eur. J. Biochem. 208, 781787. Greenlund, A. C., Schreiber, R. D., Goeddel, D. V., Pennica, D. (1993) Interferon-gamma induces receptor dimerization in solution and on cells. J. Biol. Chem. 268, 1810318110. Briscoe, J., Rogers, N. C., Witthuhn, B. A., Watling, D., Harpur, A. G., Wilks, A. F., Stark, G. R., Ihle, J. N., Kerr, I. M. (1996) Kinase-negative mutants of JAK1 can sustain interferon-gamma-inducible gene expression but not an antiviral state. EMBO J. 15, 799 809.

99. Igarashi, K., Garotta, G., Ozmen, L., Ziemiecki, A., Wilks, A. F., Harpur, A. G., Larner, A. C., Finbloom, D. S. (1994) Interferon-gamma induces tyrosine phosphorylation of interferon-gamma receptor and regulated association of protein tyrosine kinases, Jak1 and Jak2, with its receptor. J. Biol. Chem. 269, 1433314336. 100. Greenlund, A. C., Morales, M. O., Viviano, B. L., Yan, H., Krolewski, J., Schreiber, R. D. (1995) Stat recruitment by tyrosine-phosphorylated cytokine receptors: an ordered reversible afnity-driven process. Immunity 2, 677 687. 101. Heim, M. H., Kerr, I. M., Stark, G. R., Darnell Jr., J. E. (1995) Contribution of STAT SH2 groups to specic interferon signaling by the Jak-STAT pathway. Science 267, 13471349. 102. Darnell Jr., J. E., Kerr, I. M., Stark, G. R. (1994) Jak-STAT pathways and transcriptional activation in response to IFNs and other extracellular signaling proteins. Science 264, 14151421. 103. Schindler, C., Darnell Jr., J. E. (1995) Transcriptional responses to polypeptide ligands: the JAK-STAT pathway. Annu. Rev. Biochem. 64, 621 651. 104. Ramana, C. V., Chatterjee-Kishore, M., Nguyen, H., Stark, G. R. (2000) Complex roles of Stat1 in regulating gene expression. Oncogene 19, 2619 2627. 105. Decker, T., Kovarik, P., Meinke, A. (1997) GAS elements: a few nucleotides with a major impact on cytokine-induced gene expression. J. Interferon Cytokine Res. 17, 121134. 106. Meraz, M. A., White, J. M., Sheehan, K. C., Bach, E. A., Rodig, S. J., Dighe, A. S., Kaplan, D. H., Riley, J. K., Greenlund, A. C., Campbell, D., Carver-Moore, K., DuBois, R. N., Clark, R., Aguet, M., Schreiber, R. D. (1996) Targeted disruption of the Stat1 gene in mice reveals unexpected physiologic specicity in the JAK-STAT signaling pathway. Cell 84, 431 442. 107. Kerr, I. M., Stark, G. R. (1991) The control of interferon-inducible gene expression. FEBS Lett. 285, 194 198. 108. Paludan, S. R. (1998) Interleukin-4 and interferon-gamma: the quintessence of a mutual antagonistic relationship. Scand. J. Immunol. 48, 459 468. 109. Chen, X., Vinkemeier, U., Zhao, Y., Jeruzalmi, D., Darnell Jr., J. E., Kuriyan, J. (1998) Crystal structure of a tyrosine phosphorylated STAT-1 dimer bound to DNA. Cell 93, 827 839. 110. Kovarik, P., Stoiber, D., Eyers, P. A., Menghini, R., Neininger, A., Gaestel, M., Cohen, P., Decker, T. (1999) Stress-induced phosphorylation of STAT1 at Ser727 requires p38 mitogen-activated protein kinase whereas IFN-gamma uses a different signaling pathway. Proc. Natl. Acad. Sci. USA 96, 13956 13961. 111. Goh, K. C., Haque, S. J., Williams, B. R. (1999) p38 MAP kinase is required for STAT1 serine phosphorylation and transcriptional activation induced by interferons. EMBO J. 18, 56015608. 112. Wen, Z., Zhong, Z., Darnell Jr., J. E. (1995) Maximal activation of transcription by Stat1 and Stat3 requires both tyrosine and serine phosphorylation. Cell 82, 241250. 113. Zhang, X., Blenis, J., Li, H. C., Schindler, C., Chen-Kiang, S. (1995) Requirement of serine phosphorylation for formation of STAT-promoter complexes. Science 267, 1990 1994. 114. Shuai, K., Stark, G. R., Kerr, I. M., Darnell Jr., J. E. (1993) A single phosphotyrosine residue of Stat91 required for gene activation by interferon-gamma. Science 261, 1744 1746. 115. Shuai, K., Horvath, C. M., Huang, L. H., Qureshi, S. A., Cowburn, D., Darnell Jr., J. E. (1994) Interferon activation of the transcription factor Stat91 involves dimerization through SH2-phosphotyrosyl peptide interactions. Cell 76, 821 828. 116. Matsumoto, M., Tanaka, N., Harada, H., Kimura, T., Yokochi, T., Kitagawa, M., Schindler, C., Taniguchi, T. (1999) Activation of the transcription factor ISGF3 by interferon-gamma. Biol. Chem. 380, 699 703. 117. Stark, G. R., Kerr, I. M., Williams, B. R., Silverman, R. H., Schreiber, R. D. (1998) How cells respond to interferons. Annu. Rev. Biochem. 67, 227264. 118. Bluyssen, A. R., Durbin, J. E., Levy, D. E. (1996) ISGF3 gamma p48, a specicity switch for interferon activated transcription factors. Cytokine Growth Factor Rev. 7, 1117. 119. Chatterjee-Kishore, M., van den Akker, F., Stark, G. R. (2000) Association of STATs with relatives and friends. Trends Cell Biol. 10, 106 111. 120. Decker, T., Kovarik, P. (2000) Serine phosphorylation of STATs. Oncogene 19, 2628 2637. 121. Zhu, X., Wen, Z., Xu, L. Z., Darnell Jr., J. E. (1997) Stat1 serine phosphorylation occurs independently of tyrosine phosphorylation and requires an activated Jak2 kinase. Mol. Cell. Biol. 17, 6618 6623.

Schroder et al. Signals, mechanisms, and functions of IFN- 181

122. Horvath, C. M., Darnell Jr., J. E. (1996) The antiviral state induced by alpha interferon and gamma interferon requires transcriptionally active Stat1 protein. J. Virol. 70, 647 650. 123. Kovarik, P., Stoiber, D., Novy, M., Decker, T. (1998) Stat1 combines signals derived from IFN-gamma and LPS receptors during macrophage activation. EMBO J. 17, 3660 3668. 124. Gollob, J. A., Schnipper, C. P., Murphy, E. A., Ritz, J., Frank, D. A. (1999) The functional synergy between IL-12 and IL-2 involves p38 mitogen-activated protein kinase and is associated with the augmentation of STAT serine phosphorylation. J. Immunol. 162, 4472 4481. 125. Singh, K., Balligand, J. L., Fischer, T. A., Smith, T. W., Kelly, R. A. (1996) Regulation of cytokine-inducible nitric oxide synthase in cardiac myocytes and microvascular endothelial cells. Role of extracellular signal-regulated kinases 1 and 2 (ERK1/ERK2) and STAT1 alpha. J. Biol. Chem. 271, 11111117. 126. Chan, E. D., Riches, D. W. (2001) IFN-gamma LPS induction of iNOS is modulated by ERK, JNK/SAPK, and p38(mapk) in a mouse macrophage cell line. Am. J. Physiol. Cell Physiol. 280, C441C450. 127. Mahboubi, K., Pober, J. S. (2002) Activation of signal transducer and activator of transcription 1 (STAT1) is not sufcient for the induction of STAT1-dependent genes in endothelial cells. Comparison of interferongamma and oncostatin M. J. Biol. Chem. 277, 8012 8021. 128. Ouchi, T., Lee, S. W., Ouchi, M., Aaronson, S. A., Horvath, C. M. (2000) Collaboration of signal transducer and activator of transcription 1 (STAT1) and BRCA1 in differential regulation of IFN-gamma target genes. Proc. Natl. Acad. Sci. USA 97, 5208 5213. 129. Zhang, J. J., Zhao, Y., Chait, B. T., Lathem, W. W., Ritzi, M., Knippers, R., Darnell Jr., J. E., (1998) Ser727-dependent recruitment of MCM5 by Stat1alpha in IFN-gamma-induced transcriptional activation. EMBO J. 17, 6963 6971. 130. Vinkemeier, U., Cohen, S. L., Moare, I., Chait, B. T., Kuriyan, J., Darnell Jr., J. E. (1996) DNA binding of in vitro activated Stat1 alpha, Stat1 beta and truncated Stat1: interaction between NH2-terminal domains stabilizes binding of two dimers to tandem DNA sites. EMBO J. 15, 5616 5626. 131. Xu, X., Sun, Y. L., Hoey, T. (1996) Cooperative DNA binding and sequence-selective recognition conferred by the STAT amino-terminal domain. Science 273, 794 797. 132. Kumar, A., Commane, M., Flickinger, T. W., Horvath, C. M., Stark, G. R. (1997) Defective TNF-alpha-induced apoptosis in STAT1-null cells due to low constitutive levels of caspases. Science 278, 1630 1632. 133. Lee, C. K., Smith, E., Gimeno, R., Gertner, R., Levy, D. E. (2000) STAT1 affects lymphocyte survival and proliferation partially independent of its role downstream of IFN-gamma. J. Immunol. 164, 1286 1292. 134. Chatterjee-Kishore, M., Wright, K. L., Ting, J. P., Stark, G. R. (2000) How Stat1 mediates constitutive gene expression: a complex of unphosphorylated Stat1 and IRF1 supports transcription of the LMP2 gene. EMBO J. 19, 4111 4122. 135. Sekimoto, T., Imamoto, N., Nakajima, K., Hirano, T., Yoneda, Y. (1997) Extracellular signal-dependent nuclear import of Stat1 is mediated by nuclear pore-targeting complex formation with NPI-1, but not Rch1. EMBO J. 16, 70677077. 136. Lillemeier, B. F., Koster, M., Kerr, I. M. (2001) STAT1 from the cell membrane to the DNA. EMBO J. 20, 2508 2517. 137. McBride, K. M., McDonald, C., Reich, N. C. (2000) Nuclear export signal located within the DNA-binding domain of the STAT1 transcription factor. EMBO J. 19, 6196 6206. 138. Koster, M., Hauser, H. (1999) Dynamic redistribution of STAT1 protein in IFN signaling visualized by GFP fusion proteins. Eur. J. Biochem. 260, 137144. 139. Dingwall, C., Laskey, R. A. (1991) Nuclear targeting sequencesa consensus? Trends Biochem. Sci. 16, 478 481. 140. MacDonald, H. S., Kushnaryov, V. M., Sedmak, J. J., Grossberg, S. E. (1986) Transport of gamma-interferon into the cell nucleus may be mediated by nuclear membrane receptors. Biochem. Biophys. Res. Commun. 138, 254 260. 141. Bader, T., Weitzerbin, J. (1994) Nuclear accumulation of interferon gamma. Proc. Natl. Acad. Sci. USA 91, 1183111835. 142. Subramaniam, P. S., Mujtaba, M. G., Paddy, M. R., Johnson, H. M. (1999) The carboxyl terminus of interferon-gamma contains a functional polybasic nuclear localization sequence. J. Biol. Chem. 274, 403 407. 143. Subramaniam, P. S., Larkin III, J., Mujtaba, M. G., Walter, M. R., Johnson, H. M. (2000) The COOH-terminal nuclear localization sequence of interferon gamma regulates STAT1 alpha nuclear translocation at an intracellular site. J. Cell Sci. 113, 27712781. 144. Wetzel, R., Perry, L. J., Veilleux, C., Chang, G. (1990) Mutational analysis of the C-terminus of human interferon-gamma. Protein Eng. 3, 611 623.

145. Lundell, D., Lunn, C., Dalgarno, D., Fossetta, J., Greenberg, R., Reim, R., Grace, M., Narula, S. (1991) The carboxyl-terminal region of human interferon gamma is important for biological activity: mutagenic and NMR analysis. Protein Eng. 4, 335341. 146. Fidler, I. J., Fogler, W. E., Kleinerman, E. S., Saiki, I. (1985) Abrogation of species specicity for activation of tumoricidal properties in macrophages by recombinant mouse or human interferon-gamma encapsulated in liposomes. J. Immunol. 135, 4289 4296. 147. Sanceau, J., Sondermeyer, P., Beranger, F., Falcoff, R., Vaquero, C. (1987) Intracellular human gamma-interferon triggers an antiviral state in transformed murine L cells. Proc. Natl. Acad. Sci. USA 84, 2906 2910. 148. Smith, M. R., Muegge, K., Keller, J. R., Kung, H. F., Young, H. A., Durum, S. K. (1990) Direct evidence for an intracellular role for IFNgamma. Microinjection of human IFN-gamma induces Ia expression on murine macrophages. J. Immunol. 144, 17771782. 149. Subramaniam, P. S., Green, M. M., Larkin III, J., Torres, B. A., Johnson, H. M. (2001) Nuclear translocation of IFN-gamma is an intrinsic requirement for its biologic activity and can be driven by a heterologous nuclear localization sequence. J. Interferon Cytokine Res. 21, 951959. 150. Larkin III, J., Johnson, H. M., Subramaniam, P. S. (2000) Differential nuclear localization of the IFNGR-1 and IFNGR-2 subunits of the IFN-gamma receptor complex following activation by IFN-gamma. J. Interferon Cytokine Res. 20, 565576. 151. Durbin, J. E., Hackenmiller, R., Simon, M. C., Levy, D. E. (1996) Targeted disruption of the mouse Stat1 gene results in compromised innate immunity to viral disease. Cell 84, 443 450. 152. Ramana, C. V., Gil, M. P., Han, Y., Ransohoff, R. M., Schreiber, R. D., Stark, G. R. (2001) Stat1-independent regulation of gene expression in response to IFN- gamma. Proc. Natl. Acad. Sci. USA 98, 6674 6679. 153. Bromberg, J. F., Horvath, C. M., Wen, Z., Schreiber, R. D., Darnell Jr., J. E. (1996) Transcriptionally active Stat1 is required for the antiproliferative effects of both interferon alpha and interferon gamma. Proc. Natl. Acad. Sci. USA 93, 76737678. 154. Ramana, C. V., Grammatikakis, N., Chernov, M., Nguyen, H., Goh, K. C., Williams, B. R., Stark, G. R. (2000) Regulation of c-myc expression by IFN-gamma through Stat1-dependent and -independent pathways. EMBO J. 19, 263272. 155. Gil, M. P., Bohn, E., OGuin, A. K., Ramana, C. V., Levine, B., Stark, G. R., Virgin, H. W., Schreiber, R. D. (2001) Biologic consequences of Stat1-independent IFN signaling. Proc. Natl. Acad. Sci. USA 98, 6680 6685. 156. Dupuis, S., Dargemont, C., Fieschi, C., Thomassin, N., Rosenzweig, S., Harris, J., Holland, S. M., Schreiber, R. D., Casanova, J. L. (2001) Impairment of mycobacterial but not viral immunity by a germline human STAT1 mutation. Science 293, 300 303. 157. Taniguchi, T., Ogasawara, K., Takaoka, A., Tanaka, N. (2001) IRF family of transcription factors as regulators of host defense. Annu. Rev. Immunol. 19, 623 655. 158. Sims, S. H., Cha, Y., Romine, M. F., Gao, P. Q., Gottlieb, K., Deisseroth, A. B. (1993) A novel interferon-inducible domain: structural and functional analysis of the human interferon regulatory factor 1 gene promoter. Mol. Cell. Biol. 13, 690 702. 159. Harada, H., Takahashi, E., Itoh, S., Harada, K., Hori, T. A., Taniguchi, T. (1994) Structure and regulation of the human interferon regulatory factor 1 (IRF-1) and IRF-2 genes: implications for a gene network in the interferon system. Mol. Cell. Biol. 14, 1500 1509. 160. Pine, R. (1997) Convergence of TNFalpha and IFNgamma signalling pathways through synergistic induction of IRF-1/ISGF-2 is mediated by a composite GAS/kappaB promoter element. Nucleic Acids Res. 25, 4346 4354. 161. Mamane, Y., Heylbroeck, C., Genin, P., Algarte, M., Servant, M. J., LePage, C., DeLuca, C., Kwon, H., Lin, R., Hiscott, J. (1999) Interferon regulatory factors: the next generation. Gene 237, 114. 162. Salkowski, C. A., Kopydlowski, K., Blanco, J., Cody, M. J., McNally, R., Vogel, S. N. (1999) IL-12 is dysregulated in macrophages from IRF-1 and IRF-2 knockout mice. J. Immunol. 163, 1529 1536. 163. Hobart, M., Ramassar, V., Goes, N., Urmson, J., Halloran, P. F. (1996) The induction of class I and II major histocompatibility complex by allogeneic stimulation is dependent on the transcription factor interferon regulatory factor 1 (IRF-1): observations in IRF-1 knockout mice. Transplantation 62, 18951901. 164. Tanaka, N., Kawakami, T., Taniguchi, T. (1993) Recognition DNA sequences of interferon regulatory factor 1 (IRF-1) and IRF-2, regulators of cell growth and the interferon system. Mol. Cell. Biol. 13, 4531 4538. 165. Tanaka, N., Ishihara, M., Kitagawa, M., Harada, H., Kimura, T., Matsuyama, T., Lamphier, M. S., Aizawa, S., Mak, T. W., Taniguchi, T.

182

Journal of Leukocyte Biology Volume 75, February 2004

http://www.jleukbio.org

166.

167.

168.

169. 170.

171.

172.

173.

174.

175.

176.

177.

178.

179.

180.

181.

182.

183.

(1994) Cellular commitment to oncogene-induced transformation or apoptosis is dependent on the transcription factor IRF-1. Cell 77, 829 839. Kimura, T., Nakayama, K., Penninger, J., Kitagawa, M., Harada, H., Matsuyama, T., Tanaka, N., Kamijo, R., Vilcek, J., Mak, T. W., et al. (1994) Involvement of the IRF-1 transcription factor in antiviral responses to interferons. Science 264, 19211924. Ohmori, Y., Hamilton, T. A. (1994) IFN-gamma selectively inhibits lipopolysaccharide-inducible JE/monocyte chemoattractant protein-1 and KC/GRO/melanoma growth-stimulating activity gene expression in mouse peritoneal macrophages. J. Immunol. 153, 2204 2212. Cook, J. R., Emanuel, S. L., Donnelly, R. J., Soh, J., Mariano, T. M., Schwartz, B., Rhee, S., Pestka, S. (1994) Sublocalization of the human interferon-gamma receptor accessory factor gene and characterization of accessory factor activity by yeast articial chromosomal fragmentation. J. Biol. Chem. 269, 70137018. Darnell Jr., J. E. (1997) STATs and gene regulation. Science 277, 1630 1635. Schreiber, R. D., Farrar, M. A. (1993) The biology and biochemistry of interferon-gamma and its receptor. Gastroenterol Jpn 28 (Suppl. 4), 88 94. Anderson, P., Yip, Y. K., Vilcek, J. (1983) Human interferon-gamma is internalized and degraded by cultured broblasts. J. Biol. Chem. 258, 6497 6502. Celada, A., Schreiber, R. D. (1987) Internalization and degradation of receptor-bound interferon-gamma by murine macrophages. Demonstration of receptor recycling. J. Immunol. 139, 147153. Endo, T. A., Masuhara, M., Yokouchi, M., Suzuki, R., Sakamoto, H., Mitsui, K., Matsumoto, A., Tanimura, S., Ohtsubo, M., Misawa, H., Miyazaki, T., Leonor, N., Taniguchi, T., Fujita, T., Kanakura, Y., Komiya, S., Yoshimura, A. (1997) A new protein containing an SH2 domain that inhibits JAK kinases. Nature 387, 921924. Starr, R., Willson, T. A., Viney, E. M., Murray, L. J., Rayner, J. R., Jenkins, B. J., Gonda, T. J., Alexander, W. S., Metcalf, D., Nicola, N. A., Hilton, D. J. (1997) A family of cytokine-inducible inhibitors of signalling. Nature 387, 917921. Suzuki, R., Sakamoto, H., Yasukawa, H., Masuhara, M., Wakioka, T., Sasaki, A., Yuge, K., Komiya, S., Inoue, A., Yoshimura, A. (1998) CIS3 and JAB have different regulatory roles in interleukin-6 mediated differentiation and STAT3 activation in M1 leukemia cells. Oncogene 17, 22712278. Sakamoto, H., Yasukawa, H., Masuhara, M., Tanimura, S., Sasaki, A., Yuge, K., Ohtsubo, M., Ohtsuka, A., Fujita, T., Ohta, T., Furukawa, Y., Iwase, S., Yamada, H., Yoshimura, A. (1998) A Janus kinase inhibitor, JAB, is an interferon-gamma-inducible gene and confers resistance to interferons. Blood 92, 1668 1676. Yasukawa, H., Misawa, H., Sakamoto, H., Masuhara, M., Sasaki, A., Wakioka, T., Ohtsuka, S., Imaizumi, T., Matsuda, T., Ihle, J. N., Yoshimura, A. (1999) The JAK-binding protein JAB inhibits Janus tyrosine kinase activity through binding in the activation loop. EMBO J. 18, 1309 1320. Zhang, J. G., Farley, A., Nicholson, S. E., Willson, T. A., Zugaro, L. M., Simpson, R. J., Moritz, R. L., Cary, D., Richardson, R., Hausmann, G., Kile, B. J., Kent, S. B., Alexander, W. S., Metcalf, D., Hilton, D. J., Nicola, N. A., Baca, M. (1999) The conserved SOCS box motif in suppressors of cytokine signaling binds to elongins B and C and may couple bound proteins to proteasomal degradation. Proc. Natl. Acad. Sci. USA 96, 20712076. Fujimoto, M., Naka, T., Nakagawa, R., Kawazoe, Y., Morita, Y., Tateishi, A., Okumura, K., Narazaki, M., Kishimoto, T. (2000) Defective thymocyte development and perturbed homeostasis of T cells in STAT-induced STAT inhibitor-1/suppressors of cytokine signaling-1 transgenic mice. J. Immunol. 165, 1799 1806. Alexander, W. S., Starr, R., Fenner, J. E., Scott, C. L., Handman, E., Sprigg, N. S., Corbin, J. E., Cornish, A. L., Darwiche, R., Owczarek, C. M., Kay. T. W., Nicola, N. A. Hertzog, P. J., Metcalf, D., Hilton, D. J. (1999) SOCS1 is a critical inhibitor of interferon gamma signaling and prevents the potentially fatal neonatal actions of this cytokine. Cell 98, 597 608. Marine, J. C., Topham, D. J., McKay, C., Wang, D., Parganas, E., Stravopodis, D., Yoshimura, A., Ihle, J. N. (1999) SOCS1 deciency causes a lymphocyte-dependent perinatal lethality. Cell 98, 609 616. Song, M. M., Shuai, K. (1998) The suppressor of cytokine signaling (SOCS) 1 and SOCS3 but not SOCS2 proteins inhibit interferon-mediated antiviral and antiproliferative activities. J. Biol. Chem. 273, 35056 35062. Haque, S. J., Wu, Q., Kammer, W., Friedrich, K., Smith, J. M., Kerr, I. M., Stark, G. R., Williams, B. R. (1997) Receptor-associated consti-

184.

185.

186.

187.

188.

189. 190.

191. 192.

193.

194.

195.

196.

197.

198.

199.

200.

201.

202.

203.

204.

tutive protein tyrosine phosphatase activity controls the kinase function of JAK1. Proc. Natl. Acad. Sci. USA 94, 8563 8568. You, M., Yu, D. H., Feng, G. S. (1999) Shp-2 tyrosine phosphatase functions as a negative regulator of the interferon-stimulated Jak/STAT pathway. Mol. Cell. Biol. 19, 2416 2424. Ruff, S. J., Chen, K., Cohen, S. (1997) Peroxovanadate induces tyrosine phosphorylation of multiple signaling proteins in mouse liver and kidney. J. Biol. Chem. 272, 12631267. David, M., Chen, H. E., Goelz, S., Larner, A. C., Neel, B. G. (1995) Differential regulation of the alpha/beta interferon-stimulated Jak/Stat pathway by the SH2 domain-containing tyrosine phosphatase SHPTP1. Mol. Cell. Biol. 15, 7050 7058. Yu, C. L., Jin, Y. J., Burakoff, S. J. (2000) Cytosolic tyrosine dephosphorylation of STAT5. Potential role of SHP-2 in STAT5 regulation. J. Biol. Chem. 275, 599 604. Shultz, L. D., Schweitzer, P. A., Rajan, T. V., Yi, T., Ihle, J. N., Matthews, R. J., Thomas, M. L., Beier, D. R. (1993) Mutations at the murine motheaten locus are within the hematopoietic cell protein-tyrosine phosphatase (Hcph) gene. Cell 73, 14451454. Mowen, K., David, M. (2000) Regulation of STAT1 nuclear export by Jak1. Mol. Cell. Biol. 20, 72737281. Fornerod, M., Ohno, M., Yoshida, M., Mattaj, I. W. (1997) CRM1 is an export receptor for leucine-rich nuclear export signals. Cell 90, 1051 1060. Stade, K., Ford, C. S., Guthrie, C., Weis, K. (1997) Exportin 1 (Crm1p) is an essential nuclear export factor. Cell 90, 10411050. Haspel, R. L., Darnell Jr., J. E. (1999) A nuclear protein tyrosine phosphatase is required for the inactivation of Stat1. Proc. Natl. Acad. Sci. USA 96, 10188 10193. Haspel, R. L., Salditt-Georgieff, M., Darnell Jr., J. E. (1996) The rapid inactivation of nuclear tyrosine phosphorylated Stat1 depends upon a protein tyrosine phosphatase. EMBO J. 15, 6262 6268. Muller, M., Briscoe, J., Laxton, C., Guschin, D., Ziemiecki, A., Silvennoinen, O., Harpur, A. G., Barbieri, G., Witthuhn, B. A., Schindler, C., et al. (1993) The protein tyrosine kinase JAK1 complements defects in interferon-alpha/beta and -gamma signal transduction. Nature 366, 129 135. Watling, D., Guschin, D., Muller, M., Silvennoinen, O., Witthuhn, B. A., Quelle, F. W., Rogers, N. C., Schindler, C., Stark, G. R., Ihle, J. N., et al. (1993) Complementation by the protein tyrosine kinase JAK2 of a mutant cell line defective in the interferon-gamma signal transduction pathway. Nature 366, 166 170. Rodig, S. J., Meraz, M. A., White, J. M., Lampe, P. A., Riley, J. K., Arthur, C. D., King, K. L., Sheehan, K. C., Yin, L., Pennica, D., Johnson Jr., E. M., Schreiber, R. D. (1998) Disruption of the Jak1 gene demonstrates obligatory and nonredundant roles of the Jaks in cytokine-induced biologic responses. Cell 93, 373383. Parganas, E., Wang, D., Stravopodis, D., Topham, D. J., Marine, J. C., Teglund, S., Vanin, E. F., Bodner, S., Colamonici, O. R., van Deursen, J. M., Grosveld, G., Ihle, J. N. (1998) Jak2 is essential for signaling through a variety of cytokine receptors. Cell 93, 385395. Neubauer, H., Cumano, A., Muller, M., Wu, H., Huffstadt, U., Pfeffer, K. (1998) Jak2 deciency denes an essential developmental checkpoint in denitive hematopoiesis. Cell 93, 397 409. Kotenko, S. V., Izotova, L. S., Pollack, B. P., Muthukumaran, G., Paukku, K., Silvennoinen, O., Ihle, J. N., Pestka, S. (1996) Other kinases can substitute for Jak2 in signal transduction by interferon-gamma. J. Biol. Chem. 271, 17174 17182. Kaplan, M. H., Sun, Y. L., Hoey, T., Grusby, M. J. (1996) Impaired IL-12 responses and enhanced development of Th2 cells in Stat4-decient mice. Nature 382, 174 177. Thierfelder, W. E., van Deursen, J. M., Yamamoto, K., Tripp, R. A., Sarawar, S. R., Carson, R. T., Sangster, M. Y., Vignali, D. A., Doherty, P. C., Grosveld, G. C., Ihle, J. N. (1996) Requirement for Stat4 in interleukin-12-mediated responses of natural killer and T cells. Nature 382, 171174. Takeda, K., Tanaka, T., Shi, W., Matsumoto, M., Minami, M., Kashiwamura, S., Nakanishi, K., Yoshida, N., Kishimoto, T., Akira, S. (1996) Essential role of Stat6 in IL-4 signalling. Nature 380, 627 630. Shimoda, K., van Deursen, J., Sangster, M. Y., Sarawar, S. R., Carson, R. T., Tripp, R. A., Chu, C., Quelle, F. W., Nosaka, T., Vignali, D. A., Doherty, P. C., Grosveld, G., Paul, W. E., Ihle, J. N. (1996) Lack of IL-4-induced Th2 response and IgE class switching in mice with disrupted Stat6 gene. Nature 380, 630 633. Liu, X., Robinson, G. W., Wagner, K. U., Garrett, L., Wynshaw-Boris, A., Hennighausen, L. (1997) Stat5a is mandatory for adult mammary gland development and lactogenesis. Genes Dev. 11, 179 186.

Schroder et al. Signals, mechanisms, and functions of IFN- 183

205. Takeda, K., Kaisho, T., Yoshida, N., Takeda, J., Kishimoto, T., Akira, S. (1998) Stat3 activation is responsible for IL-6-dependent T cell proliferation through preventing apoptosis: generation and characterization of T cell-specic Stat3-decient mice. J. Immunol. 161, 4652 4660. 206. Kawakami, T., Matsumoto, M., Sato, M., Harada, H., Taniguchi, T., Kitagawa, M. (1995) Possible involvement of the transcription factor ISGF3 gamma in virus-induced expression of the IFN-beta gene. FEBS Lett. 358, 225229. 207. Yoneyama, M., Suhara, W., Fukuhara, Y., Sato, M., Ozato, K., Fujita, T. (1996) Autocrine amplication of type I interferon gene expression mediated by interferon stimulated gene factor 3 (ISGF3). J. Biochem. (Tokyo) 120, 160 169. 208. Min, W., Pober, J. S., Johnson, D. R. (1998) Interferon induction of TAP1: the phosphatase SHP-1 regulates crossover between the IFNalpha/beta and the IFN-gamma signal-transduction pathways. Circ. Res. 83, 815 823. 209. Takaoka, A., Mitani, Y., Suemori, H., Sato, M., Yokochi, T., Noguchi, S., Tanaka, N., Taniguchi, T. (2000) Cross talk between interferon-gamma and -alpha/beta signaling components in caveolar membrane domains. Science 288, 23572360. 210. Belich, M. P., Glynne, R. J., Senger, G., Sheer, D., Trowsdale, J. (1994) Proteasome components with reciprocal expression to that of the MHCencoded LMP proteins. Curr. Biol. 4, 769 776. 211. Kelly, A., Powis, S. H., Glynne, R., Radley, E., Beck, S., Trowsdale, J. (1991) Second proteasome-related gene in the human MHC class II region. Nature 353, 667 668. 212. Groettrup, M., Kraft, R., Kostka, S., Standera, S., Stohwasser, R., Kloetzel, P. M. (1996) A third interferon-gamma-induced subunit exchange in the 20S proteasome. Eur. J. Immunol. 26, 863 869. 213. Nandi, D., Jiang, H., Monaco, J. J. (1996) Identication of MECL-1 (LMP-10) as the third IFN-gamma-inducible proteasome subunit. J. Immunol. 156, 23612364. 214. Hisamatsu, H., Shimbara, N., Saito, Y., Kristensen, P., Hendil, K. B., Fujiwara, T., Takahashi, E., Tanahashi, N., Tamura, T., Ichihara, A., Tanaka, K. (1996) Newly identied pair of proteasomal subunits regulated reciprocally by interferon gamma. J. Exp. Med. 183, 18071816. 215. Strobl, B., Arulampalam, V., Isharc, H., Newman, S. J., Schlaak, J. F., Watling, D., Costa-Pereira, A. P., Schaper, F., Behrmann, I., Sheehan, K. C., Schreiber, R. D., Horn, F., Heinrich, P. C., Kerr, I. M. (2001) A completely foreign receptor can mediate an interferon-gamma-like response. EMBO J. 20, 54315442. 216. Groettrup, M., Khan, S., Schwarz, K., Schmidtke, G. (2001) Interferongamma inducible exchanges of 20S proteasome active site subunits: why? Biochimie 83, 367372. 217. Dick, T. P., Ruppert, T., Groettrup, M., Kloetzel, P. M., Kuehn, L., Koszinowski, U. H., Stevanovic, S., Schild, H., Rammensee, H. G. (1996) Coordinated dual cleavages induced by the proteasome regulator PA28 lead to dominant MHC ligands. Cell 86, 253262. 218. Groettrup, M., Soza, A., Eggers, M., Kuehn, L., Dick, T. P., Schild, H., Rammensee, H. G., Koszinowski, U. H., Kloetzel, P. M. (1996) A role for the proteasome regulator PA28alpha in antigen presentation. Nature 381, 166 168. 219. Boes, B., Hengel, H., Ruppert, T., Multhaup, G., Koszinowski, U. H., Kloetzel, P. M. (1994) Interferon gamma stimulation modulates the proteolytic activity and cleavage site preference of 20S mouse proteasomes. J. Exp. Med. 179, 901909. 220. Trowsdale, J., Hanson, I., Mockridge, I., Beck, S., Townsend, A., Kelly, A. (1990) Sequences encoded in the class II region of the MHC related to the ABC superfamily of transporters. Nature 348, 741744. 221. Epperson, D. E., Arnold, D., Spies, T., Cresswell, P., Pober, J. S., Johnson, D. R. (1992) Cytokines increase transporter in antigen processing-1 expression more rapidly than HLA class I expression in endothelial cells. J. Immunol. 149, 32973301. 222. Decker, T., Stockinger, S., Karaghiosoff, M., Muller, M., Kovarik, P. (2002) IFNs and STATs in innate immunity to microorganisms. J. Clin. Invest. 109, 12711277. 223. Cramer, L. A., Nelson, S. L., Klemsz, M. J. (2000) Synergistic induction of the Tap-1 gene by IFN-gamma and lipopolysaccharide in macrophages is regulated by STAT1. J. Immunol. 165, 3190 3197. 224. Shirayoshi, Y., Burke, P. A., Appella, E., Ozato, K. (1988) Interferoninduced transcription of a major histocompatibility class I gene accompanies binding of inducible nuclear factors to the interferon consensus sequence. Proc. Natl. Acad. Sci. USA 85, 5884 5888. 225. Pestka, S., Langer, J. A., Zoon, K. C., Samuel, C. E. (1987) Interferons and their actions. Annu. Rev. Biochem. 56, 727777. 226. Scheurich, P., Kronke, M., Schluter, C., Ucer, U., Pzenmaier, K. (1986) Noncytocidal mechanisms of action of tumor necrosis factor-alpha on

227.

228.

229.

230.

231.

232.

233.

234.

235.

236.

237.

238. 239.

240.

241.

242.

243.

244.

245.

246.

247.

248.

human tumor cells: enhancement of HLA gene expression synergistic with interferon-gamma. Immunobiology 172, 291300. Johnson, D. R., Pober, J. S. (1990) Tumor necrosis factor and immune interferon synergistically increase transcription of HLA class I heavyand light-chain genes in vascular endothelium. Proc. Natl. Acad. Sci. USA 87, 51835187. Reis, L. F., Harada, H., Wolchok, J. D., Taniguchi, T., Vilcek, J. (1992) Critical role of a common transcription factor, IRF-1, in the regulation of IFN-beta and IFN-inducible genes. EMBO J. 11, 185193. Chang, C. H., Hammer, J., Loh, J. E., Fodor, W. L., Flavell, R. A. (1992) The activation of major histocompatibility complex class I genes by interferon regulatory factor-1 (IRF-1). Immunogenetics 35, 378 384. Wallach, D., Fellous, M., Revel, M. (1982) Preferential effect of gamma interferon on the synthesis of HLA antigens and their mRNAs in human cells. Nature 299, 833 836. Fellous, M., Nir, U., Wallach, D., Merlin, G., Rubinstein, M., Revel, M. (1982) Interferon-dependent induction of mRNA for the major histocompatibility antigens in human broblasts and lymphoblastoid cells. Proc. Natl. Acad. Sci. USA 79, 30823086. Seder, R. A., Gazzinelli, R. T. (1999) Cytokines are critical in linking the innate and adaptive immune responses to bacterial, fungal, and parasitic infection. Adv. Intern. Med. 44, 353388. Seliger, B., Schreiber, K., Delp, K., Meissner, M., Hammers, S., Reichert, T., Pawlischko, K., Tampe, R., Huber, C. (2001) Downregulation of the constitutive tapasin expression in human tumor cells of distinct origin and its transcriptional upregulation by cytokines. Tissue Antigens 57, 39 45. Mach, B., Steimle, V., Martinez-Soria, E., Reith, W. (1996) Regulation of MHC class II genes: lessons from a disease. Annu. Rev. Immunol. 14, 301331. Kern, I., Steimle, V., Siegrist, C. A., Mach, B. (1995) The two novel MHC class II transactivators RFX5 and CIITA both control expression of HLA-DM genes. Int. Immunol. 7, 12951299. Figueiredo, F., Koerner, T. J., Adams, D. O. (1989) Molecular mechanisms regulating the expression of class II histocompatibility molecules on macrophages. Effects of inductive and suppressive signals on gene transcription. J. Immunol. 143, 37813786. Chang, C. H., Flavell, R. A. (1995) Class II transactivator regulates the expression of multiple genes involved in antigen presentation. J. Exp. Med. 181, 765767. Cresswell, P. (1994) Assembly, transport, and function of MHC class II molecules. Annu. Rev. Immunol. 12, 259 293. Wolf, P. R., Ploegh, H. L. (1995) How MHC class II molecules acquire peptide cargo: biosynthesis and trafcking through the endocytic pathway. Annu. Rev. Cell Dev. Biol. 11, 267306. Lah, T. T., Hawley, M., Rock, K. L., Goldberg, A. L. (1995) Gammainterferon causes a selective induction of the lysosomal proteases, cathepsins B and L, in macrophages. FEBS Lett. 363, 85 89. Lafuse, W. P., Brown, D., Castle, L., Zwilling, B. S. (1995) IFN-gamma increases cathepsin H mRNA levels in mouse macrophages. J. Leukoc. Biol. 57, 663 669. Meurs, E., Chong, K., Galabru, J., Thomas, N. S., Kerr, I. M., Williams, B. R., Hovanessian, A. G. (1990) Molecular cloning and characterization of the human double-stranded RNA-activated protein kinase induced by interferon. Cell 62, 379 390. Zamanian-Daryoush, M., Mogensen, T. H., DiDonato, J. A., Williams, B. R. (2000) NF-kappaB activation by double-stranded-RNA-activated protein kinase (PKR) is mediated through NF-kappaB-inducing kinase and IkappaB kinase. Mol. Cell. Biol. 20, 1278 1290. Osman, F., Jarrous, N., Ben-Asouli, Y., Kaempfer, R. (1999) A cis-acting element in the 3-untranslated region of human TNF-alpha mRNA renders splicing dependent on the activation of protein kinase PKR. Genes Dev. 13, 3280 3293. Donze, O., Dostie, J., Sonenberg, N. (1999) Regulatable expression of the interferon-induced double-stranded RNA dependent protein kinase PKR induces apoptosis and fas receptor expression. Virology 256, 322329. Balachandran, S., Kim, C. N., Yeh, W. C., Mak, T. W., Bhalla, K., Barber, G. N. (1998) Activation of the dsRNA-dependent protein kinase, PKR, induces apoptosis through FADD-mediated death signaling. EMBO J. 17, 6888 6902. Wong, A. H., Tam, N. W., Yang, Y. L., Cuddihy, A. R., Li, S., Kirchhoff, S., Hauser, H., Decker, T., Koromilas, A. E. (1997) Physical association between STAT1 and the interferon-inducible protein kinase PKR and implications for interferon and double-stranded RNA signaling pathways. EMBO J. 16, 12911304. Patterson, J. B., Thomis, D. C., Hans, S. L., Samuel, C. E. (1995) Mechanism of interferon action: double-stranded RNA-specic adeno-

184

Journal of Leukocyte Biology Volume 75, February 2004

http://www.jleukbio.org

249.

250.

251.

252.

253.

254.

255.

256.

257.

258. 259. 260. 261. 262.

263. 264.

265. 266. 267.

268. 269. 270.

sine deaminase from human cells is inducible by alpha and gamma interferons. Virology 210, 508 511. Anderson, S. L., Carton, J. M., Lou, J., Xing, L., Rubin, B. Y. (1999) Interferon-induced guanylate binding protein-1 (GBP-1) mediates an antiviral effect against vesicular stomatitis virus and encephalomyocarditis virus. Virology 256, 8 14. Vestal, D. J., Gorbacheva, V. Y., Sen, G. C. (2000) Different subcellular localizations for the related interferon-induced GTPases, MuGBP-1 and MuGBP-2: implications for different functions? J. Interferon Cytokine Res. 20, 9911000. Raveh, T., Hovanessian, A. G., Meurs, E. F., Sonenberg, N., Kimchi, A. (1996) Double-stranded RNA-dependent protein kinase mediates c-Myc suppression induced by type I interferons. J. Biol. Chem. 271, 25479 25484. Xaus, J., Cardo, M., Valledor, A. F., Soler, C., Lloberas, J., Celada, A. (1999) Interferon gamma induces the expression of p21waf-1 and arrests macrophage cell cycle, preventing induction of apoptosis. Immunity 11, 103113. Harvat, B. L., Seth, P., Jetten, A. M. (1997) The role of p27Kip1 in gamma interferon-mediated growth arrest of mammary epithelial cells and related defects in mammary carcinoma cells. Oncogene 14, 2111 2122. Matsuoka, M., Nishimoto, I., Asano, S. (1999) Interferon-gamma impairs physiologic downregulation of cyclin-dependent kinase inhibitor, p27Kip1, during G1 phase progression in macrophages. Exp. Hematol. 27, 203209. Kominsky, S., Johnson, H. M., Bryan, G., Tanabe, T., Hobeika, A. C., Subramaniam, P. S., Torres, B. (1998) IFNgamma inhibition of cell growth in glioblastomas correlates with increased levels of the cyclin dependent kinase inhibitor p21WAF1/CIP1. Oncogene 17, 29732979. Mandal, M., Bandyopadhyay, D., Goepfert, T. M., Kumar, R. (1998) Interferon-induces expression of cyclin-dependent kinase-inhibitors p21WAF1 and p27Kip1 that prevent activation of cyclin-dependent kinase by CDK-activating kinase (CAK). Oncogene 16, 217225. Harvat, B. L., Jetten, A. M. (1996) Gamma-interferon induces an irreversible growth arrest in mid-G1 in mammary epithelial cells which correlates with a block in hyperphosphorylation of retinoblastoma. Cell Growth Differ. 7, 289 300. Choubey, D., Gutterman, J. U. (1997) Inhibition of E2F-4/DP-1-stimulated transcription by p202. Oncogene 15, 291301. Ayer, D. E., Kretzner, L., Eisenman, R. N. (1993) Mad: a heterodimeric partner for Max that antagonizes Myc transcriptional activity. Cell 72, 211222. Ayer, D. E., Eisenman, R. N. (1993) A switch from Myc:Max to Mad:Max heterocomplexes accompanies monocyte/macrophage differentiation. Genes Dev. 7, 2110 2119. Cultraro, C. M., Bino, T., Segal, S. (1997) Function of the c-Myc antagonist Mad1 during a molecular switch from proliferation to differentiation. Mol. Cell. Biol. 17, 23532359. Vastrik, I., Kaipainen, A., Penttila, T. L., Lymboussakis, A., Alitalo, R., Parvinen, M., Alitalo, K. (1995) Expression of the mad gene during cell differentiation in vivo and its inhibition of cell growth in vitro. J. Cell Biol. 128, 11971208. Obaya, A. J., Mateyak, M. K., Sedivy, J. M. (1999) Mysterious liaisons: the relationship between c-Myc and the cell cycle. Oncogene 18, 2934 2941. Vairo, G., Vadiveloo, P. K., Royston, A. K., Rockman, S. P., Rock, C. O., Jackowski, S., Hamilton, J. A. (1995) Deregulated c-myc expression overrides IFN gamma-induced macrophage growth arrest. Oncogene 10, 1969 1976. Grandori, C., Eisenman, R. N. (1997) Myc target genes. Trends Biochem. Sci. 22, 177181. Taniguchi, T., Lamphier, M. S., Tanaka, N. (1997) IRF-1: the transcription factor linking the interferon response and oncogenesis. Biochim. Biophys. Acta 1333, M9 17. Tamura, T., Ishihara, M., Lamphier, M. S., Tanaka, N., Oishi, I., Aizawa, S., Matsuyama, T., Mak, T. W., Taki, S., Taniguchi, T. (1997) DNA damage-induced apoptosis and ice gene induction in mitogenically activated T lymphocytes require IRF-1. Leukemia 11, (Suppl. 3), 439 440. Villa, P., Kaufmann, S. H., Earnshaw, W. C. (1997) Caspases and caspase inhibitors. Trends Biochem. Sci. 22, 388 393. Chin, Y. E., Kitagawa, M., Kuida, K., Flavell, R. A., Fu, X. Y. (1997) Activation of the STAT signaling pathway can cause expression of caspase 1 and apoptosis. Mol. Cell. Biol. 17, 5328 5337. Takizawa, T., Ohashi, K., Nakanishi, Y. (1996) Possible involvement of double-stranded RNA-activated protein kinase in cell death by inuenza virus infection. J. Virol. 70, 8128 8132.

271. Yeung, M. C., Liu, J., Lau, A. S. (1996) An essential role for the interferon-inducible, double-stranded RNA-activated protein kinase PKR in the tumor necrosis factor-induced apoptosis in U937 cells. Proc. Natl. Acad. Sci. USA 93, 1245112455. 272. Lee, S. B., Esteban, M. (1994) The interferon-induced double-stranded RNA-activated protein kinase induces apoptosis. Virology 199, 491 496. 273. Inbal, B., Cohen, O., Polak-Charcon, S., Kopolovic, J., Vadai, E., Eisenbach, L., Kimchi, A. (1997) DAP kinase links the control of apoptosis to metastasis. Nature 390, 180 184. 274. Deiss, L. P., Feinstein, E., Berissi, H., Cohen, O., Kimchi, A. (1995) Identication of a novel serine/threonine kinase and a novel 15-kD protein as potential mediators of the gamma interferon-induced cell death. Genes Dev. 9, 1530. 275. Kissil, J. L., Deiss, L. P., Bayewitch, M., Raveh, T., Khaspekov, G., Kimchi, A. (1995) Isolation of DAP3, a novel mediator of interferongamma-induced cell death. J. Biol. Chem. 270, 2793227936. 276. Levy-Strumpf, N., Deiss, L. P., Berissi, H., Kimchi, A. (1997) DAP-5, a novel homolog of eukaryotic translation initiation factor 4G isolated as a putative modulator of gamma interferon-induced programmed cell death. Mol. Cell. Biol. 17, 16151625. 277. Deiss, L. P., Galinka, H., Berissi, H., Cohen, O., Kimchi, A. (1996) Cathepsin D protease mediates programmed cell death induced by interferon-gamma, Fas/APO-1 and TNF-alpha. EMBO J. 15, 3861 3870. 278. Xu, X., Fu, X. Y., Plate, J., Chong, A. S. (1998) IFN-gamma induces cell growth inhibition by Fas-mediated apoptosis: requirement of STAT1 protein for up-regulation of Fas and FasL expression. Cancer Res. 58, 28322837. 279. Zheng, H., Luo, R. C., Zhang, L. S., Mai, G. F. (2002) Interferon-gamma up-regulates Fas expression and increases Fas-mediated apoptosis in tumor cell lines. Di Yi Jun Yi Da Xue Xue Bao 22, 1090 1092. 280. Tsujimoto, M., Yip, Y. K., Vilcek, J. (1986) Interferon-gamma enhances expression of cellular receptors for tumor necrosis factor. J. Immunol. 136, 24412444. 281. MacMicking, J., Xie, Q. W., Nathan, C. (1997) Nitric oxide and macrophage function. Annu. Rev. Immunol. 15, 323350. 282. Eizirik, D. L., Flodstrom, M., Karlsen, A. E., Welsh, N. (1996) The harmony of the spheres: inducible nitric oxide synthase and related genes in pancreatic beta cells. Diabetologia 39, 875 890. 283. Kawahara, K., Gotoh, T., Oyadomari, S., Kajizono, M., Kuniyasu, A., Ohsawa, K., Imai, Y., Kohsaka, S., Nakayama, H., Mori, M. (2001) Co-induction of argininosuccinate synthetase, cationic amino acid transporter-2, and nitric oxide synthase in activated murine microglial cells. Brain Res. Mol. Brain Res. 90, 165173. 284. Nussler, A. K., Billiar, T. R., Liu, Z. Z., Morris Jr., S. M. (1994) Coinduction of nitric oxide synthase and argininosuccinate synthetase in a murine macrophage cell line. Implications for regulation of nitric oxide production. J. Biol. Chem. 269, 12571261. 285. Di Silvio, M., Geller, D. A., Gross, S. S., Nussler, A., Freeswick, P., Simmons, R. L., Billiar, T. R. (1993) Inducible nitric oxide synthase activity in hepatocytes is dependent on the coinduction of tetrahydrobiopterin synthesis. Adv. Exp. Med. Biol. 338, 305308. 286. Baek, K. J., Thiel, B. A., Lucas, S., Stuehr, D. J. (1993) Macrophage nitric oxide synthase subunits. Purication, characterization, and role of prosthetic groups and substrate in regulating their association into a dimeric enzyme. J. Biol. Chem. 268, 21120 21129. 287. Cassatella, M. A., Bazzoni, F., Flynn, R. M., Dusi, S., Trinchieri, G., Rossi, F. (1990) Molecular basis of interferon-gamma and lipopolysaccharide enhancement of phagocyte respiratory burst capability. Studies on the gene expression of several NADPH oxidase components. J. Biol. Chem. 265, 2024120246. 288. Newburger, P. E., Ezekowitz, R. A., Whitney, C., Wright, J., Orkin, S. H. (1988) Induction of phagocyte cytochrome b heavy chain gene expression by interferon gamma. Proc. Natl. Acad. Sci. USA 85, 52155219. 289. Gupta, J. W., Kubin, M., Hartman, L., Cassatella, M., Trinchieri, G. (1992) Induction of expression of genes encoding components of the respiratory burst oxidase during differentiation of human myeloid cell lines induced by tumor necrosis factor and gamma-interferon. Cancer Res. 52, 2530 2537. 290. Govoni, G., Vidal, S., Cellier, M., Lepage, P., Malo, D., Gros, P. (1995) Genomic structure, promoter sequence, and induction of expression of the mouse Nramp1 gene in macrophages. Genomics 27, 9 19. 291. Erbe, D. V., Collins, J. E., Shen, L., Graziano, R. F., Fanger, M. W. (1990) The effect of cytokines on the expression and function of Fc receptors for IgG on human myeloid cells. Mol. Immunol. 27, 57 67. 292. Strunk, R. C., Cole, F. S., Perlmutter, D. H., Colten, H. R. (1985) Gamma-interferon increases expression of class III complement genes

Schroder et al. Signals, mechanisms, and functions of IFN- 185

293.

294.

295.

296.

297.

298.

299.

300.

301.

302. 303.

304.

305. 307.

308.

309.

310.

311.

312.

313.

314.

315.

C2 and factor B in human monocytes and in murine broblasts transfected with human C2 and factor B genes. J. Biol. Chem. 260, 15280 15285. Vincent, F., de la Salle, H., Bohbot, A., Bergerat, J. P., Hauptmann, G., Oberling, F. (1993) Synthesis and regulation of complement components by human monocytes/macrophages and by acute monocytic leukemia. DNA Cell Biol. 12, 415 423. Colten, H. R., Strunk, R. C., Perlmutter, D. H., Cole, F. S. (1986) Regulation of complement protein biosynthesis in mononuclear phagocytes. Ciba Found. Symp. 118, 141154. Drevets, D. A., Leenen, P. J., Campbell, P. A. (1996) Complement receptor type 3 mediates phagocytosis and killing of Listeria monocytogenes by a TNF-alpha- and IFN-gamma-stimulated macrophage precursor hybrid. Cell. Immunol. 169, 1 6. Yoshida, A., Koide, Y., Uchijima, M., Yoshida, T. O. (1994) IFN-gamma induces IL-12 mRNA expression by a murine macrophage cell line, J774. Biochem. Biophys. Res. Commun. 198, 857 861. Kubin, M., Chow, J. M., Trinchieri, G. (1994) Differential regulation of interleukin-12 (IL-12), tumor necrosis factor alpha, and IL-1 beta production in human myeloid leukemia cell lines and peripheral blood mononuclear cells. Blood 83, 18471855. Taub, D. D., Lloyd, A. R., Conlon, K., Wang, J. M., Ortaldo, J. R., Harada, A., Matsushima, K., Kelvin, D. J., Oppenheim, J. J. (1993) Recombinant human interferon-inducible protein 10 is a chemoattractant for human monocytes and T lymphocytes and promotes T cell adhesion to endothelial cells. J. Exp. Med. 177, 1809 1814. Rollins, B. J., Yoshimura, T., Leonard, E. J., Pober, J. S. (1990) Cytokine-activated human endothelial cells synthesize and secrete a monocyte chemoattractant, MCP-1/JE. Am. J. Pathol. 136, 1229 1233. Liao, F., Rabin, R. L., Yannelli, J. R., Koniaris, L. G., Vanguri, P., Farber, J. M. (1995) Human Mig chemokine: biochemical and functional characterization. J. Exp. Med. 182, 13011314. Taub, D. D., Conlon, K., Lloyd, A. R., Oppenheim, J. J., Kelvin, D. J. (1993) Preferential migration of activated CD4 and CD8 T cells in response to MIP-1 alpha and MIP-1 beta. Science 260, 355358. Appay, V., Rowland-Jones, S. L. (2001) RANTES: a versatile and controversial chemokine. Trends Immunol. 22, 83 87. Hou, J., Baichwal, V., Cao, Z. (1994) Regulatory elements and transcription factors controlling basal and cytokine-induced expression of the gene encoding intercellular adhesion molecule 1. Proc. Natl. Acad. Sci. USA 91, 1164111645. Jesse, T. L., LaChance, R., Iademarco, M. F., Dean, D. C. (1998) Interferon regulatory factor-2 is a transcriptional activator in muscle where it regulates expression of vascular cell adhesion molecule-1. J. Cell Biol. 140, 12651276. Billiau, A. (1996) Interferon-gamma: biology and role in pathogenesis. Adv. Immunol. 62, 61130. York, I. A., Rock, K. L. (1996) Antigen processing and presentation by the class I major histocompatibility complex. Annu. Rev. Immunol. 14, 369 396. Grandea III, A. G., Androlewicz, M. J., Athwal, R. S., Geraghty, D. E., Spies, T. (1995) Dependence of peptide binding by MHC class I molecules on their interaction with TAP. Science 270, 105108. Anderson, S. L., Shen, T., Lou, J., Xing, L., Blachere, N. E., Srivastava, P. K., Rubin, B. Y. (1994) The endoplasmic reticular heat shock protein gp96 is transcriptionally upregulated in interferon-treated cells. J. Exp. Med. 180, 15651569. Suto, R., Srivastava, P. K. (1995) A mechanism for the specic immunogenicity of heat shock protein-chaperoned peptides. Science 269, 15851588. Chang, C. H., Guerder, S., Hong, S. C., van Ewijk, W., Flavell, R. A. (1996) Mice lacking the MHC class II transactivator (CIITA) show tissue-specic impairment of MHC class II expression. Immunity 4, 167178. Griscelli, C., Lisowska-Grospierre, B., Mach, B. (1989) Combined immunodeciency with defective expression in MHC class II genes. Immunodec. Rev. 1, 135153. Chang, C. H., Fontes, J. D., Peterlin, M., Flavell, R. A. (1994) Class II transactivator (CIITA) is sufcient for the inducible expression of major histocompatibility complex class II genes. J. Exp. Med. 180, 1367 1374. Waldburger, J. M., Masternak, K., Muhlethaler-Mottet, A., Villard, J., Peretti, M., Landmann, S., Reith, W. (2000) Lessons from the bare lymphocyte syndrome: molecular mechanisms regulating MHC class II expression. Immunol. Rev. 178, 148 165. Boss, J. M. (1997) Regulation of transcription of MHC class II genes. Curr. Opin. Immunol. 9, 107113.

316. Heuer, C., Koch, F., Stanzl, U., Topar, G., Wysocka, M., Trinchieri, G., Enk, A., Steinman, R. M., Romani, N., Schuler, G. (1996) Interleukin-12 is produced by dendritic cells and mediates T helper 1 development as well as interferon-gamma production by T helper 1 cells. Eur. J. Immunol. 26, 659 668. 317. Coutelier, J. P., Van Broeck, J., Wolf, S. F. (1995) Interleukin-12 gene expression after viral infection in the mouse. J. Virol. 69, 19551958. 318. Bliss, S. K., Marshall, A. J., Zhang, Y., Denkers, E. Y. (1999) Human polymorphonuclear leukocytes produce IL-12, TNF-alpha, and the chemokines macrophage-inammatory protein-1 alpha and -1 beta in response to Toxoplasma gondii antigens. J. Immunol. 162, 7369 7375. 319. Cassatella, M. A., Meda, L., Gasperini, S., DAndrea, A., Ma, X., Trinchieri, G. (1995) Interleukin-12 production by human polymorphonuclear leukocytes. Eur. J. Immunol. 25, 15. 320. Macatonia, S. E., Hosken, N. A., Litton, M., Vieira, P., Hsieh, C. S., Culpepper, J. A., Wysocka, M., Trinchieri, G., Murphy, K. M., OGarra, A. (1995) Dendritic cells produce IL-12 and direct the development of Th1 cells from naive CD4 T cells. J. Immunol. 154, 50715079. 321. Lederer, J. A., Perez, V. L., DesRoches, L., Kim, S. M., Abbas, A. K., Lichtman, A. H. (1996) Cytokine transcriptional events during helper T cell subset differentiation. J. Exp. Med. 184, 397 406. 322. Trinchieri, G. (1995) Interleukin-12: a proinammatory cytokine with immunoregulatory functions that bridge innate resistance and antigenspecic adaptive immunity. Annu. Rev. Immunol. 13, 251276. 323. Morinobu, A., Kumagai, S. (1998) Cytokine measurement at a single-cell level to analyze human Th1 and Th2 cells. Rinsho Byori 46, 908 914. 324. Lehn, M., Weiser, W. Y., Engelhorn, S., Gillis, S., Remold, H. G. (1989) IL-4 inhibits H2O2 production and antileishmanial capacity of human cultured monocytes mediated by IFN-gamma. J. Immunol. 143, 3020 3024. 325. Liew, F. Y., Li, Y., Severn, A., Millott, S., Schmidt, J., Salter, M., Moncada, S. (1991) A possible novel pathway of regulation by murine T helper type-2 (Th2) cells of a Th1 cell activity via the modulation of the induction of nitric oxide synthase on macrophages. Eur. J. Immunol. 21, 2489 2494. 326. Rousset, F., Malejt, R. W., Slierendregt, B., Aubry, J. P., Bonnefoy, J. Y., Defrance, T., Banchereau, J., de Vries, J. E. (1988) Regulation of Fc receptor for IgE (CD23) and class II MHC antigen expression on Burkitts lymphoma cell lines by human IL-4 and IFN-gamma. J. Immunol. 140, 26252632. 327. Taniguchi, T., Takaoka, A. (2002) The interferon-alpha/beta system in antiviral responses: a multimodal machinery of gene regulation by the IRF family of transcription factors. Curr. Opin. Immunol. 14, 111116. 328. Lu, B., Ebensperger, C., Dembic, Z., Wang, Y., Kvatyuk, M., Lu, T., Coffman, R. L., Pestka, S., Rothman, P. B. (1998) Targeted disruption of the interferon-gamma receptor 2 gene results in severe immune defects in mice. Proc. Natl. Acad. Sci. USA 95, 8233 8238. 329. Cantin, E., Tanamachi, B., Openshaw, H. (1999) Role for gamma interferon in control of herpes simplex virus type 1 reactivation. J. Virol. 73, 3418 3423. 330. Beretta, L., Gabbay, M., Berger, R., Hanash, S. M., Sonenberg, N. (1996) Expression of the protein kinase PKR in modulated by IRF-1 and is reduced in 5q- associated leukemias. Oncogene 12, 15931596. 331. Hovanessian, A. G., Galabru, J. (1987) The double-stranded RNAdependent protein kinase is also activated by heparin. Eur. J. Biochem. 167, 467 473. 332. Goodbourn, S., Didcock, L., Randall, R. E. (2000) Interferons: cell signalling, immune modulation, antiviral response and virus countermeasures. J. Gen. Virol. 81, 23412364. 333. Lee, S. B., Melkova, Z., Yan, W., Williams, B. R., Hovanessian, A. G., Esteban, M. (1993) The interferon-induced double-stranded RNA-activated human p68 protein kinase potently inhibits protein synthesis in cultured cells. Virology 192, 380 385. 334. Chu, W. M., Ostertag, D., Li, Z. W., Chang, L., Chen, Y., Hu, Y., Williams, B., Perrault, J., Karin, M. (1999) JNK2 and IKKbeta are required for activating the innate response to viral infection. Immunity 11, 721731. 335. Zhou, A., Paranjape, J. M., Der, S. D., Williams, B. R., Silverman, R. H. (1999) Interferon action in triply decient mice reveals the existence of alternative antiviral pathways. Virology 258, 435 440. 336. Presti, R. M., Popkin, D. L., Connick, M., Paetzold, S., Virgin IV, H. W. (2001) Novel cell type-specic antiviral mechanism of interferon gamma action in macrophages. J. Exp. Med. 193, 483 496. 337. Xaus, J., Comalada, M., Valledor, A. F., Cardo, M., Herrero, C., Soler, C., Lloberas, J., Celada, A. (2001) Molecular mechanisms involved in macrophage survival, proliferation, activation or apoptosis. Immunobiology 204, 543550.

186

Journal of Leukocyte Biology Volume 75, February 2004

http://www.jleukbio.org

338. Breen, F. N., Hume, D. A., Weidemann, M. J. (1991) Interactions among granulocyte-macrophage colony-stimulating factor, macrophage colonystimulating factor, and IFN-gamma lead to enhanced proliferation of murine macrophage progenitor cells. J. Immunol. 147, 15421547. 339. Balkwill, F., Taylor-Papadimitriou, J. (1978) Interferon affects both G1 and SG2 in cells stimulated from quiescence to growth. Nature 274, 798 800. 340. Roos, G., Leanderson, T., Lundgren, E. (1984) Interferon-induced cell cycle changes in human hematopoietic cell lines and fresh leukemic cells. Cancer Res. 44, 2358 2362. 341. Lundblad, D., Lundgren, E. (1981) Block of glioma cell line in S by interferon. Int. J. Cancer 27, 749 754. 342. Tiefenbrun, N., Melamed, D., Levy, N., Resnitzky, D., Hoffman, I., Reed, S. I., Kimchi, A. (1996) Alpha interferon suppresses the cyclin D3 and cdc25A genes, leading to a reversible G0-like arrest. Mol. Cell. Biol. 16, 3934 3944. 343. Sangfelt, O., Erickson, S., Einhorn, S., Grander, D. (1997) Induction of Cip/Kip and Ink4 cyclin dependent kinase inhibitors by interferon-alpha in hematopoietic cell lines. Oncogene 14, 415 423. 344. Hobeika, A. C., Subramaniam, P. S., Johnson, H. M. (1997) IFNalpha induces the expression of the cyclin-dependent kinase inhibitor p21 in human prostate cancer cells. Oncogene 14, 11651170. 345. Chin, Y. E., Kitagawa, M., Su, W. C., You, Z. H., Iwamoto, Y., Fu, X. Y. (1996) Cell growth arrest and induction of cyclin-dependent kinase inhibitor p21 WAF1/CIP1 mediated by STAT1. Science 272, 719 722. 346. Takami, K., Takuwa, N., Okazaki, H., Kobayashi, M., Ohtoshi, T., Kawasaki, S., Dohi, M., Yamamoto, K., Nakamura, T., Tanaka, M., Nakahara, K., Takuwa, Y., Takizawa, H. (2002) Interferon-gamma inhibits hepatocyte growth factor-stimulated cell proliferation of human bronchial epithelial cells: upregulation of p27(kip1) cyclin-dependent kinase inhibitor. Am. J. Respir. Cell Mol. Biol. 26, 231238. 347. Dey, A., Kim, L., Li, W. (1999) Gamma interferon induces expression of Mad1 gene in macrophage, which inhibits colony-stimulating factor-1dependent mitogenesis. J. Cell. Biochem. 72, 232241. 348. Thornberry, N. A. (1998) Caspases: key mediators of apoptosis. Chem. Biol. 5, R97103. 349. Novelli, F., Bernabei, P., Ozmen, L., Rigamonti, L., Allione, A., Pestka, S., Garotta, G., Forni, G. (1996) Switching on of the proliferation or apoptosis of activated human T lymphocytes by IFN-gamma is correlated with the differential expression of the alpha- and beta-chains of its receptor. J. Immunol. 157, 19351943. 350. Bernabei, P., Allione, A., Rigamonti, L., Bosticardo, M., Losana, G., Borghi, I., Forni, G., Novelli, F. (2001) Regulation of interferon-gamma receptor (INF-gammaR) chains: a peculiar way to rule the life and death of human lymphocytes. Eur. Cytokine Netw. 12, 6 14. 351. Los, M., Van de Craen, M., Penning, L. C., Schenk, H., Westendorp, M., Baeuerle, P. A., Droge, W., Krammer, P. H., Fiers, W., Schulze-Osthoff, K. (1995) Requirement of an ICE/CED-3 protease for Fas/APO-1-mediated apoptosis. Nature 375, 81 83. 352. Miura, M., Zhu, H., Rotello, R., Hartwieg, E. A., Yuan, J. (1993) Induction of apoptosis in broblasts by IL-1 beta-converting enzyme, a mammalian homolog of the C. elegans cell death gene ced-3. Cell 75, 653 660. 353. Mathiak, G., Grass, G., Herzmann, T., Luebke, T., Zetina, C. C., Boehm, S. A., Bohlen, H., Neville, L. F., Hoelscher, A. H. (2000) Caspase-1inhibitor ac-YVAD-cmk reduces LPS-lethality in rats without affecting haematology or cytokine responses. Br. J. Pharmacol. 131, 383386. 354. Fantuzzi, G., Dinarello, C. A. (1999) Interleukin-18 and interleukin-1 beta: two cytokine substrates for ICE (caspase-1). J. Clin. Immunol. 19, 111. 355. Roos, D., Bolscher, B. G. J., de Boer, M. (1992) Generation of reactive oxygen species by phagocytes. In Mononuclear Phagocytes: Biology of Monocytes and Macrophages (R. van Furth, ed.), Dordrecht, The Netherlands, Kluwer Academic, 243253. 356. Nathan, C. F., Gabay, J. (1992) Antimicrobial mechanisms of macrophages. In Mononuclear Phagocytes: Biology of Monocytes and Macrophages (R. van Furth, ed.), Dordrecht, The Netherlands, Kluwer Academic, 259 267. 357. Shiloh, M. U., MacMicking, J. D., Nicholson, S., Brause, J. E., Potter, S., Marino, M., Fang, F., Dinauer, M., Nathan, C. (1999) Phenotype of mice and macrophages decient in both phagocyte oxidase and inducible nitric oxide synthase. Immunity 10, 29 38. 358. Karupiah, G., Xie, Q. W., Buller, R. M., Nathan, C., Duarte, C., MacMicking, J. D. (1993) Inhibition of viral replication by interferon-gammainduced nitric oxide synthase. Science 261, 14451448. 359. Green, S. J., Nacy, C. A., Meltzer, M. S. (1991) Cytokine-induced synthesis of nitrogen oxides in macrophages: a protective host response

360.

361.

362.

363.

364.

365.

366.

367. 368.

369.

370.

371.

372.

373.

374.

375.

376.

377.

378.

to Leishmania and other intracellular pathogens. J. Leukoc. Biol. 50, 93103. Green, S. J., Crawford, R. M., Hockmeyer, J. T., Meltzer, M. S., Nacy, C. A. (1990) Leishmania major amastigotes initiate the L-argininedependent killing mechanism in IFN-gamma-stimulated macrophages by induction of tumor necrosis factor-alpha. J. Immunol. 145, 4290 4297. Denis, M. (1991) Tumor necrosis factor and granulocyte macrophagecolony stimulating factor stimulate human macrophages to restrict growth of virulent Mycobacterium avium and to kill avirulent M. avium: killing effector mechanism depends on the generation of reactive nitrogen intermediates. J. Leukoc. Biol. 49, 380 387. Wei, X. Q., Charles, I. G., Smith, A., Ure, J., Feng, G. J., Huang, F. P., Xu, D., Muller, W., Moncada, S., Liew, F. Y. (1995) Altered immune responses in mice lacking inducible nitric oxide synthase. Nature 375, 408 411. Karupiah, G., Chen, J. H., Nathan, C. F., Mahalingam, S., MacMicking, J. D. (1998) Identication of nitric oxide synthase 2 as an innate resistance locus against ectromelia virus infection. J. Virol. 72, 7703 7706. Guidotti, L. G., McClary, H., Loudis, J. M., Chisari, F. V. (2000) Nitric oxide inhibits hepatitis B virus replication in the livers of transgenic mice. J. Exp. Med. 191, 12471252. Kwon, N. S., Nathan, C. F., Stuehr, D. J. (1989) Reduced biopterin as a cofactor in the generation of nitrogen oxides by murine macrophages. J. Biol. Chem. 264, 20496 20501. Drapier, J. C., Wietzerbin, J., Hibbs Jr., J. B. (1988) Interferon-gamma and tumor necrosis factor induce the L-arginine-dependent cytotoxic effector mechanism in murine macrophages. Eur. J. Immunol. 18, 15871592. Morris Jr., S. M., Billiar, T. R. (1994) New insights into the regulation of inducible nitric oxide synthesis. Am. J. Physiol. 266, E829 E839. Kamijo, R., Shapiro, D., Le, J., Huang, S., Aguet, M., Vilcek, J. (1993) Generation of nitric oxide and induction of major histocompatibility complex class II antigen in macrophages from mice lacking the interferon gamma receptor. Proc. Natl. Acad. Sci. USA 90, 6626 6630. Kamijo, R., Harada, H., Matsuyama, T., Bosland, M., Gerecitano, J., Shapiro, D., Le, J., Koh, S. I., Kimura, T., Green, S. J., et al. (1994) Requirement for transcription factor IRF-1 in NO synthase induction in macrophages. Science 263, 16121615. Radi, R., Beckman, J. S., Bush, K. M., Freeman, B. A. (1991) Peroxynitrite-induced membrane lipid peroxidation: the cytotoxic potential of superoxide and nitric oxide. Arch. Biochem. Biophys. 288, 481 487. Radi, R., Beckman, J. S., Bush, K. M., Freeman, B. A. (1991) Peroxynitrite oxidation of sulfhydryls. The cytotoxic potential of superoxide and nitric oxide. J. Biol. Chem. 266, 4244 4250. Cassatella, M. A., Hartman, L., Perussia, B., Trinchieri, G. (1989) Tumor necrosis factor and immune interferon synergistically induce cytochrome b-245 heavy-chain gene expression and nicotinamide-adenine dinucleotide phosphate hydrogenase oxidase in human leukemic myeloid cells. J. Clin. Invest. 83, 1570 1579. Pfefferkorn, E. R. (1984) Interferon gamma blocks the growth of Toxoplasma gondii in human broblasts by inducing the host cells to degrade tryptophan. Proc. Natl. Acad. Sci. USA 81, 908 912. Burke, F., Knowles, R. G., East, N., Balkwill, F. R. (1995) The role of indoleamine 2,3-dioxygenase in the anti-tumour activity of human interferon-gamma in vivo. Int. J. Cancer 60, 115122. Hissong, B. D., Carlin, J. M. (1997) Potentiation of interferon-induced indoleamine 2,3-dioxygenase mRNA in human mononuclear phagocytes by lipopolysaccharide and interleukin-1. J. Interferon Cytokine Res. 17, 387393. Carlin, J. M., Borden, E. C., Sondel, P. M., Byrne, G. I. (1987) Biologicresponse-modier-induced indoleamine 2,3-dioxygenase activity in human peripheral blood mononuclear cell cultures. J. Immunol. 139, 2414 2418. Hassanain, H. H., Chon, S. Y., Gupta, S. L. (1993) Differential regulation of human indoleamine 2,3-dioxygenase gene expression by interferonsgamma and -alpha. Analysis of the regulatory region of the gene and identication of an interferon-gamma-inducible DNA-binding factor. J. Biol. Chem. 268, 50775084. Fleckner, J., Martensen, P. M., Tolstrup, A. B., Kjeldgaard, N. O., Justesen, J. (1995) Differential regulation of the human, interferon inducible tryptophanyl-tRNA synthetase by various cytokines in cell lines. Cytokine 7, 70 77.

Schroder et al. Signals, mechanisms, and functions of IFN- 187

379. Kisselev, L., Frolova, L., Haenni, A. L. (1993) Interferon inducibility of mammalian tryptophanyl-tRNA synthetase: new perspectives. Trends Biochem. Sci. 18, 263267. 380. Xue, H., Wong, J. T. (1995) Interferon induction of human tryptophanyltRNA synthetase safeguards the synthesis of tryptophan-rich immunesystem proteins: a hypothesis. Gene 165, 335339. 381. Collins, J. T., Dunnick, W. A. (1993) Germline transcripts of the murine immunoglobulin gamma 2a gene: structure and induction by IFN-gamma. Int. Immunol. 5, 885 891. 382. Snapper, C. M., Paul, W. E. (1987) Interferon-gamma and B cell stimulatory factor-1 reciprocally regulate Ig isotype production. Science 236, 944 947. 383. Sweet, M. J., Hume, D. A. (1996) Endotoxin signal transduction in macrophages. J. Leukoc. Biol. 60, 8 26. 384. Akira, S. (2000) Toll-like receptors: lessons from knockout mice. Biochem. Soc. Trans. 28, 551556. 385. Aderem, A., Ulevitch, R. J. (2000) Toll-like receptors in the induction of the innate immune response. Nature 406, 782787. 386. Dobrovolskaia, M. A., Vogel, S. N. (2002) Toll receptors, CD14, and macrophage activation and deactivation by LPS. Microbes Infect. 4, 903914. 387. Kawai, T., Adachi, O., Ogawa, T., Takeda, K., Akira, S. (1999) Unresponsiveness of MyD88-decient mice to endotoxin. Immunity 11, 115 122. 388. Lorsbach, R. B., Murphy, W. J., Lowenstein, C. J., Snyder, S. H., Russell, S. W. (1993) Expression of the nitric oxide synthase gene in mouse macrophages activated for tumor cell killing. Molecular basis for the synergy between interferon-gamma and lipopolysaccharide. J. Biol. Chem. 268, 1908 1913. 389. Jurkovich, G. J., Mileski, W. J., Maier, R. V., Winn, R. K., Rice, C. L. (1991) Interferon gamma increases sensitivity to endotoxin. J. Surg. Res. 51, 197203. 390. Sweet, M. J., Stacey, K. J., Kakuda, D. K., Markovich, D., Hume, D. A. (1998) IFN-gamma primes macrophage responses to bacterial DNA. J. Interferon Cytokine Res. 18, 263271. 391. Vadiveloo, P. K., Vairo, G., Hertzog, P., Kola, I., Hamilton, J. A. (2000) Role of type I interferons during macrophage activation by lipopolysaccharide. Cytokine 12, 1639 1646. 392. Costelloe, E. A. (1997) The regulation and function of murine PAI-2. In Centre for Molecular and Cellular Biology, Brisbane, Australia, Biochemistry Department, University of Queensland, 158. 393. Williams, J. G., Jurkovich, G. J., Hahnel, G. B., Maier, R. V. (1992) Macrophage priming by interferon gamma: a selective process with potentially harmful effects. J. Leukoc. Biol. 52, 579 584. 394. Car, B. D., Eng, V. M., Schnyder, B., Ozmen, L., Huang, S., Gallay, P., Heumann, D., Aguet, M., Ryffel, B. (1994) Interferon gamma receptor decient mice are resistant to endotoxic shock. J. Exp. Med. 179, 14371444. 395. Bosisio, D., Polentarutti, N., Sironi, M., Bernasconi, S., Miyake, K., Webb, G. R., Martin, M. U., Mantovani, A., Muzio, M. (2002) Stimulation of Toll-like receptor 4 expression in human mononuclear phagocytes by interferon-gamma: a molecular basis for priming and synergism with bacterial lipopolysaccharide. Blood 99, 34273431. 396. Mita, Y., Dobashi, K., Shimizu, Y., Nakazawa, T., Mori, M. (2001) Toll-like receptor 2 and 4 surface expressions on human monocytes are modulated by interferon-gamma and macrophage colony-stimulating factor. Immunol. Lett. 78, 97101. 397. Darmani, H., Parton, J., Harwood, J. L., Jackson, S. K. (1994) Interferongamma and polyunsaturated fatty acids increase the binding of lipopolysaccharide to macrophages. Int. J. Exp. Pathol. 75, 363368. 398. Adib-Conquy, M., Cavaillon, J. M. (2002) Gamma interferon and granulocyte/monocyte colony-stimulating factor prevent endotoxin tolerance in human monocytes by promoting interleukin-1 receptor-associated kinase expression and its association to MyD88 and not by modulating TLR4 expression. J. Biol. Chem. 277, 2792727934. 399. Held, T. K., Weihua, X., Yuan, L., Kalvakolanu, D. V., Cross, A. S. (1999) Gamma interferon augments macrophage activation by lipopolysaccharide by two distinct mechanisms, at the signal transduction level and via an autocrine mechanism involving tumor necrosis factor alpha and interleukin-1. Infect. Immun. 67, 206 212. 400. de Wit, H., Hoogstraten, D., Halie, R. M., Vellenga, E. (1996) Interferongamma modulates the lipopolysaccharide-induced expression of AP-1 and NF-kappa B at the mRNA and protein level in human monocytes. Exp. Hematol. 24, 228 235. 401. Huang, Y., Krein, P. M., Muruve, D. A., Winston, B. W. (2002) Complement factor B gene regulation: synergistic effects of TNF-alpha and IFN-gamma in macrophages. J. Immunol. 169, 26272635.

402. Jahnke, A., Johnson, J. P. (1994) Synergistic activation of intercellular adhesion molecule 1 (ICAM-1) by TNF-alpha and IFN-gamma is mediated by p65/p50 and p65/c-Rel and interferon-responsive factor Stat1 alpha (p91) that can be activated by both IFN-gamma and IFN-alpha. FEBS Lett. 354, 220 226. 403. Cheshire, J. L., Baldwin Jr., A. S. (1997) Synergistic activation of NF-kappaB by tumor necrosis factor alpha and gamma interferon via enhanced I kappaB alpha degradation and de novo I kappaB beta degradation. Mol. Cell. Biol. 17, 6746 6754. 404. Ohmori, Y., Schreiber, R. D., Hamilton, T. A. (1997) Synergy between interferon-gamma and tumor necrosis factor-alpha in transcriptional activation is mediated by cooperation between signal transducer and activator of transcription 1 and nuclear factor kappaB. J. Biol. Chem. 272, 14899 14907. 405. Fujita, T., Reis, L. F., Watanabe, N., Kimura, Y., Taniguchi, T., Vilcek, J. (1989) Induction of the transcription factor IRF-1 and interferon-beta mRNAs by cytokines and activators of second-messenger pathways. Proc. Natl. Acad. Sci. USA 86, 9936 9940. 406. Han, Y., Rogers, N., Ransohoff, R. M. (1999) Tumor necrosis factoralpha signals to the IFN-gamma receptor complex to increase Stat1alpha activation. J. Interferon Cytokine Res. 19, 731740. 407. Kumar, A., Haque, J., Lacoste, J., Hiscott, J., Williams, B. R. (1994) Double-stranded RNA-dependent protein kinase activates transcription factor NF-kappa B by phosphorylating I kappa B. Proc. Natl. Acad. Sci. USA 91, 6288 6292. 408. Tannenbaum, C. S., Major, J. A., Hamilton, T. A. (1993) IFN-gamma and lipopolysaccharide differentially modulate expression of tumor necrosis factor receptor mRNA in murine peritoneal macrophages. J. Immunol. 151, 6833 6839. 409. Aggarwal, B. B., Eessalu, T. E., Hass, P. E. (1985) Characterization of receptors for human tumour necrosis factor and their regulation by gamma-interferon. Nature 318, 665 667. 410. Michishita, M., Yoshida, Y., Uchino, H., Nagata, K. (1990) Induction of tumor necrosis factor-alpha and its receptors during differentiation in myeloid leukemic cells along the monocytic pathway. A possible regulatory mechanism for TNF-alpha production. J. Biol. Chem. 265, 8751 8759. 411. Aggarwal, B. B., Eessalu, T. E. (1987) Induction of receptors for tumor necrosis factor-alpha by interferons is not a major mechanism for their synergistic cytotoxic response. J. Biol. Chem. 262, 10000 10007. 412. Rhee, S. H., Jones, B. W., Toshchakov, V., Vogel, S. N., Fenton, M. J. (2003) Toll-like receptor 2 and 4 activate STAT1 serine phosphorylation by distinct mechanisms in macrophages. J. Biol. Chem. 278, 22506 22512. 413. Gao, J. J., Filla, M. B., Fultz, M. J., Vogel, S. N., Russell, S. W., Murphy, W. J. (1998) Autocrine/paracrine IFN-alphabeta mediates the lipopolysaccharide-induced activation of transcription factor Stat1alpha in mouse macrophages: pivotal role of Stat1alpha in induction of the inducible nitric oxide synthase gene. J. Immunol. 161, 4803 4810. 414. Gessani, S., Belardelli, F., Pecorelli, A., Puddu, P., Baglioni, C. (1989) Bacterial lipopolysaccharide and gamma interferon induce transcription of beta interferon mRNA and interferon secretion in murine macrophages. J. Virol. 63, 27852789. 415. Toshchakov, V., Jones, B. W., Perera, P. Y., Thomas, K., Cody, M. J., Zhang, S., Williams, B. R., Major, J., Hamilton, T. A., Fenton, M. J., Vogel, S. N. (2002) TLR4, but not TLR2, mediates IFN-beta-induced STAT1alpha/beta-dependent gene expression in macrophages. Nat. Immunol. 3, 392398. 416. Ohmori, Y., Hamilton, T. A. (2001) Requirement for STAT1 in LPS-induced gene expression in macrophages. J. Leukoc. Biol. 69, 598 604. 417. Kawai, T., Takeuchi, O., Fujita, T., Inoue, J., Muhlradt, P. F., Sato, S., Hoshino, K., Akira, S. (2001) Lipopolysaccharide stimulates the MyD88independent pathway and results in activation of IFN-regulatory factor 3 and the expression of a subset of lipopolysaccharide-inducible genes. J. Immunol. 167, 58875894. 418. Ohmori, Y., Hamilton, T. A. (1995) The interferon-stimulated response element and a kappa B site mediate synergistic induction of murine IP-10 gene transcription by IFN-gamma and TNF-alpha. J. Immunol. 154, 52355244. 419. Gao, J., Morrison, D. C., Parmely, T. J., Russell, S. W., Murphy, W. J. (1997) An interferon-gamma-activated site (GAS) is necessary for full expression of the mouse iNOS gene in response to interferon-gamma and lipopolysaccharide. J. Biol. Chem. 272, 1226 1230.

188

Journal of Leukocyte Biology Volume 75, February 2004

http://www.jleukbio.org

420. Xie, Q. W., Kashiwabara, Y., Nathan, C. (1994) Role of transcription factor NF-kappa B/Rel in induction of nitric oxide synthase. J. Biol. Chem. 269, 4705 4708. 421. Ledebur, H. C., Parks, T. P. (1995) Transcriptional regulation of the intercellular adhesion molecule-1 gene by inammatory cytokines in human endothelial cells. Essential roles of a variant NF-kappa B site and p65 homodimers. J. Biol. Chem. 270, 933943. 422. Caldenhoven, E., Coffer, P., Yuan, J., Van de Stolpe, A., Horn, F., Kruijer, W., Van der Saag, P. T. (1994) Stimulation of the human intercellular adhesion molecule-1 promoter by interleukin-6 and inter-

feron-gamma involves binding of distinct factors to a palindromic response element. J. Biol. Chem. 269, 21146 21154. 423. Look, D. C., Pelletier, M. R., Holtzman, M. J. (1994) Selective interaction of a subset of interferon-gamma response element-binding proteins with the intercellular adhesion molecule-1 (ICAM-1) gene promoter controls the pattern of expression on epithelial cells. J. Biol. Chem. 269, 8952 8958. 424. Hume, D. A. (2000) Probability in transcriptional regulation and its implications for leukocyte differentiation and inducible gene expression. Blood 96, 23232328.

Schroder et al. Signals, mechanisms, and functions of IFN- 189

S-ar putea să vă placă și