You are on page 1of 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
Inflammatory 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 di- Macrophages respond to a range of different cell products
verse array of cellular programs through transcrip- during the innate and acquired immune response. Of these,
tional regulation of immunologically relevant IFN-␥ (originally called macrophage-activating factor) is
genes. This article reviews the current understand- among the most important. Macrophage stimulation with IFN-␥
ing of IFN-␥ ligand, receptor, signal transduction, induces direct antimicrobial and antitumor mechanisms as well
and cellular effects with a focus on macrophage as up-regulating antigen processing and presentation path-
responses and to a lesser extent, responses from ways. IFN-␥ orchestrates leukocyte attraction and directs
other cell types that influence macrophage func- growth, maturation, and differentiation of many cell types
tion during infection. The current model for IFN-␥ [2– 4], in addition to enhancing natural killer (NK) cell activity
signal transduction is discussed, as well as signal [5] and regulating B cell functions such as immunoglobulin (Ig)
regulation and factors conferring signal specificity. production and class switching [4, 6]. This review provides a
Cellular effects of IFN-␥ are described, including brief overview of IFN-␥ biology with respect to the well-
up-regulation of pathogen recognition, antigen characterized responses that alter macrophage function during
processing and presentation, the antiviral state, infectious challenge.
inhibition of cellular proliferation and effects on
apoptosis, activation of microbicidal effector func-
tions, immunomodulation, and leukocyte traffick- THE IFNs
ing. In addition, integration of signaling and re-
sponse with other cytokines and pathogen-associ-
The IFNs were originally discovered as agents that interfere
ated molecular patterns, such as tumor necrosis
with viral replication [7]. Initially, they were classified by the
factor-␣, interleukin-4, type I IFNs, and lipopoly-
secreting cell type but are now classified into type I and type
saccharide are discussed. J. Leukoc. Biol. 75:
II according to receptor specificity and sequence homology.
163–189; 2004.
The type I IFNs are comprised of multiple IFN-␣ subtypes
(14 –20, depending on species), IFN-␤, IFN-␻, and IFN-␶, all
Key Words: macrophage 䡠 cytokine 䡠 lipopolysaccharide 䡠 Toll-like
receptor 䡠 inflammation of which are structurally related and bind to a common het-
erodimeric 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 fibroblasts are
INTRODUCTION
a major cellular source of IFN-␤ [8]. IFN-␤ is also produced by
macrophages under appropriate stimulus (discussed later). Vi-
This review focuses on the interplay between interferon-␥ ral infection is the classic stimulus for IFN-␣ and IFN-␤
(IFN-␥) and macrophages in inflammation and acquired im- expression [8, 9]. Secretion of IFN-␶ has only been reported in
munity during infection. Macrophages are extremely versatile ruminants [10].
cells involved in a number of complex functions in disease and IFN-␥ is the sole type II IFN. It is structurally unrelated to
health. A pathogen encounters macrophages soon after host type I IFNs, binds to a different receptor, and is encoded by a
entry, and the result of these encounters is fundamental to the separate chromosomal locus. Initially, it was believed that
host’s ability to mount an effective immune response. Macro- CD4⫹ T helper cell type 1 (Th1) lymphocytes, CD8⫹ cyto-
phage “activation”, broadly defined as “acquisition of compe- toxic lymphocytes, and NK cells exclusively produced IFN-␥
tence to execute a complex function” [1], is among the first [8, 11]. However, there is now evidence that other cells, such
actions to occur in innate immunity to potential pathogens. In
most situations, macrophages are activated to acquire micro-
bicidal effector functions and secrete proinflammatory cyto- 1
Correspondence: Institute for Molecular Bioscience, University of Queens-
kines, resulting in inflammation and recruitment of immune land, St. Lucia, Brisbane 4072, Australia. E-mail: D.Hume@imb.uq.edu.au
cells and subsequent elimination of the microbe by phagocy- Received June 2, 2003; revised July 25, 2003; accepted July 27, 2003; doi:
tosis or release of toxic metabolites. 10.1189/jlb.0603252.

Journal of Leukocyte Biology Volume 75, February 2004 163


as B cells, NKT cells, and professional antigen-presenting cells may induce the apoptotic program [35, 38, 39]. This mecha-
(APCs) secrete IFN-␥ (reviewed in refs. [5, 12–16]). IFN-␥ nism may aid in the Th2-to-Th1 phenotype switch apparent
production by professional APCs [monocyte/macrophage, den- when CD4⫹ cells are treated with IFN-␥: the Th2 population
dritic cells (DCs)] acting locally may be important in cell decreases as a result of growth-inhibitory and proapoptotic
self-activation and activation of nearby cells [12, 13]. IFN-␥ effects of IFN-␥, and the Th1 population continues to prolif-
secretion by NK cells and possibly professional APCs is likely erate as a result of blockade of IFN-␥ function. As a conse-
to be important in early host defense against infection, whereas quence, IFN-␥ signaling causes IFNGR2 down-regulation and
T lymphocytes become the major source of IFN-␥ in the switching to a Th1 phenotype as a T cell population but not at
adaptive immune response [12, 17]. a single-cell level [37].
IFN-␥ production is controlled by cytokines secreted by Both IFNGR chains lack intrinsic kinase/phosphatase ac-
APCs, most notably interleukin (IL)-12 and IL-18. These cy- tivity and so must associate with signaling machinery for signal
tokines serve as a bridge to link infection with IFN-␥ produc- transduction. The IFNGR1 intracellular domain contains bind-
tion in the innate immune response [18 –24]. Macrophage ing motifs for the Janus tyrosine kinase (Jak)1 and the latent
recognition of many pathogens induces secretion of IL-12 and cytosolic factor, signal transducer and activator of transcription
chemokines [e.g., macrophage-inflammatory protein-1␣ (MIP- (Stat)1. In humans, the Jak1-binding motif LPKS is a mem-
1␣); ref. 25]. These chemokines attract NK cells to the site of brane-proximal sequence located at residues 266 –269 [40 –
inflammation, and IL-12 promotes IFN-␥ synthesis in these 42]. The human Stat1-binding site YDKPH is positioned at
cells [25, 26]. In macrophages, NK and T cells, the combina- residues 440 – 444 [43]. This motif contains an essential Y440
tion of IL-12 and IL-18 stimulation further increases IFN-␥ phosphorylation site that is phosphorylated during signal trans-
production [20, 23, 24, 27, 28]. duction to allow Stat1 recruitment to the receptor [42– 44].
Negative regulators of IFN-␥ production include IL-4, IL- Residues 441DKPH444 are responsible for the binding speci-
10, transforming growth factor-␤, and glucocorticoids [17, 21, ficity of IFNGR1 and Stat1 [41, 43, 44]. The Jak1- and Stat1-
27–29]. Given the complexity of IFN-␥ regulation, it is not binding motifs are required for receptor phosphorylation, sig-
surprising that inbred mouse strains vary in their ability to nal transduction, and induction of biological response. Recep-
secrete this cytokine; for example, T lymphocytes of C57BL/6 tor-ligand internalization is mediated by an isoleucine-leucine
and C3H mice secrete significantly higher amounts of IFN-␥ sequence at residues 270 and 271 of the mature IFNGR1 chain
compared with the T lymphocytes of BALB/c and B10.D2 in humans. Mutant receptors with deletion or substitution of
mice. Increased IFN-␥ production in these strains is associated this sequence cannot internalize ligand, although signal trans-
with greater resistance to bacteria and viruses [30 –32]. Many duction is not affected [42, 45].
excellent reviews on the regulation of IFN-␥ production have The intracellular region of IFNGR2 contains a noncontigu-
been published recently and the reader is referred to these for ous binding motif for recruitment of Jak2 kinase for participa-
further information [11–13]. tion 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].
THE IFN-␥ RECEPTOR Cross-linking experiments with labeled human IFN-␥ demon-
strated that IFN-␥ only associates with IFNGR2 when the
Functional IFN-␥ receptor (IFNGR) is comprised of two li- IFNGR1 chain is present, indicating that the primary interac-
gand-binding IFNGR1 chains associated with two signal-trans- tion between the IFNGR2 chain and the IFN-␥:IFNGR1 com-
ducing IFNGR2 chains and associated signaling machinery. plex is found between IFNGR1 and IFNGR2, although it is
IFNGR1 and IFNGR2 chains belong to the class II cytokine likely that IFNGR2 also interacts weakly with the ligand [46,
receptor family, a class of receptors that bind ligand in the 48].
small angle of a V formed by the two Ig-like folds that consti- IFN-␥:IFNGR1 and IFNGR1:IFNGR2 interactions are spe-
tute the extracellular domain. It is likely that the IFNGR cies-specific. A number of studies with chimeric receptors in
chains and other family members (type I IFN receptor and which the intracellular, transmembrane, or extracellular do-
tissue factor) evolved from primitive adhesive molecules [33, mains were swapped between the human and murine IFNGR
34]. chains showed that species specificity in interactions of the
The IFNGR2 chain is generally the limiting factor in IFN-␥ ligand:receptor complex is restricted to the receptor extracel-
responsiveness, as the IFNGR1 chain is usually in surplus [8, lular domains [49 –53].
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 IFN-␥ SYSTEM DYSFUNCTION
CD4⫹ Th1 populations have very low levels of cell-surface
expression of the IFNGR2 chain and thus low expression of IFN-␥⫺/⫺ and IFNGR1⫺/⫺ mice showed no overt developmen-
active IFNGR, leading to a functional blockade of some as- tal defects, and their immune system appeared to develop
pects of IFN-␥ signaling [36, 37]. As the growth-inhibitory normally [54]. However, these mice show deficiencies in nat-
effects of IFN-␥ are blocked, T cells expressing low levels of ural resistance to bacterial, parasitic, and viral infections such
the IFNGR2 chain continue to proliferate during IFN-␥ treat- as vaccinia virus, Theiler’s murine encephalomyelitis virus,
ment. Conversely, IFN-␥ exposure to CD4⫹ Th2 populations Leishmania major, Toxoplasma gondii, Listeria monocytogenes,
displaying high levels of IFNGR2 inhibits proliferation and and several poorly virulent mycobacteria species [54 –59].

164 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


Some viruses (e.g., vaccinia virus, Theiler’s murine encepha- SIGNAL TRANSDUCTION
lomyelitis virus, and lymphocytic choriomeningitis virus) ap-
pear to require both type I and II IFN pathways, whereas IFN-␥ primarily signals through the Jak-Stat pathway, a path-
viruses such as Semliki forest virus and vesicular stomatitis way used by over 50 cytokines, growth factors, and hormones to
virus require a predominantly type I IFN response for efficient affect gene regulation [90]. Jak-Stat signaling involves sequen-
clearing [60]. These observations have prompted the sugges- tial receptor recruitment and activation of members of the
tion that the two IFN systems may have evolved to complement Janus family of kinases (Jaks: Jaks 1–3 and Tyk2) and the Stats
each other in overlapping but nonredundant activities to de- (Stats 1– 6, including Stat5a and Stat5b) to control transcrip-
fend against a broad spectrum of pathogens [60, 61]. Histori- tion of target genes via specific response elements. As this
cally, the type I IFNs were considered to be primarily antiviral signaling mechanism is a recurring theme amongst members of
agents with limited immunomodulatory activity, whereas IFN-␥ the cytokine receptor superfamily, IFN-␥-induced Jak-Stat sig-
was considered to be primarily an immunomodulator with naling has emerged as the global paradigm for class II cytokine
limited antiviral activity [8]. It is now apparent that both types receptor signal transduction.
of IFN possess antiviral and immunomodulatory activities to The IFNGR1 and IFNGR2 subunits of the IFN-␥ receptor
some degree [62]. The antiviral activity of both types of IFNs is were thought not to strongly associate with each other in the
often important in the early stages of viral infection, but their absence of ligand [13, 40, 47, 91, 92], but new techniques
immunomodulatory activities become important in later stages allowing the study of receptor chain interactions in intact cells
of infection when the adaptive immune response becomes have shown the receptor complex is assembled before ligand
critical [60]. binding [93]. Upon ligand binding, the intracellular domains of
Patients with inactivating mutations of the human IFNGR1 the receptor chains open out to allow association of downstream
or IFNGR2 chains show clinical presentation similar to the signaling components. Biologically active IFN-␥ is a noncova-
mouse models. Human loss-of-function mutations in the lent homodimer formed by the self-association of two mature
IFNGR1 or IFNGR2 chain are closely associated with severe polypeptides in an antiparallel orientation [94, 95] and thus
susceptibility to poorly virulent mycobacteria, often presenting binds IFNGR1 in 2:2 binding stoichiometry [47, 95–97]. Li-
as early onset bacillus Calmette-Guerin infection and fatality gand binding induces Jak2 autophosphorylation and activation,
in childhood [63–71]. Partial deficiency of either chain gives a which allows Jak1 transphosphorylation by Jak2 (Fig. 1) [98].
milder phenotype and a much better prognosis [63– 65]. Loss of The activated Jak1 phosphorylates functionally critical ty-
functional IFNGR1 appears to be associated only with in- rosines on residue 440 of each IFNGR1 chain to form two
creased susceptibility to infection with some viruses and in- adjacent docking sites for the SH2 domains of latent Stat1 [41,
tracellular bacteria; increased susceptibility to other common 99 –101]. The receptor-recruited Stat1 pair is phosphorylated
bacterial and fungi pathogens have not been reported [72]. near the C terminus at Y701, probably by Jak2 [98]. Phos-
In addition to recurrent infection, infants with deficient phorylation induces dissociation of a Stat1 homodimer (also
production of IFN-␥ exhibited decreased neutrophil mobility known as ␥-IFN activation factor) from the receptor [100]. The
and NK cell activity, highlighting the importance of IFN-␥ in four critical tyrosines (contained by Jak1, Jak2, IFNGR1, and
the inflammatory response and immunoregulation [73]. It is Stat1) are phosphorylated within 1 min of IFN-␥ treatment [41,
interesting that natural IFN-␥ polymorphisms have been cor- 99].
related with increased longevity [74]. It has been proposed that The dissociated Stat1 homodimer enters the nucleus and
a slightly dampened inflammatory status caused by an IFN-␥ binds to promoter elements to initiate or suppress transcription
polymorphism, while not enough to significantly impact on the of IFN-␥-regulated genes (reviewed in refs. [102–104]). Stat1
individual’s ability to clear infection, may prevent or defer homodimers bind DNA at GAS elements of consensus se-
quence TTCN(2-4)GAA [105]. The key function of Stat1 in
inflammation-related diseases such as cardiovascular disease,
mediating IFN-␥ signal transduction is indicated by the phe-
neurodegeneration, osteoarthritis, osteoporosis, and diabetes
notype of Stat1⫺/⫺ mice, which largely phenocopy IFNGR1⫺/⫺
[74].
mice during IFN-␥ stimulation [106].
Aside from functions in host defense, IFN-␥ may also con-
The first wave of IFN-␥-induced transcription occurs within
tribute to autoimmune pathology. Although IFN-␥ production
15–30 min of IFN-␥ treatment [107]. Many of the induced
was shown to be disease-limiting in autoimmune models such
genes are in fact transcription factors (for example, IRF-1),
as murine experimental allergic encephalomyelitis (EAE) [75,
which are activated by IFN-␥ and are able to further drive
76], it may contribute to autoimmune nephritis [77, 78]. In
regulation of the next wave of transcription. Transcription of
humans, IFN-␥ is implicated in pathology of diseases such as
many IFN-␥-responsive genes is controlled by a GAS element
systemic lupus erythematosus [79, 80], multiple sclerosis [81],
or an ISRE [108].
and insulin-dependent diabetes mellitus [82– 84].
IFN-␥ function is significant in tumor surveillance. IFNGR1
knockout (KO) mice, cells with dominant-negative IFNGR1
mutations, and cells treated with IFN-␥-neutralizing antibodies Stat1
display compromised tumor rejection [85– 87]. IFN-␥ protects
against tumor development and directs the immunogenic phe- The crystal structure for tyrosine phosphorylated Stat1 bound
notype of tumors that arise in an immunocompetent host (the to DNA has been determined [109]. The structure verified that
“cancer immunoediting” concept; reviewed in refs. [88, 89]). the SH2 domain necessary for dimerization is also involved in

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


IFNγ

Cytosol
JAK2
JAK1

JAK1
JAK2

JAK2
JAK1

JAK1
JAK2
P P

1
AT
ST

T1
STA
IRF-1

STAT1

STAT2
STAT1

STAT1
T1
STA
P P P
STA

P P P
IRF-1
T1

IRF-9 IRF-9

Nucleus

eg. ICAM-1
MIG eg. IRF-2
STAT1
STAT1

P iNOS
P IFNβ
IFN-regulated gene IFN-regulated gene
GAS ISRE

IRF-1

1
AT
STAT1

Stat1
STAT1

ST
P
P IRF-E/GAS/IRF-E
IRF-1 IRF-1
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 Stat2:IRF-9 complex (known as ISGF3) was thought to be the
the Y701 phosphorylation site and also a S727 phosphorylation typical type I IFN transcription factor. It is now evident that
site, both of which are functionally important for efficient type I IFN signals primarily through ISGF3 but also uses Stat1
signaling [110 –113]. All Stat proteins possess a SH2 domain homodimers and conversely, type II IFN signals primarily
that when tyrosine phosphorylated, mediates homodimerization through Stat1 homodimers but also uses ISGF3 to activate
or heterodimerization with other Stat molecules [114, 115]. transcription [116]. This may explain many of the overlapping
Although the simplistic model described above suggests that effects of the two types of IFNs.
IFN-␥ signaling produces active Stat1 homodimers alone, other Phosphorylation of Stat1 at S727 is essential for maximal
active complexes such as Stat1 heterodimers (e.g., Stat1:Stat2) ability to activate transcription of target genes [111–113, 120 –
and heterotrimers (e.g., Stat1:Stat1:IRF-9, Stat1:Stat2:IRF-9) 122]. A number of different stimuli induce Stat1 serine phos-
form during signaling [102, 116 –118]. Stat2 is the only Stat phorylation, including types I and II IFN, lipopolysaccharide
that does not contain a DNA-binding domain [119]. The Stat1: (LPS), IL-2, IL-12, tumor necrosis factor ␣ (TNF-␣), and

166 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


platelet-derived growth factor [110 –112, 123, 124]. This may sequence very similar to the Simian virus-40 T antigen NLS,
126
be a mechanism whereby S727 serves as an avenue for mod- RKRKRSR132 in mice and 128KTGKRKR134 in humans
ulation of IFN-␥ signaling by independent extracellular cues. [139, 142–145]. This sequence is a functional NLS and is
For example, LPS signaling increases Stat1 S727 phosphory- essential for the full biological activity of IFN-␥ [144 –148]. A
lation independently of Y701 phosphorylation in macrophages, deletion mutant of the putative NLS in human IFN-␥ is still
thereby augmenting cellular responses to IFN-␥ [110]. able to bind with high affinity to the IFNGR1 chain, but
The ability of Stat1 to activate or repress gene transcription induction of biological response is absent [149]. The addition
depends on the presence of other transcription factors binding of a heterologous NLS is able to rescue biological responsive-
to the promoter element and Stat1 interaction with other fac- ness in this system. Another study tracked the cellular local-
tors. Stat1 activation is necessary but not sufficient for tran- ization of components of the IFNGR complex following ligand
scription of a number of genes [125, 126]. Likewise, factors exposure [150]. IFN-␥, Stat1, and IFNGR1 were found to be
such as oncostatin M, which participate in Jak-Stat signaling to quickly internalized, colocalized, and accumulated in the nu-
produce activated Stat1, do not induce transcription of IFN-␥- cleus, while the majority of the IFNGR2 subunit remained at
inducible genes [127]. Stat1 interaction with transcription fac- the cell surface throughout signaling. Stat1 nuclear transloca-
tors such as IRF-9, upstream stimulatory factor-1, specificity tion is mediated by NPI-1 [135], and immunoprecipitation of
protein-1, and heat shock factor-1 have been reported (re- Stat1 following IFN-␥ treatment showed that IFN-␥, IFNGR1,
viewed in ref. [119]) and are likely to influence specificity of Stat1, and NPI-1 form a complex at the nuclear pore during
DNA binding and transactivator ability. Phosphorylation of translocation [142, 143]. Subramaniam et al. [90] proposed a
S727 is necessary for interaction with some proteins, such as model suggesting a function for intracellular ligand in Stat1
BRCA1 and MCM-5, and this interaction potentiates Stat1- nuclear localization. In this model, extracellular ligand binds
mediated transcription [128, 129]. In some promoters, tran- to the IFNGR1 chain, causing endocytosis of the complex and
scription is maximal when more than one GAS-binding tran- recycling of the IFNGR2 chain to the cell surface. The intra-
scriptional activator binds to tandem GAS sites in the target cellular domain of IFNGR1 associates with activated Stat1 and
gene promoter [130, 131]. intracellular IFN-␥, and the IFN␥:IFNGR1:Stat1 complex is
In the absence of IFN-induced tyrosine phosphorylation, actively transported through the nuclear pore via IFN-␥–NLS
Stat1 drives constitutive transcription of genes such as interaction with the NPI-1 importin. Activated Stat1 may then
caspases [132, 133]. The Stat1 KO mouse showed increased bind to promoter elements to activate transcription.
lymphocyte survival and proliferation, which was attributed to Functional Stat1 is crucial to host response to infection. KO
insufficient caspase levels to carry out the apoptotic program of Stat1 renders mice extremely sensitive to infection by viral
[133]. Stat1 also constitutively regulates expression of the low and microbial pathogens (e.g., vesicular stomatitis viruses,
molecular protein 2 (LMP2) gene [134]. A complex of unphos- mouse herpes virus, L. monocytogenes) [106, 151]. This sus-
phorylated Stat1 and IRF-1 may mediate this. ceptibility has been attributed to defective immune system
The mechanism of Stat1 entry into the nucleus is still function as a result of the combined loss of types I and II IFN
controversial, but the involvement of importin-␣-1 (NPI-1) is responses [151]. NO production in Stat1⫺/⫺ macrophages
implied [135]. Two major models have been suggested, one primed with types I or II IFN and treated with LPS is greatly
being the classic nuclear importin mechanism and the other impaired compared with the wild type (WT), resulting in re-
involving a requirement for intracellular IFN-␥. Nuclear entry duced macrophage microbicidal function [106]. Impaired mac-
of Stat1 is apparent at 15 min and almost complete after 30 min rophage microbicidal function may contribute to the increased
exposure to IFN-␥ [136]. Stat1 nuclear translocation does not susceptibility to infection apparent in Stat1⫺/⫺ mice.
appear to be a result of liberation from a cytoplasmic anchor or Stat1 KO mice also demonstrated the existence of Stat1-
transport by the cellular cytoskeleton or microtubules [136]. independent mechanisms in IFN-␥ signaling. IFN-␥ may exert
Experiments with green fluorescent protein-tagged Stat1 im- pro- and antiproliferative signals through Stat1-independent
plied that it travels through the cytoplasm to the nucleus by and Stat1-dependent pathways, respectively [152–154]. Al-
random walk movement [136]. A model was proposed whereby though IFN-␥ treatment of lymphocytes and fibroblasts inhib-
Stat1 dimerization allows interaction with importin NPI-1, ited cell growth, the same treatment in Stat1⫺/⫺ cells promoted
which then mediates translocation through the nuclear pore proliferation and decreased apoptotic signals [133, 153, 154].
[135]. The nuclear membrane forms an efficient barrier to Microarray analysis revealed that many genes, including the
inactivated Stat1, and entry requires a gain of nuclear local- growth-promoting c-myc and c-jun genes, were induced in
ization sequence (NLS) function [136 –138]. To date, no “clas- response to IFN-␥ in Stat1⫺/⫺ cells [152]. This signaling was
sic” NLS (single stretch of basic residues or bipartite motif; ref. not dependent on the Stat1-docking site of IFNGR1. Although
[139]) has been identified for Stat1, indicating an as-yet un- much more susceptible than their WT counterparts, Stat1⫺/⫺
identified, novel NLS is unmasked during Stat1 homodimer- mice are significantly more resistant to viral infection than
ization, or the NLS of a Stat1 homodimer-binding partner IFNGR1/IFNAR1 double-KO mice, indicating the existence of
mediates nuclear translocation. IFN-dependent, Stat1-independent, antiviral mechanisms
The second model for Stat1 nuclear entry involves a require- [155].
ment for intracellular IFN-␥. IFN␥ enters the cell using a A loss-of-function, heterozygous mutation in the Stat1 gene
mechanism that is unclear at present. Nuclear accumulation of was recently described in humans, resulting in a dominant-
IFN-␥ is apparent shortly after cellular exposure to IFN-␥ negative effect over the WT allele for Stat1 activation [156].
[140, 141]. Nuclear translocation is mediated by a C-terminal Although these patients demonstrated increased susceptibility

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


to bacterial infection, viral infections took the normal clinical 172]. In some cell types, IFN-␥ signaling may induce degra-
course. This suggests that IFN-␥-induced, physiologically sig- dation of the internalized receptor, thereby down-regulating
nificant, Stat1-independent control of antiviral responses also IFNGR1 surface expression [45]. This may be a mechanism by
occurs in humans. which these cells can prevent overstimulation of the pathway
Mechanisms of Stat1-independent IFN-␥ signaling are cur- through cellular desensitization. Such a mechanism may also
rently unclear. One possibility is the existence of an unknown allow differential regulation of IFN-␥-induced responses in
IFN-␥ receptor or IFN-␥ receptor complex linked to an alter- different cell types [37].
native signaling pathway. This hypothesis was suggested when One of the most inducible targets of IFN-␥ is a specific
it was found that anti-IFN-␥ antibodies affected the clinical feedback inhibitor, SOCS-1, which associates with Jak1/2,
course of EAE in IFNGR1 KO mice [75]. interfering with tyrosine kinase activity and inhibiting down-
stream IFN-␥ signaling [173–177]. SOCS-1 may also promote
degradation of signaling machinery by binding to and targeting
IRF-1 signaling molecules for the ubiquitin-proteasome pathway
[178]. Overexpression of SOCS-1 in cells in vitro or in trans-
genic mice causes loss of responsiveness to IFN-␥ [176, 179].
The IRF gene family is intimately involved in mediating the Conversely, SOCS-1⫺/⫺ mice are hyper-responsive to micro-
type I and II IFN signal cascades. Members of the IRF family bial infection, exhibit enhanced macrophage cytocidal activity,
share similar structure, including an amino-terminal DNA- and die of IFN-␥-dependent, multisystem inflammatory tissue
binding domain containing multiple tryptophan residues and a destruction in the absence of infection [180, 181]. This indi-
carboxy-terminal transcriptional activation or repression do- cates that IFN-␥ is normally present and performs important
main (reviewed in ref. [157]). IRF-1, IRF-2, and IRF-9 (p48, functions in vivo, even in the absence of infection. Another
ISGF3-␥) all participate in IFN-␥ signaling. Basal IRF-1 ex- SOCS protein, SOCS-3, is also induced by IFN-␥ and nega-
pression has functions in constitutive gene expression [134], tively regulates IFN-␥ signaling, although perhaps less effec-
but Stat1 and nuclear factor (NF)-␬B interaction with promoter tively than SOCS-1 [182].
elements dramatically increases IRF-1 transcription [158 – Aside from SOCS protein function, IFN-␥ signal down-
160]. IRF-1 expression is up-regulated in response to types I regulation can occur by receptor and Jak dephosphorylation by
and II IFN, virus, or cytokines. IRF-1 transcriptional regula- the PTPs Shp1 and Shp2 [183–187]. Shp2 null cells exhibit
tory activity is regulated independently in response to IFN-␥, elevated tyrosine phosphorylation and DNA-binding ability of
virus, and dsRNA [161]. The physiological relevance of IRF-1 Stat1 and show enhanced suppression of cell growth with
function in transcriptional control of IFN-␥-regulated genes is IFN-␥ treatment [184]. Shp1 KO mice (motheaten mice, me/
highlighted by IRF-1 KO mice, in which many genes are me) die from a condition caused by dysregulation of a large
submaximally induced by IFN-␥ when compared with their WT number of cytokines, possibly including IFN-␥ [188].
littermates [162, 163]. IRF-1 functions are not limited to IFN-␥ Stat1 phosphorylation may also be controlled by nuclear
signaling; IFN-␥-independent activities in viral/bacterial in- dephosphorylation. Stat1 homodimers enter the nucleus and
fection, lymphocyte development, regulation of cell cycle, are actively retained by a mechanism dependent on the kinase
growth inhibition, apoptosis and tumor suppression have been function of Jak1 [189]. McBride et al. [137] proposed a model
reported [107, 123, 135, 159, 164 –168]. in which nuclear export is mediated by a leucine-rich nuclear
IRF-1 drives inducible expression of many target genes export signal (NES) at residues 197–205, which is hidden
through interaction with the IRF-E site, consensus G(A)AAA when Stat1 is bound to DNA. Stat1 dephosphorylation by a
G T
/C /CGAAAG/CT/C [102, 164]. This specificity overlaps with nuclear protein tyrosine phosphate (PTP) allows the release of
the ISRE consensus site A/GNGAAANNGAAACT, recognized DNA and subsequent recognition of the NES by chromosome
by ISGF3, which is induced by type I IFN and to a lesser region maintenance/exportin 1 (CRM1), an export receptor for
extent, type II IFN [102]. In this way, IRF-1 is able to induce NES-containing proteins [190, 191]. CRM1 then mediates
a subset of the full spectrum of IFN-inducible genes. IRF-2 Stat1 translocation through the nuclear pore in a Ran-GTPase-
also binds to the IRF-E site and largely functions to repress dependent manner. This model is supported by work from other
transcription of IRF-1-inducible genes through its transcrip- groups, demonstrating Stat1 tyrosine dephosphorylation in the
tional repressor domain [102, 159]. nucleus and subsequent recycling of inactive Stat1 to the
Negative regulation of IFN-␥ signaling cytoplasm, and is consistent with the crystal structure of Stat1
bound to DNA [137, 189, 192, 193].
Stat1 activation is inhibited within 1 h of IFN-␥ treatment,
despite the continued presence of extracellular IFN-␥, and so
Signaling specificity
mechanisms must exist to control the extent of ligand stimu- The many cytokines that use Jak-Stat signaling direct tran-
lation of IFN-␥ signaling (Fig. 2) [117, 169]. These mecha- scription through specific but not necessarily unique combina-
nisms involve every level of the pathway. tions of Jaks and Stats. This raises the question of how signal
Following signal transduction, the IFN-␥:IFNGR1 complex specificity is maintained when many cytokines use the same
internalizes and enters the endosomal pathway, where the signaling machinery.
complex dissociates [170]. In many cell types, the IFNGR1 Evidence for a devoted role of Jaks in signal transduction of
chain eventually recycles to the cell surface in its uncoupled, specific cytokines is controversial. Mutant cell lines lacking
dephosphorylated form, and the ligand is degraded [45, 171, Jak1 or Jak2 have demonstrated deficiency in types I and II

168 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


Receptor
expression level

IFNγ

Cytosol
JAK2
JAK1

JAK1
JAK2

JAK2
JAK1

JAK1
JAK2
SHP-2

SOCS-3
P P
SOCS-1
PTPs/SOCS limit
receptor and Jak
phosphorylation

1
AT
ST

T1
STA
IRF-1

STAT1

STAT2
STAT1

STAT1
T1
STA

P P P
STA

P P P
IRF-1
T1
T1
STA
STA

IRF-9 IRF-9
T1

Nucleus
PTP
STAT1
STAT1

P
T1
STA

P
STA

IFN-regulated gene IFN-regulated gene


T1

GAS ISRE

Nuclear PTPs
eg. IRF-1 IRF-2 Transcriptional
dephosphorylate
repression by IRF-2
Stat1 causing
competition for
inactivation and
ISRE/IRF-E sites
nuclear export

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]. in erythropoiesis [197, 198]. Other than suppression of IFN-␥
These deficiencies are reflected in the Jak1 and Jak2 KO mice, signaling, the cytokine response deficit of Jak2 KO mice is
suggesting that these Jaks are functionally specific to their largely nonoverlapping compared with Jak1 KO mice [197,
cytokine systems, and other Jaks cannot substitute for them in 198]. This indicates that other Jaks cannot functionally sub-
vivo. Jak1 gene KO mice exhibit a number of abnormalities, all stitute for Jak2 in vivo, although Jak3 can substitute for Jak2
attributable to signaling deficits in three cytokine receptor in vitro to produce Stat1 homodimers and activate transcription
families: type II cytokine receptors (i.e., types I and II IFN and of class I major histomcompatibility complex (MHC) [199].
IL-10 receptors); all cytokines that use the ␥c subunit for Although specific Stats appear to be dedicated to specific
signaling, thereby promoting lymphopoiesis (e.g., IL-2, IL-4, subsets of cytokine signaling, the same Stat or combination of
IL-7 receptors); and gp130 receptor family members (e.g., Stats can be activated in different cytokine systems to give
IL-6, IL-11, leukemia inhibitory factor receptors) [196]. These completely different biological effects (e.g., IFN-␥ and on-
results showed that in vivo, the role of Jak1 is essential and costatin M as described earlier). Possible mechanisms giving
nonredundant in signal transduction and induction of biologi- rise to specificity of response with respect to Stats include:
cal response for a specific range of cytokine receptors. Jak2 different thresholds/timeframes of Stat activation may be re-
gene KO in mice is embryonic lethal as a result of deficiencies quired for specific biological responses, and specific Stat com-

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


plexes induced by cytokines may give different biological IFN-␥ stimulation induces a replacement of the constitutive
effects. It is widely agreed that signal specificity is largely proteasome subunits with “immunoproteasome” subunits. In
conferred by the particular Stats activated in cytokine signaling unstimulated cells, the proteasome enzymatic subunits are ␤1,
[106, 151, 200 –205]. This is supported by the Stat1 KO ␤2, and ␤5, encoded outside of the MHC locus. IFN-␥ induces
mouse, which displays loss of only IFN signaling, despite expression of new subunits, LMP2, MECL-1, and LMP7, which
multiple reports of Stat1 activation in response to a range of competitively replace ␤1, ␤2, and ␤5 subunits of the protea-
other cytokines in vitro [106, 151]. Similarly, KO mice of the some, respectively [210 –216]. In this way, new species of
other Stat proteins showed a loss of biological responsiveness proteasome are formed, consisting of LMP2:MECL-1:LMP7
to a single cytokine, indicating that Stat activation may be (the immunoproteasome), LMP2:MECL-1:␤5, or ␤1:␤2:LMP7,
specific for a single cytokine in vivo if not in vitro [169]. In as well as low levels of the ␤1:␤2:␤5 constitutive proteasome
addition to the Stats themselves, other factors (e.g., intracellu- [216]. Inducible proteasome replacement is thought to be a
lar IFN-␥ as described above, or cell-specific transcription mechanism by which IFN-␥ can increase the quantity, quality,
factors), which associate with Stats may contribute to the and repertoire of peptides for class I MHC loading. The quan-
specificity of the signal. tity is increased, as overall expression levels of proteasome are
increased. The cleavage specificity of the immunoproteasome
Cross-talk between IFN-␣/␤ and IFN-␥ pathways may allow production of peptides better able to bind class I
MHC and thereby increase efficiency in the system. Peptide
The IFN-␥ and IFN-␣/␤ signal pathways cross-talk at multiple
diversity is thought to increase as a result of differences in
levels. The signal pathways and target genes used by types I
cleavage specificities between different species of proteasome.
and II IFN are partially overlapping, which gives the opportu-
As a whole, this serves to increase levels and diversity of
nity for cross-talk to synergize or antagonize particular func-
epitopes presented for CD8⫹ T cell recognition in the context
tions within the cell. This cross-talk is biologically relevant, as
of class I MHC and thus increase immune surveillance (re-
cells in vivo are not stimulated with one cytokine in isolation,
viewed in ref. [216]). This mechanism may have evolved to
rather a cytokine cocktail instructs gene expression through the
ensure LMP2/LMP7/MECL-1-dependent epitopes are only
integration of multiple signaling pathways.
produced in sites of inflammation and thus avoid autoimmunity
Although IFN-␥ primarily signals through Stat1 homodimers
without compromising appropriate T cell stimulation [216].
in the initial signal cascade, additional signaling molecules are
The efficiency of peptide generation is further increased with
activated. As noted above, the archetypal type I IFN signaling
the IFN-␥-induced PA28, which is composed of PA28␣ and
molecule ISGF3 is activated by IFN-␥, thus providing a mech-
PA29␤ subunits and associates with the proteasome and alters
anism for cross-talk between the types I and II IFN pathways
proteasome proteolytic cleavage preference. It is thought to
[116]. ISGF3 is able to induce expression of type I IFN,
gear antigen processing toward more efficient generation of
thereby further amplifying the response of IFN-␣/␤-induced
TAP- and class I MHC-compatible peptides to increase overall
genes [206, 207]. Conversely, type I IFN can elicit classic type
efficiency of class I MHC peptide delivery [217, 218].
II IFN signaling molecules such as active Stat1 homodimers,
The IFN-␥-inducible TAP transporter is vital in peptide
which are able to bind to GAS sites to activate transcription of
transport from the cytosol to the ER lumen [307]. TAP tran-
target genes [17, 208].
siently associates with class I MHC to aid in efficient peptide
Takaoka et al. [209] published data indicating that IFN-␥
loading [307, 308]. Genes encoding the TAP subunits are
signaling in mouse embryonic fibroblasts requires a constitu-
located within the class II MHC locus and are coordinately
tive subthreshold IFN-␣/␤ signal. They have suggested that
expressed with MHC class I [220, 221].
IFN-␣/␤ may promote the interaction of the IFNGR2 and
The class I MHC complex is composed of a heavy chain
IFNAR1 chains in the caveolar membrane domains of the cell
(consisting of ␣1, ␣2, and ␣3 domains) and a light chain (␤2
membrane. In this model, low level IFN-␣/␤ signal is neces-
microglobulin) and is up-regulated by IFN-␥ stimulation [4,
sary for maintaining its receptor in the phosphorylated form,
222, 224 –232]. Chaperones such as tapasin and GP96 impli-
thus providing a functional aid for efficient assembly of IFN-
cated in aiding in the efficient assembly of peptide:MHC I
␥-activated Stat1 homodimers.
complexes are also up-regulated by IFN-␥ [4, 309, 310].
Class II antigen presentation pathway
CELLULAR EFFECTS OF IFN-␥ Of the IFNs, IFN-␥ alone can efficiently up-regulate the class
II antigen presenting pathway and thus promote peptide-spe-
Class I antigen presentation pathway cific activation of CD4⫹ T cells [4, 234]. IFN-␥ treatment
further up-regulates class II MHC molecules in cells constitu-
Types I and II IFN up-regulate multiple functions within the tively expressing class II MHC, such as B cells, DCs, and cells
class I antigen presentation pathway to increase the quantity of the monocyte-macrophage lineage (professional APCs)
and diversity of peptides presented on the cell surface in the [234]. IFN-␥ is also able to induce class II MHC expression in
context of class I MHC. Up-regulation of cell-surface class I cells that do not constitutively express these genes (nonprofes-
MHC by IFN-␥ (Table 1) is important for host response to sional APCs) [76]. IFN-␥ up-regulates the quantity of peptide:
intracellular pathogens, as it increases the potential for cyto- MHC II complexes on the cell surface by promoting expression
toxic T cell recognition of foreign peptides and thus promotes of several key molecules (Table 1): the Ii chain and the
the induction of cell-mediated immunity. components of the class II MHC complex [234 –239]; cathep-

170 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


TABLE 1. IFN-␥-Regulated Genes and Their Function in Producing IFN-␥ Effects

Antigen Processing and Presentation


Class I antigen presentation pathway
Gene/protein up-regulated
by IFN-␥ Function Refs.
LMP-2, LMP-7, MECL-1 IFN-␥-inducible, enzymatic proteasome subunits, which replace the constitutive [210–216]
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
PA28␣, PA28␤ Proteasome activator (PA)28␣:PA28␤ dimer is a nonenzymatic proteasome [4, 217, 218, 219]
subunit, which alters the specificity of peptides generated to increase
efficiency of class I MHC peptide delivery.
TAP-1, TAP-2 The transporter associated with antigen processing (TAP) is a heterodimer [4, 215, 220–223]
consisting of TAP-1 and TAP-2 subunits and belongs to the adenosine 5⬘-
triphosphate-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.
Class I MHC heavy chain The heavy chain associates with ␤2-microglobulin to form the MHC class I [4, 224–231]
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.
␤2-microglobulin Light chain that associates with the class I MHC heavy chain to form the MHC [222, 230, 232]
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 Tapasin is a chaperone that aids in the retention of empty MHC I in the ER [4, 233]
and peptide loading into MHC I peptide-binding cleft.
Class II antigen presentation pathway
Gene/protein up-regulated
by IFN-␥ Function Refs.
␣1, ␣2, ␤1, ␤2 MHC II Constituents of the heterodimeric MHC II. MHC II displays foreign and self- [234–237]
chains peptides on the cell surface for immune surveillance by CD4⫹ T cells. The
MHC II ␣ and ␤ chains are encoded by the class II MHC locus.
Ii chain The invariant (Ii) chain is a transmembrane chaperone, which traffics MHC II [237–239]
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.
DMA, DMB The DMA:DMB dimer forms DM, a class II-like heterodimeric protein resident [235, 237]
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.
Cathepsins B, H, L Lysosomal proteases implicated in peptide production for class II MHC loading. [240, 241]
CIITA Class II transactivator (CIITA) is a transactivator with no DNA-binding motif. It [234]
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.
Antiviral Effect
Gene/protein up-regulated
by IFN-␥ Function Refs.
PKR Protein kinase dsRNA-regulated (PKR) is a serine/threonine kinase activated [154, 242–247]
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.
ADAR The dsRNA-specific adenosine deaminase (ADAR) catalyzes the deamination of [248]
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.
GBP1, GBP2 The guanylate-binding proteins (GBP) are GTPases with antiviral properties [61, 249, 250]
that function by an unknown mechanism.

Schroder et al. Signals, mechanisms, and functions of IFN-␥ 171


TABLE 1. (Continued)

Antiproliferative Effect
Gene/protein up-regulated
by IFN-␥ Function Refs.
PKR PKR is an antiviral enzyme, which functions as a serine/threonine kinase when activated [154, 242–247]
by dsRNA. PKR inhibits cellular proliferation by phosphorylating the ␣ subunit of [251]
eIF-2, thereby halting protein synthesis. May also suppress c-myc function.
p21, p27 p21 and p27 are cyclin-dependent kinase (CDK) inhibitors of the Cip/Kip family. [252]
p21 and p27 inhibit the activity of CDK2 and CDK4, respectively, causing cell [253–256]
cycle arrest at the G1/S checkpoint.
Rb IFN-␥ inhibits the activity of the G1 cyclin:CDK complexes [via up-regulation of [257]
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 p202 is a strong cell cycle repressor that can bind to E2F and inactivate its [155, 258]
DNA-binding activity, thereby preventing transcription of E2F-dependent
genes required for S phase.
Mad1 Mad1 antagonizes c-myc function, thereby inhibiting proliferation. Mad1 [259–262]
competes with myc for max binding, forming the mad1:max heterodimer, which
has similar DNA sequence specificity 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.
Gene/protein down-
regulated by IFN-␥ Function Refs.
c-myc c-myc controls G1/S transition by activating cyclin:CDK complexes and inducing [263]
transcription of genes required for S phase. Expression of c-myc is induced by [264]
a number of growth factors and cytokines and is down-regulated by [154]
antiproliferative agents such as IFN-␥. IFN-␥ regulates c-myc expression by [251]
Rb-dependent down-regulation of E2F activity, as well as Rb-independent [265]
pathways that may be a result of PKR action. c-myc is active when associated [259]
with max. c-myc activity is antagonized by mad1.
Apoptotic Effect
Gene/protein up-regulated
by IFN-␥ Function Refs.
IRF-1 IRF-1 is a tumor-suppressor gene required for the induction of apoptosis by [165, 266, 267]
signals such as DNA damage. Many of the proapoptotic effects of IRF-1 are [267–269]
mediated by the IRF-1-induced caspase 1.
Caspase 1 (IL-1␤-converting Caspase-1 is a cysteine protease involved in the generation of bioactive IL-1␤ [267–269]
enzyme) and IL-18 and implicated in mediating macrophage apoptosis.
PKR PKR is an antiviral enzyme that appears to mediate TNF-␣-induced apoptosis. [270–272]
Mechanisms by which PKR affects apoptosis are still unclear but may involve
induction of Fas.
DAPs The death-associated proteins (DAP1–5) are mediators of IFN-␥-induced [273–276]
apoptosis by poorly defined mechanisms.
Cathepsin D Mediates IFN-␥, Fas/apolipoprotein, and IFN-␣-induced cell death by poorly [277]
defined mechanisms.
Fas/Fas ligand IFN-␥ may increase cellular sensitivity to apoptosis by up-regulating expression [278, 279]
of Fas and Fas ligand.
TNF-␣ receptor IFN-␥ may promote cellular sensitivity to the proapoptotic effects of TNF-␣ by [280]
promoting surface expression of a TNF-␣ receptor on tumor cells.
Activation of Microbial Effector Functions
Production of NO intermediates
Gene/protein up-regulated
by IFN-␥ Function Refs.
iNOS/NOS2 The NOS enzymes (NOS1, iNOS, NOS3) catalyze the reduced nictonimide [281]
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 Argininosuccinate synthetase produces the L-arginine substrate required for the [282–284]
NO-liberating reaction catalyzed by iNOS.
GTP-cyclohydroxylase I Guanosine 5⬘-triphosphate (GTP)-cyclohydroxylase I supplies the [285, 286]
tetrahydrobiopterin cofactor required for NO production by iNOS.

172 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


TABLE 1. (Continued)

Production of reactive oxygen species (ROS)


Gene/protein up-regulated
by IFN-␥ Function Refs.
phox phox phox
gp91 gp91 is a subunit of the NADPH oxidase, which associates with gp22 , [287, 288]
gp47phox, and gp67phox to form the active complex capable of the generation
of ROS during the respiratory burst.
gp67phox gp67phox is a subunit of the NADPH oxidase found in the cytosol in resting [289]
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.
Other antimicrobial mechanisms
Gene/protein up-regulated
by IFN-␥ Function Refs.
NRAMP1 The natural resistance-associated macrophage protein (NRAMP1) confers [4, 290]
resistance to macrophage intracellular pathogens by largely unknown
mechanisms.
FcR␥I Expression of the high-affinity Fc receptor (Fc␥RI) is increased in myeloid [291]
cells by IFN-␥ stimulation. FcR␥I binds extracellular pathogens via IgG in
the adaptive phase of the immune response.
C2, C4, Factor B Complement proteins are secreted by macrophages and fibroblasts in response [292–294]
to IFN-␥. Complement functions to opsonize extracellular pathogen for
receptor-mediated phagocytosis by mononuclear phagocytes.
Complement receptor CR3 Complement receptors of mononuclear phagocytes are up-regulated by IFN-␥ to [295]
(Mac-1) promote receptor-mediated phagocytosis of opsonized extracellular pathogens.
Immunomodulation, th Development, and Leukocyte Trafficking
Gene/protein up-regulated
by IFN-␥ Function Refs.
IL-12 NK cell activator and differentiation factor driving CD4⫹ cell development to a [296, 297]
Th1 phenotype.
IFN-inducible protein 10 Chemoattractant for monocytes and T cells. [155, 298]
(IP-10; CXCL10)
Monocyte chemoattractant Chemoattractant for monocyte/macrophages. [299]
protein-1 (MCP)-1/JE [155]
(CCL2)
MIG (CXCL9) Chemoattractant for T cells. [155, 300]
MIP-1␣, MIP-1␤ (CCL3, Chemoattractant for CD4⫹, CD8⫹, and memory T cells. [301]
CCL4) [155]
Regulated on activation, Chemoattractant for memory CD4⫹ T cells and monocyte/macrophages. [155, 302]
normal T expressed and
secreted (RANTES; CCL5)
ICAM-1 Adhesion molecule-binding to lymphocyte function-associated antigen-1 and [303]
Mac1.
Vascular cell adhesion Adhesion molecule-binding to very late antigen-4. [304]
molecule-1 (VCAM-1)
B7.2 Surface molecule on APCs that provide costimulus for antigen-specific T cell [305, 306]
activation.

sins B, H, and L, lysosomal proteases implicated in production IFN-␥ and development of Th1 response
of antigenic peptides for class II MHC loading [240, 241]; and
DM, a regulator of peptide accessibility to the peptide-binding IFN-␥ is a major product of Th1 cells and further skews the
cleft of class II MHC [235, 237]. immune response toward a Th1 phenotype. IFN-␥ achieves this
The CIITA mediates coordinate transcriptional control over by promoting characteristic Th1 effector mechanisms: innate
these genes. Targeted disruption of CIITA in mice causes cell-mediated immunity (via activation of NK cell effector
complete loss of constitutive or inducible display of class II functions), specific cytotoxic immunity (via T cell:APC inter-
MHC molecules [311]. Likewise, loss of constitutive/inducible actions), and macrophage activation (discussed below) [4].
class II MHC display is found in humans with CIITA mutation IFN-␥-induced, specific cytotoxic immunity is promoted by
(Bare Lymphocyte Syndrome) [312–314]. As CIITA is the direct and indirect mechanisms. IFN-␥ promotes specific cy-
limiting factor in a complex that directs transcriptional regu- totoxic immunity by indirect mechanisms, such as growth
lation, it acts as a switch for rapid up-regulation of class II inhibition of Th2 populations and up-regulation of antigen
MHC-related genes by IFN-␥ [315]. processing, presentation, and APC costimulatory molecules,

Schroder et al. Signals, mechanisms, and functions of IFN-␥ 173


thereby increasing CD4⫹ differentiation. IFN-␥ also influ- chinery. Phosphorylation by PKR prevents recycling of eIF-2–
ences naı̈ve CD4⫹ cell differentiation toward a Th1 phenotype GTP from its guanosine 5⬘-diphosphate-bound form, thereby
more directly. The phenotype adopted by a naive T cell during inhibiting viral and cellular protein synthesis [242]. PKR is
T cell activation is heavily influenced by the cytokine milieu implicated in numerous other functions, including activation of
present at the time of T cell receptor engagement. IFN-␥ and NF-␬B [243, 334], TNF-␣ transcript splice regulation [244],
IL-12 are the prototypic cytokines directing Th1 differentiation induction of fas-mediated apoptosis [245, 246], and regulation
during the primary response to antigen, and IL-4 directs dif- of Stat1 activity [154, 247].
ferentiation of Th2 populations. IFN-␥ induces IL-12 produc- It is likely that IFNs induce many as-yet undiscovered
tion in phagocytes [296] and inhibits IL-4 secretion by Th2 antiviral proteins, as the enzymes above and described in
populations [39], which may further drive Th1 differentiation Table 1 do not account for all IFN-induced antiviral effects
in vivo. [335, 336]. Also, the host of factors mediating the profound
IL-12 and IFN-␥ coordinate the link between pathogen growth-inhibitory and proapoptotic effects of IFN-␥ would have
recognition by innate immune cells and the induction of spe- no small part in limiting viral replication within the cell and
cific immunity, by mediating a positive feedback loop to am- spread to other cells.
plify the Th1 response [4]. LPS and other pathogen-associated
molecular patterns directly trigger IL-12 production upon rec-
Cell cycle, growth, and apoptosis
ognition by macrophages, DCs, and neutrophils [316 –319], Macrophages are produced in large amounts by the bone
which in turn induces IFN-␥ secretion in antigen-stimulated, marrow, and many die shortly after production by a process of
naive CD4⫹ T cells and NK cells [320, 321]. IL-12-induced programmed cell death (“apoptosis”) [337]. After release from
IFN-␥ participates in positive feedback by further promoting the bone marrow, macrophages proliferate, differentiate, ac-
IL-12 production in macrophages [296, 297]. This amplifica- quire specialized functions (e.g., microbicidal activity), or die
tion may be important in initiation or stabilization of the Th1 by apoptosis depending on the presence or absence of extra-
response (reviewed in refs. [4, 322]). IL-18 is highly synergistic cellular cues. Many reports indicate that IFN-␥ arrests the
with IL-12 for IFN-␥ production and has important functions in macrophage cell cycle and provides a survival signal, and
the in vivo Th response [23, 24]. others suggest that it serves as a proapoptotic signal. Although
Secretion of large amounts of IFN-␥ or IL-4 is the defining recent advances in cell cycle control and apoptosis have
feature of Th1 or Th2 cells, respectively [323]. These cytokines greatly increased our understanding in these functions in iso-
function to directly promote cell-mediated immunity (IFN-␥) or lation, the inter-relationship between these intimately related
humoral immunity (IL-4) and to reciprocally antagonize each processes is still somewhat confusing. For this reason, the
other’s actions, further polarizing the response (reviewed in ref. effects of IFN-␥ on cellular proliferation and apoptosis are
[108]). For example, IL-4 inhibits IFN-␥-dependent activation described here separately.
of macrophage-effector functions [324, 325]. Conversely, One of the most easily observed effects of IFN-␥ is cell
IFN-␥ antagonizes IL-4-dependent induction of the IgE recep- growth inhibition. Types I and II IFN are able to protect against
tor, FcεRII [326]. pathogen-induced apoptosis and suppress colony stimulating
factor type 1 (CSF-1)-dependent growth of bone marrow-de-
The IFN-induced antiviral state rived macrophages (BMM) [252]. Granulocyte macrophage-
The IFN system regulates innate and adaptive immunity to CSF, which is usually released by T cells at the same time as
viral infection. Viral invasion directly triggers induction of type IFN-␥, is able to protect against the growth-inhibitory effects of
I IFNs, through a mechanism involving IRF-3 and IRF-7 IFN-␥ on macrophages, and this may be physiologically im-
(reviewed in ref. [327]). Although both types of IFN are crucial portant in vivo [338]. The IFNs most commonly arrest the cell
in the immediate cellular response to viral infection, the im- cycle at the G1/S checkpoint, although blockages of other cell
munomodulatory activities of IFN-␥ (discussed elsewhere) be- cycle stages have been reported [339 –341]. It has been sug-
come important later in the response in coordinating the im- gested that IFN-induced G1 arrest may actually reflect exit
mune response and establishing an antiviral state for longer from the cell cycle into G0 [342].
term control [54, 60, 328, 329]. IFNs inhibit proliferation primarily by increasing protein
IFN-␥-induced antiviral mechanisms include induction key levels of Ink4 and Cip/Kip CKIs [252–255, 343, 344]. IFN-␥
antiviral enzymes (Table 1), most notably PKR, which is a transcriptionally induces p21 and p27 CKIs (Table 1) [256,
serine/threonine kinase greatly induced by types I and II IFN 345, 346]. p21 and p27 inhibit the activity of cyclin E:CDK2
stimulation [242, 330]. PKR is inactive in its constitutive form and cyclin D:CDK4 complexes, respectively, thereby arresting
and requires an activating signal for autophosphorylation. the cell cycle at the G1/S boundary. Decreased cyclin:CDK
dsRNA, a necessary intermediate in replication of RNA vi- activity during G1 results in hypophosphorylation of the tumor
ruses, is the best characterized activator of PKR, although suppressor Rb. In its hypophosphorylated state, Rb sequesters
other agents such as heparin are able to activate PKR [331]. the E2F family of transcription factors from activation of genes
Association of PKR with dsRNA is likely to cause a confor- (e.g., c-myc) required for cell cycle progression from G1 to S
mational change that unmasks the catalytic domain responsible phase [292]. c-myc controls G1/S transition by activating cy-
for PKR autophosphorylation [242, 332]. PKR is then activated clin:CDK complexes and inducing transcription of genes re-
for dsRNA-independent phosphorylation of specific cellular quired for S phase (reviewed in ref. [263]). Expression of c-myc
substrates [333]. One of these substrates is the eIF-2␣ subunit, is induced rapidly by a number of growth factors and cytokines
a rate-limiting factor in the normal cellular translational ma- and is down-regulated by IFN-␥ [154, 264]. IFN-␥ influences

174 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


c-myc expression through multiple pathways, including sup- tion, tryptophan depletion, and up-regulation of lysosomal en-
pression of Rb phosphorylation, resulting in decreased E2F zymes promoting microbe destruction (Table 1) [222]. Similar
activity as well as Rb-independent pathways, which may be a microbicidal mechanisms are activated by IFN-␥ in neutro-
result of PKR action [251]. In its active form, c-myc is asso- phils [4].
ciated with max, and the myc:max heterodimer activates tran- Macrophages kill bacteria, viruses, protozoa, helminths,
scription of genes required for cell cycle progression [265]. The fungi, and tumor cells primarily by production of ROS and
level of active myc is negatively regulated by mad1, which reactive nitrogen intermediates (RNI) via induction of the
sequesters the max coactivator away from myc and suppresses NADPH oxidase system and iNOS, respectively (reviewed in
transcription of myc-inducible genes [259]. In this way, the refs. [281, 355, 356]). ROS and RNI use the advantages of low
mad1-to-c-myc ratio determines whether a cell will proliferate molecular weight, reactivity, and lipophilicity to easily pene-
(as a result of high levels of myc:max) or arrest in G1 (as a trate the microbial cell wall/coat to inflict injury. The impor-
result of high levels of mad1:max) [259, 260]. In addition to tance of these molecules in host defense is highlighted by the
decreasing the abundance of myc, IFN-␥ increases mad1 lev- highly pathogen-susceptible phenotype of mice doubly defi-
els, thereby further antagonizing myc activity and inhibiting cient in NADPH oxidase and iNOS enzymes [357]. The time-
CSF-1-dependent proliferation in BMM [347]. frames and situations in which macrophage use these cytocidal
Apoptosis is a process in which the cell directs its own activities are generally different. ROS production becomes
demise by activating intrinsic suicide machinery (reviewed in apparent within 1 h after stimulus, whereas RNI generation
refs. [337, 348]). In macrophages, apoptosis may be desirable requires ⬃24 h after encountering a pathogen product. In
to prevent colonization by intracellular pathogens. Conversely, general, ROS are used to target extracellular pathogens during
many pathogens actually induce host-cell death through secre- phagocytosis or that are too large for phagocytosis, whereas
tion of macrophage-specific toxins. Induction of apoptosis by RNI target intracellular pathogens and upon appropriate stim-
signals such as DNA damage requires the IRF-1 tumor-sup- ulus, extracellular pathogens and tumor cells.
pressor gene [165, 266, 267]. Levels of IRF-1 may be a NO is produced in the NADPH-dependent conversion of
deciding factor in whether IFN-␥ induces or protects from L-arginine to L-citrulline by the NOS enzymes (NOS1–3). The
apoptosis on treated cells [35, 349, 350]. It is proposed that NOS2/iNOS isoform alone is inducible by cytokine and/or
IFN-␥ treatment of cells with high levels of functional IFNGR microbial stimulus [281]. The NO generated by this enzyme
very rapidly activates Stat1, thereby producing high levels of exists in equilibrium among a number of different redox states
IRF-1 that are able to induce apoptosis. In contrast, IFN-␥ (called RNI in this review) that have differing biological ac-
treatment of cells with low levels of functional IFNGR may tivities. It should be noted that much of the work presented
activate Stat1 more slowly, thereby producing lower levels of below applies to mice rather than humans, and there is at
IRF-1 that are not sufficient to induce apoptosis [35]. Exper- present much controversy of the physiological relevance of NO
iments in which overexpression of functional IFNGR on nor- production in human host defense (reviewed in ref. [281]).
mally low-level IFNGR-expressing cells changed the IFN-␥ A wealth of evidence supports a crucial function of iNOS
response of these cells from an antiapoptotic/proliferative phe- and RNI in host defense (reviewed in ref. [281]). IFN-␥-
notype to a proapoptotic phenotype are consistent with this dependent RNI production is associated with increased ability
hypothesis [35]. This may also explain why myeloid cells are of phagocytic cells to kill ingested pathogens, and this ability
more sensitive to the proapoptotic actions of IFN-␥ than other is inhibited by the addition of a substrate analog inhibitor of
cells such as T cells, as myeloid cells express relatively high iNOS such as N-methyl-L-arginine [358 –361]. Furthermore,
numbers of functional IFNGR on the cell surface [35]. mice in which the iNOS gene has been mutated show greater
Many of the proapoptotic effects of IRF-1 are mediated by susceptibility to parasitic infection and a dampened, nonspe-
the IRF-1-induced caspase 1 (IL-1␤-converting enzyme) [267– cific inflammatory response [362]. Production of RNI also
269]. Caspase 1 is a cysteine protease implicated in mediating inhibits replication of a range of viruses, through ill-defined
macrophage apoptosis by various stimuli including LPS [351– mechanisms [358, 363, 364].
353] and is involved in generation of bioactive IL-1␤ and IL-18 IFN-␥ induces RNI production by up-regulating expression
[354]. Caspase 1 expression and resulting apoptosis can be of substrate, cofactor, and catalyst required for NO generation.
IFN-␥-inducible [269] or can occur through IFN-␥-indepen- IFN-␥ up-regulates argininosuccinate synthetase (which pro-
dent IRF-1 activity [267]. IFN-␥ also induces a number of duces the L-arginine substrate), GTP-cyclohydroxylase I
other proapoptotic molecules. These include PKR, the DAPs, (which supplies the tetrahydrobiopterin cofactor required for
cathepsin D, and surface expression of Fas and the TNF-␣ NO production), and the iNOS enzyme [285, 286, 365–367].
receptor (Table 1). Maximal induction of iNOS transcription requires “priming”
and “triggering” stimuli such as priming with IFN-␥ and
Activation of microbicidal effector functions subsequent triggering with LPS or TNF-␣ [368, 369]. The type
One of the most important effects of IFN-␥ on macrophages is I IFNs cannot fully substitute for IFN-␥ in triggering NO
the activation of microbicidal effector functions. Macrophages production [368].
activated by IFN-␥ display increased pinocytosis and receptor- Superoxide and its reactive products are also important toxic
mediated phagocytosis as well as enhanced microbial killing effector molecules of the nonspecific cytotoxic response. The
ability. IFN-␥-activated microbicidal ability includes induc- superoxide anion, O2–, is generated by the multicomponent
tion of the NADPH-dependent phagocyte oxidase (NADPH flavocytochrome enzyme phagocyte oxidase during a process
oxidase) system (“respiratory burst”), priming for NO produc- 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, concen- chemokine milieu to extravasate into the tissue via interactions
trated in the phagosome membrane, and two cytosolic compo- between adhesion molecules presented on leukocyte and en-
nents, p47phox and p67phox. Upon appropriate stimulus (e.g., dothelial surfaces (“diapedesis”) [4]. IFN-␥ regulates this pro-
phagocytosis), the cytosolic components translocate to the cess by up-regulating expression of chemokines (e.g., IP-10,
membrane to form the active complex, which generates super- MCP-1, MIG, MIP-1␣/␤, RANTES; see Table 1) and adhesion
oxide in the phagosome through transfer of a transported elec- molecules (e.g., ICAM-1, VCAM-1; see Table 1). TNF-␣ and
tron to molecular oxygen. The superoxide anion generated by IL-1␤ synergistically regulate many of these molecules [4].
the respiratory burst spontaneously reacts to form hydrogen
peroxide (H2O2), hydroxyl radicals (·OH), and hypochlorous
IFN-␥ priming of the macrophage LPS response
acid (HOCl) [281]. The toxic oxidants produced by the respi- LPS/endotoxin is a cell wall constituent of gram-negative bac-
ratory burst are also able to react with those produced by iNOS, teria that activates macrophage microbicidal effector functions
thereby forming a large number of different toxic species (e.g., and the production of proinflammatory cytokines (e.g., TNF-␣,
peroxynitrite) to mediate cytotoxicity by a wide variety of IL-1, IL-6). LPS:macrophage interaction can be beneficial or
mechanisms [281, 370, 371]. deleterious to the host. Macrophages sense and are activated to
The primary mechanism of IFN-␥-induced up-regulation of fight and clear bacterial infection through LPS recognition, a
ROS production in phagocytes is transcriptional induction of mechanism obviously advantageous to the host; however, high
the gp91phox and p67phox subunits of the NADPH oxidase doses of LPS that trigger excessive production of inflammatory
complex [287–289, 372]. The biological importance of this mediators can result in the potentially lethal systemic disorder,
pathway in host defense is emphasized by the human genetic septic shock.
disease, chronic granulomatous disease, caused by defects in Macrophages recognition of lipid A (the active moiety of
the NADPH oxidase system and characterized by recurrent and LPS), requires Toll-like receptor (TLR) family member TLR4.
often fatal infections. The TLR family is a homolog of the Drosophila antifungal
IFN-␥ also promotes microbe destruction by augmenting protein Toll. To date, 10 mammalian TLRs have been identi-
surface expression of the high-affinity Fc␥RI on mononuclear fied (TLR1–10), and many have proposed or demonstrated
phagocytes, thereby promoting antibody-dependent, cell-medi- functions in innate immunity. LPS:TLR4 interaction triggers a
ated cytotoxicity [291]. Complement-mediated phagocytosis is signaling cascade (Fig. 3), which ultimately results in activa-
also up-regulated by IFN-␥ through increased complement tion of NF-␬B and activated protein-1 (AP-1) and transcrip-
secretion and complement receptor surface expression on tional control over genes directing immune function in macro-
mononuclear phagocytes [292]. phages (reviewed in refs. [383–386]). MyD88 functions as an
Tryptophan depletion from cells in response to IFN-␥ may adaptor to connect the intracellular portion of TLR4 to down-
be an antiparasitic mechanism in humans [373]. IFN-␥ up- stream signaling components and is required for many, but not
regulates the expression of indoeamine 2,3 dioxygenase, an all, LPS-induced effects [387]. For this reason, a TLR4 signal-
enzyme responsible for the generation of N-formylkynurenine ing paradigm has emerged in which LPS signals through
from tryptophan [374 –377]. It is interesting that IFN-␥ also MyD88-dependent and MyD88-independent pathways. MyD88
induces expression of tryptophanyl-tRNA synthetase, a house- gene KO mice demonstrated that the MyD88-dependent path-
keeping gene that catalyzes tRNA Trp aminoacetylation during way is necessary for rapid activation of MAPK, AP-1, and
protein synthesis [378, 379]. It is proposed that this enzyme NF-␬B, the production of inflammatory cytokines, and septic
may also aid in tryptophan depletion by incorporating free shock [387]. The MyD88-independent pathway results in
tryptophan into translating protein and may ensure synthesis of slower activation of MAPK, AP-1, and NF-␬B and does not
Trp-rich proteins with immunologically important functions contribute to the same extent to the inflammatory response.
(e.g., the IRF family and many MHC molecules) during cellular IFN-␥ primes macrophages for more rapid and heightened
tryptophan depletion [380]. responses to LPS [58, 388, 389] as well as other TLR agonists,
such as unmethylated CpG motifs present in bacterial DNA
Immunomodulation and leukocyte trafficking (CpG DNA) [390]. Pretreatment with IFN-␥ is necessary for the
The ability of IFN-␥ to coordinate the transition from innate induction of some genes in response to LPS. iNOS is the
immunity to adaptive immunity distinguishes it from the other archetypal example of such a gene, although the need for
IFNs. Mechanisms by which IFN-␥ coordinates this transition exogenous priming by IFN-␥ can be overcome by adequate
include aiding in the development of a Th1-type response production of endogenous type I IFN in response to high doses
(described earlier), directly promoting B cell isotype switching of LPS [391]. For other genes, IFN-␥ is able to shift the
to IgG2a [54, 381, 382], and regulation of local leukocyte- dose-response curve such that transcription is induced at lower
endothelial interactions (Table 1). concentrations of LPS, thereby superinducing transcription in
IFN-␥ orchestrates the trafficking of specific immune cells to IFN-␥-primed cells treated with LPS compared with the same
sites of inflammation through up-regulating expression of ad- concentration of LPS alone [392]. IFN-␥ priming does not
hesion molecules and chemokines. Unstimulated leukocytes occur for all LPS-induced responses; IFN-␥ appears to prime
cycle continuously between the blood and the lymph. IFN-␥ for a select subset of LPS-induced genes. For example, IFN-␥
and NO produced at the site of inflammation cause local pretreatment promotes LPS-induced TNF-␣ but not procoagu-
dilation of the blood vessels, thereby decreasing the local blood lant activity enzyme production [393].
flow rate and causing gathering of blood in leaky vessels. IFN-␥ receptor KO mice are highly resistant to LPS-induced
Specific leukocyte subsets are instructed by the cytokine/ toxicity [394]. This highlights the physiological significance of

176 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


LBP MD-2
LPS

1
CD14 TLR4

Cytosol

MyD88
MyD88-dependent
pathway
IRAK 2 TRIF? MyD88-independent
pathway
TRAF6 IRF-3
NIK 3
IKK p38 MAPK
JNK/SAPK
ERK

Rapid activation of Slow activation of


NFκB and AP-1 NFκB and AP-1

4 Nucleus
5
NF-κB AP-1 Transcription of IFNα/β

Transcription of target genes Transcription of target genes


eg. IL-1β, IL-6, TNFα eg. GARG16, IRG-1, IP-10

6 6

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 I␬B 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-defined
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 proinflammatory 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 figure): (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 nor- inhibits the LPS-dependent down-regulation of TLR4 surface
mally produced during the response to LPS and acts to amplify expression apparent in unprimed cells [395]. IFN-␥ stimula-
LPS-induced cellular responses. The priming phenomenon re- tion may also promote LPS-dependent signaling by promoting
lies on IFN-␥ and LPS reciprocal cross-regulation of signaling expression of the MD-2 accessory molecule, the MyD88 adap-
molecules of the other pathway. tor, and the IRAK signaling molecule [395, 398].
IFN-␥ influences LPS-dependent signaling capabilities by LPS mediates many of its effects through the NF-␬B tran-
promoting ligand-receptor interactions as well as downstream scription factor. In the macrophage-like cell line RAW264.7, it
signaling machinery. Efficient LPS:TLR4 interaction requires was found that IFN-␥ pretreatment promoted NF-␬B activation
the MD-2 and CD14 accessory molecules, and subsequent upon LPS exposure, as well as more rapid DNA-binding ki-
maximal signaling requires MyD88 signaling adaptor. In a netics and faster degradation of the NF-␬B inhibitor, I␬B-␣
number of experiments, IFN-␥ was shown to promote transcrip- [399]. In human monocytes also, IFN-␥ pretreatment causes
tion of TLR4, subsequent TLR4 surface expression, and LPS- superinduction of active NF-␬B upon LPS treatment [400].
binding ability in macrophages [395–397]. Furthermore, IFN-␥ 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-␥ 416]. Stat1 is tyrosine-phosphorylated as a result of autocrine/
priming caused an increase in p65 mRNA as a result of paracrine type I IFN and is required for the full LPS response
increased transcript stability. in IFN-␥-unprimed macrophages [416]. LPS-induced Stat1
Many IFN-␥-inducible genes are also TNF-␣-inducible, and Y701 phosphorylation was found to be essential for iNOS
these genes are often superinduced by the combination of these expression [413], and macrophages from mice unable to re-
factors [366, 372, 401, 402]. TNF-␣ is a macrophage-derived spond to type I IFN through targeted disruption of the IFNAR1
cytokine secreted in response to LPS, which can act in an receptor subunit exhibited limited capacity to induce iNOS in
autocrine manner to mediate many LPS-induced effects via response to LPS [391]. iNOS-producing capacity in response to
NF-␬B [403]. IFN-␥ priming of TNF-␣ responses is thus LPS was able to be rescued by exposure to IFN-␥, which
responsible for the priming of a subset of LPS-induced genes. presumably restored Stat1 activation [391]. LPS induces tran-
In many cases, synergistic induction by the combination of scription of type I IFN through an ill-defined, MyD88-inde-
these factors may be a result of the combined presence of Stat1 pendent pathway involving IRF-3 [417]. The contribution of
and NF-␬B-binding sites in the promoter elements of respon- type I IFN to IFN-␥ priming of the LPS response has not been
sive genes [160, 404]. Synergy may be also be a result of formally studied; however, as IFN-␥ up-regulates expression of
cross-talk between the IFN-␥ and TNF-␣ signaling pathways: IFNAR in other cell types (P. J. Hertzog, unpublished obser-
TNF-␣ stimulation is able to induce transcription of IRF-1 and vations) and promotes IFN-␣/␤ expression, it is possible that
promotes IFN-␥-induced Stat1 activation [405, 406]; IFN-␥- IFN-␥ sensitizes the cells to LPS-induced IFN-␣/␤. When the
induced PKR is able to signal to NF-␬B [407]; and IFN-␥ many levels of cross-talk between the IFN-␥ and IFN-␣/␤
increases surface expression of the TNF-␣ receptor [280, 408, pathway are considered, it is likely that in vivo IFN-␥ poten-
409], although the significance of this is uncertain, as an tiates LPS-induced IFN-␣/␤ effects and conversely, LPS aug-
increased receptor expression level has not been linked with ments IFN-␥-mediated effects (e.g., by Stat1 Y701 and S727
increased response to TNF-␣ [410, 411]. phosphorylation).
Although probably not complete, these mechanisms of in- Synergy between LPS- and IFN-␥-induced transcription fac-
creasing cellular sensitivity to LPS may start to explain the tors in expression of target genes also contributes to the prim-
shift in the LPS dose-response curve seen with IFN-␥ priming. ing phenotype. Genes such as IRF-1, IP-10, ICAM-1, and
Genes that need IFN-␥ and LPS treatment for induction may iNOS contain Stat1 and NF-␬B-binding sites in their promoter,
require a certain threshold of LPS signaling for transactivation, and maximal transcription requires both signals [158, 160,
and so IFN-␥ treatment shifts the LPS dose-response curve by 402, 404, 418 – 423]. This is a demonstrated mechanism in
increasing cellular sensitivity. Alternatively, these genes may IFN-␥-dependent gene superinduction in response to LPS for a
require specific cross-talk between pathways before transcrip- number of genes and is likely to be a global mechanism for
tional activation can occur. Examples of such cross-talk may synergistic, coordinate regulation of large synexpression
include LPS-induced Stat1 serine phosphorylation or cross- groups. Genes that require IFN-␥ and LPS treatment for in-
talk between IFN-␥- and LPS-induced type I IFN pathways. duction (e.g., iNOS) may also be explained by this mechanism,
LPS is able to promote the IFN-␥ signaling pathway through as it is likely that the IFN-␥- and LPS-induced transcription
Stat1. As noted earlier, phosphorylation of S727 of Stat1 occurs factors are limiting for gene induction at a single-cell level.
in response to IFN-␥ or LPS exposure and is necessary for This theory is consistent with the all-or-nothing transcription of
maximal Stat1 activation. LPS treatment following IFN-␥ ex- iNOS and many other IFN-␥/LPS-regulated genes observed at
posure increases the percentage of molecules phosphorylated a single-cell level (ref. [424] and references therein).
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-␥ CONCLUDING REMARKS
and LPS may use different pathways for Stat1 serine phosphor-
ylation, as LPS-induced Stat1 serine phosphorylation is sensi- IFN-␥ is a remarkable cytokine that orchestrates many distinct
tive to a p38 MAPK inhibitor, whereas the serine phosphory- cellular programs through transcriptional control over large
lation induced by IFN-␥ is not [110]. Another study suggested numbers of genes. Many IFN-␥-induced effects resulting in
that although p38 MAPK is necessary, it is not sufficient for heightened immune surveillance and immune system function
Stat1 serine phosphorylation by LPS, and protein kinase C-␦ is during infection have been discussed in this review. As patho-
also required to mediate phosphorylation [412]. In addition to gen products such as LPS and CpG DNA augment local IFN-␥
promoting serine phosphorylation, LPS treatment following production, and IFN-␥ augments the immune system response
IFN-␥ exposure augments Stat1 tyrosine phosphorylation of to these agonists, an important function of IFN-␥ during in vivo
residue 701 [399]. Macrophage exposure to LPS triggers pro- infection is suggested, whereby IFN-␥ participates in an am-
duction of endogenous type I IFN that can act on the macro- plification loop to increase immune system sensitivity and
phage population in an autocrine/paracrine manner. This has response to pathogens. This model is supported by the in vivo
been identified as the causal agent of LPS-induced Stat1 and in vitro priming phenotype apparent when LPS-stimulated
tyrosine phosphorylation [413]. macrophages are pretreated with IFN-␥, and also by the high
It has long been recognized that LPS induces transcription of resistance of IFNGR⫺/⫺ mice to LPS-induced septic shock.
type I IFN [414], but only in recent years has the significance The ability of IFN-␥ to synergize or antagonize the effects of
of autocrine/paracrine functions of type I IFN in mediating cytokines, growth factors, and pathogen-associated molecular
LPS-induced gene expression been acknowledged [391, 415, pattern (PAMP)-signaling pathways (e.g., TNF-␣, IL-4, CSF-1,

178 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


IFN-␣/␤, LPS, and CpG DNA) is particularly important in regulation of polarized cytokine production by effector B and T cells.
Nat. Immunol. 1, 475– 482.
macrophage biology, as macrophages constantly receive mul- 17. Sen, G. C. (2001) Viruses and interferons. Annu. Rev. Microbiol. 55,
tiple signals and need to integrate them to give a response 255–281.
appropriate to the extracellular milieu. The reductionist ap- 18. Golab, J., Zagozdzon, Stoklosal, T., Kaminski, R., Kozar, K., Jakobisiak,
proach used by many scientists has greatly furthered our M. (2000) Direct stimulation of macrophages by IL-12 and IL-18 –a
bridge too far? Immunol. Lett. 72, 153–157.
understanding of mechanisms by which IFN-␥ coordinates its 19. Munder, M., Mallo, M., Eichmann, K., Modolell, M. (2001) Direct
pleiotropic effects. However, a cell is not exposed to one stimulation of macrophages by IL-12 and IL-18 —a bridge built on solid
stimulus in isolation in vivo, and so, one of the great challenges ground. Immunol. Lett. 75, 159 –160.
20. Munder, M., Mallo, M., Eichmann, K., Modolell, M. (1998) Murine
to IFN biologists today is to understand the mechanisms by macrophages secrete interferon gamma upon combined stimulation with
which IFN-␥ signaling and cellular effects integrate with other interleukin (IL)-12 and IL-18: a novel pathway of autocrine macrophage
growth factor, cytokine, and PAMP signaling pathways. This activation. J. Exp. Med. 187, 2103–2108.
21. Fukao, T., Matsuda, S., Koyasu, S. (2000) Synergistic effects of IL-4 and
understanding will enable a more realistic view of in vivo IL-18 on IL-12-dependent IFN-gamma production by dendritic cells.
macrophage function during naturally occurring human infec- J. Immunol. 164, 64 –71.
tion. 22. Otani, T., Nakamura, S., Toki, M., Motoda, R., Kurimoto, M., Orita, K.
(1999) Identification of IFN-gamma-producing cells in IL-12/IL-18-
treated mice. Cell. Immunol. 198, 111–119.
23. Akira, S. (2000) The role of IL-18 in innate immunity. Curr. Opin.
Immunol. 12, 59 – 63.
ACKNOWLEDGMENT 24. Dinarello, C. A. (1999) IL-18: a TH1-inducing, proinflammatory cytokine
and new member of the IL-1 family. J. Allergy Clin. Immunol. 103,
The authors thank Katharine Irvine for critical reading of the 11–24.
25. Salazar-Mather, T. P., Hamilton, T. A., Biron, C. A. (2000) A chemokine-
manuscript and helpful suggestions. to-cytokine-to-chemokine cascade critical in antiviral defense. J. Clin.
Invest. 105, 985–993.
26. Pien, G. C., Satoskar, A. R., Takeda, K., Akira, S., Biron, C. A. (2000)
Cutting edge: selective IL-18 requirements for induction of compartmen-
REFERENCES tal IFN-gamma responses during viral infection. J. Immunol. 165,
4787– 4791.
27. Schindler, H., Lutz, M. B., Rollinghoff, M., Bogdan, C. (2001) The
1. Adams, D. O. (1989) Molecular interactions in macrophage activation. production of IFN-gamma by IL-12/IL-18-activated macrophages re-
Immunol. Today 10, 33–35. quires STAT4 signaling and is inhibited by IL-4. J. Immunol. 166,
2. Perussia, B., Dayton, E. T., Fanning, V., Thiagarajan, P., Hoxie, J., 3075–3082.
Trinchieri, G. (1983) Immune interferon and leukocyte-conditioned me- 28. Fukao, T., Frucht, D. M., Yap, G., Gadina, M., O’Shea, J. J., Koyasu, S.
dium induce normal and leukemic myeloid cells to differentiate along the (2001) Inducible expression of Stat4 in dendritic cells and macrophages
monocytic pathway. J. Exp. Med. 158, 2058 –2080. and its critical role in innate and adaptive immune responses. J. Immu-
3. Young, H. A., Hardy, K. J. (1995) Role of interferon-gamma in immune nol. 166, 4446 – 4455.
cell regulation. J. Leukoc. Biol. 58, 373–381. 29. Hochrein, H., Shortman, K., Vremec, D., Scott, B., Hertzog, P., O’Keeffe,
4. Boehm, U., Klamp, T., Groot, M., Howard, J. C. (1997) Cellular re- M. (2001) Differential production of IL-12, IFN-alpha, and IFN-gamma
sponses to interferon-gamma. Annu. Rev. Immunol. 15, 749 –795. by mouse dendritic cell subsets. J. Immunol. 166, 5448 –5455.
5. Carnaud, C., Lee, D., Donnars, O., Park, S. H., Beavis, A., Koezuka, Y., 30. Major, A. S., Cuff, C. F. (1996) Effects of the route of infection on
Bendelac, A. (1999) Cutting edge: cross-talk between cells of the innate immunoglobulin G subclasses and specificity of the reovirus-specific
immune system: NKT cells rapidly activate NK cells. J. Immunol. 163, humoral immune response. J. Virol. 70, 5968 –5974.
4647– 4650. 31. Autenrieth, I. B., Reissbrodt, R., Saken, E., Berner, R., Vogel, U.,
6. Finkelman, F. D., Katona, I. M., Mosmann, T. R., Coffman, R. L. (1988)
Rabsch, W., Heesemann, J. (1994) Desferrioxamine-promoted virulence
IFN-gamma regulates the isotypes of Ig secreted during in vivo humoral
of Yersinia enterocolitica in mice depends on both desferrioxamine type
immune responses. J. Immunol. 140, 1022–1027.
and mouse strain. J. Infect. Dis. 169, 562–567.
7. Isaacs, A., Lindermann, J. (1957) Virus interference. I. The interferon.
32. Autenrieth, I. B., Beer, M., Bohn, E., Kaufmann, S. H., Heesemann, J.
Proc. R. Soc. Lond. B Biol. Sci. 147, 258 –267.
(1994) Immune responses to Yersinia enterocolitica in susceptible
8. Bach, E. A., Aguet, M., Schreiber, R. D. (1997) The IFN gamma receptor:
a paradigm for cytokine receptor signaling. Annu. Rev. Immunol. 15, BALB/c and resistant C57BL/6 mice: an essential role for gamma inter-
563–591. feron. Infect. Immun. 62, 2590 –2599.
9. Jonasch, E., Haluska, F. G. (2001) Interferon in oncological practice: 33. Bazan, J. F. (1990) Structural design and molecular evolution of a
review of interferon biology, clinical applications, and toxicities. Oncol- cytokine receptor superfamily. Proc. Natl. Acad. Sci. USA 87, 6934 –
ogist 6, 34 –55. 6938.
10. Bazer, F. W., Spencer, T. E., Ott, T. L. (1997) Interferon tau: a novel 34. Thoreau, E., Petridou, B., Kelly, P. A., Djiane, J., Mornon, J. P. (1991)
pregnancy recognition signal. Am. J. Reprod. Immunol. 37, 412– 420. Structural symmetry of the extracellular domain of the cytokine/growth
11. Young, H. A. (1996) Regulation of interferon-gamma gene expression. hormone/prolactin receptor family and interferon receptors revealed by
J. Interferon Cytokine Res. 16, 563–568. hydrophobic cluster analysis. FEBS Lett. 282, 26 –31.
12. Frucht, D. M., Fukao, T., Bogdan, C., Schindler, H., O’Shea, J. J., 35. Bernabei, P., Coccia, E. M., Rigamonti, L., Bosticardo, M., Forni, G.,
Koyasu, S. (2001) IFN-gamma production by antigen-presenting cells: Pestka, S., Krause, C. D., Battistini, A., Novelli, F. (2001) Interferon-
mechanisms emerge. Trends Immunol. 22, 556 –560. gamma receptor 2 expression as the deciding factor in human T, B, and
13. Gessani, S., Belardelli, F. (1998) IFN-gamma expression in macrophages myeloid cell proliferation or death. J. Leukoc. Biol. 70, 950 –960.
and its possible biological significance. Cytokine Growth Factor Rev. 9, 36. Pernis, A., Gupta, S., Gollob, K. J., Garfein, E., Coffman, R. L., Schind-
117–123. ler, C., Rothman, P. (1995) Lack of interferon gamma receptor beta chain
14. Yoshimoto, T., Takeda, K., Tanaka, T., Ohkusu, K., Kashiwamura, S., and the prevention of interferon gamma signaling in TH1 cells. Science
Okamura, H., Akira, S., Nakanishi, K. (1998) IL-12 up-regulates IL-18 269, 245–247.
receptor expression on T cells, Th1 cells, and B cells: synergism with 37. Bach, E. A., Szabo, S. J., Dighe, A. S., Ashkenazi, A., Aguet, M., Murphy,
IL-18 for IFN-gamma production. J. Immunol. 161, 3400 –3407. K. M., Schreiber, R. D. (1995) Ligand-induced autoregulation of IFN-
15. Flaishon, L., Hershkoviz, R., Lantner, F., Lider, O., Alon, R., Levo, Y., gamma receptor beta chain expression in T helper cell subsets. Science
Flavell, R. A., Shachar, I. (2000) Autocrine secretion of interferon 270, 1215–1218.
gamma negatively regulates homing of immature B cells. J. Exp. Med. 38. Maggi, E., Parronchi, P., Manetti, R., Simonelli, C., Piccinni, M. P.,
192, 1381–1388. Rugiu, F. S., De Carli, M., Ricci, M., Romagnani, S. (1992) Reciprocal
16. Harris, D. P., Haynes, L., Sayles, P. C., Duso, D. K., Eaton, S. M., Lepak, regulatory effects of IFN-gamma and IL-4 on the in vitro development of
N. M., Johnson, L. L., Swain, S. L., Lund, F. E. (2000) Reciprocal human Th1 and Th2 clones. J. Immunol. 148, 2142–2147.

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


39. Gajewski, T. F., Fitch, F. W. (1988) Anti-proliferative effect of IFN- 59. Suzuki, Y., Orellana, M. A., Schreiber, R. D., Remington, J. S. (1988)
gamma in immune regulation. I. IFN-gamma inhibits the proliferation of Interferon-gamma: the major mediator of resistance against Toxoplasma
Th2 but not Th1 murine helper T lymphocyte clones. J. Immunol. 140, gondii. Science 240, 516 –518.
4245– 4252. 60. Muller, U., Steinhoff, U., Reis, L. F., Hemmi, S., Pavlovic, J., Zinker-
40. Kaplan, D. H., Greenlund, A. C., Tanner, J. W., Shaw, A. S., Schreiber, nagel, R. M., Aguet, M. (1994) Functional role of type I and type II
R. D. (1996) Identification of an interferon-gamma receptor alpha chain interferons in antiviral defense. Science 264, 1918 –1921.
sequence required for JAK-1 binding. J. Biol. Chem. 271, 9 –12. 61. Kimura, T., Kadokawa, Y., Harada, H., Matsumoto, M., Sato, M., Kashi-
41. Greenlund, A. C., Farrar, M. A., Viviano, B. L., Schreiber, R. D. (1994) wazaki, Y., Tarutani, M., Tan, R. S., Takasugi, T., Matsuyama, T., Mak,
Ligand-induced IFN gamma receptor tyrosine phosphorylation couples T. W., Noguchi, S., Taniguchi, T. (1996) Essential and non-redundant
the receptor to its signal transduction system (p91). EMBO J. 13, roles of p48 (ISGF3 gamma) and IRF-1 in both type I and type II
1591–1600. interferon responses, as revealed by gene targeting studies. Genes Cells
42. Farrar, M. A., Fernandez-Luna, J., Schreiber, R. D. (1991) Identification 1, 115–124.
of two regions within the cytoplasmic domain of the human interferon- 62. Biron, C. A. (2001) Interferons alpha and beta as immune regulators–a
gamma receptor required for function. J. Biol. Chem. 266, 19626 – new look. Immunity 14, 661– 664.
19635. 63. Doffinger, R., Jouanguy, E., Dupuis, S., Fondaneche, M. C., Stephan,
43. Farrar, M. A., Campbell, J. D., Schreiber, R. D. (1992) Identification of J. L., Emile, J. F., Lamhamedi-Cherradi, S., Altare, F., Pallier, A.,
a functionally important sequence in the C terminus of the interferon- Barcenas-Morales, G., Meinl, E., Krause, C., Pestka, S., Schreiber, R. D.,
gamma receptor. Proc. Natl. Acad. Sci. USA 89, 11706 –11710. Novelli, F., Casanova, J. L. (2000) Partial interferon-gamma receptor
44. Cook, J. R., Jung, V., Schwartz, B., Wang, P., Pestka, S. (1992) Structural signaling chain deficiency in a patient with bacille Calmette-Guerin and
analysis of the human interferon gamma receptor: a small segment of the Mycobacterium abscessus infection. J. Infect. Dis. 181, 379 –384.
intracellular domain is specifically required for class I major histocom- 64. Jouanguy, E., Lamhamedi-Cherradi, S., Lammas, D., Dorman, S. E.,
patibility complex antigen induction and antiviral activity. Proc. Natl. Fondaneche, M. C., Dupuis, S., Doffinger, R., Altare, F., Girdlestone, J.,
Acad. Sci. USA 89, 11317–11321. Emile, J. F., Ducoulombier, H., Edgar, D., Clarke, J., Oxelius, V. A.,
45. Farrar, M. A., Schreiber, R. D. (1993) The molecular cell biology of Brai, M., Novelli, V., Heyne, K., Fischer, A., Holland, S. M., Ku-
interferon-gamma and its receptor. Annu. Rev. Immunol. 11, 571– 611. mararatne, D. S., Schreiber, R. D., Casanova, J. L. (1999) A human
46. Kotenko, S. V., Izotova, L. S., Pollack, B. P., Mariano, T. M., Donnelly, IFNGR1 small deletion hotspot associated with dominant susceptibility
R. J., Muthukumaran, G., Cook, J. R., Garotta, G., Silvennoinen, O., Ihle, to mycobacterial infection. Nat. Genet. 21, 370 –378.
J. N., et al. (1995) Interaction between the components of the interferon 65. Jouanguy, E., Lamhamedi-Cherradi, S., Altare, F., Fondaneche, M. C.,
gamma receptor complex. J. Biol. Chem. 270, 20915–20921. Tuerlinckx, D., Blanche, S., Emile, J. F., Gaillard, J. L., Schreiber, R.,
47. Bach, E. A., Tanner, J. W., Marsters, S., Ashkenazi, A., Aguet, M., Shaw, Levin, M., Fischer, A., Hivroz, C., Casanova, J. L. (1997) Partial inter-
A. S., Schreiber, R. D. (1996) Ligand-induced assembly and activation of feron-gamma receptor 1 deficiency in a child with tuberculoid bacillus
the gamma interferon receptor in intact cells. Mol. Cell. Biol. 16, Calmette-Guerin infection and a sibling with clinical tuberculosis.
3214 –3221. J. Clin. Invest. 100, 2658 –2664.
66. Jouanguy, E., Altare, F., Lamhamedi-Cherradi, S., Casanova, J. L. (1997)
48. Pestka, S., Kotenko, S. V., Muthukumaran, G., Izotova, L. S., Cook, J. R.,
Infections in IFNGR-1-deficient children. J. Interferon Cytokine Res. 17,
Garotta, G. (1997) The interferon gamma (IFN-gamma) receptor: a par-
583–587.
adigm for the multichain cytokine receptor. Cytokine Growth Factor Rev.
67. Newport, M. J., Huxley, C. M., Huston, S., Hawrylowicz, C. M., Oostra,
8, 189 –206.
B. A., Williamson, R., Levin, M. (1996) A mutation in the interferon-
49. Muthukumaran, G., Donnelly, R. J., Ebensperger, C., Mariano, T. M.,
gamma-receptor gene and susceptibility to mycobacterial infection.
Garotta, G., Dembic, Z., Poast, J., Baron, S., Pestka, S. (1996) The
N. Engl. J. Med. 335, 1941–1949.
intracellular domain of the second chain of the interferon-gamma recep-
68. Pierre-Audigier, C., Jouanguy, E., Lamhamedi, S., Altare, F., Rauzier, J.,
tor is interchangeable between species. J. Interferon Cytokine Res. 16,
Vincent, V., Canioni, D., Emile, J. F., Fischer, A., Blanche, S., Gaillard,
1039 –1045. J. L., Casanova, J. L. (1997) Fatal disseminated Mycobacterium smeg-
50. Gibbs, V. C., Williams, S. R., Gray, P. W., Schreiber, R. D., Pennica, D., matis infection in a child with inherited interferon gamma receptor
Rice, G., Goeddel, D. V. (1991) The extracellular domain of the human deficiency. Clin. Infect. Dis. 24, 982–984.
interferon gamma receptor interacts with a species-specific signal trans- 69. Roesler, J., Kofink, B., Wendisch, J., Heyden, S., Paul, D., Friedrich, W.,
ducer. Mol. Cell. Biol. 11, 5860 –5866. Casanova, J. L., Leupold, W., Gahr, M., Rosen-Wolff, A. (1999) Listeria
51. Hibino, Y., Kumar, C. S., Mariano, T. M., Lai, D. H., Pestka, S. (1992) monocytogenes and recurrent mycobacterial infections in a child with
Chimeric interferon-gamma receptors demonstrate that an accessory complete interferon-gamma-receptor (IFNgammaR1) deficiency: muta-
factor required for activity interacts with the extracellular domain. tional analysis and evaluation of therapeutic options. Exp. Hematol. 27,
J. Biol. Chem. 267, 3741–3749. 1368 –1374.
52. Hemmi, S., Merlin, G., Aguet, M. (1992) Functional characterization of 70. Jouanguy, E., Altare, F., Lamhamedi, S., Revy, P., Emile, J. F., Newport,
a hybrid human-mouse interferon gamma receptor: evidence for species- M., Levin, M., Blanche, S., Seboun, E., Fischer, A., Casanova, J. L.
specific interaction of the extracellular receptor domain with a putative (1996) Interferon-gamma-receptor deficiency in an infant with fatal ba-
signal transducer. Proc. Natl. Acad. Sci. USA 89, 2737–2741. cille Calmette-Guerin infection. N. Engl. J. Med. 335, 1956 –1961.
53. Kalina, U., Ozmen, L., Di Padova, K., Gentz, R., Garotta, G. (1993) The 71. Dorman, S. E., Holland, S. M. (1998) Mutation in the signal-transducing
human gamma interferon receptor accessory factor encoded by chromo- chain of the interferon-gamma receptor and susceptibility to mycobac-
some 21 transduces the signal for the induction of 2⬘,5⬘-oligoadenylate- terial infection. J. Clin. Invest. 101, 2364 –2369.
synthetase, resistance to virus cytopathic effect, and major histocompat- 72. Dorman, S. E., Uzel, G., Roesler, J., Bradley, J. S., Bastian, J., Billman,
ibility complex class I antigens. J. Virol. 67, 1702–1706. G., King, S., Filie, A., Schermerhorn, J., Holland, S. M. (1999) Viral
54. Huang, S., Hendriks, W., Althage, A., Hemmi, S., Bluethmann, H., infections in interferon-gamma receptor deficiency. J. Pediatr. 135,
Kamijo, R., Vilcek, J., Zinkernagel, R. M., Aguet, M. (1993) Immune 640 – 643.
response in mice that lack the interferon-gamma receptor. Science 259, 73. Davies, E. G., Isaacs, D., Levinsky, R. J. (1982) Defective immune
1742–1745. interferon production and natural killer activity associated with poor
55. Pearl, J. E., Saunders, B., Ehlers, S., Orme, I. M., Cooper, A. M. (2001) neutrophil mobility and delayed umbilical cord separation. Clin. Exp.
Inflammation and lymphocyte activation during mycobacterial infection Immunol. 50, 454 – 460.
in the interferon-gamma-deficient mouse. Cell. Immunol. 211, 43–50. 74. Lio, D., Scola, L., Crivello, A., Bonafe, M., Franceschi, C., Olivieri, F.,
56. van den Broek, M. F., Muller, U., Huang, S., Zinkernagel, R. M., Aguet, Colonna-Romano, G., Candore, G., Caruso, C. (2002) Allele frequencies
M. (1995) Immune defence in mice lacking type I and/or type II inter- of ⫹874T3 A single nucleotide polymorphism at the first intron of
feron receptors. Immunol. Rev. 148, 5–18. interferon-gamma gene in a group of Italian centenarians. Exp. Gerontol.
57. Buchmeier, N. A., Schreiber, R. D. (1985) Requirement of endogenous 37, 315–319.
interferon-gamma production for resolution of Listeria monocytogenes 75. Espejo, C., Penkowa, M., Saez-Torres, I., Xaus, J., Celada, A., Montal-
infection. Proc. Natl. Acad. Sci. USA 82, 7404 –7408. ban, X., Martinez-Caceres, E. M. (2001) Treatment with anti-interferon-
58. Kamijo, R., Le, J., Shapiro, D., Havell, E. A., Huang, S., Aguet, M., gamma monoclonal antibodies modifies experimental autoimmune en-
Bosland, M., Vilcek, J. (1993) Mice that lack the interferon-gamma cephalomyelitis in interferon-gamma receptor knockout mice. Exp. Neu-
receptor have profoundly altered responses to infection with Bacillus rol. 172, 460 – 468.
Calmette-Guerin and subsequent challenge with lipopolysaccharide. J. 76. Billiau, A., Heremans, H., Vandekerckhove, F., Dijkmans, R., Sobis, H.,
Exp. Med. 178, 1435–1440. Meulepas, E., Carton, H. (1988) Enhancement of experimental allergic

180 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


encephalomyelitis in mice by antibodies against IFN-gamma. J. Immu- 99. Igarashi, K., Garotta, G., Ozmen, L., Ziemiecki, A., Wilks, A. F., Harpur,
nol. 140, 1506 –1510. A. G., Larner, A. C., Finbloom, D. S. (1994) Interferon-gamma induces
77. Heremans, H., Billiau, A., Colombatti, A., Hilgers, J., de Somer, P. tyrosine phosphorylation of interferon-gamma receptor and regulated
(1978) Interferon treatment of NZB mice: accelerated progression of association of protein tyrosine kinases, Jak1 and Jak2, with its receptor.
autoimmune disease. Infect. Immun. 21, 925–930. J. Biol. Chem. 269, 14333–14336.
78. Jacob, C. O., McDevitt, H. O. (1988) Tumour necrosis factor-alpha in 100. Greenlund, A. C., Morales, M. O., Viviano, B. L., Yan, H., Krolewski, J.,
murine autoimmune ‘lupus’ nephritis. Nature 331, 356 –358. Schreiber, R. D. (1995) Stat recruitment by tyrosine-phosphorylated
79. Lee, J. Y., Goldman, D., Piliero, L. M., Petri, M., Sullivan, K. E. (2001) cytokine receptors: an ordered reversible affinity-driven process. Immu-
Interferon-gamma polymorphisms in systemic lupus erythematosus. nity 2, 677– 687.
Genes Immun. 2, 254 –257. 101. Heim, M. H., Kerr, I. M., Stark, G. R., Darnell Jr., J. E. (1995)
80. Baechler, E. C., Batliwalla, F. M., Karypis, G., Gaffney, P. M., Ortmann, Contribution of STAT SH2 groups to specific interferon signaling by the
W. A., Espe, K. J., Shark, K. B., Grande, W. J., Hughes, K. M., Kapur, Jak-STAT pathway. Science 267, 1347–1349.
V., Gregersen, P. K., Behrens, T. W. (2003) Interferon-inducible gene 102. Darnell Jr., J. E., Kerr, I. M., Stark, G. R. (1994) Jak-STAT pathways and
expression signature in peripheral blood cells of patients with severe transcriptional activation in response to IFNs and other extracellular
lupus. Proc. Natl. Acad. Sci. USA 100, 2610 –2615. signaling proteins. Science 264, 1415–1421.
81. Panitch, H. S., Hirsch, R. L., Haley, A. S., Johnson, K. P. (1987) 103. Schindler, C., Darnell Jr., J. E. (1995) Transcriptional responses to
Exacerbations of multiple sclerosis in patients treated with gamma polypeptide ligands: the JAK-STAT pathway. Annu. Rev. Biochem. 64,
interferon. Lancet 1, 893– 895.
621– 651.
82. Sarvetnick, N., Liggitt, D., Pitts, S. L., Hansen, S. E., Stewart, T. A.
104. Ramana, C. V., Chatterjee-Kishore, M., Nguyen, H., Stark, G. R. (2000)
(1988) Insulin-dependent diabetes mellitus induced in transgenic mice
Complex roles of Stat1 in regulating gene expression. Oncogene 19,
by ectopic expression of class II MHC and interferon-gamma. Cell 52,
2619 –2627.
773–782.
105. Decker, T., Kovarik, P., Meinke, A. (1997) GAS elements: a few nucle-
83. Awata, T., Matsumoto, C., Urakami, T., Hagura, R., Amemiya, S., Ka-
nazawa, Y. (1994) Association of polymorphism in the interferon gamma otides with a major impact on cytokine-induced gene expression. J. In-
gene with IDDM. Diabetologia 37, 1159 –1162. terferon Cytokine Res. 17, 121–134.
84. Wang, B., Andre, I., Gonzalez, A., Katz, J. D., Aguet, M., Benoist, C., 106. Meraz, M. A., White, J. M., Sheehan, K. C., Bach, E. A., Rodig, S. J.,
Mathis, D. (1997) Interferon-gamma impacts at multiple points during Dighe, A. S., Kaplan, D. H., Riley, J. K., Greenlund, A. C., Campbell, D.,
the progression of autoimmune diabetes. Proc. Natl. Acad. Sci. USA 94, Carver-Moore, K., DuBois, R. N., Clark, R., Aguet, M., Schreiber, R. D.
13844 –13849. (1996) Targeted disruption of the Stat1 gene in mice reveals unexpected
85. Kaplan, D. H., Shankaran, V., Dighe, A. S., Stockert, E., Aguet, M., Old, physiologic specificity in the JAK-STAT signaling pathway. Cell 84,
L. J., Schreiber, R. D. (1998) Demonstration of an interferon gamma- 431– 442.
dependent tumor surveillance system in immunocompetent mice. Proc. 107. Kerr, I. M., Stark, G. R. (1991) The control of interferon-inducible gene
Natl. Acad. Sci. USA 95, 7556 –7561. expression. FEBS Lett. 285, 194 –198.
86. Dighe, A. S., Richards, E., Old, L. J., Schreiber, R. D. (1994) Enhanced 108. Paludan, S. R. (1998) Interleukin-4 and interferon-gamma: the quintes-
in vivo growth and resistance to rejection of tumor cells expressing sence of a mutual antagonistic relationship. Scand. J. Immunol. 48,
dominant negative IFN gamma receptors. Immunity 1, 447– 456. 459 – 468.
87. Tannenbaum, C. S., Hamilton, T. A. (2000) Immune-inflammatory mech- 109. Chen, X., Vinkemeier, U., Zhao, Y., Jeruzalmi, D., Darnell Jr., J. E.,
anisms in IFNgamma-mediated anti-tumor activity. Semin. Cancer Biol. Kuriyan, J. (1998) Crystal structure of a tyrosine phosphorylated STAT-1
10, 113–123. dimer bound to DNA. Cell 93, 827– 839.
88. Dunn, G. P., Bruce, A. T., Ikeda, H., Old, L. J., Schreiber, R. D. (2002) 110. Kovarik, P., Stoiber, D., Eyers, P. A., Menghini, R., Neininger, A.,
Cancer immunoediting: from immunosurveillance to tumor escape. Nat. Gaestel, M., Cohen, P., Decker, T. (1999) Stress-induced phosphoryla-
Immunol. 3, 991–998. tion of STAT1 at Ser727 requires p38 mitogen-activated protein kinase
89. Ikeda, H., Old, L. J., Schreiber, R. D. (2002) The roles of IFN gamma in whereas IFN-gamma uses a different signaling pathway. Proc. Natl.
protection against tumor development and cancer immunoediting. Cyto- Acad. Sci. USA 96, 13956 –13961.
kine Growth Factor Rev. 13, 95–109. 111. Goh, K. C., Haque, S. J., Williams, B. R. (1999) p38 MAP kinase is
90. Subramaniam, P. S., Torres, B. A., Johnson, H. M. (2001) So many required for STAT1 serine phosphorylation and transcriptional activation
ligands, so few transcription factors: a new paradigm for signaling induced by interferons. EMBO J. 18, 5601–5608.
through the STAT transcription factors. Cytokine 15, 175–187. 112. Wen, Z., Zhong, Z., Darnell Jr., J. E. (1995) Maximal activation of
91. Sakatsume, M., Igarashi, K., Winestock, K. D., Garotta, G., Larner, A. C., transcription by Stat1 and Stat3 requires both tyrosine and serine phos-
Finbloom, D. S. (1995) The Jak kinases differentially associate with the phorylation. Cell 82, 241–250.
alpha and beta (accessory factor) chains of the interferon gamma receptor 113. Zhang, X., Blenis, J., Li, H. C., Schindler, C., Chen-Kiang, S. (1995)
to form a functional receptor unit capable of activating STAT transcrip- Requirement of serine phosphorylation for formation of STAT-promoter
tion factors. J. Biol. Chem. 270, 17528 –17534. complexes. Science 267, 1990 –1994.
92. Marsters, S. A., Pennica, D., Bach, E., Schreiber, R. D., Ashkenazi, A. 114. Shuai, K., Stark, G. R., Kerr, I. M., Darnell Jr., J. E. (1993) A single
(1995) Interferon gamma signals via a high-affinity multisubunit receptor phosphotyrosine residue of Stat91 required for gene activation by inter-
complex that contains two types of polypeptide chain. Proc. Natl. Acad. feron-gamma. Science 261, 1744 –1746.
Sci. USA 92, 5401–5405.
115. Shuai, K., Horvath, C. M., Huang, L. H., Qureshi, S. A., Cowburn, D.,
93. Krause, C. D., Mei, E., Xie, J., Jia, Y., Bopp, M. A., Hochstrasser, R. M.,
Darnell Jr., J. E. (1994) Interferon activation of the transcription factor
Pestka, S. (2002) Seeing the light: preassembly and ligand-induced
Stat91 involves dimerization through SH2-phosphotyrosyl peptide inter-
changes of the interferon gamma receptor complex in cells. Mol. Cell.
actions. Cell 76, 821– 828.
Proteomics 1, 805– 815.
94. Ealick, S. E., Cook, W. J., Vijay-Kumar, S., Carson, M., Nagabhushan, 116. Matsumoto, M., Tanaka, N., Harada, H., Kimura, T., Yokochi, T., Kita-
T. L., Trotta, P. P., Bugg, C. E. (1991) Three-dimensional structure of gawa, M., Schindler, C., Taniguchi, T. (1999) Activation of the transcrip-
recombinant human interferon-gamma. Science 252, 698 –702. tion factor ISGF3 by interferon-gamma. Biol. Chem. 380, 699 –703.
95. Walter, M. R., Windsor, W. T., Nagabhushan, T. L., Lundell, D. J., Lunn, 117. Stark, G. R., Kerr, I. M., Williams, B. R., Silverman, R. H., Schreiber,
C. A., Zauodny, P. J., Narula, S. K. (1995) Crystal structure of a complex R. D. (1998) How cells respond to interferons. Annu. Rev. Biochem. 67,
between interferon-gamma and its soluble high-affinity receptor. Nature 227–264.
376, 230 –235. 118. Bluyssen, A. R., Durbin, J. E., Levy, D. E. (1996) ISGF3 gamma p48, a
96. Fountoulakis, M., Zulauf, M., Lustig, A., Garotta, G. (1992) Stoichiom- specificity switch for interferon activated transcription factors. Cytokine
etry of interaction between interferon gamma and its receptor. Eur. Growth Factor Rev. 7, 11–17.
J. Biochem. 208, 781–787. 119. Chatterjee-Kishore, M., van den Akker, F., Stark, G. R. (2000) Associ-
97. Greenlund, A. C., Schreiber, R. D., Goeddel, D. V., Pennica, D. (1993) ation of STATs with relatives and friends. Trends Cell Biol. 10, 106 –
Interferon-gamma induces receptor dimerization in solution and on cells. 111.
J. Biol. Chem. 268, 18103–18110. 120. Decker, T., Kovarik, P. (2000) Serine phosphorylation of STATs. Onco-
98. Briscoe, J., Rogers, N. C., Witthuhn, B. A., Watling, D., Harpur, A. G., gene 19, 2628 –2637.
Wilks, A. F., Stark, G. R., Ihle, J. N., Kerr, I. M. (1996) Kinase-negative 121. Zhu, X., Wen, Z., Xu, L. Z., Darnell Jr., J. E. (1997) Stat1 serine
mutants of JAK1 can sustain interferon-gamma-inducible gene expres- phosphorylation occurs independently of tyrosine phosphorylation and
sion but not an antiviral state. EMBO J. 15, 799 – 809. 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 145. Lundell, D., Lunn, C., Dalgarno, D., Fossetta, J., Greenberg, R., Reim,
alpha interferon and gamma interferon requires transcriptionally active R., Grace, M., Narula, S. (1991) The carboxyl-terminal region of human
Stat1 protein. J. Virol. 70, 647– 650. interferon gamma is important for biological activity: mutagenic and
123. Kovarik, P., Stoiber, D., Novy, M., Decker, T. (1998) Stat1 combines NMR analysis. Protein Eng. 4, 335–341.
signals derived from IFN-gamma and LPS receptors during macrophage 146. Fidler, I. J., Fogler, W. E., Kleinerman, E. S., Saiki, I. (1985) Abrogation
activation. EMBO J. 17, 3660 –3668. of species specificity for activation of tumoricidal properties in macro-
124. Gollob, J. A., Schnipper, C. P., Murphy, E. A., Ritz, J., Frank, D. A. phages by recombinant mouse or human interferon-gamma encapsulated
(1999) The functional synergy between IL-12 and IL-2 involves p38 in liposomes. J. Immunol. 135, 4289 – 4296.
mitogen-activated protein kinase and is associated with the augmentation 147. Sanceau, J., Sondermeyer, P., Beranger, F., Falcoff, R., Vaquero, C.
of STAT serine phosphorylation. J. Immunol. 162, 4472– 4481. (1987) Intracellular human gamma-interferon triggers an antiviral state
125. Singh, K., Balligand, J. L., Fischer, T. A., Smith, T. W., Kelly, R. A. in transformed murine L cells. Proc. Natl. Acad. Sci. USA 84, 2906 –
(1996) Regulation of cytokine-inducible nitric oxide synthase in cardiac 2910.
myocytes and microvascular endothelial cells. Role of extracellular sig- 148. Smith, M. R., Muegge, K., Keller, J. R., Kung, H. F., Young, H. A.,
nal-regulated kinases 1 and 2 (ERK1/ERK2) and STAT1 alpha. J. Biol. Durum, S. K. (1990) Direct evidence for an intracellular role for IFN-
Chem. 271, 1111–1117. gamma. Microinjection of human IFN-gamma induces Ia expression on
126. Chan, E. D., Riches, D. W. (2001) IFN-gamma ⫹ LPS induction of iNOS murine macrophages. J. Immunol. 144, 1777–1782.
is modulated by ERK, JNK/SAPK, and p38(mapk) in a mouse macro- 149. Subramaniam, P. S., Green, M. M., Larkin III, J., Torres, B. A., Johnson,
phage cell line. Am. J. Physiol. Cell Physiol. 280, C441–C450. H. M. (2001) Nuclear translocation of IFN-gamma is an intrinsic require-
127. Mahboubi, K., Pober, J. S. (2002) Activation of signal transducer and ment for its biologic activity and can be driven by a heterologous nuclear
activator of transcription 1 (STAT1) is not sufficient for the induction of localization sequence. J. Interferon Cytokine Res. 21, 951–959.
STAT1-dependent genes in endothelial cells. Comparison of interferon- 150. Larkin III, J., Johnson, H. M., Subramaniam, P. S. (2000) Differential
gamma and oncostatin M. J. Biol. Chem. 277, 8012– 8021. nuclear localization of the IFNGR-1 and IFNGR-2 subunits of the
128. Ouchi, T., Lee, S. W., Ouchi, M., Aaronson, S. A., Horvath, C. M. (2000) IFN-gamma receptor complex following activation by IFN-gamma. J. In-
Collaboration of signal transducer and activator of transcription 1 terferon Cytokine Res. 20, 565–576.
(STAT1) and BRCA1 in differential regulation of IFN-gamma target 151. Durbin, J. E., Hackenmiller, R., Simon, M. C., Levy, D. E. (1996)
genes. Proc. Natl. Acad. Sci. USA 97, 5208 –5213. Targeted disruption of the mouse Stat1 gene results in compromised
129. Zhang, J. J., Zhao, Y., Chait, B. T., Lathem, W. W., Ritzi, M., Knippers, innate immunity to viral disease. Cell 84, 443– 450.
R., Darnell Jr., J. E., (1998) Ser727-dependent recruitment of MCM5 by 152. Ramana, C. V., Gil, M. P., Han, Y., Ransohoff, R. M., Schreiber, R. D.,
Stat1alpha in IFN-gamma-induced transcriptional activation. EMBO J. Stark, G. R. (2001) Stat1-independent regulation of gene expression in
17, 6963– 6971. response to IFN- gamma. Proc. Natl. Acad. Sci. USA 98, 6674 – 6679.
130. Vinkemeier, U., Cohen, S. L., Moarefi, I., Chait, B. T., Kuriyan, J., 153. Bromberg, J. F., Horvath, C. M., Wen, Z., Schreiber, R. D., Darnell Jr.,
Darnell Jr., J. E. (1996) DNA binding of in vitro activated Stat1 alpha, J. E. (1996) Transcriptionally active Stat1 is required for the antiprolif-
Stat1 beta and truncated Stat1: interaction between NH2-terminal do- erative effects of both interferon alpha and interferon gamma. Proc. Natl.
mains stabilizes binding of two dimers to tandem DNA sites. EMBO J.
Acad. Sci. USA 93, 7673–7678.
15, 5616 –5626.
154. Ramana, C. V., Grammatikakis, N., Chernov, M., Nguyen, H., Goh, K. C.,
131. Xu, X., Sun, Y. L., Hoey, T. (1996) Cooperative DNA binding and
Williams, B. R., Stark, G. R. (2000) Regulation of c-myc expression by
sequence-selective recognition conferred by the STAT amino-terminal
IFN-gamma through Stat1-dependent and -independent pathways.
domain. Science 273, 794 –797.
EMBO J. 19, 263–272.
132. Kumar, A., Commane, M., Flickinger, T. W., Horvath, C. M., Stark, G. R.
155. Gil, M. P., Bohn, E., O’Guin, A. K., Ramana, C. V., Levine, B., Stark,
(1997) Defective TNF-alpha-induced apoptosis in STAT1-null cells due
G. R., Virgin, H. W., Schreiber, R. D. (2001) Biologic consequences of
to low constitutive levels of caspases. Science 278, 1630 –1632.
Stat1-independent IFN signaling. Proc. Natl. Acad. Sci. USA 98, 6680 –
133. Lee, C. K., Smith, E., Gimeno, R., Gertner, R., Levy, D. E. (2000) STAT1
affects lymphocyte survival and proliferation partially independent of its 6685.
role downstream of IFN-gamma. J. Immunol. 164, 1286 –1292. 156. Dupuis, S., Dargemont, C., Fieschi, C., Thomassin, N., Rosenzweig, S.,
134. Chatterjee-Kishore, M., Wright, K. L., Ting, J. P., Stark, G. R. (2000) Harris, J., Holland, S. M., Schreiber, R. D., Casanova, J. L. (2001)
How Stat1 mediates constitutive gene expression: a complex of unphos- Impairment of mycobacterial but not viral immunity by a germline human
phorylated Stat1 and IRF1 supports transcription of the LMP2 gene. STAT1 mutation. Science 293, 300 –303.
EMBO J. 19, 4111– 4122. 157. Taniguchi, T., Ogasawara, K., Takaoka, A., Tanaka, N. (2001) IRF family
135. Sekimoto, T., Imamoto, N., Nakajima, K., Hirano, T., Yoneda, Y. (1997) of transcription factors as regulators of host defense. Annu. Rev. Immu-
Extracellular signal-dependent nuclear import of Stat1 is mediated by nol. 19, 623– 655.
nuclear pore-targeting complex formation with NPI-1, but not Rch1. 158. Sims, S. H., Cha, Y., Romine, M. F., Gao, P. Q., Gottlieb, K., Deisseroth,
EMBO J. 16, 7067–7077. A. B. (1993) A novel interferon-inducible domain: structural and func-
136. Lillemeier, B. F., Koster, M., Kerr, I. M. (2001) STAT1 from the cell tional analysis of the human interferon regulatory factor 1 gene promoter.
membrane to the DNA. EMBO J. 20, 2508 –2517. Mol. Cell. Biol. 13, 690 –702.
137. McBride, K. M., McDonald, C., Reich, N. C. (2000) Nuclear export signal 159. Harada, H., Takahashi, E., Itoh, S., Harada, K., Hori, T. A., Taniguchi,
located within the DNA-binding domain of the STAT1 transcription T. (1994) Structure and regulation of the human interferon regulatory
factor. EMBO J. 19, 6196 – 6206. factor 1 (IRF-1) and IRF-2 genes: implications for a gene network in the
138. Koster, M., Hauser, H. (1999) Dynamic redistribution of STAT1 protein interferon system. Mol. Cell. Biol. 14, 1500 –1509.
in IFN signaling visualized by GFP fusion proteins. Eur. J. Biochem. 160. Pine, R. (1997) Convergence of TNFalpha and IFNgamma signalling
260, 137–144. pathways through synergistic induction of IRF-1/ISGF-2 is mediated by
139. Dingwall, C., Laskey, R. A. (1991) Nuclear targeting sequences–a con- a composite GAS/kappaB promoter element. Nucleic Acids Res. 25,
sensus? Trends Biochem. Sci. 16, 478 – 481. 4346 – 4354.
140. MacDonald, H. S., Kushnaryov, V. M., Sedmak, J. J., Grossberg, S. E. 161. Mamane, Y., Heylbroeck, C., Genin, P., Algarte, M., Servant, M. J.,
(1986) Transport of gamma-interferon into the cell nucleus may be LePage, C., DeLuca, C., Kwon, H., Lin, R., Hiscott, J. (1999) Interferon
mediated by nuclear membrane receptors. Biochem. Biophys. Res. Com- regulatory factors: the next generation. Gene 237, 1–14.
mun. 138, 254 –260. 162. Salkowski, C. A., Kopydlowski, K., Blanco, J., Cody, M. J., McNally, R.,
141. Bader, T., Weitzerbin, J. (1994) Nuclear accumulation of interferon Vogel, S. N. (1999) IL-12 is dysregulated in macrophages from IRF-1
gamma. Proc. Natl. Acad. Sci. USA 91, 11831–11835. and IRF-2 knockout mice. J. Immunol. 163, 1529 –1536.
142. Subramaniam, P. S., Mujtaba, M. G., Paddy, M. R., Johnson, H. M. 163. Hobart, M., Ramassar, V., Goes, N., Urmson, J., Halloran, P. F. (1996)
(1999) The carboxyl terminus of interferon-gamma contains a functional The induction of class I and II major histocompatibility complex by
polybasic nuclear localization sequence. J. Biol. Chem. 274, 403– 407. allogeneic stimulation is dependent on the transcription factor interferon
143. Subramaniam, P. S., Larkin III, J., Mujtaba, M. G., Walter, M. R., regulatory factor 1 (IRF-1): observations in IRF-1 knockout mice. Trans-
Johnson, H. M. (2000) The COOH-terminal nuclear localization se- plantation 62, 1895–1901.
quence of interferon gamma regulates STAT1 alpha nuclear translocation 164. Tanaka, N., Kawakami, T., Taniguchi, T. (1993) Recognition DNA
at an intracellular site. J. Cell Sci. 113, 2771–2781. sequences of interferon regulatory factor 1 (IRF-1) and IRF-2, regulators
144. Wetzel, R., Perry, L. J., Veilleux, C., Chang, G. (1990) Mutational of cell growth and the interferon system. Mol. Cell. Biol. 13, 4531– 4538.
analysis of the C-terminus of human interferon-gamma. Protein Eng. 3, 165. Tanaka, N., Ishihara, M., Kitagawa, M., Harada, H., Kimura, T., Mat-
611– 623. suyama, T., Lamphier, M. S., Aizawa, S., Mak, T. W., Taniguchi, T.

182 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


(1994) Cellular commitment to oncogene-induced transformation or apo- tutive protein tyrosine phosphatase activity controls the kinase function
ptosis is dependent on the transcription factor IRF-1. Cell 77, 829 – 839. of JAK1. Proc. Natl. Acad. Sci. USA 94, 8563– 8568.
166. Kimura, T., Nakayama, K., Penninger, J., Kitagawa, M., Harada, H., 184. You, M., Yu, D. H., Feng, G. S. (1999) Shp-2 tyrosine phosphatase
Matsuyama, T., Tanaka, N., Kamijo, R., Vilcek, J., Mak, T. W., et al. functions as a negative regulator of the interferon-stimulated Jak/STAT
(1994) Involvement of the IRF-1 transcription factor in antiviral re- pathway. Mol. Cell. Biol. 19, 2416 –2424.
sponses to interferons. Science 264, 1921–1924. 185. Ruff, S. J., Chen, K., Cohen, S. (1997) Peroxovanadate induces tyrosine
167. Ohmori, Y., Hamilton, T. A. (1994) IFN-gamma selectively inhibits phosphorylation of multiple signaling proteins in mouse liver and kidney.
lipopolysaccharide-inducible JE/monocyte chemoattractant protein-1 J. Biol. Chem. 272, 1263–1267.
and KC/GRO/melanoma growth-stimulating activity gene expression in 186. David, M., Chen, H. E., Goelz, S., Larner, A. C., Neel, B. G. (1995)
mouse peritoneal macrophages. J. Immunol. 153, 2204 –2212. Differential regulation of the alpha/beta interferon-stimulated Jak/Stat
168. Cook, J. R., Emanuel, S. L., Donnelly, R. J., Soh, J., Mariano, T. M., pathway by the SH2 domain-containing tyrosine phosphatase SHPTP1.
Schwartz, B., Rhee, S., Pestka, S. (1994) Sublocalization of the human Mol. Cell. Biol. 15, 7050 –7058.
interferon-gamma receptor accessory factor gene and characterization of 187. Yu, C. L., Jin, Y. J., Burakoff, S. J. (2000) Cytosolic tyrosine dephos-
accessory factor activity by yeast artificial chromosomal fragmentation. phorylation of STAT5. Potential role of SHP-2 in STAT5 regulation.
J. Biol. Chem. 269, 7013–7018. J. Biol. Chem. 275, 599 – 604.
169. Darnell Jr., J. E. (1997) STATs and gene regulation. Science 277, 188. Shultz, L. D., Schweitzer, P. A., Rajan, T. V., Yi, T., Ihle, J. N.,
1630 –1635. Matthews, R. J., Thomas, M. L., Beier, D. R. (1993) Mutations at the
170. Schreiber, R. D., Farrar, M. A. (1993) The biology and biochemistry of murine motheaten locus are within the hematopoietic cell protein-ty-
interferon-gamma and its receptor. Gastroenterol Jpn 28 (Suppl. 4), rosine phosphatase (Hcph) gene. Cell 73, 1445–1454.
88 –94. 189. Mowen, K., David, M. (2000) Regulation of STAT1 nuclear export by
171. Anderson, P., Yip, Y. K., Vilcek, J. (1983) Human interferon-gamma is Jak1. Mol. Cell. Biol. 20, 7273–7281.
internalized and degraded by cultured fibroblasts. J. Biol. Chem. 258, 190. Fornerod, M., Ohno, M., Yoshida, M., Mattaj, I. W. (1997) CRM1 is an
6497– 6502. export receptor for leucine-rich nuclear export signals. Cell 90, 1051–
172. Celada, A., Schreiber, R. D. (1987) Internalization and degradation of 1060.
receptor-bound interferon-gamma by murine macrophages. Demonstra- 191. Stade, K., Ford, C. S., Guthrie, C., Weis, K. (1997) Exportin 1 (Crm1p)
tion of receptor recycling. J. Immunol. 139, 147–153. is an essential nuclear export factor. Cell 90, 1041–1050.
173. Endo, T. A., Masuhara, M., Yokouchi, M., Suzuki, R., Sakamoto, H., 192. Haspel, R. L., Darnell Jr., J. E. (1999) A nuclear protein tyrosine
Mitsui, K., Matsumoto, A., Tanimura, S., Ohtsubo, M., Misawa, H., phosphatase is required for the inactivation of Stat1. Proc. Natl. Acad.
Miyazaki, T., Leonor, N., Taniguchi, T., Fujita, T., Kanakura, Y., Sci. USA 96, 10188 –10193.
Komiya, S., Yoshimura, A. (1997) A new protein containing an SH2 193. Haspel, R. L., Salditt-Georgieff, M., Darnell Jr., J. E. (1996) The rapid
domain that inhibits JAK kinases. Nature 387, 921–924. inactivation of nuclear tyrosine phosphorylated Stat1 depends upon a
174. Starr, R., Willson, T. A., Viney, E. M., Murray, L. J., Rayner, J. R., protein tyrosine phosphatase. EMBO J. 15, 6262– 6268.
Jenkins, B. J., Gonda, T. J., Alexander, W. S., Metcalf, D., Nicola, N. A., 194. Muller, M., Briscoe, J., Laxton, C., Guschin, D., Ziemiecki, A., Silven-
Hilton, D. J. (1997) A family of cytokine-inducible inhibitors of signal- noinen, O., Harpur, A. G., Barbieri, G., Witthuhn, B. A., Schindler, C.,
ling. Nature 387, 917–921. et al. (1993) The protein tyrosine kinase JAK1 complements defects in
175. Suzuki, R., Sakamoto, H., Yasukawa, H., Masuhara, M., Wakioka, T., interferon-alpha/beta and -gamma signal transduction. Nature 366,
Sasaki, A., Yuge, K., Komiya, S., Inoue, A., Yoshimura, A. (1998) CIS3 129 –135.
and JAB have different regulatory roles in interleukin-6 mediated dif- 195. Watling, D., Guschin, D., Muller, M., Silvennoinen, O., Witthuhn, B. A.,
ferentiation and STAT3 activation in M1 leukemia cells. Oncogene 17, Quelle, F. W., Rogers, N. C., Schindler, C., Stark, G. R., Ihle, J. N., et
2271–2278. al. (1993) Complementation by the protein tyrosine kinase JAK2 of a
176. Sakamoto, H., Yasukawa, H., Masuhara, M., Tanimura, S., Sasaki, A., mutant cell line defective in the interferon-gamma signal transduction
Yuge, K., Ohtsubo, M., Ohtsuka, A., Fujita, T., Ohta, T., Furukawa, Y., pathway. Nature 366, 166 –170.
Iwase, S., Yamada, H., Yoshimura, A. (1998) A Janus kinase inhibitor, 196. Rodig, S. J., Meraz, M. A., White, J. M., Lampe, P. A., Riley, J. K.,
JAB, is an interferon-gamma-inducible gene and confers resistance to Arthur, C. D., King, K. L., Sheehan, K. C., Yin, L., Pennica, D., Johnson
interferons. Blood 92, 1668 –1676. Jr., E. M., Schreiber, R. D. (1998) Disruption of the Jak1 gene demon-
177. Yasukawa, H., Misawa, H., Sakamoto, H., Masuhara, M., Sasaki, A., strates obligatory and nonredundant roles of the Jaks in cytokine-induced
Wakioka, T., Ohtsuka, S., Imaizumi, T., Matsuda, T., Ihle, J. N., Yo- biologic responses. Cell 93, 373–383.
shimura, A. (1999) The JAK-binding protein JAB inhibits Janus tyrosine 197. Parganas, E., Wang, D., Stravopodis, D., Topham, D. J., Marine, J. C.,
kinase activity through binding in the activation loop. EMBO J. 18, Teglund, S., Vanin, E. F., Bodner, S., Colamonici, O. R., van Deursen,
1309 –1320. J. M., Grosveld, G., Ihle, J. N. (1998) Jak2 is essential for signaling
178. Zhang, J. G., Farley, A., Nicholson, S. E., Willson, T. A., Zugaro, L. M., through a variety of cytokine receptors. Cell 93, 385–395.
Simpson, R. J., Moritz, R. L., Cary, D., Richardson, R., Hausmann, G., 198. Neubauer, H., Cumano, A., Muller, M., Wu, H., Huffstadt, U., Pfeffer, K.
Kile, B. J., Kent, S. B., Alexander, W. S., Metcalf, D., Hilton, D. J., (1998) Jak2 deficiency defines an essential developmental checkpoint in
Nicola, N. A., Baca, M. (1999) The conserved SOCS box motif in definitive hematopoiesis. Cell 93, 397– 409.
suppressors of cytokine signaling binds to elongins B and C and may 199. Kotenko, S. V., Izotova, L. S., Pollack, B. P., Muthukumaran, G., Paukku,
couple bound proteins to proteasomal degradation. Proc. Natl. Acad. Sci. K., Silvennoinen, O., Ihle, J. N., Pestka, S. (1996) Other kinases can
USA 96, 2071–2076. substitute for Jak2 in signal transduction by interferon-gamma. J. Biol.
179. Fujimoto, M., Naka, T., Nakagawa, R., Kawazoe, Y., Morita, Y., Tateishi, Chem. 271, 17174 –17182.
A., Okumura, K., Narazaki, M., Kishimoto, T. (2000) Defective thymo- 200. Kaplan, M. H., Sun, Y. L., Hoey, T., Grusby, M. J. (1996) Impaired IL-12
cyte development and perturbed homeostasis of T cells in STAT-induced responses and enhanced development of Th2 cells in Stat4-deficient
STAT inhibitor-1/suppressors of cytokine signaling-1 transgenic mice. mice. Nature 382, 174 –177.
J. Immunol. 165, 1799 –1806. 201. Thierfelder, W. E., van Deursen, J. M., Yamamoto, K., Tripp, R. A.,
180. Alexander, W. S., Starr, R., Fenner, J. E., Scott, C. L., Handman, E., Sarawar, S. R., Carson, R. T., Sangster, M. Y., Vignali, D. A., Doherty,
Sprigg, N. S., Corbin, J. E., Cornish, A. L., Darwiche, R., Owczarek, P. C., Grosveld, G. C., Ihle, J. N. (1996) Requirement for Stat4 in
C. M., Kay. T. W., Nicola, N. A. Hertzog, P. J., Metcalf, D., Hilton, D. J. interleukin-12-mediated responses of natural killer and T cells. Nature
(1999) SOCS1 is a critical inhibitor of interferon gamma signaling and 382, 171–174.
prevents the potentially fatal neonatal actions of this cytokine. Cell 98, 202. Takeda, K., Tanaka, T., Shi, W., Matsumoto, M., Minami, M., Kashi-
597– 608. wamura, S., Nakanishi, K., Yoshida, N., Kishimoto, T., Akira, S. (1996)
181. Marine, J. C., Topham, D. J., McKay, C., Wang, D., Parganas, E., Essential role of Stat6 in IL-4 signalling. Nature 380, 627– 630.
Stravopodis, D., Yoshimura, A., Ihle, J. N. (1999) SOCS1 deficiency 203. Shimoda, K., van Deursen, J., Sangster, M. Y., Sarawar, S. R., Carson,
causes a lymphocyte-dependent perinatal lethality. Cell 98, 609 – 616. R. T., Tripp, R. A., Chu, C., Quelle, F. W., Nosaka, T., Vignali, D. A.,
182. Song, M. M., Shuai, K. (1998) The suppressor of cytokine signaling Doherty, P. C., Grosveld, G., Paul, W. E., Ihle, J. N. (1996) Lack of
(SOCS) 1 and SOCS3 but not SOCS2 proteins inhibit interferon-mediated IL-4-induced Th2 response and IgE class switching in mice with dis-
antiviral and antiproliferative activities. J. Biol. Chem. 273, 35056 – rupted Stat6 gene. Nature 380, 630 – 633.
35062. 204. Liu, X., Robinson, G. W., Wagner, K. U., Garrett, L., Wynshaw-Boris, A.,
183. Haque, S. J., Wu, Q., Kammer, W., Friedrich, K., Smith, J. M., Kerr, Hennighausen, L. (1997) Stat5a is mandatory for adult mammary gland
I. M., Stark, G. R., Williams, B. R. (1997) Receptor-associated consti- 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. human tumor cells: enhancement of HLA gene expression synergistic
(1998) Stat3 activation is responsible for IL-6-dependent T cell prolif- with interferon-gamma. Immunobiology 172, 291–300.
eration through preventing apoptosis: generation and characterization of 227. Johnson, D. R., Pober, J. S. (1990) Tumor necrosis factor and immune
T cell-specific Stat3-deficient mice. J. Immunol. 161, 4652– 4660. interferon synergistically increase transcription of HLA class I heavy-
206. Kawakami, T., Matsumoto, M., Sato, M., Harada, H., Taniguchi, T., and light-chain genes in vascular endothelium. Proc. Natl. Acad. Sci.
Kitagawa, M. (1995) Possible involvement of the transcription factor USA 87, 5183–5187.
ISGF3 gamma in virus-induced expression of the IFN-beta gene. FEBS 228. Reis, L. F., Harada, H., Wolchok, J. D., Taniguchi, T., Vilcek, J. (1992)
Lett. 358, 225–229. Critical role of a common transcription factor, IRF-1, in the regulation of
207. Yoneyama, M., Suhara, W., Fukuhara, Y., Sato, M., Ozato, K., Fujita, T. IFN-beta and IFN-inducible genes. EMBO J. 11, 185–193.
(1996) Autocrine amplification of type I interferon gene expression 229. Chang, C. H., Hammer, J., Loh, J. E., Fodor, W. L., Flavell, R. A. (1992)
mediated by interferon stimulated gene factor 3 (ISGF3). J. Biochem. The activation of major histocompatibility complex class I genes by
(Tokyo) 120, 160 –169. interferon regulatory factor-1 (IRF-1). Immunogenetics 35, 378 –384.
208. Min, W., Pober, J. S., Johnson, D. R. (1998) Interferon induction of 230. Wallach, D., Fellous, M., Revel, M. (1982) Preferential effect of gamma
TAP1: the phosphatase SHP-1 regulates crossover between the IFN- interferon on the synthesis of HLA antigens and their mRNAs in human
alpha/beta and the IFN-gamma signal-transduction pathways. Circ. Res. cells. Nature 299, 833– 836.
83, 815– 823. 231. Fellous, M., Nir, U., Wallach, D., Merlin, G., Rubinstein, M., Revel, M.
209. Takaoka, A., Mitani, Y., Suemori, H., Sato, M., Yokochi, T., Noguchi, S., (1982) Interferon-dependent induction of mRNA for the major histocom-
Tanaka, N., Taniguchi, T. (2000) Cross talk between interferon-gamma patibility antigens in human fibroblasts and lymphoblastoid cells. Proc.
and -alpha/beta signaling components in caveolar membrane domains. Natl. Acad. Sci. USA 79, 3082–3086.
Science 288, 2357–2360. 232. Seder, R. A., Gazzinelli, R. T. (1999) Cytokines are critical in linking the
210. Belich, M. P., Glynne, R. J., Senger, G., Sheer, D., Trowsdale, J. (1994) innate and adaptive immune responses to bacterial, fungal, and parasitic
Proteasome components with reciprocal expression to that of the MHC- infection. Adv. Intern. Med. 44, 353–388.
encoded LMP proteins. Curr. Biol. 4, 769 –776. 233. Seliger, B., Schreiber, K., Delp, K., Meissner, M., Hammers, S.,
211. Kelly, A., Powis, S. H., Glynne, R., Radley, E., Beck, S., Trowsdale, J. Reichert, T., Pawlischko, K., Tampe, R., Huber, C. (2001) Downregula-
(1991) Second proteasome-related gene in the human MHC class II tion of the constitutive tapasin expression in human tumor cells of
region. Nature 353, 667– 668. distinct origin and its transcriptional upregulation by cytokines. Tissue
212. Groettrup, M., Kraft, R., Kostka, S., Standera, S., Stohwasser, R., Kloet- Antigens 57, 39 – 45.
zel, P. M. (1996) A third interferon-gamma-induced subunit exchange in 234. Mach, B., Steimle, V., Martinez-Soria, E., Reith, W. (1996) Regulation of
the 20S proteasome. Eur. J. Immunol. 26, 863– 869. MHC class II genes: lessons from a disease. Annu. Rev. Immunol. 14,
213. Nandi, D., Jiang, H., Monaco, J. J. (1996) Identification of MECL-1 301–331.
(LMP-10) as the third IFN-gamma-inducible proteasome subunit. J. Im- 235. Kern, I., Steimle, V., Siegrist, C. A., Mach, B. (1995) The two novel MHC
munol. 156, 2361–2364. class II transactivators RFX5 and CIITA both control expression of
214. Hisamatsu, H., Shimbara, N., Saito, Y., Kristensen, P., Hendil, K. B., HLA-DM genes. Int. Immunol. 7, 1295–1299.
Fujiwara, T., Takahashi, E., Tanahashi, N., Tamura, T., Ichihara, A., 236. Figueiredo, F., Koerner, T. J., Adams, D. O. (1989) Molecular mecha-
Tanaka, K. (1996) Newly identified pair of proteasomal subunits regu- nisms regulating the expression of class II histocompatibility molecules
lated reciprocally by interferon gamma. J. Exp. Med. 183, 1807–1816. on macrophages. Effects of inductive and suppressive signals on gene
215. Strobl, B., Arulampalam, V., Is’harc, H., Newman, S. J., Schlaak, J. F., transcription. J. Immunol. 143, 3781–3786.
Watling, D., Costa-Pereira, A. P., Schaper, F., Behrmann, I., Sheehan, 237. Chang, C. H., Flavell, R. A. (1995) Class II transactivator regulates the
K. C., Schreiber, R. D., Horn, F., Heinrich, P. C., Kerr, I. M. (2001) A expression of multiple genes involved in antigen presentation. J. Exp.
completely foreign receptor can mediate an interferon-gamma-like re- Med. 181, 765–767.
sponse. EMBO J. 20, 5431–5442. 238. Cresswell, P. (1994) Assembly, transport, and function of MHC class II
216. Groettrup, M., Khan, S., Schwarz, K., Schmidtke, G. (2001) Interferon- molecules. Annu. Rev. Immunol. 12, 259 –293.
gamma inducible exchanges of 20S proteasome active site subunits: why? 239. Wolf, P. R., Ploegh, H. L. (1995) How MHC class II molecules acquire
Biochimie 83, 367–372. peptide cargo: biosynthesis and trafficking through the endocytic path-
217. Dick, T. P., Ruppert, T., Groettrup, M., Kloetzel, P. M., Kuehn, L., way. Annu. Rev. Cell Dev. Biol. 11, 267–306.
Koszinowski, U. H., Stevanovic, S., Schild, H., Rammensee, H. G. (1996) 240. Lah, T. T., Hawley, M., Rock, K. L., Goldberg, A. L. (1995) Gamma-
Coordinated dual cleavages induced by the proteasome regulator PA28 interferon causes a selective induction of the lysosomal proteases, cathe-
lead to dominant MHC ligands. Cell 86, 253–262. psins B and L, in macrophages. FEBS Lett. 363, 85– 89.
218. Groettrup, M., Soza, A., Eggers, M., Kuehn, L., Dick, T. P., Schild, H., 241. Lafuse, W. P., Brown, D., Castle, L., Zwilling, B. S. (1995) IFN-gamma
Rammensee, H. G., Koszinowski, U. H., Kloetzel, P. M. (1996) A role for increases cathepsin H mRNA levels in mouse macrophages. J. Leukoc.
the proteasome regulator PA28alpha in antigen presentation. Nature Biol. 57, 663– 669.
381, 166 –168. 242. Meurs, E., Chong, K., Galabru, J., Thomas, N. S., Kerr, I. M., Williams,
219. Boes, B., Hengel, H., Ruppert, T., Multhaup, G., Koszinowski, U. H., B. R., Hovanessian, A. G. (1990) Molecular cloning and characterization
Kloetzel, P. M. (1994) Interferon gamma stimulation modulates the of the human double-stranded RNA-activated protein kinase induced by
proteolytic activity and cleavage site preference of 20S mouse protea- interferon. Cell 62, 379 –390.
somes. J. Exp. Med. 179, 901–909. 243. Zamanian-Daryoush, M., Mogensen, T. H., DiDonato, J. A., Williams,
220. Trowsdale, J., Hanson, I., Mockridge, I., Beck, S., Townsend, A., Kelly, B. R. (2000) NF-kappaB activation by double-stranded-RNA-activated
A. (1990) Sequences encoded in the class II region of the MHC related protein kinase (PKR) is mediated through NF-kappaB-inducing kinase
to the ‘ABC’ superfamily of transporters. Nature 348, 741–744. and IkappaB kinase. Mol. Cell. Biol. 20, 1278 –1290.
221. Epperson, D. E., Arnold, D., Spies, T., Cresswell, P., Pober, J. S., 244. Osman, F., Jarrous, N., Ben-Asouli, Y., Kaempfer, R. (1999) A cis-acting
Johnson, D. R. (1992) Cytokines increase transporter in antigen process- element in the 3⬘-untranslated region of human TNF-alpha mRNA
ing-1 expression more rapidly than HLA class I expression in endothelial renders splicing dependent on the activation of protein kinase PKR.
cells. J. Immunol. 149, 3297–3301. Genes Dev. 13, 3280 –3293.
222. Decker, T., Stockinger, S., Karaghiosoff, M., Muller, M., Kovarik, P. 245. Donze, O., Dostie, J., Sonenberg, N. (1999) Regulatable expression of the
(2002) IFNs and STATs in innate immunity to microorganisms. J. Clin. interferon-induced double-stranded RNA dependent protein kinase PKR
Invest. 109, 1271–1277. induces apoptosis and fas receptor expression. Virology 256, 322–329.
223. Cramer, L. A., Nelson, S. L., Klemsz, M. J. (2000) Synergistic induction 246. Balachandran, S., Kim, C. N., Yeh, W. C., Mak, T. W., Bhalla, K.,
of the Tap-1 gene by IFN-gamma and lipopolysaccharide in macrophages Barber, G. N. (1998) Activation of the dsRNA-dependent protein kinase,
is regulated by STAT1. J. Immunol. 165, 3190 –3197. PKR, induces apoptosis through FADD-mediated death signaling.
224. Shirayoshi, Y., Burke, P. A., Appella, E., Ozato, K. (1988) Interferon- EMBO J. 17, 6888 – 6902.
induced transcription of a major histocompatibility class I gene accom- 247. Wong, A. H., Tam, N. W., Yang, Y. L., Cuddihy, A. R., Li, S., Kirchhoff,
panies binding of inducible nuclear factors to the interferon consensus S., Hauser, H., Decker, T., Koromilas, A. E. (1997) Physical association
sequence. Proc. Natl. Acad. Sci. USA 85, 5884 –5888. between STAT1 and the interferon-inducible protein kinase PKR and
225. Pestka, S., Langer, J. A., Zoon, K. C., Samuel, C. E. (1987) Interferons implications for interferon and double-stranded RNA signaling pathways.
and their actions. Annu. Rev. Biochem. 56, 727–777. EMBO J. 16, 1291–1304.
226. Scheurich, P., Kronke, M., Schluter, C., Ucer, U., Pfizenmaier, K. (1986) 248. Patterson, J. B., Thomis, D. C., Hans, S. L., Samuel, C. E. (1995)
Noncytocidal mechanisms of action of tumor necrosis factor-alpha on Mechanism of interferon action: double-stranded RNA-specific adeno-

184 Journal of Leukocyte Biology Volume 75, February 2004 http://www.jleukbio.org


sine deaminase from human cells is inducible by alpha and gamma 271. Yeung, M. C., Liu, J., Lau, A. S. (1996) An essential role for the
interferons. Virology 210, 508 –511. interferon-inducible, double-stranded RNA-activated protein kinase
249. Anderson, S. L., Carton, J. M., Lou, J., Xing, L., Rubin, B. Y. (1999) PKR in the tumor necrosis factor-induced apoptosis in U937 cells. Proc.
Interferon-induced guanylate binding protein-1 (GBP-1) mediates an Natl. Acad. Sci. USA 93, 12451–12455.
antiviral effect against vesicular stomatitis virus and encephalomyocar- 272. Lee, S. B., Esteban, M. (1994) The interferon-induced double-stranded
ditis virus. Virology 256, 8 –14. RNA-activated protein kinase induces apoptosis. Virology 199, 491–
250. Vestal, D. J., Gorbacheva, V. Y., Sen, G. C. (2000) Different subcellular 496.
localizations for the related interferon-induced GTPases, MuGBP-1 and 273. Inbal, B., Cohen, O., Polak-Charcon, S., Kopolovic, J., Vadai, E., Eisen-
MuGBP-2: implications for different functions? J. Interferon Cytokine bach, L., Kimchi, A. (1997) DAP kinase links the control of apoptosis to
Res. 20, 991–1000. metastasis. Nature 390, 180 –184.
251. Raveh, T., Hovanessian, A. G., Meurs, E. F., Sonenberg, N., Kimchi, A. 274. Deiss, L. P., Feinstein, E., Berissi, H., Cohen, O., Kimchi, A. (1995)
(1996) Double-stranded RNA-dependent protein kinase mediates c-Myc Identification of a novel serine/threonine kinase and a novel 15-kD
suppression induced by type I interferons. J. Biol. Chem. 271, 25479 – protein as potential mediators of the gamma interferon-induced cell
25484. death. Genes Dev. 9, 15–30.
252. Xaus, J., Cardo, M., Valledor, A. F., Soler, C., Lloberas, J., Celada, A. 275. Kissil, J. L., Deiss, L. P., Bayewitch, M., Raveh, T., Khaspekov, G.,
(1999) Interferon gamma induces the expression of p21waf-1 and arrests Kimchi, A. (1995) Isolation of DAP3, a novel mediator of interferon-
macrophage cell cycle, preventing induction of apoptosis. Immunity 11, gamma-induced cell death. J. Biol. Chem. 270, 27932–27936.
103–113. 276. Levy-Strumpf, N., Deiss, L. P., Berissi, H., Kimchi, A. (1997) DAP-5, a
253. Harvat, B. L., Seth, P., Jetten, A. M. (1997) The role of p27Kip1 in novel homolog of eukaryotic translation initiation factor 4G isolated as a
gamma interferon-mediated growth arrest of mammary epithelial cells putative modulator of gamma interferon-induced programmed cell death.
and related defects in mammary carcinoma cells. Oncogene 14, 2111– Mol. Cell. Biol. 17, 1615–1625.
2122. 277. Deiss, L. P., Galinka, H., Berissi, H., Cohen, O., Kimchi, A. (1996)
254. Matsuoka, M., Nishimoto, I., Asano, S. (1999) Interferon-gamma impairs Cathepsin D protease mediates programmed cell death induced by
physiologic downregulation of cyclin-dependent kinase inhibitor, interferon-gamma, Fas/APO-1 and TNF-alpha. EMBO J. 15, 3861–
p27Kip1, during G1 phase progression in macrophages. Exp. Hematol. 3870.
27, 203–209. 278. Xu, X., Fu, X. Y., Plate, J., Chong, A. S. (1998) IFN-gamma induces cell
255. Kominsky, S., Johnson, H. M., Bryan, G., Tanabe, T., Hobeika, A. C., growth inhibition by Fas-mediated apoptosis: requirement of STAT1
Subramaniam, P. S., Torres, B. (1998) IFNgamma inhibition of cell protein for up-regulation of Fas and FasL expression. Cancer Res. 58,
growth in glioblastomas correlates with increased levels of the cyclin 2832–2837.
dependent kinase inhibitor p21WAF1/CIP1. Oncogene 17, 2973–2979. 279. Zheng, H., Luo, R. C., Zhang, L. S., Mai, G. F. (2002) Interferon-gamma
256. Mandal, M., Bandyopadhyay, D., Goepfert, T. M., Kumar, R. (1998) up-regulates Fas expression and increases Fas-mediated apoptosis in
Interferon-induces expression of cyclin-dependent kinase-inhibitors tumor cell lines. Di Yi Jun Yi Da Xue Xue Bao 22, 1090 –1092.
p21WAF1 and p27Kip1 that prevent activation of cyclin-dependent 280. Tsujimoto, M., Yip, Y. K., Vilcek, J. (1986) Interferon-gamma enhances
kinase by CDK-activating kinase (CAK). Oncogene 16, 217–225. expression of cellular receptors for tumor necrosis factor. J. Immunol.
257. Harvat, B. L., Jetten, A. M. (1996) Gamma-interferon induces an irre- 136, 2441–2444.
versible growth arrest in mid-G1 in mammary epithelial cells which 281. MacMicking, J., Xie, Q. W., Nathan, C. (1997) Nitric oxide and macro-
correlates with a block in hyperphosphorylation of retinoblastoma. Cell phage function. Annu. Rev. Immunol. 15, 323–350.
Growth Differ. 7, 289 –300. 282. Eizirik, D. L., Flodstrom, M., Karlsen, A. E., Welsh, N. (1996) The
258. Choubey, D., Gutterman, J. U. (1997) Inhibition of E2F-4/DP-1-stimu- harmony of the spheres: inducible nitric oxide synthase and related genes
lated transcription by p202. Oncogene 15, 291–301. in pancreatic beta cells. Diabetologia 39, 875– 890.
259. Ayer, D. E., Kretzner, L., Eisenman, R. N. (1993) Mad: a heterodimeric 283. Kawahara, K., Gotoh, T., Oyadomari, S., Kajizono, M., Kuniyasu, A.,
partner for Max that antagonizes Myc transcriptional activity. Cell 72, Ohsawa, K., Imai, Y., Kohsaka, S., Nakayama, H., Mori, M. (2001)
211–222. Co-induction of argininosuccinate synthetase, cationic amino acid trans-
260. Ayer, D. E., Eisenman, R. N. (1993) A switch from Myc:Max to Mad:Max porter-2, and nitric oxide synthase in activated murine microglial cells.
heterocomplexes accompanies monocyte/macrophage differentiation. Brain Res. Mol. Brain Res. 90, 165–173.
Genes Dev. 7, 2110 –2119. 284. Nussler, A. K., Billiar, T. R., Liu, Z. Z., Morris Jr., S. M. (1994)
261. Cultraro, C. M., Bino, T., Segal, S. (1997) Function of the c-Myc Coinduction of nitric oxide synthase and argininosuccinate synthetase in
antagonist Mad1 during a molecular switch from proliferation to differ- a murine macrophage cell line. Implications for regulation of nitric oxide
entiation. Mol. Cell. Biol. 17, 2353–2359. production. J. Biol. Chem. 269, 1257–1261.
262. Vastrik, I., Kaipainen, A., Penttila, T. L., Lymboussakis, A., Alitalo, R., 285. Di Silvio, M., Geller, D. A., Gross, S. S., Nussler, A., Freeswick, P.,
Parvinen, M., Alitalo, K. (1995) Expression of the mad gene during cell Simmons, R. L., Billiar, T. R. (1993) Inducible nitric oxide synthase
differentiation in vivo and its inhibition of cell growth in vitro. J. Cell activity in hepatocytes is dependent on the coinduction of tetrahydro-
Biol. 128, 1197–1208. biopterin synthesis. Adv. Exp. Med. Biol. 338, 305–308.
263. Obaya, A. J., Mateyak, M. K., Sedivy, J. M. (1999) Mysterious liaisons: 286. Baek, K. J., Thiel, B. A., Lucas, S., Stuehr, D. J. (1993) Macrophage
the relationship between c-Myc and the cell cycle. Oncogene 18, 2934 – nitric oxide synthase subunits. Purification, characterization, and role of
2941. prosthetic groups and substrate in regulating their association into a
264. Vairo, G., Vadiveloo, P. K., Royston, A. K., Rockman, S. P., Rock, C. O., dimeric enzyme. J. Biol. Chem. 268, 21120 –21129.
Jackowski, S., Hamilton, J. A. (1995) Deregulated c-myc expression 287. Cassatella, M. A., Bazzoni, F., Flynn, R. M., Dusi, S., Trinchieri, G.,
overrides IFN gamma-induced macrophage growth arrest. Oncogene 10, Rossi, F. (1990) Molecular basis of interferon-gamma and lipopolysac-
1969 –1976. charide enhancement of phagocyte respiratory burst capability. Studies
265. Grandori, C., Eisenman, R. N. (1997) Myc target genes. Trends Biochem. on the gene expression of several NADPH oxidase components. J. Biol.
Sci. 22, 177–181. Chem. 265, 20241–20246.
266. Taniguchi, T., Lamphier, M. S., Tanaka, N. (1997) IRF-1: the transcrip- 288. Newburger, P. E., Ezekowitz, R. A., Whitney, C., Wright, J., Orkin, S. H.
tion factor linking the interferon response and oncogenesis. Biochim. (1988) Induction of phagocyte cytochrome b heavy chain gene expression
Biophys. Acta 1333, M9 –17. by interferon gamma. Proc. Natl. Acad. Sci. USA 85, 5215–5219.
267. Tamura, T., Ishihara, M., Lamphier, M. S., Tanaka, N., Oishi, I., Aizawa, 289. Gupta, J. W., Kubin, M., Hartman, L., Cassatella, M., Trinchieri, G.
S., Matsuyama, T., Mak, T. W., Taki, S., Taniguchi, T. (1997) DNA (1992) Induction of expression of genes encoding components of the
damage-induced apoptosis and ice gene induction in mitogenically ac- respiratory burst oxidase during differentiation of human myeloid cell
tivated T lymphocytes require IRF-1. Leukemia 11, (Suppl. 3), 439 – lines induced by tumor necrosis factor and gamma-interferon. Cancer
440. Res. 52, 2530 –2537.
268. Villa, P., Kaufmann, S. H., Earnshaw, W. C. (1997) Caspases and 290. Govoni, G., Vidal, S., Cellier, M., Lepage, P., Malo, D., Gros, P. (1995)
caspase inhibitors. Trends Biochem. Sci. 22, 388 –393. Genomic structure, promoter sequence, and induction of expression of
269. Chin, Y. E., Kitagawa, M., Kuida, K., Flavell, R. A., Fu, X. Y. (1997) the mouse Nramp1 gene in macrophages. Genomics 27, 9 –19.
Activation of the STAT signaling pathway can cause expression of 291. Erbe, D. V., Collins, J. E., Shen, L., Graziano, R. F., Fanger, M. W.
caspase 1 and apoptosis. Mol. Cell. Biol. 17, 5328 –5337. (1990) The effect of cytokines on the expression and function of Fc
270. Takizawa, T., Ohashi, K., Nakanishi, Y. (1996) Possible involvement of receptors for IgG on human myeloid cells. Mol. Immunol. 27, 57– 67.
double-stranded RNA-activated protein kinase in cell death by influenza 292. Strunk, R. C., Cole, F. S., Perlmutter, D. H., Colten, H. R. (1985)
virus infection. J. Virol. 70, 8128 – 8132. Gamma-interferon increases expression of class III complement genes

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


C2 and factor B in human monocytes and in murine fibroblasts trans- 316. Heufler, C., Koch, F., Stanzl, U., Topar, G., Wysocka, M., Trinchieri, G.,
fected with human C2 and factor B genes. J. Biol. Chem. 260, 15280 – Enk, A., Steinman, R. M., Romani, N., Schuler, G. (1996) Interleukin-12
15285. is produced by dendritic cells and mediates T helper 1 development as
293. Vincent, F., de la Salle, H., Bohbot, A., Bergerat, J. P., Hauptmann, G., well as interferon-gamma production by T helper 1 cells. Eur. J. Immu-
Oberling, F. (1993) Synthesis and regulation of complement components nol. 26, 659 – 668.
by human monocytes/macrophages and by acute monocytic leukemia. 317. Coutelier, J. P., Van Broeck, J., Wolf, S. F. (1995) Interleukin-12 gene
DNA Cell Biol. 12, 415– 423. expression after viral infection in the mouse. J. Virol. 69, 1955–1958.
294. Colten, H. R., Strunk, R. C., Perlmutter, D. H., Cole, F. S. (1986) 318. Bliss, S. K., Marshall, A. J., Zhang, Y., Denkers, E. Y. (1999) Human
Regulation of complement protein biosynthesis in mononuclear phago- polymorphonuclear leukocytes produce IL-12, TNF-alpha, and the che-
cytes. Ciba Found. Symp. 118, 141–154. mokines macrophage-inflammatory protein-1 alpha and -1 beta in re-
295. Drevets, D. A., Leenen, P. J., Campbell, P. A. (1996) Complement sponse to Toxoplasma gondii antigens. J. Immunol. 162, 7369 –7375.
receptor type 3 mediates phagocytosis and killing of Listeria monocyto- 319. Cassatella, M. A., Meda, L., Gasperini, S., D’Andrea, A., Ma, X., Trin-
genes by a TNF-alpha- and IFN-gamma-stimulated macrophage precur- chieri, G. (1995) Interleukin-12 production by human polymorphonu-
sor hybrid. Cell. Immunol. 169, 1– 6. clear leukocytes. Eur. J. Immunol. 25, 1–5.
296. Yoshida, A., Koide, Y., Uchijima, M., Yoshida, T. O. (1994) IFN-gamma 320. Macatonia, S. E., Hosken, N. A., Litton, M., Vieira, P., Hsieh, C. S.,
induces IL-12 mRNA expression by a murine macrophage cell line, Culpepper, J. A., Wysocka, M., Trinchieri, G., Murphy, K. M., O’Garra,
J774. Biochem. Biophys. Res. Commun. 198, 857– 861. A. (1995) Dendritic cells produce IL-12 and direct the development of
297. Kubin, M., Chow, J. M., Trinchieri, G. (1994) Differential regulation of Th1 cells from naive CD4⫹ T cells. J. Immunol. 154, 5071–5079.
interleukin-12 (IL-12), tumor necrosis factor alpha, and IL-1 beta pro- 321. Lederer, J. A., Perez, V. L., DesRoches, L., Kim, S. M., Abbas, A. K.,
duction in human myeloid leukemia cell lines and peripheral blood Lichtman, A. H. (1996) Cytokine transcriptional events during helper T
mononuclear cells. Blood 83, 1847–1855. cell subset differentiation. J. Exp. Med. 184, 397– 406.
298. Taub, D. D., Lloyd, A. R., Conlon, K., Wang, J. M., Ortaldo, J. R., 322. Trinchieri, G. (1995) Interleukin-12: a proinflammatory cytokine with
Harada, A., Matsushima, K., Kelvin, D. J., Oppenheim, J. J. (1993) immunoregulatory functions that bridge innate resistance and antigen-
Recombinant human interferon-inducible protein 10 is a chemoattractant specific adaptive immunity. Annu. Rev. Immunol. 13, 251–276.
for human monocytes and T lymphocytes and promotes T cell adhesion 323. Morinobu, A., Kumagai, S. (1998) Cytokine measurement at a single-cell
to endothelial cells. J. Exp. Med. 177, 1809 –1814. level to analyze human Th1 and Th2 cells. Rinsho Byori 46, 908 –914.
299. Rollins, B. J., Yoshimura, T., Leonard, E. J., Pober, J. S. (1990) Cyto- 324. Lehn, M., Weiser, W. Y., Engelhorn, S., Gillis, S., Remold, H. G. (1989)
kine-activated human endothelial cells synthesize and secrete a mono- IL-4 inhibits H2O2 production and antileishmanial capacity of human
cyte chemoattractant, MCP-1/JE. Am. J. Pathol. 136, 1229 –1233. cultured monocytes mediated by IFN-gamma. J. Immunol. 143, 3020 –
300. Liao, F., Rabin, R. L., Yannelli, J. R., Koniaris, L. G., Vanguri, P., 3024.
Farber, J. M. (1995) Human Mig chemokine: biochemical and functional 325. Liew, F. Y., Li, Y., Severn, A., Millott, S., Schmidt, J., Salter, M.,
characterization. J. Exp. Med. 182, 1301–1314. Moncada, S. (1991) A possible novel pathway of regulation by murine T
301. Taub, D. D., Conlon, K., Lloyd, A. R., Oppenheim, J. J., Kelvin, D. J. helper type-2 (Th2) cells of a Th1 cell activity via the modulation of the
(1993) Preferential migration of activated CD4⫹ and CD8⫹ T cells in induction of nitric oxide synthase on macrophages. Eur. J. Immunol. 21,
response to MIP-1 alpha and MIP-1 beta. Science 260, 355–358. 2489 –2494.
302. Appay, V., Rowland-Jones, S. L. (2001) RANTES: a versatile and con- 326. Rousset, F., Malefijt, R. W., Slierendregt, B., Aubry, J. P., Bonnefoy,
troversial chemokine. Trends Immunol. 22, 83– 87. J. Y., Defrance, T., Banchereau, J., de Vries, J. E. (1988) Regulation of
303. Hou, J., Baichwal, V., Cao, Z. (1994) Regulatory elements and transcrip- Fc receptor for IgE (CD23) and class II MHC antigen expression on
tion factors controlling basal and cytokine-induced expression of the Burkitt’s lymphoma cell lines by human IL-4 and IFN-gamma. J. Im-
gene encoding intercellular adhesion molecule 1. Proc. Natl. Acad. Sci. munol. 140, 2625–2632.
USA 91, 11641–11645. 327. Taniguchi, T., Takaoka, A. (2002) The interferon-alpha/beta system in
304. Jesse, T. L., LaChance, R., Iademarco, M. F., Dean, D. C. (1998) antiviral responses: a multimodal machinery of gene regulation by the
Interferon regulatory factor-2 is a transcriptional activator in muscle IRF family of transcription factors. Curr. Opin. Immunol. 14, 111–116.
where it regulates expression of vascular cell adhesion molecule-1. 328. Lu, B., Ebensperger, C., Dembic, Z., Wang, Y., Kvatyuk, M., Lu, T.,
J. Cell Biol. 140, 1265–1276. Coffman, R. L., Pestka, S., Rothman, P. B. (1998) Targeted disruption of
305. Billiau, A. (1996) Interferon-gamma: biology and role in pathogenesis. the interferon-gamma receptor 2 gene results in severe immune defects
Adv. Immunol. 62, 61–130. in mice. Proc. Natl. Acad. Sci. USA 95, 8233– 8238.
307. York, I. A., Rock, K. L. (1996) Antigen processing and presentation by 329. Cantin, E., Tanamachi, B., Openshaw, H. (1999) Role for gamma inter-
the class I major histocompatibility complex. Annu. Rev. Immunol. 14, feron in control of herpes simplex virus type 1 reactivation. J. Virol. 73,
369 –396. 3418 –3423.
308. Grandea III, A. G., Androlewicz, M. J., Athwal, R. S., Geraghty, D. E., 330. Beretta, L., Gabbay, M., Berger, R., Hanash, S. M., Sonenberg, N. (1996)
Spies, T. (1995) Dependence of peptide binding by MHC class I mole- Expression of the protein kinase PKR in modulated by IRF-1 and is
cules on their interaction with TAP. Science 270, 105–108. reduced in 5q- associated leukemias. Oncogene 12, 1593–1596.
309. Anderson, S. L., Shen, T., Lou, J., Xing, L., Blachere, N. E., Srivastava, 331. Hovanessian, A. G., Galabru, J. (1987) The double-stranded RNA-
P. K., Rubin, B. Y. (1994) The endoplasmic reticular heat shock protein dependent protein kinase is also activated by heparin. Eur. J. Biochem.
gp96 is transcriptionally upregulated in interferon-treated cells. J. Exp. 167, 467– 473.
Med. 180, 1565–1569. 332. Goodbourn, S., Didcock, L., Randall, R. E. (2000) Interferons: cell
310. Suto, R., Srivastava, P. K. (1995) A mechanism for the specific immu- signalling, immune modulation, antiviral response and virus countermea-
nogenicity of heat shock protein-chaperoned peptides. Science 269, sures. J. Gen. Virol. 81, 2341–2364.
1585–1588. 333. Lee, S. B., Melkova, Z., Yan, W., Williams, B. R., Hovanessian, A. G.,
311. Chang, C. H., Guerder, S., Hong, S. C., van Ewijk, W., Flavell, R. A. Esteban, M. (1993) The interferon-induced double-stranded RNA-acti-
(1996) Mice lacking the MHC class II transactivator (CIITA) show vated human p68 protein kinase potently inhibits protein synthesis in
tissue-specific impairment of MHC class II expression. Immunity 4, cultured cells. Virology 192, 380 –385.
167–178. 334. Chu, W. M., Ostertag, D., Li, Z. W., Chang, L., Chen, Y., Hu, Y.,
312. Griscelli, C., Lisowska-Grospierre, B., Mach, B. (1989) Combined im- Williams, B., Perrault, J., Karin, M. (1999) JNK2 and IKKbeta are
munodeficiency with defective expression in MHC class II genes. Im- required for activating the innate response to viral infection. Immunity
munodefic. Rev. 1, 135–153. 11, 721–731.
313. Chang, C. H., Fontes, J. D., Peterlin, M., Flavell, R. A. (1994) Class II 335. Zhou, A., Paranjape, J. M., Der, S. D., Williams, B. R., Silverman, R. H.
transactivator (CIITA) is sufficient for the inducible expression of major (1999) Interferon action in triply deficient mice reveals the existence of
histocompatibility complex class II genes. J. Exp. Med. 180, 1367– alternative antiviral pathways. Virology 258, 435– 440.
1374. 336. Presti, R. M., Popkin, D. L., Connick, M., Paetzold, S., Virgin IV, H. W.
314. Waldburger, J. M., Masternak, K., Muhlethaler-Mottet, A., Villard, J., (2001) Novel cell type-specific antiviral mechanism of interferon gamma
Peretti, M., Landmann, S., Reith, W. (2000) Lessons from the bare action in macrophages. J. Exp. Med. 193, 483– 496.
lymphocyte syndrome: molecular mechanisms regulating MHC class II 337. Xaus, J., Comalada, M., Valledor, A. F., Cardo, M., Herrero, C., Soler, C.,
expression. Immunol. Rev. 178, 148 –165. Lloberas, J., Celada, A. (2001) Molecular mechanisms involved in mac-
315. Boss, J. M. (1997) Regulation of transcription of MHC class II genes. rophage survival, proliferation, activation or apoptosis. Immunobiology
Curr. Opin. Immunol. 9, 107–113. 204, 543–550.

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 to Leishmania and other intracellular pathogens. J. Leukoc. Biol. 50,
granulocyte-macrophage colony-stimulating factor, macrophage colony- 93–103.
stimulating factor, and IFN-gamma lead to enhanced proliferation of 360. Green, S. J., Crawford, R. M., Hockmeyer, J. T., Meltzer, M. S., Nacy,
murine macrophage progenitor cells. J. Immunol. 147, 1542–1547. C. A. (1990) Leishmania major amastigotes initiate the L-arginine-
339. Balkwill, F., Taylor-Papadimitriou, J. (1978) Interferon affects both G1 dependent killing mechanism in IFN-gamma-stimulated macrophages
and S⫹G2 in cells stimulated from quiescence to growth. Nature 274, by induction of tumor necrosis factor-alpha. J. Immunol. 145, 4290 –
798 – 800. 4297.
340. Roos, G., Leanderson, T., Lundgren, E. (1984) Interferon-induced cell 361. Denis, M. (1991) Tumor necrosis factor and granulocyte macrophage-
cycle changes in human hematopoietic cell lines and fresh leukemic colony stimulating factor stimulate human macrophages to restrict growth
cells. Cancer Res. 44, 2358 –2362. of virulent Mycobacterium avium and to kill avirulent M. avium: killing
341. Lundblad, D., Lundgren, E. (1981) Block of glioma cell line in S by effector mechanism depends on the generation of reactive nitrogen
interferon. Int. J. Cancer 27, 749 –754. intermediates. J. Leukoc. Biol. 49, 380 –387.
342. Tiefenbrun, N., Melamed, D., Levy, N., Resnitzky, D., Hoffman, I., Reed, 362. Wei, X. Q., Charles, I. G., Smith, A., Ure, J., Feng, G. J., Huang, F. P.,
S. I., Kimchi, A. (1996) Alpha interferon suppresses the cyclin D3 and Xu, D., Muller, W., Moncada, S., Liew, F. Y. (1995) Altered immune
cdc25A genes, leading to a reversible G0-like arrest. Mol. Cell. Biol. 16, responses in mice lacking inducible nitric oxide synthase. Nature 375,
3934 –3944. 408 – 411.
343. Sangfelt, O., Erickson, S., Einhorn, S., Grander, D. (1997) Induction of
363. Karupiah, G., Chen, J. H., Nathan, C. F., Mahalingam, S., MacMicking,
Cip/Kip and Ink4 cyclin dependent kinase inhibitors by interferon-alpha
J. D. (1998) Identification of nitric oxide synthase 2 as an innate
in hematopoietic cell lines. Oncogene 14, 415– 423.
resistance locus against ectromelia virus infection. J. Virol. 72, 7703–
344. Hobeika, A. C., Subramaniam, P. S., Johnson, H. M. (1997) IFNalpha
7706.
induces the expression of the cyclin-dependent kinase inhibitor p21 in
364. Guidotti, L. G., McClary, H., Loudis, J. M., Chisari, F. V. (2000) Nitric
human prostate cancer cells. Oncogene 14, 1165–1170.
345. Chin, Y. E., Kitagawa, M., Su, W. C., You, Z. H., Iwamoto, Y., Fu, X. Y. oxide inhibits hepatitis B virus replication in the livers of transgenic
(1996) Cell growth arrest and induction of cyclin-dependent kinase mice. J. Exp. Med. 191, 1247–1252.
inhibitor p21 WAF1/CIP1 mediated by STAT1. Science 272, 719 –722. 365. Kwon, N. S., Nathan, C. F., Stuehr, D. J. (1989) Reduced biopterin as a
346. Takami, K., Takuwa, N., Okazaki, H., Kobayashi, M., Ohtoshi, T., cofactor in the generation of nitrogen oxides by murine macrophages.
Kawasaki, S., Dohi, M., Yamamoto, K., Nakamura, T., Tanaka, M., J. Biol. Chem. 264, 20496 –20501.
Nakahara, K., Takuwa, Y., Takizawa, H. (2002) Interferon-gamma in- 366. Drapier, J. C., Wietzerbin, J., Hibbs Jr., J. B. (1988) Interferon-gamma
hibits hepatocyte growth factor-stimulated cell proliferation of human and tumor necrosis factor induce the L-arginine-dependent cytotoxic
bronchial epithelial cells: upregulation of p27(kip1) cyclin-dependent effector mechanism in murine macrophages. Eur. J. Immunol. 18,
kinase inhibitor. Am. J. Respir. Cell Mol. Biol. 26, 231–238. 1587–1592.
347. Dey, A., Kim, L., Li, W. (1999) Gamma interferon induces expression of 367. Morris Jr., S. M., Billiar, T. R. (1994) New insights into the regulation of
Mad1 gene in macrophage, which inhibits colony-stimulating factor-1- inducible nitric oxide synthesis. Am. J. Physiol. 266, E829 –E839.
dependent mitogenesis. J. Cell. Biochem. 72, 232–241. 368. Kamijo, R., Shapiro, D., Le, J., Huang, S., Aguet, M., Vilcek, J.
348. Thornberry, N. A. (1998) Caspases: key mediators of apoptosis. Chem. (1993) Generation of nitric oxide and induction of major histocom-
Biol. 5, R97–103. patibility complex class II antigen in macrophages from mice lacking
349. Novelli, F., Bernabei, P., Ozmen, L., Rigamonti, L., Allione, A., Pestka, the interferon gamma receptor. Proc. Natl. Acad. Sci. USA 90,
S., Garotta, G., Forni, G. (1996) Switching on of the proliferation or 6626 – 6630.
apoptosis of activated human T lymphocytes by IFN-gamma is correlated 369. Kamijo, R., Harada, H., Matsuyama, T., Bosland, M., Gerecitano, J.,
with the differential expression of the alpha- and beta-chains of its Shapiro, D., Le, J., Koh, S. I., Kimura, T., Green, S. J., et al. (1994)
receptor. J. Immunol. 157, 1935–1943. Requirement for transcription factor IRF-1 in NO synthase induction in
350. Bernabei, P., Allione, A., Rigamonti, L., Bosticardo, M., Losana, G., macrophages. Science 263, 1612–1615.
Borghi, I., Forni, G., Novelli, F. (2001) Regulation of interferon-gamma 370. Radi, R., Beckman, J. S., Bush, K. M., Freeman, B. A. (1991) Peroxyni-
receptor (INF-gammaR) chains: a peculiar way to rule the life and death trite-induced membrane lipid peroxidation: the cytotoxic potential of
of human lymphocytes. Eur. Cytokine Netw. 12, 6 –14. superoxide and nitric oxide. Arch. Biochem. Biophys. 288, 481– 487.
351. Los, M., Van de Craen, M., Penning, L. C., Schenk, H., Westendorp, M., 371. Radi, R., Beckman, J. S., Bush, K. M., Freeman, B. A. (1991) Peroxyni-
Baeuerle, P. A., Droge, W., Krammer, P. H., Fiers, W., Schulze-Osthoff, trite oxidation of sulfhydryls. The cytotoxic potential of superoxide and
K. (1995) Requirement of an ICE/CED-3 protease for Fas/APO-1-medi- nitric oxide. J. Biol. Chem. 266, 4244 – 4250.
ated apoptosis. Nature 375, 81– 83. 372. Cassatella, M. A., Hartman, L., Perussia, B., Trinchieri, G. (1989) Tumor
352. Miura, M., Zhu, H., Rotello, R., Hartwieg, E. A., Yuan, J. (1993) necrosis factor and immune interferon synergistically induce cytochrome
Induction of apoptosis in fibroblasts by IL-1 beta-converting enzyme, a b-245 heavy-chain gene expression and nicotinamide-adenine dinucle-
mammalian homolog of the C. elegans cell death gene ced-3. Cell 75, otide phosphate hydrogenase oxidase in human leukemic myeloid cells.
653– 660. J. Clin. Invest. 83, 1570 –1579.
353. Mathiak, G., Grass, G., Herzmann, T., Luebke, T., Zetina, C. C., Boehm, 373. Pfefferkorn, E. R. (1984) Interferon gamma blocks the growth of Toxo-
S. A., Bohlen, H., Neville, L. F., Hoelscher, A. H. (2000) Caspase-1- plasma gondii in human fibroblasts by inducing the host cells to degrade
inhibitor ac-YVAD-cmk reduces LPS-lethality in rats without affecting
tryptophan. Proc. Natl. Acad. Sci. USA 81, 908 –912.
haematology or cytokine responses. Br. J. Pharmacol. 131, 383–386.
374. Burke, F., Knowles, R. G., East, N., Balkwill, F. R. (1995) The role of
354. Fantuzzi, G., Dinarello, C. A. (1999) Interleukin-18 and interleukin-1
indoleamine 2,3-dioxygenase in the anti-tumour activity of human inter-
beta: two cytokine substrates for ICE (caspase-1). J. Clin. Immunol. 19,
feron-gamma in vivo. Int. J. Cancer 60, 115–122.
1–11.
355. Roos, D., Bolscher, B. G. J., de Boer, M. (1992) Generation of reactive 375. Hissong, B. D., Carlin, J. M. (1997) Potentiation of interferon-induced
oxygen species by phagocytes. In Mononuclear Phagocytes: Biology of indoleamine 2,3-dioxygenase mRNA in human mononuclear phagocytes
Monocytes and Macrophages (R. van Furth, ed.), Dordrecht, The Neth- by lipopolysaccharide and interleukin-1. J. Interferon Cytokine Res. 17,
erlands, Kluwer Academic, 243–253. 387–393.
356. Nathan, C. F., Gabay, J. (1992) Antimicrobial mechanisms of macro- 376. Carlin, J. M., Borden, E. C., Sondel, P. M., Byrne, G. I. (1987) Biologic-
phages. In Mononuclear Phagocytes: Biology of Monocytes and Macro- response-modifier-induced indoleamine 2,3-dioxygenase activity in hu-
phages (R. van Furth, ed.), Dordrecht, The Netherlands, Kluwer Aca- man peripheral blood mononuclear cell cultures. J. Immunol. 139,
demic, 259 –267. 2414 –2418.
357. Shiloh, M. U., MacMicking, J. D., Nicholson, S., Brause, J. E., Potter, S., 377. Hassanain, H. H., Chon, S. Y., Gupta, S. L. (1993) Differential regulation
Marino, M., Fang, F., Dinauer, M., Nathan, C. (1999) Phenotype of mice of human indoleamine 2,3-dioxygenase gene expression by interferons-
and macrophages deficient in both phagocyte oxidase and inducible gamma and -alpha. Analysis of the regulatory region of the gene and
nitric oxide synthase. Immunity 10, 29 –38. identification of an interferon-gamma-inducible DNA-binding factor.
358. Karupiah, G., Xie, Q. W., Buller, R. M., Nathan, C., Duarte, C., Mac- J. Biol. Chem. 268, 5077–5084.
Micking, J. D. (1993) Inhibition of viral replication by interferon-gamma- 378. Fleckner, J., Martensen, P. M., Tolstrup, A. B., Kjeldgaard, N. O.,
induced nitric oxide synthase. Science 261, 1445–1448. Justesen, J. (1995) Differential regulation of the human, interferon in-
359. Green, S. J., Nacy, C. A., Meltzer, M. S. (1991) Cytokine-induced ducible tryptophanyl-tRNA synthetase by various cytokines in cell lines.
synthesis of nitrogen oxides in macrophages: a protective host response 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 402. Jahnke, A., Johnson, J. P. (1994) Synergistic activation of intercellular
mammalian tryptophanyl-tRNA synthetase: new perspectives. Trends adhesion molecule 1 (ICAM-1) by TNF-alpha and IFN-gamma is medi-
Biochem. Sci. 18, 263–267. ated by p65/p50 and p65/c-Rel and interferon-responsive factor Stat1
380. Xue, H., Wong, J. T. (1995) Interferon induction of human tryptophanyl- alpha (p91) that can be activated by both IFN-gamma and IFN-alpha.
tRNA synthetase safeguards the synthesis of tryptophan-rich immune- FEBS Lett. 354, 220 –226.
system proteins: a hypothesis. Gene 165, 335–339. 403. Cheshire, J. L., Baldwin Jr., A. S. (1997) Synergistic activation of
381. Collins, J. T., Dunnick, W. A. (1993) Germline transcripts of the murine NF-kappaB by tumor necrosis factor alpha and gamma interferon via
immunoglobulin gamma 2a gene: structure and induction by IFN-gamma. enhanced I kappaB alpha degradation and de novo I kappaB beta
Int. Immunol. 5, 885– 891. degradation. Mol. Cell. Biol. 17, 6746 – 6754.
382. Snapper, C. M., Paul, W. E. (1987) Interferon-gamma and B cell stim- 404. Ohmori, Y., Schreiber, R. D., Hamilton, T. A. (1997) Synergy between
ulatory factor-1 reciprocally regulate Ig isotype production. Science 236, interferon-gamma and tumor necrosis factor-alpha in transcriptional ac-
944 –947. tivation is mediated by cooperation between signal transducer and acti-
383. Sweet, M. J., Hume, D. A. (1996) Endotoxin signal transduction in vator of transcription 1 and nuclear factor kappaB. J. Biol. Chem. 272,
macrophages. J. Leukoc. Biol. 60, 8 –26. 14899 –14907.
384. Akira, S. (2000) Toll-like receptors: lessons from knockout mice. Bio- 405. Fujita, T., Reis, L. F., Watanabe, N., Kimura, Y., Taniguchi, T., Vilcek,
chem. Soc. Trans. 28, 551–556. J. (1989) Induction of the transcription factor IRF-1 and interferon-beta
385. Aderem, A., Ulevitch, R. J. (2000) Toll-like receptors in the induction of
mRNAs by cytokines and activators of second-messenger pathways.
the innate immune response. Nature 406, 782–787.
Proc. Natl. Acad. Sci. USA 86, 9936 –9940.
386. Dobrovolskaia, M. A., Vogel, S. N. (2002) Toll receptors, CD14, and
406. Han, Y., Rogers, N., Ransohoff, R. M. (1999) Tumor necrosis factor-
macrophage activation and deactivation by LPS. Microbes Infect. 4,
alpha signals to the IFN-gamma receptor complex to increase Stat1alpha
903–914.
activation. J. Interferon Cytokine Res. 19, 731–740.
387. Kawai, T., Adachi, O., Ogawa, T., Takeda, K., Akira, S. (1999) Unre-
sponsiveness of MyD88-deficient mice to endotoxin. Immunity 11, 115– 407. Kumar, A., Haque, J., Lacoste, J., Hiscott, J., Williams, B. R. (1994)
122. Double-stranded RNA-dependent protein kinase activates transcription
388. Lorsbach, R. B., Murphy, W. J., Lowenstein, C. J., Snyder, S. H., Russell, factor NF-kappa B by phosphorylating I kappa B. Proc. Natl. Acad. Sci.
S. W. (1993) Expression of the nitric oxide synthase gene in mouse USA 91, 6288 – 6292.
macrophages activated for tumor cell killing. Molecular basis for the 408. Tannenbaum, C. S., Major, J. A., Hamilton, T. A. (1993) IFN-gamma and
synergy between interferon-gamma and lipopolysaccharide. J. Biol. lipopolysaccharide differentially modulate expression of tumor necrosis
Chem. 268, 1908 –1913. factor receptor mRNA in murine peritoneal macrophages. J. Immunol.
389. Jurkovich, G. J., Mileski, W. J., Maier, R. V., Winn, R. K., Rice, C. L. 151, 6833– 6839.
(1991) Interferon gamma increases sensitivity to endotoxin. J. Surg. Res. 409. Aggarwal, B. B., Eessalu, T. E., Hass, P. E. (1985) Characterization of
51, 197–203. receptors for human tumour necrosis factor and their regulation by
390. Sweet, M. J., Stacey, K. J., Kakuda, D. K., Markovich, D., Hume, D. A. gamma-interferon. Nature 318, 665– 667.
(1998) IFN-gamma primes macrophage responses to bacterial DNA. 410. Michishita, M., Yoshida, Y., Uchino, H., Nagata, K. (1990) Induction of
J. Interferon Cytokine Res. 18, 263–271. tumor necrosis factor-alpha and its receptors during differentiation in
391. Vadiveloo, P. K., Vairo, G., Hertzog, P., Kola, I., Hamilton, J. A. (2000) myeloid leukemic cells along the monocytic pathway. A possible regu-
Role of type I interferons during macrophage activation by lipopolysac- latory mechanism for TNF-alpha production. J. Biol. Chem. 265, 8751–
charide. Cytokine 12, 1639 –1646. 8759.
392. Costelloe, E. A. (1997) The regulation and function of murine PAI-2. In 411. Aggarwal, B. B., Eessalu, T. E. (1987) Induction of receptors for tumor
Centre for Molecular and Cellular Biology, Brisbane, Australia, Biochem- necrosis factor-alpha by interferons is not a major mechanism for their
istry Department, University of Queensland, 158. synergistic cytotoxic response. J. Biol. Chem. 262, 10000 –10007.
393. Williams, J. G., Jurkovich, G. J., Hahnel, G. B., Maier, R. V. (1992) 412. Rhee, S. H., Jones, B. W., Toshchakov, V., Vogel, S. N., Fenton, M. J.
Macrophage priming by interferon gamma: a selective process with (2003) Toll-like receptor 2 and 4 activate STAT1 serine phosphorylation
potentially harmful effects. J. Leukoc. Biol. 52, 579 –584. by distinct mechanisms in macrophages. J. Biol. Chem. 278, 22506 –
394. Car, B. D., Eng, V. M., Schnyder, B., Ozmen, L., Huang, S., Gallay, P., 22512.
Heumann, D., Aguet, M., Ryffel, B. (1994) Interferon gamma receptor 413. Gao, J. J., Filla, M. B., Fultz, M. J., Vogel, S. N., Russell, S. W.,
deficient mice are resistant to endotoxic shock. J. Exp. Med. 179, Murphy, W. J. (1998) Autocrine/paracrine IFN-alphabeta mediates
1437–1444. the lipopolysaccharide-induced activation of transcription factor
395. Bosisio, D., Polentarutti, N., Sironi, M., Bernasconi, S., Miyake, K., Stat1alpha in mouse macrophages: pivotal role of Stat1alpha in in-
Webb, G. R., Martin, M. U., Mantovani, A., Muzio, M. (2002) Stimulation duction of the inducible nitric oxide synthase gene. J. Immunol. 161,
of Toll-like receptor 4 expression in human mononuclear phagocytes by 4803– 4810.
interferon-gamma: a molecular basis for priming and synergism with 414. Gessani, S., Belardelli, F., Pecorelli, A., Puddu, P., Baglioni, C. (1989)
bacterial lipopolysaccharide. Blood 99, 3427–3431. Bacterial lipopolysaccharide and gamma interferon induce transcription
396. Mita, Y., Dobashi, K., Shimizu, Y., Nakazawa, T., Mori, M. (2001) of beta interferon mRNA and interferon secretion in murine macro-
Toll-like receptor 2 and 4 surface expressions on human monocytes are phages. J. Virol. 63, 2785–2789.
modulated by interferon-gamma and macrophage colony-stimulating fac-
415. Toshchakov, V., Jones, B. W., Perera, P. Y., Thomas, K., Cody, M. J.,
tor. Immunol. Lett. 78, 97–101.
Zhang, S., Williams, B. R., Major, J., Hamilton, T. A., Fenton, M. J.,
397. Darmani, H., Parton, J., Harwood, J. L., Jackson, S. K. (1994) Interferon-
Vogel, S. N. (2002) TLR4, but not TLR2, mediates IFN-beta-induced
gamma and polyunsaturated fatty acids increase the binding of lipopoly-
STAT1alpha/beta-dependent gene expression in macrophages. Nat. Im-
saccharide to macrophages. Int. J. Exp. Pathol. 75, 363–368.
398. Adib-Conquy, M., Cavaillon, J. M. (2002) Gamma interferon and gran- munol. 3, 392–398.
ulocyte/monocyte colony-stimulating factor prevent endotoxin tolerance 416. Ohmori, Y., Hamilton, T. A. (2001) Requirement for STAT1 in
in human monocytes by promoting interleukin-1 receptor-associated LPS-induced gene expression in macrophages. J. Leukoc. Biol. 69,
kinase expression and its association to MyD88 and not by modulating 598 – 604.
TLR4 expression. J. Biol. Chem. 277, 27927–27934. 417. Kawai, T., Takeuchi, O., Fujita, T., Inoue, J., Muhlradt, P. F., Sato, S.,
399. Held, T. K., Weihua, X., Yuan, L., Kalvakolanu, D. V., Cross, A. S. Hoshino, K., Akira, S. (2001) Lipopolysaccharide stimulates the MyD88-
(1999) Gamma interferon augments macrophage activation by lipopoly- independent pathway and results in activation of IFN-regulatory factor 3
saccharide by two distinct mechanisms, at the signal transduction level and the expression of a subset of lipopolysaccharide-inducible genes.
and via an autocrine mechanism involving tumor necrosis factor alpha J. Immunol. 167, 5887–5894.
and interleukin-1. Infect. Immun. 67, 206 –212. 418. Ohmori, Y., Hamilton, T. A. (1995) The interferon-stimulated response
400. de Wit, H., Hoogstraten, D., Halie, R. M., Vellenga, E. (1996) Interferon- element and a kappa B site mediate synergistic induction of murine
gamma modulates the lipopolysaccharide-induced expression of AP-1 IP-10 gene transcription by IFN-gamma and TNF-alpha. J. Immunol.
and NF-kappa B at the mRNA and protein level in human monocytes. 154, 5235–5244.
Exp. Hematol. 24, 228 –235. 419. Gao, J., Morrison, D. C., Parmely, T. J., Russell, S. W., Murphy, W. J.
401. Huang, Y., Krein, P. M., Muruve, D. A., Winston, B. W. (2002) Com- (1997) An interferon-gamma-activated site (GAS) is necessary for full
plement factor B gene regulation: synergistic effects of TNF-alpha and expression of the mouse iNOS gene in response to interferon-gamma and
IFN-gamma in macrophages. J. Immunol. 169, 2627–2635. 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 feron-gamma involves binding of distinct factors to a palindromic re-
factor NF-kappa B/Rel in induction of nitric oxide synthase. J. Biol. sponse element. J. Biol. Chem. 269, 21146 –21154.
Chem. 269, 4705– 4708. 423. Look, D. C., Pelletier, M. R., Holtzman, M. J. (1994) Selective interaction
421. Ledebur, H. C., Parks, T. P. (1995) Transcriptional regulation of the of a subset of interferon-gamma response element-binding proteins with
intercellular adhesion molecule-1 gene by inflammatory cytokines in the intercellular adhesion molecule-1 (ICAM-1) gene promoter controls
human endothelial cells. Essential roles of a variant NF-kappa B site and the pattern of expression on epithelial cells. J. Biol. Chem. 269, 8952–
p65 homodimers. J. Biol. Chem. 270, 933–943. 8958.
422. Caldenhoven, E., Coffer, P., Yuan, J., Van de Stolpe, A., Horn, F., 424. Hume, D. A. (2000) Probability in transcriptional regulation and its
Kruijer, W., Van der Saag, P. T. (1994) Stimulation of the human implications for leukocyte differentiation and inducible gene expression.
intercellular adhesion molecule-1 promoter by interleukin-6 and inter- Blood 96, 2323–2328.

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

You might also like