Skip to content
BY-NC-ND 4.0 license Open Access Published by De Gruyter May 30, 2018

Interferon-gamma (IFN-γ): Exploring its implications in infectious diseases

  • Gunjan Kak EMAIL logo , Mohsin Raza and Brijendra K Tiwari
From the journal Biomolecular Concepts

Abstract

A key player in driving cellular immunity, IFN-γ is capable of orchestrating numerous protective functions to heighten immune responses in infections and cancers. It can exhibit its immunomodulatory effects by enhancing antigen processing and presentation, increasing leukocyte trafficking, inducing an anti-viral state, boosting the anti-microbial functions and affecting cellular proliferation and apoptosis. A complex interplay between immune cell activity and IFN-γ through coordinated integration of signals from other pathways involving cytokines and Pattern Recognition Receptors (PRRs) such as Interleukin (IL)-4, TNF-α, Lipopolysaccharide (LPS), Type-I Interferons (IFNS) etc. leads to initiation of a cascade of pro-inflammatory responses. Microarray data has unraveled numerous genes whose transcriptional regulation is influenced by IFN-γ. Consequently, IFN-γ stimulated cells display altered expression of many such target genes which mediate its downstream effector functions. The importance of IFN-γ is further reinforced by the fact that mice possessing disruptions in the IFN-γ gene or its receptor develop extreme susceptibility to infectious diseases and rapidly succumb to them. In this review, we attempt to elucidate the biological functions and physiological importance of this versatile cytokine. The functional implications of its biological activity in several infectious diseases and autoimmune pathologies are also discussed. As a counter strategy, many virulent pathogenic species have devised ways to thwart IFN-γ endowed immune-protection. Thus, IFN-γ mediated host-pathogen interactions are critical for our understanding of disease mechanisms and these aspects also manifest enormous therapeutic importance for the annulment of various infections and autoimmune conditions.

IFN-γ: An Introduction

The human immune system is evolved to eradicate or contain any pathogenic challenge and eliminate self-altered cancerous cells. In this regard, IFN-γ has a critical role in recognizing and eliminating pathogens. IFN-γ, being the central effector of cell mediated immunity, can coordinate a plethora of anti-microbial functions. It can serve to amplify antigen presentation through antigen presenting cells (APCs) by enhancing antigen recognition via cognate T-cell interaction, increase the production of Reactive Oxygen Species (ROS) and Reactive Nitrogen Intermediates (RNIs) and induce anti-viral responses [1]. Additionally, cancer cells are destroyed by IFN-γ activity via induction of an anti-proliferative state. Immunity to several pathogens is mainly governed by IFN-γ activity. For example, the role of IFN-γ in endowing protection against Chlamydial infections is quite immense [2]. Release of IFN-γ by the CD4+ helper T cell population contributes to necessary effector cell activation and the generation of antibody mediated responses to Chlamydia infections [3]. The presence of IFN-γ is of absolute necessity in combating mycobacterial infections through its ability to regulate various protective functions and sustain both CD4+ and CD8+ cell activity [4]. The protective benefits of IFN-γ can also been seen in the context of viral infections, as enhanced survival of neurons infected with varicella zoster virus is observed post IFN-γ treatment [5]. Natural Killer (NK) cell -mediated IFN-γ production can successfully limit Hepatitis C virus proliferation in HIV+ (Human Immunodeficiency Virus) patients [6]. Enhanced anti-bacterial and immune protective effects simultaneous with pro-inflammatory responses leading to protection of epithelial monolayers from pathogen mediated injury are also conferred by IFN-γ during Staphylococcus aureus infections [7]. Thus, IFN-γ is a robust protective effector molecule mediating protection against a wide array of pathogenic entities.

IFN-γ increases the efficiency of immune system by enhancing its competence to deliver anti-microbial effector functions. Undoubtedly, such an important immune mediator is under stringent regulatory control. Over-activity of IFN-γ has been reported to cause excessive tissue damage, necrosis and inflammation and may contribute to disease pathology. Hyper-activity of both IFN-γ and IL-18 can exacerbate pathogenesis in Burkholderia infections [8]. Aberrant production of IFN-γ has been linked with autoimmunity and alterations in gut flora [9,10]. Therefore, IFN-γ activity is a double edged sword and immune regulatory mechanisms strive to strike a delicate balance between infection control and disease pathology in this regard. Foxp3+ (Forkhead box P3) -expressing regulatory T cells (Tregs) and suppressor of cytokine signaling protein-1 (SOCS-1) are critical controllers of IFN-γ activity in this regard.

IFN-γ: Production and signaling

Interferons were initially described as agents interfering with viral multiplication and eliciting potent anti-viral activity. However, with the passage of time, they have been attributed with a vast array of physiological activities and are more than just anti-viral agents. The main producers of IFN-γ include T-cells, NK cells, NKT cells, professional APCs such as macrophages, dendritic cells (DCs) and B cells. The role of IFN-γ secreted by NK cells and APCs has been implicated in early host defense and autocrine regulation [11]. T cells are the major source of IFN-γ during adaptive immunity [12].

IFN-γ responses are an offshoot of receptor-mediated signaling. Its biological effects are manifested through its interaction with the receptor (IFN-γR) present on target cells such as macrophages, DCs and many other cell types. The receptor is a heterodimeric complex composed of a ligand binding alpha subunit (IFNGR1) and a signal transducing beta subunit (IFNGR2). IFN-γ binds to IFNGR1 with relatively higher affinity compared with IFNGR2. Receptor-ligand interaction initiates a signaling cascade that in turn produces a series of protective responses [13]. Binding of IFN-γ crosslinks IFNGR1 and IFNGR2 and activates Janus Kinases JAK1 and JAK2 through phosphorylation. Activated JAKs in turn phosphorylate inactive cytosolic transcription factor STAT1 (signal transducer and activator of transcription 1) [14]. Activated and phosphorylated STAT1 homodimerizes and is translocated to the nucleus where it binds to GAS (gamma activated sequences) elements in the promoter region and mediates transcription of IFN-γ associated genes required for protective immunity. IFN-γ-IFN-γR signaling triggers STAT1 phosphorylation and subsequent activation of transcription factor T-bet which depicts the onset of T cell polarization for commitment to TH1 lineage [15]. Recruitment of an additional factor, Runx3, which binds to the IFN-γ promoter, drives the expression of IFN-γ with concurrent silencing of the IL-4 gene [16]. Thus, IFN-γ plays a pivotal role in dampening the IL-4 -producing ability of TH1 cells and in maintaining sustained expression of T-bet [17].

An overview of IFN-γ activity

Macrophages represent a class of APCs which are versatile sentinels of the immune system. A pathogen is most likely to encounter a macrophage soon after its entry into the host and IFN-γ-stimulated macrophages show enhanced anti-microbial activity. Thus, IFN-γ activates macrophages and makes them better able to mount an effective immune response, such as enhanced antigen processing and presentation through upregulation of class II MHC, increased ROS and NOS production, induction of autophagy for clearance of intracellular pathogens and increased secretion of pro-inflammatory cytokines. Additionally, it can activate NK cells, increase their tumoricidal activity and also regulate antibody production to modulate B cell responses. Growth and maturation of other cell types and leukocyte migration are also facilitated by IFN-γ. The heightened immune activation ultimately leads to effective clearance of pathogens through enhanced phagocytosis, pro-inflammatory responses and lymphocyte recruitment.

Anti-microbial effector functions of IFN-γ

IFN-γ is endowed with the exquisite ability to promote rapid acidification of phagolysosomes within infected macrophages. This low pH within the phagolysosome ameliorates RNS production and this cooperation leads to elimination of the pathogen [18]. IFN-γ is a potent inducer of autophagy which has now emerged as a novel antimicrobial host response [19,20]. Immunity-related GTPases (IRGs) comprise another very interesting class of IFN-inducible proteins. IFN-γ stimulation upregulates several IRGs which facilitate recruitment of autophagic machinery onto the bacterial phagosomes. Human IRGM helps in trafficking various components of autophagy pathway to vacuoles containing Mycobacterium tuberculosis (M. tb) [20]. The IRGs primarily orchestrate vacuolar trafficking and facilitate pathogen delivery to lysosomes for enzymemediated degradation. These may function in conjunction with certain other proteins, namely sequestosome (SQSTM1/p62), NDP52 and optineurin. These proteins mainly recognize ubiquitin-bearing bacteria that have successfully escaped the confines of the phagosomal compartment. On the contrary, galactins, another set of IFN-inducible proteins, recognize bacterial glycan moieties. The recognition and subsequent binding recruits the autophagy proteins and directs the bound components towards lysosomes [21]. Activation of M. tb-infected macrophages by IFN-γ stimulates the expression of IRGs that eventually coordinate vacuolar traffic and subsequent targeting of bacterial cargo for lysosomal hydrolysis [22]. Orchestration of cell autonomous immune responses by coordinating vacuolar compartmentalization is effective against several intracellular pathogens including Chlamydia psittaci, C. trachomatis and Salmonella Typhimurium. Additionally, Guanylate binding proteins (GBPs), expressed as a functional IFN-γ response also bind escaped bacteria and channelize them towards lysosomes [23]. Their salient structural aspects enable them to identify both phagosomal entrapped bacteria as well as free bacteria. For example, GBP1 can facilitate bacterial binding to lysosomes via interactions with SQSTM1. On the contrary, GBP7 engages autophagy proteins (ATG4B) driving extension of the elongation membrane around bacteria to sequester them completely within the autophagosomal compartment for degradation [21]. Recently, a protective house-keeping role of GBP1 has been observed at the intestinal tight junctions. Being expressed at these places, it serves to control mucosal immunity by exerting regulatory effects on apoptosis [24].

Interestingly, IFN-γ can also deplete tryptophan, touted to be an effective anti-parasitic mechanism in humans [25]. The anti-viral effects invoked by IFN-γ are most notably through the induction of RNA-activated protein kinase R (PKR) and adenosine deaminase RNA specific-1 (ADAR-1). The anti-viral effectors either impact viral multiplication or genomic stability. Interferon-induced transmembrane proteins (IFITMs) and tripartite motif-containing proteins (TRIMs) are another set of effector proteins that also elicit marked anti-viral properties. IFITMs restrict viral uncoating or entry into the host cells and confer protection against a number of viruses including Influenza-A, Flaviviruses, HIV-1, Ebola virus and Corona virus [26]. More recently it was shown that IFITM3 restricts virions inside the endosome to block their access to the cytoplasm [27]. On the other hand, TRIMs are induced in macrophages and myeloid DCs and function to limit viral entry, most notably that of retroviruses [28]. IFN-γ-inducible proteins can also employ components of the autophagic machinery to curb viral replication. For example, IFN-inducible GTPases utilize the microtubule-associated-protein-1-light chain-3 (LC3) protein of the autophagy pathway to arrest murine norovirus replication. The concerted action of IFN-inducible GTPases and GBPs leads to marked inhibition of viral growth in murine cells [29]. IFN-γ also regulates the production of proteins, viz. tethrin or viperin, involved in limiting and interfering with virion egress from cells or viral assembly. Tethrin acts as a restrictive factor against viruses such as Filo virus, HIV-1 or Arena virus by blocking viral exit from the infected macrophages [30].

IFN-γ regulates the production of various lysosomal constituents, granules or exudating substances (β-defensins, α-defensins and cathelicidins) mediating inflammation that display robust killing action against microbes. IFN-γ is a potent inducer of autophagy upon mycobacterial induction by eliciting production of hCAP18/LL-37 cathelicidin in a Vitamin D3-dependent manner [31]. Additionally, IFN-γ also induces of several efflux systems that deprive pathogens of essential cations, consequently limiting their growth within host. NRAMP1, natural resistance associated macrophage protein 1, is one such immune effector protein that is upregulated by IFN-γ. NRAMP1 expels Mn2+ or Fe2+, lowering cationic concentration. It can effectively compete with pathogens for essential cations and reduce pathogenic possession of these ions within the phagosome [32]. Alternatively, IFN-γ triggers the upregulation and relocation of P-type Cu2+ onto phagosomes facilitating higher inflow of Cu2+ cations to ameliorate ROS generation [33]. In a similar mechanism, IFN-γ significantly enhances the expression of iron-exuding ferroprotein 1 with concomitant down-modulation of transferrin receptor to diminish Fe2+ concentration, thereby restricting Salmonella growth [34].

Implications in infectious diseases

IFN-γ has emerged as an extremely versatile cytokine that can carry out countless biological activities that are non-redundant with other interferons. The literature is inundated with reports that emphasize its importance in disease pathologies. Consequently, any failure in the IFN-γ-IFN-γR system severely hampers host immune responses to infections. Thus, its relevance with regard to conferring protective immunity in infectious diseases and cancers is undisputable. In this section we discuss and highlight its protective role in mediating immune responses in infectious diseases, cancer or autoimmunity. Numerous evasive strategies have been devised by pathogens to overcome its curbing effects and circumvent the protection conferred by IFN-γ.

Anti-microbial effects of IFN-γ : role in mycobacterial infection

The role of IFN-γ is almost undoubted in mediating protection against tuberculosis (TB) [35]. The hallmark of TB infection is the induction of cellular immunity and inflammation orchestrated by IFN-γ. Quite plausibly, mice lacking functional IFN-γ or the IFN-γR gene are most susceptible to mycobacterial infections. Mice with a disrupted IFN-γ gene show rapid, fulminant mycobacterial growth and disseminated TB [36]. Dysfunctional IFN-γ or IFN-γR signaling increases susceptibility to even mildly virulent strains such as attenuated Mycobacterium bovis Bacillus Calmette-Guérin (BCG). Children with nonfunctional IFN-γR show disseminated infection with BCG [37]. Quite recently, a role of a novel IFN-γR adaptor has been implicated in the killing of M. tb. The Mal-dependent IFN-γR-mediated signaling is independent of the Tolllike Receptor (TLR) pathway and initiates a cascade of downstream responses involving phosphorylation by MAPK-p38 (Mitogen activated protein kinase) and induction of autophagy. Mutations in the Mal adaptor diminished its IFN-γR binding capacity which ultimately led to compromised signaling in response to IFN-γ and impaired immune responses to M. tb [38].

IFN-γ knock out mice manifest uncontrolled bacterial growth and rapidly develop necrotic granulomas. Such mice have severely hampered cell-mediated immunity and Nitric Oxide Synthase 2 (NOS2) expression. IFN-γ favors ROS, RNI production, revokes the blockade of phagosome-lysosome fusion imposed by M. tb as part of its survival strategy, and also induces autophagy to favor clearance of M. tb [39]. Apoptosis has been linked with effective anti-microbial host defense mechanisms against intracellular pathogens, especially M. tb. IFN-γ also promotes apoptosis in a NO-dependent manner in human macrophages [40]. IFN-γ in combination with TLR2 stimulation also facilitates the production antimicrobial peptides and induction of autophagy in a Vitamin-D-dependent pathway [41,42]. As a counter strategy, M. tb has evolved ways to subvert the deleterious effects of IFN-γ. M. tb averts the activating effects of IFN-γ on macrophages in a TLR-dependent manner. Many crucial immune protective responses exerted by IFN-γ are inhibited by M. tb. Various mycobacterial cell wall lipoproteins abrogate IFN-γ-mediated anti-bacterial effects. For example, the M. tb 19 kDa lipoprotein inhibits MHC class II upregulation and antigen processing in IFN-γ treated macrophages to dampen the protective effects of IFN-γ. This occurs due to repressed expression of class II transactivator (CIITA) in a TLR-dependent manner and is believed to be due to inhibitory effects imposed on the chromatin remodeling of CIITA involving the TLR2 and MAPK pathways [43,44]. Prolonged signaling induced by 19kDa protein through the TLR pathway has also been implicated in the downregulated expression of several IFN-γ-induced genes in macrophages [45]. Downstream transcriptional responses induced by IFN-γ are also inhibited by M. tb, although the proximal steps involved in IFN-γR signaling such as STAT1 phosphorylation and dimerization are unaffected [46]. Diminished IFN-γ-induced responses are attributed to reduced association of STAT1 dimers with its associated co-activators such as cAMP response element binding protein (CREB) and p300 in M. tb infected macrophages. This disrupted association leads to diluted macrophage activity, such as reduced killing of Toxoplasma gondii [47]. Interestingly, the role of TLR signaling has been linked with the induction of inhibitory responses to IFN-γ in mycobacterium infected macrophages. It was reported that TLR2 stimulation can abrogate IFN-γ responsive effects through stabilization and expression of a dominant negative form of STAT1β. Higher levels of STAT1β are induced upon M. avium infection which is attributed to increased STAT1β mRNA stability. Further, overexpression of STAT1β in macrophages leads to decreased IFN-γ gene expression [48]. STAT1 knock out (KO) mice are highly susceptible to TB and succumb very rapidly to the disease. These KO mice develop multiple necrotic and granulomatous lesions in the lungs with defective Tumor Necrosis Factor (TNF)-α, inducible NOS and IL-12 production [49].

Anti-microbial effects of IFN-γ: role in Salmonella infection

IFN-γ is critical in mediating intestinal immunity to Salmonella typhimurium [50,51]. Gene KO mice have revealed the importance of IFN-γ in promoting protective immunity. These mice show aberrant immune responses and disseminated septicemia as compared with wild type (WT) mice which are characterized by high CD4+, CD8+, Class II MHC and Vascular Cell Adhesion Molecule I (VCAM-1) expression [50]. IL-12 dependent IFN-γ production plays a key role in endowing protection against Salmonella infections. IL-23 is also a major regulatory cytokine governing IFN-γ production [52].

IL-18 is a major contributor in imparting protective immunity by being a potent inducer of IFN-γ [53]. However, in many instances such as those involving colitis and inflammation, IFN-γ can have a dual role. In such scenarios, IFN-γ can promote inflammation and exacerbate inflammatory responses in the intestines to aggravate Salmonella-induced colitis. In this case neutrophils become an important source of IFN-γ. As a corollary, depletion of neutrophils relieves many of the IFN-γ-induced disease symptoms; it curbs excessive inflammation and decreases the severity of intestinal lesions [54]. IFN-γ has also been implicated in generating memory responses essential in vaccine protection against Salmonella. Following infection with mutant avirulent S. typhimurium Aro strains, mice deficient in IFN-γ are more susceptible than WT mice. These mice elicit highly impaired immunity and widespread septicemia [50]. A host of IFN-γ-induced genes are upregulated in response to Salmonella. Significant changes occur in the expression profile of many of the vital players of the immune system, such as chemokines, signaling molecules, transcription factors and surface receptors within the infected macrophage. Amongst the genes upregulated in macrophages or epithelial cells by S. typhimurium or S. dublin the prominent ones include IRF-1, iNOS and MHC class I and II [55]. This phenomenon can be recapitulated with LPS stimulations in uninfected macrophages. IFN-γ treated cells may exhibit a changed expression profile compared with pathogen-induced gene expression, pointing toward altered host cell responsive and differential activation state. Such profound alterations in the gene expression profile are also evident in in other bacterial infections. IFN-γ also facilitates internalization and promotes early killing of Salmonella without employing oxidative burst, accompanied by a decrease in IL-10 levels with a concomitant increase in IL-12 levels [56]. Another novel anti-microbial effect of IFN-γ is the depletion of intracellular iron levels in S. typhimurium-infected macrophages. Depleting iron starves the pathogen of this essential nutrient and boosts the bactericidal mechanism of macrophages. This occurs via transferrin receptor 1 which leads to an enhanced iron efflux by increasing the iron exporter ferroprotein. This event ultimately enhances the expression of hemoxygenase I and the siderophore entrapping antimicrobial peptide, lipocalin 2, that further sequesters iron from intracellular pools [57]. The role of hemoxygenase I is pivotal in regulating the initial stages of innate immunity to S. typhimurium infections. Iron export from tissue macrophages occurs in conjunction with increased serum levels of IL-12 and IFN-γ [58]. However, relocation of iron from intracellular pools may have negative outcomes as this often leads to anemia pertinent in Salmonella infections. Nevertheless, IFN-γ mediated iron scavenging undoubtedly bolsters host survival during acute infections [59].

Anti-microbial effects of IFN-γ: role in Listeria infection

Just like in other infections, IFN-γ forms an integral part of immunity against Listeria as well. Administration of anti-IFN-γ antibodies reverses the protective effects during Listeria infection and increases disease-associated fatality. The literature is replete with evidence stating the requirement of endogenous IFN-γ for the resolution of Listeria monocytogenes infection. NK cells and T cells are the key producers of IFN-γ in this regard. IFN-γ helps in bacterial clearance and control of primary infections with L. monocytogenes [60]. An essential role of IFN-γ-mediated signaling in inducing protective responses is evident from the fact that mice lacking functional STAT1 exhibit profound propensity to Listeria infections and manifest dampened macrophage activity and pathogen clearance [61]. STAT1-mediating signaling was imperative in T cells and DCs only at the level of generating primary adaptive immune responses. The fact that IFN-γ curbs bacterial spread is evident following IFN-γ blockade with neutralizing antibodies resulting in higher bacterial burdens in the mesenteric lymph nodes [62]. The prominent anti-Listerial effects mediated by IFN-γ include the generation of ROS and RNS via induction of NOX2/CYBB (NAPDH oxidase), DUOX (dual oxidase) and iNOS/NOS2. Additionally, induction of inducible GTPases (which facilitate oxidative pathways, autophagy and orchestrate vesicular traffic) and elicitation of mitochondrial ROS through regulated expression of several nuclear genes that encode respiratory machinery through the activity of ERRα (Estrogen related receptor α) are some other examples of IFN-γ-mediated anti-Listerial effects. The role of ERRα has been associated with effective pathogen clearance through concerted mitochondrial ROS production. 63 IFN-γ also regulates the expression of indole 2,3 dioxygenase (IDO), an important enzyme involved in tryptophan metabolism known to limit infection spread and control enormous T cell activity [64]. Furthermore, in Listeria or BCG infected mice, NK cells and T cells show a transient downfall of microRNA, miR29 that is essential for sustained production of IFN-γ. Thus, miR29 negatively regulates IFN-γ expression. Based on these observations researchers were able to devise a unique transgenic mice model (GS29) bearing ‘sponge’ targets that could compete with miR29 targets. Such mice displayed higher T helper responses, higher serum concentrations of IFN-γ and lower L. monocytogenes burden [65]. The pathogen has evolved exquisite evasive strategies to counter immune activation [66]. Listeria can modulate expression of many interferon stimulatory genes (ISGs) in epithelial cells. This is accomplished by targeting a chromatin repressing complex BAHD1 at the promoter level which can dampen expression of multiple ISGs and subsequently downregulate IFN-γ induced responses [67]. Interestingly, L. monocytogenes can also produce a protein called LntA which is a nucleomodulin. This protein binds to BAHD1 to nullify its effect and functions to ameliorate IFN functions [67,68]. Type-I IFNs are often linked with suppression of innate immune responses in L. monocytogenes infection and escalate predisposition to the disease. IFN-α/β production affords some degree of subversion from the macrophage activating effects of IFN-γ which can also be achieved in part through IFN-γR downregulation [69].

Protective role of IFN-γ during parasite infection

IFN-γ plays a pivotal role in host resistance to Leishmania infections [70]. Conversely, IL-4 mitigates ROS generation and stifles IFN-γ-mediated anti-pathogenic responses of the host. NK cells are the prime source of IFN-γ production which facilitates CD4+ T cell polarization and differentiation to induce early phases of resistance [70]. In the absence of IFN-γ, TH2 responses are linked to ablated anti-parasitic immunity.

It is reported that both IL-12 and IFN-γ help in limiting visceral leishmaniasis and counteract IL-6, IL-24 and IL-27 activity [71]. Neutralization of TNF-α leads to impaired IFN-γ production by spleen cells in visceral leishmaniasis patients, pointing towards a regulatory role of TNF-α [72]. Another important anti-parasitic mechanism is the degradation of tryptohan. This strategy constitutes an important host defense against obligate parasites such as Leishmania, Toxoplasma, Chlamydia etc. However, some reports have also emphasized on the harmful effects of IFN-γ in having negative repercussions on the disease outcome. For example, IFN-γ exacerbates disease symptoms through recruitment of inflammatory monocytes which diminish iNOS-dependent parasite killing, leading to expansion of the parasite niche. This in turn prolongs host cell survival to support parasite multiplication [73]. Similarly, IFN-γ promotes L. amazonensis replication in macrophages in vitro by inducing cationic amino acid transporter 2B (CAT2B) that facilitates L-arginine transport, which can be directly utilized by the parasite for its growth and multiplication [74]. Similar to other pathogens, L. donovani can also attenuate IFN-γR mediated signaling in macrophages by inhibiting STAT1 phosphorylation at tyrosine residues catalyzed by JAKs. This downregulates IFN-γ-induced genes and consequently, the downstream effector functions are quenched. Furthermore, the Leishmaina parasite can effectively thwart STAT1 translocation to the nucleus and also impede JAK activation necessary to achieve full-fledged downstream signaling emanating from IFN-γR [75].

IFN-γ: therapeutic and clinical implications

Administration of cytokines is an important and effective therapeutic strategy to stimulate the immune system and initiate protective responses. This procedure shapes and orchestrates the host cytokine milieu to modulate various facets of immune responses most notably by influencing tumor microenvironment or pathogenesis. Cytokine administration has gained immense popularity in cancer immunotherapy and is beneficial in manipulating tumor behavior. The usability of such an approach is limited by its fatal side effects as systemic administration may lead to “toxic shock” leading to untoward consequences. However, some patients have achieved immense benefits and elicited better prognosis in certain conditions [76,77]. As discussed, owing to the numerous and vital immune-protective functions coordinated by IFN-γ, it is one such potent cytokine which can be administered either systemically or locally. Due to its profound immune modulatory abilities, IFN-γ has become an attractive therapeutic alternative and has been used as a prophylactic measure to cure several infections including cancers, fungal infections and chronic granulomatous disease (CGD) [78,79,80].

Experimental data point towards myriad protective effects of IFN-γ therapy. IFN-γ administration can successfully impede Ebola virus infectivity and can be an attractive prophylactic alternative. It effectively reduced viral viability and serum titers [81]. Similarly, adjunctive immune therapy including IFN-γ as one of the pivotal components also drastically improved the outcome of invasive fungal infections or sepsis. Recombinant IFN-γ administration proved beneficial in restoring immune functionality thereby making IFN-γ as a promising therapeutic contender [82]. In animal infection models, administration of IFN-γ has led to better survival rate and immune responses. For example, enhanced resistance against invasive aspergillosis and disseminated Candida albicans infections is elicited in IFN-γ treated mice [83,84]. Similarly, IFN-γ therapy bolstered pulmonary immune responses in corticosteroid-treated rats in experimental legionellosis [85]. Improved survival and decreased pathogenic burden in lungs was observed upon IFN-γ administration in mice infected with Cryptococcus neoformans [86]. IFN-γ can prove as an effective therapeutic candidate against as many as 22 infectious agents including bacteria, fungi, protozoa and helminths. Clinically IFN-γ is an FDA approved drug for the treatment of CGD. CGD is a form of immune-deficiency characterized by a series of recurrent infections by pyogenic bacteria. It is majorly an inherited disorder wherein most of the cases arise due to mutatation of the CYBB gene (Cytochrome B beta/NAPDH oxidase) encoding gp91phox, the remaining being autosomal recessive. The disease is primarily due to dysfunctional NADPH oxidase machinery caused by defective or missing components of the phagocyte oxidase [87]. IFN-γ promotes respiratory burst by facilitating superoxide release in these patients and partially makes up for the anomalies in the oxidative metabolism. Interestingly, in some cases IFN-γ has been shown to improve splicing efficiency of the CYBB gene, which is normally malfunctional in CGD. It augments microbicidal activities and reduces the relative risk of infections by 70% [88]. Thus, IFN-γ is a beneficial prophylactic agent in CGD with minimal side effects such as toxicity, delayed or impaired growth and development.

Figure 1 IFN-γ: Production and regulation. A complex interplay between various cytokines determine the transcription and production of IFN-γ. Release of IFN-γ either by NK cells or T cells triggers a series of anti-microbial responses. Following pathogenic recognition and uptake by an APC such as DC, the antigenic peptides are presented to the naïve T-cell population for initiation of an adaptive immune response. The cytokine milieu comprised of IL-12 and IL-2 triggers T cell proliferation to create conditions conducive for a pro-inflammatory phenotype. Exposure to IL-12 aids naïve cells to polarize to a TH1 phenotype. IFN-γ is a predominant cytokine in the TH1 response. Following transcription and production of the IFN-γ protein in the TH1 lineage committed and clonally expanded T cell population, this cytokine primes macrophages to upregulate their protective infections. The activated macrophages in turn produce IL-12 and IL-23 that further ameliorate pro-inflammatory conditions and facilitate IFN-γ production. Thus, IFN-γ production is self-sustained and occurs in a positive feedback manner. Macrophages are also known to produce IFN-γ that leads to autocrine activation further bolstering their anti-microbial functions.
Figure 1

IFN-γ: Production and regulation. A complex interplay between various cytokines determine the transcription and production of IFN-γ. Release of IFN-γ either by NK cells or T cells triggers a series of anti-microbial responses. Following pathogenic recognition and uptake by an APC such as DC, the antigenic peptides are presented to the naïve T-cell population for initiation of an adaptive immune response. The cytokine milieu comprised of IL-12 and IL-2 triggers T cell proliferation to create conditions conducive for a pro-inflammatory phenotype. Exposure to IL-12 aids naïve cells to polarize to a TH1 phenotype. IFN-γ is a predominant cytokine in the TH1 response. Following transcription and production of the IFN-γ protein in the TH1 lineage committed and clonally expanded T cell population, this cytokine primes macrophages to upregulate their protective infections. The activated macrophages in turn produce IL-12 and IL-23 that further ameliorate pro-inflammatory conditions and facilitate IFN-γ production. Thus, IFN-γ production is self-sustained and occurs in a positive feedback manner. Macrophages are also known to produce IFN-γ that leads to autocrine activation further bolstering their anti-microbial functions.

Figure 2 Protective effects of IFN-γ. IFN-γ primes the macrophage to improve its anti-microbial capacity. The biological protective effects commence upon binding of a dimeric IFN-γ to its cognate receptor on the macrophage surface. This binding leads to a series of phosphorylation reactions that ultimately phosphorylate STAT1 proteins. Following phosphorylation of STAT1 molecules, these homodimeric molecules translocate to the nucleus where they bind to GAS elements on the DNA. This triggers a plethora of downstream effector genes that encode products necessary for pathogen containment. IFN-γ -regulated proteins include Reactive Oxygen Species, Reactive Nitrogen Intermediates, autophagy proteins, efflux pumps or antiviral mediators. Besides, controlling infection, IFN-γ can serve as a double-edged sword in an immune dysfunctional state, such as autoimmunity, wherein it can exacerbate disease symptoms due to prolonged proinflammatory effects.
Figure 2

Protective effects of IFN-γ. IFN-γ primes the macrophage to improve its anti-microbial capacity. The biological protective effects commence upon binding of a dimeric IFN-γ to its cognate receptor on the macrophage surface. This binding leads to a series of phosphorylation reactions that ultimately phosphorylate STAT1 proteins. Following phosphorylation of STAT1 molecules, these homodimeric molecules translocate to the nucleus where they bind to GAS elements on the DNA. This triggers a plethora of downstream effector genes that encode products necessary for pathogen containment. IFN-γ -regulated proteins include Reactive Oxygen Species, Reactive Nitrogen Intermediates, autophagy proteins, efflux pumps or antiviral mediators. Besides, controlling infection, IFN-γ can serve as a double-edged sword in an immune dysfunctional state, such as autoimmunity, wherein it can exacerbate disease symptoms due to prolonged proinflammatory effects.

The therapeutic and immunomodulatory effects of IFN-γ have also been utilized to treat immunodefeciency syndromes. For example, hyperimmunoglobulinemia E (hyper-IgE) is a rare and unique immunodeficiency syndrome which predisposes an individual to severe and recurrent Stapylococcus aurues infections. These patients exhibit skin eczema, abscesses, elevated eosinophil production, hyper-production of IgE with limited lymphocyte chemotaxis and reduced responsiveness to IL-12. Systemic administration of recombinant IFN-γ attenuates the course of clinical symptoms in such patients [89]. Efficacy of recombinant IFN-γ as an adjunct increases manifolds when it is used in conjunction with the conventional anti-microbials. This is especially true in the case of visceral or cutaneous leishmaniasis, atypical mycobacterial infections and lepromatous leprosy [90,91]. Alveolar macrophages exhibit heightened activity and improved effector responses in a limited number of multi-drug resistant TB (MDR-TB) patients who received aerosolized IFN-γ therapy. In a small group of 5 patients, all the individuals who had undergone the therapy (thrice a week for 2 months) showed improved prognosis, reduction in lesion size and tested negative for sputum acid fast bacillus [92]. However, therapy termination led to reversion of sputum smears to positive. Nevertheless, the data indicate that aerosolized IFN-γ may be a potential and effective prophylactic agent for the treatment of MDR-TB patients, who are normally refractory to other lines of treatment. Systemic aerosolic IFN-γ administration also yields promising results for the treatment of respiratory infections caused by atypical mycobacteria [93]. IFN-γ can alleviate symptoms of sepsis by restoring LPS-induced TNF-α production and HLA-DR expression in patients. This recovers macrophage activity and helps in clearance of sepsis [94]. IFN-γ therapy has been shown to successfully contain the damaging consequences of fibrosis. It ameliorates the total lung capacity and partial pressure of lung oxygen with concomitant suppression of pro-inflammatory cytokines in fibrotic disease [95]. As discussed earlier, mycobacteria subvert host immunity for their own benefit, interestingly, many host genes also play a suppressive role during M. tb infection and expression of these genes modulates TH1 cytokines including IFN-γ [96]. Our own findings have shown that IFN-γ plays a suppressive role during mycobacterial infection. Infection with M. tb induces production of IFN-γ that suppress many host protective functions. Thus, insights into host-pathogen interaction and dynamics of infection would strengthen our understanding of several disease mechanisms. These host-specific preventive therapeutic approaches can emerge as a robust alternative to the conventional pathogen-centric intervention to mitigate the ever-rising incidents of drug resistance. This can aid the development of better drugs and improve our understanding about host immune responses during complex infections.

Table 1

Biological functions of IFN-γ: An overview.

Immunological pathwayProteins affectedBiological implicationsReferences
Antigen processing and presentationCTIIA, MHC class II, Cathepsins (L, B, S and H), MHC class I, trimming peptidases (endoplasmic reticulum aminopeptidases, ERAP-1), β32-microglobulinMainly targets and boosts the expression of major effector molecules and chaperones involved in the processing and presentation of antigens in both class I and class II pathways. This pathway is central to the inception of adaptive immunity through the priming of naïve T cells by antigenic entities, leading to robust and efficient T cell activation.[107,108,109,110,111,112]
Reactive Oxygen Speciesgp91phox, p22phox, p67phox and GTPase Gbp7Upregulates the expression and facilitates the assembly of subunits comprising the NADPH oxidase. The enzyme is critical for the generation of oxidative burst and production of ROS that show potent microbicidal effects.[21,113]
Reactive Nitrogen IntermediatesNOS2, Argininosuccinate sythetase and GTP cyclohyrdoxylase IIncreases the production of the key enzyme involved in the catalytic production of reactive nitrogen species or acts indirectly via upregulation of accessory molecules that enhance production of substrate or enzymatic cofactors.[1,114]
AutophagyLC3-II, IRGM (LRG-47), bioactive Vitamin D3 (via CYP27b1 hydroxylase), antimicrobial cathelicidin and ubiquitin-binding proteins (SQSTM1, NDP52, optineurin and galectins)A highly coordinated frame work of IFN-γ -inducible proteins including effector molecules, immunity-related GTPases and GBPs proteins trigger and facilitate the assembly of autophagic machinery and delivery of autophagic cargo to lysosomes. Additionally, to counter bacteria that have escaped into the cytosol, IFN-γ may stimulate production of proteins that direct cytosolic ubiquitinylated bacteria to the lysosome for degradation by combining them with the autophagy pathway.[21,31,115,116]
ApoptosisCaspases (8, 9 and 1), Fas, FasL, Bax, increased Cytochrome c release, IRF-1 and TNFaIFN-γ stimulation may trigger both extrinsic and intrinsic death pathways. Concomitantly the levels of pro-apoptotic factors and mediators are increased that facilitate apoptotic death.[116,117,118]
Cell migration and leucocyte trafficCXCL9, MIP-1, MIP-1, IP-10, VCAM-1, ICAM-1, RANTES, CCL2Expression of many essential chemokines and inflammatory mediators are under the influence of IFN-γ stimulation. These effector and adhesion molecules serve as chemoattractants and aid in immune cell migration or lymphocyte extravasation, thus helping in the initiation of an inflammatory response by recruiting diverse immune cells.[1,119]
Antiviral mediatorsPKR, ADAR, IFITM, TRIM Viperin and TethrinMany antiviral proteins induced by IFN-γ help in countering numerous viral infections at several stages like viral entry, uncoating, blocking viral translation or virion assembly. IFN inducible viral RNA editing enzymes may edit RNA bases that may introduce severe genomic instabilities leading to lethal mutations[21,120]
Lysosome mediated killing/Phagosome maturationProton ATPases, IRGM/LRG-47 (via induction of autophagy) GBPs and Rab proteinsIFN-γ elicits lysosomal killing of many intracellular pathogens by facilitating rapid acidification of phagosome and helping it recruit the necessary fusogenic machinery that aids its fusion with the lysosome. Alternatively, induction of several effectors of autophagy by IFN-γ can further facilitate lysosome-mediated enzymatic degradation of pathogens.[121,122,123,124]
Complement pathwayComplement proteins (C2, C3, C4 etc.) and CR3Complement receptors mediators or proteins (secreted by macrophages or fibroblasts) are upregulated in response to IFN-γ. The surge in complement pathway activity enhances opsonic uptake of extracellular pathogen via receptor mediated phagocytosis.[125,126,127]
Fc ReceptorsFCγRIFN-γ-mediated enhanced expression of high affinity FcγR1 on myeloid cells ameliorates antimicrobial defences by facilitating opsonisation of IgG-coated extracellular pathogens via upregulated receptors. Thus, increasing opsonic uptake by phagocytic cells.[128,129]
Other antimicrobial effectorsCathelicidin LL-37/hCAP18 Defensins, cationic efflux molecules such as NRAMP1, P type ATPase Cu++ Pump ATP7A and Ferroprotein (SLC40A1), IDOMany antimicrobial effectors such as antimicrobial peptides, cationic efflux pumps or molecules causing nutrient limitation to the pathogen are under the control of IFN-γ stimulation. These antimicrobial effectors display potent activity and can either directly or indirectly kill the pathogen either by depriving it of essential ions or nutrients or by disrupting its membrane integrity.[21,130,131,132]

Recombinant IFN-γ elicits potent anti-proliferative, anti-angiogenic and anti-tumorigenic effects. It is known to trigger apoptotic death in tumor cells and leads to a positive disease outcome. The earliest usage of IFN-γ as a therapeutic agent was for acute leukemia and since then it has been used on and off experimentally to induce anti-proliferative tendencies in multiple cell lines. IFN-γ was successful in limiting carcinogenesis in numerous cancerous cell lines [97,98,99]. Growth modulatory properties of IFN-γ have also made it an interesting therapeutic option for hematologic conditions and human stem cell (HSC) transplantation. Recent findings have suggested that IFN-γ can negatively regulate the expansion of HSC pool and lead to progressive loss of such cells in bone marrows and peripheral HSCs in the context of infections [100].

The therapeutic potential of IFN-γ can be limited owing to huge cost of treatments and side effects such as flu, lethargy, suppressed nervous system, cough etc [101]. Although these accompanied adverse effects do minimize the efficacy and potential of such immunomodulatory therapies, IFN-γ administration still proves to be beneficial and an effective adjunct to the existing drug regimen in this era of rapid and formidable drug resistance. Thus, cytokine intervention can control several of the life-threatening and debilitating infections. Targeting IFN-γ has been seen as a new therapeutic approach for combating atherosclerosis, as it is reported to be key player involved in the development and disease manifestation of atherosclerosis [102]. Clinical investigations have further substantiated the role of IFN-γ in progression of cardiac diseases. Patients suffering from chronic heart failure reported to have a higher serum IFN-γ levels as compared to healthy controls [103]. Moreover, myocardial tissues of patients suffering from Chagas’ cardiomyopathy displayed significant up regulation of IFN-γ-inducible genes, suggesting a direct link of IFN-γ signaling in this parasite-induced heart problem. Individuals with peripartum cardiomyopathy upon treatment with ACE-inhibitors, beta-blockers and diuretics resulted in improved cardiac function along with lowering down of serum IFN-γ levels which was high otherwise [104]. Additionally, levels of IFN-γ in serum and cerebrospinal fluid have also been found up-regulated in the neurodegenerative disease, Amyotrophic lateral sclerosis (ALS). Hence, linking this key pro-inflammatory cytokine with the progression of the disease [105].. Elevated levels of serum targeting IFN-γ has been seen as a new therapeutic approach for combating atherosclerosis, as it is reported to be key player involved in the development and disease manifestation of atherosclerosis [102]. Clinical investigations have further substantiated the role of IFN-γ in progression of cardiac diseases. Patients suffering from chronic heart failure are reported to have a higher serum IFN-γ levels as compared to healthy controls [103]. Moreover, myocardial tissues of patients suffering from Chagas’ cardiomyopathy displayed significant up regulation of IFN-γ-inducible genes, suggesting a direct link of IFN-γ signaling in this parasite-induced heart problem. Individuals with peripartum cardiomyopathy upon treatment with ACE-inhibitors, beta-blockers and diuretics resulted in improved cardiac function along with lowering of serum IFN-γ levels which was high otherwise [104]. Additionally, levels of IFN-γ in serum and cerebrospinal fluid have been found to be up-regulated in ALS, linking this key pro-inflammatory cytokine with the progression of the disease [105]. Elevated levels of serum IFN-γ has also been reported in patients with Parkinson’s disease [106]. Thus, IFN-γ can be a relevant disease marker in a clinical set-up, and therapeutic targeting of IFN-γ may emerge as a potential alternative strategy to alleviate disease pathologies.

Conclusion and future perspective

Owing to a plethora of biological activities mediated by IFN-γ, this cytokine remains an imperative and obligate protective factor in the host immune system. This versatile immune factor directly or indirectly regulates the expression of numerous genes that supposedly confer protection and mediate anti-microbial activities. Thus, comprehending its biological functionality or implications in infections or immune dysregulatory conditions, such as autoimmunity, becomes critical. Deeper understanding of the molecular basis of these IFN-γ-mediated mechanisms vis-à-vis diseases and associated pathologies would unravel valuable information about intricate host-pathogen encounters. A more in-depth understanding of the complex pattern of host-pathogen interactions would facilitate research in the areas of vaccine and drug development. Deciphering mechanisms of host mediated responses can also present a possibility of devising host-directed therapies that would alleviate the scourge of widespread drug resistance and limited drug usability. Likewise, the protective effects of IFN-γ have been harnessed for therapeutic purposes as well. This further reiterates its protective role in the host system. Comprehending the evasive mechanisms and tolerance to IFN-γ in various immune compromised conditions and cancers would further delineate complex patterns of pathogenesis and increase our understanding about disease mechanisms.

Acknowledgments

GK is recipient of CSIR-Senior Research Fellowship. MR is a recipient of UGC-Senior Research Fellowship. BKT is recipient of UGC-DS Kothari Post-Doctoral Fellowship.

  1. Funding: Research funding was provided by the University of Delhi and CSIR.

  2. Authors’ contribution: GK and MR conceptualized and wrote the manuscript. GK, BKT and MR helped with data compilation.

  3. Competing Financial Interests: The authors declare no competing financial interests

  4. Ethics approval and consent to Participate: Not applicable

References

1 Schroder K, Hertzog PJ, Ravasi T, Hume DA. Interferon-γ: an overview of signals, mechanisms and functions. Journal of leukocyte biology 2004;75:163-89.10.1189/jlb.0603252Search in Google Scholar PubMed

2 Rothfuchs AG, Trumstedt C, Wigzell H, Rottenberg ME. Intracellular Bacterial Infection-Induced IFN-γ Is Critically but Not Solely Dependent on Toll-Like Receptor 4-Myeloid Differentiation Factor 88-IFN-αβ-STAT1 Signaling. The Journal of Immunology 2004;172:6345-53.10.4049/jimmunol.172.10.6345Search in Google Scholar PubMed

3 Naglak EK, Morrison SG, Morrison RP. Gamma interferon is required for optimal antibody-mediated immunity against genital chlamydia infection. Infection and immunity 2016;84:3232-42.10.1128/IAI.00749-16Search in Google Scholar PubMed PubMed Central

4 Green AM, DiFazio R, Flynn JL. IFN-γ from CD4 T cells is essential for host survival and enhances CD8 T cell function during Mycobacterium tuberculosis infection. The Journal of Immunology 2013;190:270-7.10.4049/jimmunol.1200061Search in Google Scholar PubMed PubMed Central

5 Baird NL, Bowlin JL, Hotz TJ, Cohrs RJ, Gilden D. Interferon gamma prolongs survival of varicella-zoster virus-infected human neurons in vitro. Journal of virology 2015;89:7425-7.10.1128/JVI.00594-15Search in Google Scholar PubMed PubMed Central

6 Kokordelis P, Krämer B, Körner C, et al. An effective interferon gamma-mediated inhibition of hepatitis C virus replication by natural killer cells is associated with spontaneous clearance of acute hepatitis C in human immunodeficiency virus-positive patients. Hepatology 2014;59:814-27.10.1002/hep.26782Search in Google Scholar PubMed

7 Beekhuizen H, van de Gevel JS. Gamma interferon confers resistance to infection with Staphylococcus aureus in human vascular endothelial cells by cooperative proinflammatory and enhanced intrinsic antibacterial activities. Infection and immunity 2007;75:5615-26.10.1128/IAI.00530-07Search in Google Scholar PubMed PubMed Central

8 Koo GC, Gan Y-H. The innate interferon gamma response of BALB/c and C57BL/6 mice to in vitro Burkholderia pseudomallei infection. BMC immunology 2006;7:19.10.1186/1471-2172-7-19Search in Google Scholar PubMed PubMed Central

9 Miller CH, Maher SG, Young HA. Clinical use of interferon-γ. Annals of the New York Academy of Sciences 2009;1182:69-79.10.1111/j.1749-6632.2009.05069.xSearch in Google Scholar PubMed PubMed Central

10 Belkaid Y, Rouse BT. Natural regulatory T cells in infectious disease. Nature immunology 2005;6:353.10.1038/ni1181Search in Google Scholar PubMed

11 Frucht DM, Fukao T, Bogdan C, Schindler H, O’Shea JJ, Koyasu S. IFN-gamma production by antigen-presenting cells: mechanisms emerge. Trends Immunol 2001;22:556-60.10.1016/S1471-4906(01)02005-1Search in Google Scholar

12 Boehm U, Guethlein L, Klamp T, et al. Two families of GTPases dominate the complex cellular response to IFN-γ. The Journal of Immunology 1998;161:6715-23.Search in Google Scholar

13 Bach EA, Aguet M, Schreiber RD. The IFNγ receptor: a paradigm for cytokine receptor signaling. Annual review of immunology 1997;15:563-91.10.1146/annurev.immunol.15.1.563Search in Google Scholar

14 Ramana CV, Gil MP, Schreiber RD, Stark GR. Stat1-dependent and-independent pathways in IFN-γ-dependent signaling. Trends in immunology 2002;23:96-101.10.1016/S1471-4906(01)02118-4Search in Google Scholar

15 Lighvani AA, Frucht DM, Jankovic D, et al. T-bet is rapidly induced by interferon-γ in lymphoid and myeloid cells. Proceedings of the National Academy of Sciences 2001;98:15137-42.10.1073/pnas.261570598Search in Google Scholar PubMed PubMed Central

16 Djuretic IM, Levanon D, Negreanu V, Groner Y, Rao A, Ansel KM. Transcription factors T-bet and Runx3 cooperate to activate Ifng and silence Il4 in T helper type 1 cells. Nature immunology 2007;8:145.10.1038/ni1424Search in Google Scholar PubMed

17 Afkarian M, Sedy JR, Yang J, et al. T-bet is a STAT1-induced regulator of IL-12R expression in naive CD4+ T cells. Nature immunology 2002;3:549.10.1038/ni794Search in Google Scholar PubMed

18 Flannagan RS, Cosío G, Grinstein S. Antimicrobial mechanisms of phagocytes and bacterial evasion strategies. Nature reviews Microbiology 2009;7:355.10.1038/nrmicro2128Search in Google Scholar PubMed

19 Li P, Du Q, Cao Z, et al. Interferon-gamma induces autophagy with growth inhibition and cell death in human hepatocellular carcinoma (HCC) cells through interferon-regulatory factor-1 (IRF-1). Cancer letters 2012;314:213-22.10.1016/j.canlet.2011.09.031Search in Google Scholar PubMed PubMed Central

20 Singh SB, Davis AS, Taylor GA, Deretic V. Human IRGM induces autophagy to eliminate intracellular mycobacteria. Science 2006;313:1438-41.10.1126/science.1129577Search in Google Scholar PubMed

21 MacMicking JD. Interferon-inducible effector mechanisms in cell-autonomous immunity. Nature reviews Immunology 2012;12:367.10.1038/nri3210Search in Google Scholar PubMed PubMed Central

22 MacMicking JD, Taylor GA, McKinney JD. Immune control of tuberculosis by IFN-γ-inducible LRG-47. Science 2003;302:654-9.10.1126/science.1088063Search in Google Scholar PubMed

23 Kim B-H, Shenoy AR, Kumar P, Das R, Tiwari S, MacMicking JD. A family of IFN-γ-inducible 65-kD GTPases protects against bacterial infection. Science 2011;332:717-21.10.1126/science.1201711Search in Google Scholar PubMed

24 Schnoor M, Betanzos A, Weber D, Parkos C. Guanylate-binding protein-1 is expressed at tight junctions of intestinal epithelial cells in response to interferon-γ and regulates barrier function through effects on apoptosis. Mucosal immunology 2009;2:33-42.10.1038/mi.2008.62Search in Google Scholar PubMed PubMed Central

25 Murray PJ. Amino Acid Auxotrophy as Immunological Control Nodes. Nature immunology 2016;17:132.10.1038/ni.3323Search in Google Scholar PubMed PubMed Central

26 Day PM, Thompson CD, Lowy DR, Schiller JT. Interferon Gamma Prevents Infectious Entry of Human Papillomavirus 16 via an L2-Dependent Mechanism. Journal of Virology 2017;91:e00168-17.10.1128/JVI.00168-17Search in Google Scholar PubMed PubMed Central

27 Feeley EM, Sims JS, John SP, et al. IFITM3 inhibits influenza A virus infection by preventing cytosolic entry. PLoS pathogens 2011;7:e1002337.10.1371/journal.ppat.1002337Search in Google Scholar PubMed PubMed Central

28 Hatakeyama S. TRIM family proteins: roles in autophagy, immunity, and carcinogenesis. Trends in biochemical sciences 2017.10.1016/j.tibs.2017.01.002Search in Google Scholar PubMed

29 Biering SB, Choi J, Halstrom RA, et al. Viral Replication Complexes Are Targeted by LC3-Guided Interferon-Inducible GTPases. Cell Host & Microbe 2017;22:74-85.e7.10.1016/j.chom.2017.06.005Search in Google Scholar PubMed PubMed Central

30 Dotson D, Woodruff EA, Villalta F, Dong X. Filamin A is involved in HIV-1 Vpu-mediated evasion of host restriction by modulating Tetherin expression. Journal of Biological Chemistry 2016;291:4236-46.10.1074/jbc.M115.708123Search in Google Scholar PubMed PubMed Central

31 Fabri M, Stenger S, Shin D-M, et al. Vitamin D is required for IFN-γ-mediated antimicrobial activity of human macrophages. Science translational medicine 2011;3:104ra2-ra2.10.1126/scitranslmed.3003045Search in Google Scholar PubMed PubMed Central

32 Jabado N, Jankowski A, Dougaparsad S, Picard V, Grinstein S, Gros P. Natural resistance to intracellular infections. Journal of Experimental Medicine 2000;192:1237-48.10.1084/jem.192.9.1237Search in Google Scholar

33 White C, Lee J, Kambe T, Fritsche K, Petris MJ. A role for the ATP7A copper-transporting ATPase in macrophage bactericidal activity. Journal of Biological Chemistry 2009;284:33949-56.10.1074/jbc.M109.070201Search in Google Scholar

34 Deriu E, Liu JZ, Raffatellu M. 13 Salmonella and the Host in the Battle for Iron. Stress Response in Pathogenic Bacteria 2011;19:283.10.1079/9781845937607.0283Search in Google Scholar

35 Flynn JL, Chan J, Triebold KJ, Dalton DK, Stewart TA, Bloom BR. An essential role for interferon gamma in resistance to Mycobacterium tuberculosis infection. Journal of Experimental Medicine 1993;178:2249-54.10.1084/jem.178.6.2249Search in Google Scholar

36 Flynn JL, Chan J. Immunology of tuberculosis. Annual review of immunology 2001;19:93-129.10.1146/annurev.immunol.19.1.93Search in Google Scholar

37 Altare F, Jouanguy E, Lamhamedi S, Doffinger R, Fischer A, Casanova J-L. Mendelian susceptibility to mycobacterial infection in man. Current opinion in immunology 1998;10:413-7.10.1016/S0952-7915(98)80114-3Search in Google Scholar

38 Cheallaigh CN, Sheedy FJ, Harris J, et al. A common variant in the adaptor Mal regulates interferon gamma signaling. Immunity 2016;44:368-79.10.1016/j.immuni.2016.01.019Search in Google Scholar PubMed PubMed Central

39 MacMicking JD. Cell-autonomous effector mechanisms against Mycobacterium tuberculosis. Cold Spring Harbor perspectives in medicine 2014;4:a018507.10.1101/cshperspect.a018507Search in Google Scholar PubMed PubMed Central

40 Herbst S, Schaible UE, Schneider BE. Interferon gamma activated macrophages kill mycobacteria by nitric oxide induced apoptosis. PloS one 2011;6:e19105.10.1371/journal.pone.0019105Search in Google Scholar PubMed PubMed Central

41 Shin DM, Yuk JM, Lee HM, et al. Mycobacterial lipoprotein activates autophagy via TLR2/1/CD14 and a functional vitamin D receptor signalling. Cellular microbiology 2010;12:1648-65.10.1111/j.1462-5822.2010.01497.xSearch in Google Scholar PubMed PubMed Central

42 Yuk J-M, Yoshimori T, Jo E-K. Autophagy and bacterial infectious diseases. Experimental & molecular medicine 2012;44:99.10.3858/emm.2012.44.2.032Search in Google Scholar

43 Harding CV, Boom WH. Regulation of antigen presentation by Mycobacterium tuberculosis: a role for Toll-like receptors. Nature reviews Microbiology 2010;8:296.10.1038/nrmicro2321Search in Google Scholar

44 Pennini ME, Pai RK, Schultz DC, Boom WH, Harding CV. Mycobacterium tuberculosis 19-kDa lipoprotein inhibits IFN-γ-induced chromatin remodeling of MHC2TA by TLR2 and MAPK signaling. The Journal of Immunology 2006;176:4323-30.10.4049/jimmunol.176.7.4323Search in Google Scholar

45 Pai RK, Pennini ME, Tobian AA, Canaday DH, Boom WH, Harding CV. Prolonged toll-like receptor signaling by Mycobacterium tuberculosis and its 19-kilodalton lipoprotein inhibits gamma interferon-induced regulation of selected genes in macrophages. Infection and immunity 2004;72:6603-14.10.1128/IAI.72.11.6603-6614.2004Search in Google Scholar

46 Kincaid EZ, Ernst JD. Mycobacterium tuberculosis exerts gene-selective inhibition of transcriptional responses to IFN-gamma without inhibiting STAT1 function. Journal of immunology (Baltimore, Md : 1950) 2003;171:2042-9.10.4049/jimmunol.171.4.2042Search in Google Scholar

47 Olias P, Etheridge RD, Zhang Y, Holtzman MJ, Sibley LD. Toxoplasma effector recruits the Mi-2/NuRD complex to repress STAT1 transcription and block IFN-γ-dependent gene expression. Cell host & microbe 2016;20:72-82.10.1016/j.chom.2016.06.006Search in Google Scholar

48 Alvarez GR, Zwilling BS, Lafuse WP. Mycobacterium avium inhibition of IFN-γ signaling in mouse macrophages: toll-like receptor 2 stimulation increases expression of dominantnegative STAT1β by mRNA stabilization. The Journal of Immunology 2003;171:6766-73.10.4049/jimmunol.171.12.6766Search in Google Scholar

49 Sugawara I, Yamada H, Mizuno S. STAT1 knockout mice are highly susceptible to pulmonary mycobacterial infection. The Tohoku journal of experimental medicine 2004;202:41-50.10.1620/tjem.202.41Search in Google Scholar

50 Bao S, Beagley KW, France MP, Shen J, Husband AJ. Interferongamma plays a critical role in intestinal immunity against Salmonella typhimurium infection. Immunology 2000;99:464-72.10.1046/j.1365-2567.2000.00955.xSearch in Google Scholar

51 Muotiala A, Makela PH. The role of IFN-gamma in murine Salmonella typhimurium infection. Microbial pathogenesis 1990;8:135-41.10.1016/0882-4010(90)90077-4Search in Google Scholar

52 Gilchrist JJ, MacLennan CA, Hill AV. Genetic susceptibility to invasive Salmonella disease. Nature reviews Immunology 2015;15:452.10.1038/nri3858Search in Google Scholar

53 Jouanguy E, Döffinger R, Dupuis S, Pallier A, Altare F, Casanova J-L. IL-12 and IFN-γ in host defense against mycobacteria and salmonella in mice and men. Current opinion in immunology 1999;11:346-51.10.1016/S0952-7915(99)80055-7Search in Google Scholar

54 Netea MG, Fantuzzi G, Kullberg BJ, et al. Neutralization of IL-18 Reduces Neutrophil Tissue Accumulation and Protects Mice Against Lethal <em>Escherichia coli</em> and <em>Salmonella typhimurium</em> Endotoxemia. The Journal of Immunology 2000;164:2644-9.10.4049/jimmunol.164.5.2644Search in Google Scholar PubMed

55 Rosenberger CM, Scott MG, Gold MR, Hancock RE, Finlay BB. Salmonella typhimurium infection and lipopolysaccharide stimulation induce similar changes in macrophage gene expression. The Journal of immunology 2000;164:5894-904.10.4049/jimmunol.164.11.5894Search in Google Scholar PubMed

56 Gordon MA, Jack DL, Dockrell DH, Lee ME, Read RC. Gamma interferon enhances internalization and early nonoxidative killing of Salmonella enterica serovar Typhimurium by human macrophages and modifies cytokine responses. Infection and immunity 2005;73:3445-52.10.1128/IAI.73.6.3445-3452.2005Search in Google Scholar PubMed PubMed Central

57 Nairz M, Fritsche G, Brunner P, Talasz H, Hantke K, Weiss G. Interferon-gamma limits the availability of iron for intramacrophage Salmonella typhimurium. Eur J Immunol 2008;38:1923-36.10.1002/eji.200738056Search in Google Scholar PubMed

58 Mitterstiller AM, Haschka D, Dichtl S, et al. Heme oxygenase 1 controls early innate immune response of macrophages to Salmonella Typhimurium infection. Cell Microbiol 2016;18:1374-89.10.1111/cmi.12578Search in Google Scholar PubMed PubMed Central

59 Brown DE, Nick HJ, McCoy MW, et al. Increased Ferroportin-1 Expression and Rapid Splenic Iron Loss Occur with Anemia Caused by Salmonella enterica Serovar Typhimurium Infection in Mice. Infection and Immunity 2015;83:2290-9.10.1128/IAI.02863-14Search in Google Scholar PubMed PubMed Central

60 Zenewicz LA, Shen H. Innate and adaptive immune responses to Listeria monocytogenes: a short overview. Microbes and Infection 2007;9:1208-15.10.1016/j.micinf.2007.05.008Search in Google Scholar PubMed PubMed Central

61 Kernbauer E, Maier V, Stoiber D, et al. Conditional Stat1 Ablation Reveals the Importance of Interferon Signaling for Immunity to Listeria monocytogenes Infection. PLoS Pathogens 2012;8:e1002763.10.1371/journal.ppat.1002763Search in Google Scholar PubMed PubMed Central

62 Reynders A, Yessaad N, Vu Manh TP, et al. Identity, regulation and in vivo function of gut NKp46+RORgammat+ and NKp46+RORgammat-lymphoid cells. Embo j 2011;30:2934-47.10.1038/emboj.2011.201Search in Google Scholar PubMed PubMed Central

63 Sonoda J, Laganière J, Mehl IR, et al. Nuclear receptor ERRα and coactivator PGC-1β are effectors of IFN-γ-induced host defense. Genes & development 2007;21:1909-20.10.1101/gad.1553007Search in Google Scholar PubMed PubMed Central

64 Popov A, Abdullah Z, Wickenhauser C, et al. Indoleamine 2, 3-dioxygenase–expressing dendritic cells form suppurative granulomas following Listeria monocytogenes infection. Journal of Clinical Investigation 2006;116:3160.10.1172/JCI28996Search in Google Scholar PubMed PubMed Central

65 Ma F, Xu S, Liu X, et al. The microRNA miR-29 controls innate and adaptive immune responses to intracellular bacterial infection by targeting interferon-[gamma].. Nature immunology 2011;12:861-9.10.1038/ni.2073Search in Google Scholar PubMed

66 Wan M, Zhou Y, Zhu Y. Subversion of Macrophage Functions by Bacterial Protein Toxins and Effectors. Current issues in molecular biology 2017;25:61-80.10.21775/9781910190692.03Search in Google Scholar

67 Lebreton A, Job V, Ragon M, et al. Structural basis for the inhibition of the chromatin repressor BAHD1 by the bacterial nucleomodulin LntA. MBio 2014;5:e00775-13.10.1128/mBio.00775-13Search in Google Scholar PubMed PubMed Central

68 Lebreton A, Lakisic G, Job V, et al. A bacterial protein targets the BAHD1 chromatin complex to stimulate type III interferon response. Science 2011;331:1319-21.10.1126/science.1200120Search in Google Scholar PubMed

69 Rayamajhi M, Humann J, Kearney S, Hill KK, Lenz LL. Antagonistic crosstalk between type I and II interferons and increased host susceptibility to bacterial infections. Virulence 2010;1:418-22.10.4161/viru.1.5.12787Search in Google Scholar PubMed PubMed Central

70 Muller K, van Zandbergen G, Hansen B, et al. Chemokines, natural killer cells and granulocytes in the early course of Leishmania major infection in mice. Medical microbiology and immunology 2001;190:73-6.10.1007/s004300100084Search in Google Scholar PubMed

71 dos Santos PL, de Oliveira FA, Santos MLB, et al. The severity of visceral leishmaniasis correlates with elevated levels of serum IL-6, IL-27 and sCD14. PLoS neglected tropical diseases 2016;10:e0004375.10.1371/journal.pntd.0004375Search in Google Scholar PubMed PubMed Central

72 Singh N, Kumar R, Engwerda C, Sacks D, Nylen S, Sundar S. Tumor necrosis factor alpha neutralization has no direct effect on parasite burden, but causes impaired IFN-γ production by spleen cells from human visceral leishmaniasis patients. Cytokine 2016;85:184-90.10.1016/j.cyto.2016.06.013Search in Google Scholar PubMed PubMed Central

73 Carneiro MB, de Moura Lopes ME, Romano A, et al. IFN-gamma mediated inflammatory monocyte recruitment neutralizes iNOS-dependent parasite killing by expanding the permissive host cell reservoir during early Leishmania amazonensis infection. Am Assoc Immnol; 2016.Search in Google Scholar

74 Wanasen N, MacLeod CL, Ellies LG, Soong L. L-arginine and cationic amino acid transporter 2B regulate growth and survival of Leishmania amazonensis amastigotes in macrophages. Infection and immunity 2007;75:2802-10.10.1128/IAI.00026-07Search in Google Scholar PubMed PubMed Central

75 Gupta G, Oghumu S, Satoskar AR. Mechanisms of Immune Evasion in Leishmaniasis. Advances in applied microbiology 2013;82:155-84.10.1016/B978-0-12-407679-2.00005-3Search in Google Scholar PubMed PubMed Central

76 Ardolino M, Raulet DH. Cytokine therapy restores antitumor responses of NK cells rendered anergic in MHC I-deficient tumors. Oncoimmunology 2016;5:e1002725.10.1080/2162402X.2014.1002725Search in Google Scholar PubMed PubMed Central

77 Bhutiani N, Li Q, Anderson CD, Gu T, Egilmez NK. Combined oral cytokine therapy effectively treats colon cancer in a murine model. AACR; 2017.10.1158/1538-7445.AM2017-1605Search in Google Scholar

78 Yang P-M, Chou C-J, Tseng S-H, Hung C-F. Bioinformatics and in vitro experimental analyses identify the selective therapeutic potential of interferon gamma and apigenin against cervical squamous cell carcinoma and adenocarcinoma. Oncotarget 2017;8:46145.10.18632/oncotarget.17574Search in Google Scholar PubMed PubMed Central

79 Brar K, Leung DY. Recent considerations in the use of recombinant interferon gamma for biological therapy of atopic dermatitis. Expert opinion on biological therapy 2016;16:507-14.10.1517/14712598.2016.1135898Search in Google Scholar PubMed PubMed Central

80 Leiding JW, Holland SM. Chronic granulomatous disease. 2016.Search in Google Scholar

81 Rhein BA, Powers LS, Rogers K, et al. Interferon-γ inhibits Ebola virus infection. PLoS pathogens 2015;11:e1005263.10.1371/journal.ppat.1005263Search in Google Scholar PubMed PubMed Central

82 Delsing CE, Gresnigt MS, Leentjens J, et al. Interferon-gamma as adjunctive immunotherapy for invasive fungal infections: a case series. BMC infectious diseases 2014;14:166.10.1186/1471-2334-14-166Search in Google Scholar PubMed PubMed Central

83 Segal BH, Walsh TJ. Current approaches to diagnosis and treatment of invasive aspergillosis. American journal of respiratory and critical care medicine 2006;173:707-17.10.1164/rccm.200505-727SOSearch in Google Scholar PubMed

84 Malmvall BE, Follin P. Successful interferon-gamma therapy in a chronic granulomatous disease (CGD) patient suffering from Staphylococcus aureus hepatic abscess and invasive Candida albicans infection. Scandinavian journal of infectious diseases 1993;25:61-6.10.1080/00365549309169671Search in Google Scholar

85 Skerrett SJ, Martin TR. Intratracheal interferon-gamma augments pulmonary defenses in experimental legionellosis. Am J Respir Crit Care Med 1994;149:50-8.10.1164/ajrccm.149.1.8111597Search in Google Scholar

86 Coelho C, Casadevall A. Cryptococcal therapies and drug targets: the old, the new and the promising. Cellular microbiology 2016;18:792-9.10.1111/cmi.12590Search in Google Scholar

87 Marciano BE, Spalding C, Fitzgerald A, et al. Common severe infections in chronic granulomatous disease. Clinical Infectious Diseases 2014;60:1176-83.10.1093/cid/ciu1154Search in Google Scholar

88 Seger RA. Modern management of chronic granulomatous disease. British Journal of Haematology 2008;140:255-66.10.1111/j.1365-2141.2007.06880.xSearch in Google Scholar

89 Hashemi H, Mohebbi M, Mehravaran S, Mazloumi M, Jahanbani-Ardakani H, Abtahi S-H. Hyperimmunoglobulin E syndrome: Genetics, immunopathogenesis, clinical findings, and treatment modalities. Journal of Research in Medical Sciences : The Official Journal of Isfahan University of Medical Sciences 2017;22:53.10.4103/jrms.JRMS_1050_16Search in Google Scholar

90 Badaro R, Falcoff E, Badaro FS, et al. Treatment of visceral leishmaniasis with pentavalent antimony and interferon gamma. New England Journal of Medicine 1990;322:16-21.10.1056/NEJM199001043220104Search in Google Scholar

91 Holland SM. Immunotherapy of mycobacterial infections. Seminars in respiratory infections; 2001. p. 47-59.10.1053/srin.2001.22728Search in Google Scholar

92 Condos R, Rom WN, Schluger NW. Treatment of multidrug-resistant pulmonary tuberculosis with interferon-γ via aerosol. The Lancet 1997;349:1513-5.10.1016/S0140-6736(96)12273-XSearch in Google Scholar

93 Milanés-Virelles MT, García-García I, Santos-Herrera Y, et al. Adjuvant interferon gamma in patients with pulmonary atypical Mycobacteriosis: a randomized, double-blind, placebo-controlled study. BMC infectious diseases 2008;8:17.10.1186/1471-2334-8-17Search in Google Scholar PubMed PubMed Central

94 Vincent J-L, Sun Q, Dubois M-J. Clinical trials of immunomodulatory therapies in severe sepsis and septic shock. Clinical Infectious Diseases 2002;34:1084-93.10.1086/339549Search in Google Scholar PubMed

95 Raghu G, Brown KK, Bradford WZ, et al. A placebo-controlled trial of interferon gamma-1b in patients with idiopathic pulmonary fibrosis. New England Journal of Medicine 2004;350:125-33.10.1056/NEJMoa030511Search in Google Scholar PubMed

96 Singhal J, Agrawal N, Vashishta M, et al. Suppression of dendritic cell-mediated responses by genes in calcium and cysteine protease pathways during Mycobacterium tuberculosis infection. Journal of Biological Chemistry 2012;287:11108-21.10.1074/jbc.M111.300319Search in Google Scholar PubMed PubMed Central

97 Green DS, Nunes AT, Annunziata CM, Zoon KC. Monocyte and interferon based therapy for the treatment of ovarian cancer. Cytokine & growth factor reviews 2016;29:109-15.10.1016/j.cytogfr.2016.02.006Search in Google Scholar PubMed PubMed Central

98 Windbichler G, Hausmaninger H, Stummvoll W, et al. Interferon-gamma in the first-line therapy of ovarian cancer: a randomized phase III trial. British journal of cancer 2000;82:1138-44.10.1054/bjoc.1999.1053Search in Google Scholar PubMed PubMed Central

99 Bosserhoff A, Kortylewski M, Komyod W, Kauffmann M-E, Heinrich PC, Behrmann I. Interferon-γ-mediated growth regulation of melanoma cells: involvement of STAT1-dependent and STAT1-independent signals. Journal of Investigative Dermatology 2004;122:414-22.10.1046/j.0022-202X.2004.22237.xSearch in Google Scholar PubMed

100 McCabe A, Zhang Y, Thai V, Jones M, Jordan MB, MacNamara KC. Macrophage-Lineage Cells Negatively Regulate the Hematopoietic Stem Cell Pool in Response to Interferon Gamma at Steady State and During Infection. Stem Cells 2015;33:2294-305.10.1002/stem.2040Search in Google Scholar PubMed PubMed Central

101 George PM, Badiger R, Alazawi W, Foster GR, Mitchell JA. Pharmacology and therapeutic potential of interferons. Pharmacology & Therapeutics 2012;135:44-53.10.1016/j.pharmthera.2012.03.006Search in Google Scholar PubMed

102 Moss JW, Ramji DP. Interferon-gamma: Promising therapeutic target in atherosclerosis. World journal of experimental medicine 2015;5:154-9.10.5493/wjem.v5.i3.154Search in Google Scholar PubMed PubMed Central

103 Cheng X, Ding Y, Xia C, et al. Atorvastatin modulates Th1/Th2 response in patients with chronic heart failure. Journal of cardiac failure 2009;15:158-62.10.1016/j.cardfail.2008.10.001Search in Google Scholar PubMed

104 Forster O, Hilfiker-Kleiner D, Ansari AA, et al. Reversal of IFN-gamma, oxLDL and prolactin serum levels correlate with clinical improvement in patients with peripartum cardiomyopathy. European journal of heart failure 2008;10:861-8.10.1016/j.ejheart.2008.07.005Search in Google Scholar PubMed

105 Liu J, Gao L, Zang D. Elevated levels of IFN-γ in CSF and serum of patients with amyotrophic lateral sclerosis. PloS one 2015;10:e0136937.10.1371/journal.pone.0136937Search in Google Scholar PubMed PubMed Central

106 Mount MP, Lira A, Grimes D, et al. Involvement of Interferon-γ in Microglial-Mediated Loss of Dopaminergic Neurons. The Journal of Neuroscience 2007;27:3328-37.10.1523/JNEUROSCI.5321-06.2007Search in Google Scholar PubMed PubMed Central

107 Gallardo E, de Andres I, Illa I. Cathepsins are upregulated by IFN-gamma/STAT1 in human muscle culture: a possible active factor in dermatomyositis. Journal of neuropathology and experimental neurology 2001;60:847-55.10.1093/jnen/60.9.847Search in Google Scholar PubMed

108 Geraghty P, Greene CM, O’Mahony M, O’Neill SJ, Taggart CC, McElvaney NG. Secretory leucocyte protease inhibitor inhibits interferon-γ-induced cathepsin s expression. Journal of Biological Chemistry 2007;282:33389-95.10.1074/jbc.M706884200Search in Google Scholar PubMed

109 Giroux M, Schmidt M, Descoteaux A. IFN-γ-induced MHC class II expression: transactivation of class II transactivator promoter IV by IFN regulatory factor-1 is regulated by protein kinase C-α. The Journal of Immunology 2003;171:4187-94.10.4049/jimmunol.171.8.4187Search in Google Scholar PubMed

110 Angell TE, Lechner MG, Jang JK, LoPresti JS, Epstein AL. MHC class I loss is a frequent mechanism of immune escape in papillary thyroid cancer that is reversed by interferon and selumetinib treatment in vitro. Clinical Cancer Research 2014;20:6034-44.10.1158/1078-0432.CCR-14-0879Search in Google Scholar PubMed PubMed Central

111 Saric T, Chang S-C, Hattori A, et al. An IFN-γ-induced aminopeptidase in the ER, ERAP1, trims precursors to MHC class I-presented peptides. Nature immunology 2002;3:1169-76.10.1038/ni859Search in Google Scholar PubMed

112 Wang D, Quan Y, Yan Q, Morales JE, Wetsel RA. Targeted Disruption of the β2-Microglobulin Gene Minimizes the Immunogenicity of Human Embryonic Stem Cells. Stem cells translational medicine 2015;4:1234-45.10.5966/sctm.2015-0049Search in Google Scholar PubMed PubMed Central

113 Chang JH, Kim YJ, Han SH, Kang CY. IFN-γ-STAT1 signal regulates the differentiation of inducible Treg: Potential role for ROS-mediated apoptosis. European journal of immunology 2009;39:1241-51.10.1002/eji.200838913Search in Google Scholar PubMed

114 Lambeth JD. NOX enzymes and the biology of reactive oxygen. Nat Rev Immunol 2004;4:181-9.10.1038/nri1312Search in Google Scholar PubMed

115 Rovetta AI, Pena D, Hernandez Del Pino RE, et al. IFNG-mediated immune responses enhance autophagy against Mycobacterium tuberculosis antigens in patients with active tuberculosis. Autophagy 2014;10:2109-21.10.4161/15548627.2014.981791Search in Google Scholar PubMed PubMed Central

116 Tu SP, Quante M, Bhagat G, et al. Interferon-γ inhibits gastric carcinogenesis by inducing epithelial cell autophagy and T cell apoptosis. Cancer research 2011;71:4247-59.10.1158/0008-5472.CAN-10-4009Search in Google Scholar PubMed PubMed Central

117 Kim WH, Lee JW, Gao B, Jung MH. Synergistic activation of JNK/ SAPK induced by TNF-α and IFN-γ: apoptosis of pancreatic β-cells via the p53 and ROS pathway. Cellular signalling 2005;17:1516-32.10.1016/j.cellsig.2005.03.020Search in Google Scholar PubMed

118 Watanabe Y, Suzuki O, Haruyama T, Akaike T. Interferon-γ induces reactive oxygen species and endoplasmic reticulum stress at the hepatic apoptosis. Journal of cellular biochemistry 2003;89:244-53.10.1002/jcb.10501Search in Google Scholar PubMed

119 Gil MP, Bohn E, O’Guin AK, et al. Biologic consequences of Stat1-independent IFN signaling. Proceedings of the National Academy of Sciences 2001;98:6680-5.10.1073/pnas.111163898Search in Google Scholar PubMed PubMed Central

120 McNab FW, Rajsbaum R, Stoye JP, O’Garra A. Tripartite-motif proteins and innate immune regulation. Current opinion in immunology 2011;23:46-56.10.1016/j.coi.2010.10.021Search in Google Scholar PubMed

121 Jutras I, Houde M, Currier N, et al. Modulation of the phagosome proteome by interferon-gamma. Molecular & cellular proteomics : MCP 2008;7:697-715.10.1074/mcp.M700267-MCP200Search in Google Scholar PubMed

122 Ghigo E, Capo C, Tung C-H, Raoult D, Gorvel J-P, Mege J-L. Coxiella burnetii survival in THP-1 monocytes involves the impairment of phagosome maturation: IFN-γ mediates its restoration and bacterial killing. The Journal of Immunology 2002;169:4488-95.10.4049/jimmunol.169.8.4488Search in Google Scholar PubMed

123 Santic M, Molmeret M, Abu Kwaik Y. Modulation of biogenesis of the Francisella tularensis subsp. novicida-containing phagosome in quiescent human macrophages and its maturation into a phagolysosome upon activation by IFN-γ. Cellular microbiology 2005;7:957-67.10.1111/j.1462-5822.2005.00529.xSearch in Google Scholar PubMed

124 Pei G, Bronietzki M, Gutierrez MG. Immune regulation of Rab proteins expression and intracellular transport. Journal of leukocyte biology 2012;92:41-50.10.1189/jlb.0212076Search in Google Scholar

125 Huang Y, Krein PM, Winston BW. Characterization of IFN-gamma regulation of the complement factor B gene in macrophages. Eur J Immunol 2001;31:3676-86.10.1002/1521-4141(200112)31:12<3676::AID-IMMU3676>3.0.CO;2-KSearch in Google Scholar

126 Gasque P, Dean YD, McGreal EP, VanBeek J, Morgan BP. Complement components of the innate immune system in health and disease in the CNS. Immunopharmacology 2000;49:171-86.10.1016/S0162-3109(00)80302-1Search in Google Scholar

127 Carroll MC. The complement system in regulation of adaptive immunity. Nature immunology 2004;5:981-6.10.1038/ni1113Search in Google Scholar

128 Pearse RN, Feinman R, Shuai K, Darnell JE, Ravetch JV. Interferon gamma-induced transcription of the high-affinity Fc receptor for IgG requires assembly of a complex that includes the 91-kDa subunit of transcription factor ISGF3. Proceedings of the National Academy of Sciences 1993;90:4314-8.10.1073/pnas.90.9.4314Search in Google Scholar

129 Nimmerjahn F, Ravetch JV. Divergent immunoglobulin g subclass activity through selective Fc receptor binding. Science 2005;310:1510-2.10.1126/science.1118948Search in Google Scholar

130 Klug-Micu GM, Stenger S, Sommer A, et al. CD40 ligand and interferon-γ induce an antimicrobial response against Mycobacterium tuberculosis in human monocytes. Immunology 2013;139:121-8.10.1111/imm.12062Search in Google Scholar

131 Albanesi C, Fairchild HR, Madonna S, et al. IL-4 and IL-13 negatively regulate TNF-α-and IFN-β-induced γ-defensin expression through STAT-6, suppressor of cytokine signaling (SOCS)-1, and SOCS-3. The Journal of Immunology 2007;179:984-92.10.4049/jimmunol.179.2.984Search in Google Scholar

132 Taylor MW, Feng G. Relationship between interferon-gamma, indoleamine 2, 3-dioxygenase, and tryptophan catabolism. The FASEB Journal 1991;5:2516-22.10.1096/fasebj.5.11.1907934Search in Google Scholar

Abbreviations

SNP

Single-nucleotide polymorphism

NF-κB

Nuclear factor kappa B

cAMP

Cyclic adenosine monophosphate

IRF

Interferon regulatory factor

IDO

Indoleamine 2,3 dioxygenase

CXCL

Chemokine (C-X-C motif) ligand

MIP

Macrophage inflammatory protein

IP-10

Interferongamma-induced protein 10

CCL

Chemokine (C-C motif) ligand

Received: 2018-02-13
Accepted: 2018-04-20
Published Online: 2018-05-30

© 2018 Gunjan Kak et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 License.

Downloaded on 28.3.2024 from https://www.degruyter.com/document/doi/10.1515/bmc-2018-0007/html
Scroll to top button