TREIMM-1157; No. of Pages 18

Feature review

Nitric oxide synthase in innate and adaptive immunity: an update Christian Bogdan Mikrobiologisches Institut – Klinische Mikrobiologie, Immunologie, und Hygiene, Friedrich-Alexander-Universita¨t (FAU) ErlangenNu¨rnberg, Universita¨tsklinikum Erlangen, Wasserturmstraße 3/5, 91054 Erlangen, Germany

Thirty years after the discovery of its production by activated macrophages, our appreciation of the diverse roles of nitric oxide (NO) continues to grow. Recent findings have not only expanded our understanding of the mechanisms controlling the expression of NO synthases (NOS) in innate and adaptive immune cells, but have also revealed new functions and modes of action of NO in the control and escape of infectious pathogens, in T and B cell differentiation, and in tumor defense. I discuss these findings, in the context of a comprehensive overview of the various sources and multiple reaction partners of NO, and of the regulation of NOS2 by micromilieu factors, antisense RNAs, and ‘unexpected’ cytokines. NO and the immune system: no end (yet) to new paradigms The production of large amounts of nitrite (NO2) and nitrate (NO3) by mouse macrophages stimulated with lipopolysaccharide (LPS) and interferon (IFN)-g [1,2], and the L-arginine-dependency of the NO2 generation and of the tumor cytotoxic activity of activated macrophages [3,4], were the pioneering observations that ignited the interest of immunologists in the small inorganic radical of nitric oxide (NO). With the subsequent purification and cloning of the enzyme NOS2 [5,6], the production of polyclonal and monoclonal anti-NOS2-peptide antibodies, the development of NOS2-selective inhibitors [7,8], and the generation of mice with a functionally complete [9,10] or partial deletion of the Nos2 gene [11,12], the necessary tools became available to study in detail the expression, regulation, and function of NOS2 in the immune system. While at the beginning of NOS2 research NO was mainly viewed as an antimicrobial, tumoricidal, and tissue-damaging effector molecule operating in the innate arm of the immune system, it soon became clear that it strongly affects adaptive immune responses and also exerts cytoprotective effects due to its pleiotropic and versatile signaling functions (reviewed in [13–19]). In recent years, our understanding of the role of NO in the immune system has been enriched by fascinating discoveries which include the exquisite regulation of Corresponding author: Bogdan, C. ([email protected]). Keywords: nitric oxide synthases (NOS1/nNOS; NOS2/iNOS; NOS3/eNOS); myeloid cells; microenvironment; antimicrobial activity; Th17 cells; B cells. 1471-4906/ ß 2015 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.it.2015.01.003

NOS2 expression by micromilieu factors, the function of bacterial pathogen-derived NO in modulating and evading the host defense, and the NO-dependent control of pathogens by depletion of micronutrients. In addition, recent studies have provided novel insights into the effect of exogenous or endogenous NO on the differentiation of T helper cell subsets and into the NOS2-dependent communication between mesenchymal cells and T cells. This review aims to integrate these new findings into previously existing concepts. Sources of NO Mammalian NO synthases NOS2 is a homodimeric enzyme that, like all other NOS isoforms, converts L-arginine and oxygen into L-citrulline and NO in a complex oxidoreductase reaction (Box 1 and Figure 1A). A characteristic feature of NOS2 is its lack of expression in strictly resting cells. Instead, it is induced by immunological stimuli in a calcium-independent manner, which led to its original designation as inducible NO synthase (iNOS) [13]. The host cell localization of NOS2 has been mainly investigated in macrophages and neutrophils, where enzymatically active NOS2 can be detected in the cytosol, in small vesicles (still awaiting further characterization), in primary and tertiary granules, in the vicinity of phagosomes bearing inert particles or avirulent pathogens, attached to the submembranal actin cytoskeleton, or in mitochondria [19–26]. There is evidence from non-myeloid cells for sequestration of inactive NOS2 in cytoplasmic, perinuclear inclusion-like bodies (aggresomes) [27]. NOS2 has also been reported to localize to the nucleus, but the biological role of nuclear NOS2 (and of other NOS isoforms) is currently unclear ([25,28,29] and references therein). In addition to NOS2 there are two other isoforms of NOS: neuronal NOS (nNOS, NOS1) and endothelial NOS (eNOS, NOS3), originally named according to their predominant tissue distribution. These two isoforms have also been termed cNOS, because, unlike NOS2, they are constitutively expressed and their activity is dependent on prior calcium fluxes that enable the binding of calmodulin [13]. However, NOS1 and NOS3 are now recognized to have a wider cell and tissue distribution, and to be also regulated by cytokines, microbial products, hormones, and micromilieu factors [30–33]. NOS1 and NOS3 affect the differentiation and function of immune cells in vitro and modulate immune responses and inflammatory processes in vivo [29,34–45] (for further references see also [14,17]). Trends in Immunology xx (2015) 1–18

1

TREIMM-1157; No. of Pages 18

Feature review Box 1. Type 2 nitric oxide synthase Type 2 nitric oxide synthase (NOS2) is a heme-containing flavoprotein and oxidoreductase which requires the presence of molecular oxygen, NADPH, calmodulin, and four redox-active prosthetic groups [i.e., FAD, FMN, iron protoporphyrin IX (heme), and tetrahydrobiopterin] to catalyze the conversion of the semi-essential amino acid L-arginine into L-citrulline, water, and the NO radical (in the following abbreviated as NO), with Nv-hydroxy-L-arginine being formed as transient, enzyme-bound intermediate [246]. Owing to failure to crystallize the holoenzyme, the structure of NOS2 has so far remained speculative. Recently, high-throughput single-particle electron microscopy revealed that in the dimeric NOS2 holoenzyme the two oxidase domains are dimerized in an antiparallel fashion and are flanked by the two separate reductase domains [247], thus visually confirming earlier mechanistic models [246]. Three flexible linker domains account for the high conformational mobility of the enzyme. This allows one of the reductase domains to flip over and closely interact with the oxidase domain, which is essential for electron transfer and catalysis [247].

Trends in Immunology xxx xxxx, Vol. xxx, No. x

(A)

Eukaryotes L-arginine

Asl

+ O2

Argininosuccinate

NOHA NOS

Ass1

Aspartate

Arg

L-citrulline + NO

pH↓ XO heme UV

NO targets and signaling NO radical, the primary product released from the Fe2+– heme complex within NOS2, is a labile compound. Its multiple reaction partners, which explain both the toxic and the regulatory effects of NO, include (i) the thiol groups of cysteines within peptides or proteins, leading 2

ADC

Urea + Ornithin

AGAT

CO2 + Agmane

Ornithine + Guanidinoacetate

NO2 –

Pyrimidine nucleodes NO3 –

(B)

Other sources of NO Under normoxia, NO derived from mammalian NOS is rapidly oxidized to nitrite and nitrate, which are in principle stable products. Nitrite, however, can easily become a source of NO, either non-enzymatically under acidic conditions (e.g., in the stomach or in inflamed tissues) or via the nitrite reductase activities of xanthine oxidoreductase and iron porphyrin-containing proteins during hypoxia (e.g., deoxyhemoglobin, desoxymyoglobin, cytochrome c) [46–48] (Figure 1A). The Fe3+–porphyrin-catalyzed reduction of nitrite can be efficiently assisted by thiols (e.g., glutathione) and sulfides (e.g., H2S), which then, in addition to NO, leads to the generation of nitroxyl (HNO) and thionitrous acid (HSNO), a potent S-nitrosating agent [49]. Irradiation of human skin with ultraviolet A light (315– 400 nm) or with blue light (420–453 nm) led to enzymeindependent formation of NO and other reactive nitrogen species (RNS) due to decomposition of S-nitrosothiol and nitrite stores in the epidermis and dermis of the skin [50,51]. While beneficial effects of UVA irradiation on blood pressure and blood flow have recently been reported [52], the immunological consequences and potential tissuedamaging effects of this pathway still need to be established. Finally, it has long been known that many environmental, commensal, and pathogenic bacteria reduce NOS2derived or dietary nitrate or nitrite to NO [or even further to nitrous oxide (N2O) or nitrogen] under anaerobic conditions, via a process termed nitrate respiration, nitrate dissimilation, or denitrification [53] (Figure 1B). More recently, shortened NOS-like proteins (bNOS) with high sequence similarity to the mammalian NOS-oxygenase domain, but without reductase domain, were discovered in Gram-positive bacteria [54]. Surprisingly, these bacterial NOS seem to play an important role in host–pathogen interactions, as discussed further below.

+ glycine

NOS

Creane

Polyamines Proline collagen

Prokaryotes

NO3 – NAR

NO2 – NIR

NO NOR

L-arginine bNOS

L-arginine

L-arginine

Arg

ADI

+ O2

NOHA

H2 O NH3 + H+

bNOS

L-citrulline + NO

Urea + Ornithine

Citrulline PO4 – + H+ OCT

N2 O N 2 OR

N2

Ornithine Carbamoylphosphate ADP + 2H+ CK ATP + NH3 CO2 TRENDS in Immunology

Figure 1. Arginine metabolism and the generation of NO in host cells and bacteria. (A) In mammalian cells four enzymatic pathways account for arginine metabolism –, the NO synthases (NOS), the arginases (Arg), L-arginine decarboxylase (ADC), and L-arginine:glycine amidinotransferase (AGAT). NO also arises NOS-independently by enzymatic (xanthine oxidase – XO) or nonenzymatic reduction of nitrite (i.e., by acidic pH, UV irradiation, or reaction with Fe3+ porphyrins such as heme). The arginase product ornithine and the ADC product agmatine are precursors for the synthesis of polyamines; ornithine is also a precursor for the generation of proline and collagens. In macrophages the NOS product L-citrulline can be recycled to L-arginine via argininosuccinate synthase 1 (Ass1) and argininosuccinate lyase (Asl). Arginine can also be generated from proteins or peptides carrying N- or C-terminal arginine residues via the activity of carboxypeptidases (e.g., CPD, CPM) or aminopeptidases (e.g., ERAP1). (B) Several bacterial species (e.g., E. coli, S. enterica Typhimurium, M. tuberculosis, M. catarrhalis) use nitrate and nitrite as energy sources under anaerobic conditions (nitrate respiration or dissimilation), which leads to the generation of NO, nitrous oxide (dinitrogen monoxide, N2O), or nitrogen via the activities of respiratory nitrate reductase (NAR), nitrite reductase (NIR), nitric oxide reductase (NOR), or nitrous oxide reductase (N2OR). In some prokaryotes (e.g., S. aureus), NO synthesis results from the presence of a bacterial NO synthase (bNOS). Some bacterial pathogens are able to degrade (host-derived) arginine by expressing arginase (e.g., Helicobacter pylori) or the enzymes of the ADI pathway (ADI, arginine deiminase or arginine dihydrolase; OCT, ornithine carbamoyl transferase; CK, carbamate kinase) (e.g., S. pyogenes, S. enterica). NOHA, Nv-hydroxy-L-arginine.

to the formation of S-nitrosothiols [55]; (ii) superoxide anions (O2), which gives rise to peroxynitrous acid/peroxynitrite (ONOO), capable of modifying proteins (e.g., by tyrosine nitration [56]); (iii) divalent cations such as

TREIMM-1157; No. of Pages 18

Feature review Fe2+ (in heme groups or iron–sulfur clusters) or Zn2+ (in zinc–sulfur clusters), which regulate the function of many transcription factors and enzymes, including the mammalian NO sensor sGC (soluble guanylate cyclase) that generates cGMP after activation by NO [57,58]; (iv) nucleic acids, where NO affords deamination and leads to mutagenesis [59]; and (v) unsaturated lipids, where, for example, NO-derived NO2 can attack alkene double bonds leading to nitrolipids [60]. The post-translational cysteine modification by S-nitrosation and its reversal (Box 2) is a key mechanism for the activation or inactivation of protein functions, and underlies many of the known signaling effects of NOS2-derived NO in the immune system (reviewed in [61–63]) (for examples see also text below). Based on a recent analysis using a 16, 368 protein microarray chip, the human S-nitrosocysteine proteome comprises at least 834 proteins [64]. In an earlier study, mass spectrometry of mouse tissues (e.g., lung, liver, thymus) identified more than 1000 S-nitrosylated cysteine residues in 647 proteins, many of which belong to key metabolic pathways. Depending on the organ, 56–90% of Snitrosylation events resulted from NOS3 activity [65]. A whole new avenue of research has been opened up by the possible interactions of RNS (e.g., NO, S-nitrosothiols, ONOO) with hydrogen sulfide (also termed sulfane; H2S) or its anion (HS) [66,67]. H2S/HS, which similarly to NO has been recognized as an inorganic signaling molecule in mammals, appears to modulate multiple physiological processes including immune and inflammatory responses

Box 2. S-nitrosation and denitrosation S-nitrosothiols (R-S-NO) are formed, when a nitroso (NO)-group is covalently added to the thiol (syn. sulfhydryl) group of the amino acid cysteine. The chemically correct name for this process and its reversal is S-nitrosation and S-denitrosation, but, by analogy to the nomenclature of other post-translational modifications (e.g., glycosylation), the terms S-nitrosylation or S-denitrosylation are frequently used [62,248,249]. As described in detail elsewhere [248], S-nitrosothiols can emerge by the reaction of: (i) NO radical with NO 2 (e.g., derived from the decay of peroxynitrite), leading to the formation of N2O3, which then reacts with thiolates (RS) to give rise to R-S-NO (NO autooxidation pathway), where R signifies any cysteine residue within small peptides (e.g., glutathione) or proteins; (ii) iron-nitrosyl-complexes (heme-Fe3+-NO , heme-Fe2+-+NO) with R-S to form R-S-NO and Fe2+-heme (transition metalcatalyzed pathway); (iii) a thiol (R-S-H) with another radical (X) leading to X-H and a thiyl radical (R-S) which then pairs with NO to form R-S-NO (thiyl radical recombination pathway); (iv) an already formed S-nitrosothiol (R1-S-NO) with a thiolate (R2S) leading to an R1-S-(N=O)-S-R2 intermediate with subsequent transfer of the NO-group and the generation of R1-S and R2-S-NO (the transnitrosation or exchange pathway). S-nitrosated proteins can be denitrosated by the catalytic action of proteins (‘denitrosylases’) such as thioredoxin or S-nitrosoglutathione-reductase (GSNO-R). This might relieve nitrosative stress and provide the basis for a redox-switch in protein function. GSNOR is crucial for the development and function of the immune system, as has been demonstrated with GSNO-R deficient mice [55,204] (see main text). Recently, S-nitrosylated proteins were also detected in NOexposed bacteria, where they conferred protection against further nitrosative stress [189] (see main text for further details).

Trends in Immunology xxx xxxx, Vol. xxx, No. x

[68]. It will be a major challenge in the future to unravel the full chemistry and the functional consequences of the crosstalk between H2S and RNS. Regulation of NOS2 expression Cytokines and microbial products In the mouse system interferons [IFN-g and type I IFNs (IFN-a/b)] and microbial pathogens or products (e.g., LPS) are prototypic transcriptional inducers of NOS2 which effectively stimulate macrophages for the release of high amounts of NO. IFNs and LPS elicit the dimerization of STAT1, the expression of interferon-regulatory factor (IRF)-1, and the formation of the interferon-stimulated gamma factor (ISGF) 3-complex (consisting of STAT1, STAT2, and IRF9), or the activation of NF-kB, respectively, all of which interact with binding sites in the Nos2 promoter (reviewed in [69]). Using mouse bone marrowderived macrophages infected with Listeria (L.) monocytogenes and chromatin immunoprecipitation, Farlik et al. unraveled the molecular basis for the previously described [70,71] synergistic induction of NOS2 gene transcription by type I IFN and microbial components [72]. As a first step, bacterial components interacting with membrane-bound or cytosolic pattern recognition receptors (PRR) trigger the activation of NF-kB and the production of endogenous IFNab, respectively. The binding of NF-kB to the Nos2 promoter led to the subsequent recruitment of the transcription factor (TF) IIH and its associated cyclin-dependent kinase (CDK) 7 subunit. In a second step, IFN-a/b feedback signaling caused the formation and promoter binding of the ISGF3 complex, thus enabling the co-recruitment and promoter-binding of RNA polymerase II and its phosphorylation (C-terminal domain at serine position 7) by TFIIH– CDK7 [72]. This latter process also required the binding of members of the bromodomain and extra terminal domain (BET) protein family to the Nos2 promoter [73]. Thus, NFkB and ISGF3 act sequentially and cooperatively at the Nos2 promoter. Research on NOS2 has been dominated by mouse studies. The molecular analysis of NOS2 expression in other species, however, is highly relevant to settle discussions on the extent and biological relevance of species-specific differences. Today, there is no doubt that human cells (e.g., hepatocytes, monocyte-derived macrophages or dendritic cells, tissue macrophages) are able to express NOS2 protein and activity in vitro and in vivo (reviewed in [14,15,74– 80]). Characterization of NOS2 expression in human cells has confirmed the crucial role of NF-kB and STAT1a, but also revealed differences in the promoter structure (explaining the hyporesponsiveness to IFN-g and LPS), the importance of post-transcriptional mechanisms, and the impact of cell origin, culture conditions, and stimuli [77,81,82]. In human alveolar macrophages stimulated with IFN-g plus LPS, the lack of NOS2 mRNA and protein expression was recently attributed to epigenetic gene silencing by CpG methylation, histone modifications, and chromatin compaction [83]. While these results further corroborate the hyporesponsiveness of human macrophages to IFN-g/LPS, they contrast starkly with the presence of NOS2 protein in alveolar and granulomatous macrophages of patients infected with, for example, M. 3

TREIMM-1157; No. of Pages 18

Feature review tuberculosis ([79] and references therein) and by no means question the expression of NOS2 by human cells. In hamster macrophages the low levels or expression of NOS2 in response to IFN-g and LPS could be partially explained by the absence of the NF-IL6 binding element in the proximal basal promoter [84]. In recent years, several unexpected cytokines and growth factors have been identified as (co)inducers of NOS2 in macrophages. Mouse macrophages infected with Leishmania (L.) amazonensis parasites and stimulated with interleukin (IL)-1b released remarkable amounts of NO. Roughly 40% of the NO production was dependent on inflammasome (caspase 1) activation and consecutive synthesis of endogenous IFN-g [85]. This suggests that IL-1b itself can confer NOS2-inducing signals, confirming previous findings obtained with other cell types of different species [14,86,87]. In keeping with this in vitro observation, L. amazonensis-infected mice deficient for caspase 1, the respective inflammasome components (i.e., NLRP3 and ASC), or the IL-1 receptor (IL1R) developed significantly more severe skin lesions than wild type controls [85]. IL-33, another member of the IL-1 family and best known for its potent induction of type 2 T helper (Th2) cell responses, triggered the expression of NOS2 mRNA, protein and activity in mouse macrophages. The effect of IL-33 required the ST2 ligand-binding component of the IL-33R, protein kinase B (Akt) and b-catenin activity, and partially also the IL-1R (raising the possibility of secondary induction of IL-1 by IL-33). The biological relevance of the findings was demonstrated in a Staphylococcus (S.) aureus infection model, where IL-33 inhibited the growth of the bacteria in a NOS2-dependent manner in vitro and in vivo [88]. Erythropoietin (EPO) is the key growth factor for red blood cell development and differentiation and has been successfully used for the treatment of some forms of anemia. Because non-erythroid cells and tissues can also respond to EPO, based on their expression of tissue-protective EPO heteroreceptors, several groups have investigated the function of EPO in the immune system [89]. This led to the discovery that EPO inhibited the activation and promoter-binding of NF-kB p65, and thereby blocked the induction of NOS2 and proinflammatory cytokines by microbial Toll-like receptor (TLR) ligands in macrophages, an effect which EPO shares with classical macrophage-deactivating cytokines (IL-4, IL-10, IL-13, TGF-b). As expected, EPO impaired the control of Salmonella (S.) enterica Typhimurium in vivo, but improved the outcome of cytokine-driven pathologies such as chemically (TNBS – 2,4,6trinitrobenzenesulfonic acid)-induced colitis [90] or LPSmediated endotoxin shock [91]. These results are relevant for the application of EPO in patients with anemia due to malignancies or chronic inflammatory diseases. Non-coding RNAs MicroRNAs (miRNAs) have been reported to affect the level of NOS2 mRNA and protein expression in various cell types including mouse macrophages [92–95], splenic lymphocytes [96], mouse mesenchymal stem cells (MSCs) [97], human endothelial cells [98], and hepatocytes [86]. However, in most cases the identified miRNAs (i.e., 4

Trends in Immunology xxx xxxx, Vol. xxx, No. x

miR-125a-5p, miR-146a, miR-149, miR-155, miR-301a) acted indirectly by blocking the expression of signaling or transcription factors (e.g., SOCS-1, IRAK-1, NF-kBrepressing factor, Kru¨ppel-like factor 13, Rheb) that positively or negatively regulated the expression of NOS2 mRNA. Direct interactions of miRNAs with the 30 -untranslated region (UTR) of NOS2 mRNA, with subsequent downregulation of NOS2 protein levels and NO production, have been convincingly documented for miR-939 (in human hepatocytes) [86] and miR-26a (in human T cell lymphoma cells) [99], and possibly also exist for miR146a as shown in mouse renal carcinoma cells [100]. A related form of NOS regulation has emerged from the discovery of natural non-coding antisense transcripts (NATs) that are derived from the opposite DNA strand (i.e., the strand complementary to the coding strand). NATs are generated by a large proportion of the mammalian genome [101]. In the case of the NOS genes, antisense transcripts were detected that either originated from the same gene locus as their sense counterparts (cis-NATs) [102,103] or from a different locus (trans-NATs) [104]. Depending on whether the NATs interacted with untranslated or translated regions of the mRNA, they stabilized or downregulated NOS2 mRNA or inhibited NOS3 mRNA translation ([102–104] and references therein). Micromilieu factors and metabolites Low oxygen tension is a characteristic both of sites of inflammation and of tumor tissues. Hypoxia causes stabilization of the transcription factor hypoxia-inducible factor (HIF)-1a, which then switches on genes that ultimately help to improve tissue oxygenation and resolution of inflammation [22,105,106]. Hypoxia has three major effects on NOS2. First, it impairs the synthesis of NO by NOS2 because of the lack of the substrate oxygen ([107] and references therein). Second, hypoxia disrupts the binding between NOS2 and the adaptor protein a-actinin 4, and thereby prevents the attachment of NOS2 to the actin cytoskeleton, which is a prerequisite for NOS2 activity [22]. Third, NOS2 belongs to those genes that are transcriptionally coinduced by HIF-1a in myeloid cells such as macrophages [108]. As shown in mouse dendritic cells (DC), hypoxia alone was a weak inducer of NOS2 mRNA. However, it strongly synergized with TLR3, TLR4, and TLR9 agonists for the upregulation of NOS2 mRNA and protein. Unexpectedly, under normoxic conditions TLRinduced NOS2 expression did not only require NF-kB and Myd88 signaling, but was partially also dependent on HIF-1a [109]. Using an innovative transcutaneous sensor technology, a direct correlation between tissue oxygenation, NO production, and antimicrobial activity was recently documented in mouse L. major skin infections [107]. In a Mycobacterium (M.) marinum zebrafish model, stabilization of HIF-1a was paralleled by an increased NOS2 activity and an improved pathogen control [110]. Another micromilieu factor that is poorly reflected by the standard tissue culture conditions is tonicity (osmolarity). In two pioneering studies, mouse lymphoid tissues were discovered to be hyperosmolar relative to serum [111], and macrophages were found to respond to the

TREIMM-1157; No. of Pages 18

Feature review interstitial accumulation of sodium chloride in the skin [112]. In vitro, the TLR-dependent production of proinflammatory cytokines and NOS2-derived NO was enhanced under hypertonic conditions [113,114], which offers an explanation for the local or systemic hyperosmolarity seen in particular inflammatory disease states [115], including bacterial skin infections [114]. A highsalt diet of mice was associated with an increased expression of NOS2 mRNA and protein, and improved control of L. major parasites, which was dependent on the central osmosensor, the transcription factor tonicityresponsive enhancer-binding protein (TonEBP, also known as NFAT5) [114]. Notably, NFAT5 was also a major inducer of proinflammatory cytokines and NOS2 in macrophages in the absence of hyperosmotic stress [116]. Thus, similarly to HIF-1a NFAT5 not only senses the micromilieu, but also participates in TLR-dependent NOS2 regulation. NO itself or its derivatives (e.g., S-nitrosothiols) also contribute to the cellular micromilieu and the regulation of NOS2 expression. The feedback activation or inhibition by NO involves transcriptional (e.g., S-nitrolysation or nitration of transcription factors and signaling molecules), translational (e.g., inhibition of NOS2 dimerization), or post-translational effects (e.g., S-nitrosylation of actin and disruption of its interaction with NOS2 protein) [14,19,23,117,118]. H2S, which is synthesized in mammalian tissues from homocysteine or cysteine in a pathway involving four different enzymes (cystathionine b-synthase, cystathionine g-lyase, 3-mercaptopyruvate sulfurtransferase, and cysteine aminotransferase) [68], is another microenvironmental factor with the potential to regulate NOS. Although it is currently unknown whether (and to what extent) H2S affects the expression or activity of NO synthases under physiological or inflammatory conditions, in vitro studies reported both stimulatory and inhibitory effects of H2S on NOS1, NOS2, or NOS3 [119]. In mouse endothelial cells H2S/HS caused S-sulfhydration of NOS3 at Cys 443 (forming a hydrodisulfide residue, R-S-SH), which prevented NOS3 S-nitrosylation and promoted NOS3 phosphorylation, dimerization, and activity [120]. In a mouse macrophage cell line exogenous or endogenous H2S/HS downregulated LPS-mediated NF-kB activation, NOS2 protein expression, and NO production via induction of heme oxygenase 1 (HO-1) [121,122]. HO-1, which converts heme into biliverdin, carbon monoxide (CO), and ferrous iron (Fe2+), is induced in response to reactive oxygen species (ROS) and RNS that inactivate the repressor BACH and at the same time release the transcription factor Nrf2 (also known as NFE2L2) from its negative regulator KEAP1 in a S-nitrosylation-dependent process [123,124]. Expression of HO-1 and exposure of mouse or human cells to Fe2+ or CO have been described to suppress cytokine- or LPS-induced NOS2 expression [125–127]. However, depending on the cell type, organ, and model system studied, upregulation of NOS2 by CO [125] and oxidative preprogramming of the inflammatory response by HO-1 have also been observed [123,128,129]. Thus, the HO-1 pathway should not be viewed as a general antagonist of NOS2.

Trends in Immunology xxx xxxx, Vol. xxx, No. x

Protein–NOS2 interactions and substrate availability In addition to transcriptional, post-transcriptional (mRNA stability), and translational levels of control (reviewed in [15,19,69,82]), the activity of NOS2 in many cell types is strongly influenced by post-translational events, notably by its interaction with other proteins and by the availability of the substrate L-arginine. Several proteins have been identified that either enhance or block the activity of NOS2 by acting as adaptors, scaffolds [e.g., a-actinin-4, ezrin/ radizin/moesin-binding phosphoprotein 50 (EBP50), kinase suppressor of Ras-1 (Ksr1)], allosteric activators (e.g., Hsp90, Rac2) or dimerization inhibitors (e.g., kalirin), or that direct NOS2 towards proteasomal degradation (e.g., Rpn13/ARDM1/NAP110, UCH37) [15,25,26,130– 134]. The availability and import of extracellular L-arginine is essential for NOS2 (and NOS3) activity, despite a sufficient intracellular arginine pool (the so-called arginine paradox). Three enzymatic pathways are known to affect the L-arginine supply for NOS2 in the immune system. First, specific metallo-carboxypeptidases (i.e., the membrane-bound CPD and CPM) and the secreted endoplasmic reticulum (ER)-associated aminopeptidase 1 (ERAP1) generate L-arginine from peptides with C- or N-terminal arginine residues, and thereby support the NOS2-dependent NO production by macrophages or microvascular endothelial cells [135,136] (Figure 1A). Second, cytosolic arginase (Arg) 1, a characteristic marker of macrophages or DCs stimulated with IL-4, IL-13, or TGF-b, but which is also found in NOS2-positive macrophages, cleaves L-arginine into urea and ornithine (the latter being a precursor of and collagen synthesis) [15,137,138] polyamine (Figure 1A). Arg1 depletes macrophages of L-arginine, and thereby not only impedes their generation of NO and antimicrobial activity [139], but also their expression of NOS2 protein [140]. Arg1-mediated arginine depletion also helps to restrain T cell proliferation and T cell-mediated immunopathology, especially in the absence of NOS2 activity [141]. At least in endothelial cells, the intracellular localization of Arg1 (or Arg2) relative to NOS3 appeared to be irrelevant for the degree of L-arginine depletion and inhibition of NO production [142]. Third, whereas macrophages are unable to convert ornithine into L-arginine, they can regenerate L-arginine from reimported L-citrulline via the activity of argininosuccinate synthase 1 (Ass1) and argininosuccinate lyase (Asl) (Figure 1A). The relevance of this citrulline recycling pathway under argininerestricted conditions was recently demonstrated in macrophages and mice deficient for Ass1, which presented a reduced capacity for sustained NO production and for controlling M. bovis BCG or M. tuberculosis in vitro and in vivo [143]. There is also evidence that citrulline supplementation helps to restore arginine availability and NO production in endotoxemic mice [144]. By contrast, the enhanced pathogen replication observed in herpes simplex virus-1 infected human fibroblasts after RNAi of Ass1 was not due to the lack of L-arginine and NO, but resulted from the accumulation of aspartate and an improved synthesis of pyrimidine nucleotides [145] (Figure 1A). In addition to NOS2 and the arginases, there are two other L-arginine degrading pathways in mammalian 5

TREIMM-1157; No. of Pages 18

Feature review

Trends in Immunology xxx xxxx, Vol. xxx, No. x

organisms. These are initiated by the mitochondrial enzymes L-arginine decarboxylase (ADC; leading to the formation of agmatine and consecutively to polyamines) and L-arginine:glycine amidinotransferase (AGAT; giving rise to guanidinoacetate and subsequently to creatine) [146] (Figure 1A). However, with the exception of an inhibitory effect of agmatine on macrophage NOS2 activity, and an expression analysis of ADC and AGAT during the interaction of the intestinal protozoan parasite Giardia lamblia with human epithelial cells [147–149], the role of these pathways in the immune system is largely unexplored. Regulation at the post-activity level The denitrosylation of S-nitrosothiols (see above) can reverse the (signaling) effects of NOS2-derived NO, and therefore must be viewed as an additional level of NOS2 regulation. There is now intriguing evidence that the nitration of tyrosine residues, another biochemical footprint of NOS2 activity, can also be reversed in a controlled enzymatic process. A denitrase activity capable of specifically denitrating tyrosine-nitrated cyclooxygenase (Cox)-1 was discovered in mouse macrophages, human endothelial cells, and various mouse and rat tissues (e.g., spleen, liver, lung, brain). Upon partial purification, glutathione-S-transferase and copper/zinc-superoxide dismutase were identified as two enzymes with denitrase activity ([150] and references therein). Understanding the regulation of denitrosylation and denitration will be of great importance to judge the relevance of these processes during immune responses.

Taken together, the induction of NOS2 transcription and mRNA stability by interferons and microbial ligands remains the hallmark of NOS2 expression. However, microenvironmental factors (hypoxia, salt, substrate availability), other gasotransmitters (e.g., H2S, CO), new cytokine regulators of NOS2 (e.g., IL-33, EPO), and additional levels of control (post-translational, post-activity) have emerged as regulators of NO production. Antimicrobial and antiviral activity of NOS2 The antimicrobial and antiviral activity of NOS2 has been mostly studied with macrophages [14,15,19,151,152], but other myeloid (e.g., dendritic cells, neutrophils, and eosinophils) as well as non-myeloid cells (e.g., hepatocytes) can also exert NOS2-dependent effector functions as shown by in vitro killing assays, in vivo expression analyses, application of NOS2 inhibitors, the use of NOS2-deficient mice, or cell transfer studies [14,25,153–157]. It is important to emphasize that the expression of NOS2 and certain cytokines by myeloid cells, which led to eponymous designations (e.g., TNF/iNOS-producing dendritic cells – TIPDCs), defines a reversible activation state with the potential to kill infectious pathogens rather than a new subset of myeloid cells. Direct effects against infectious pathogens NO and other RNS can react with structural elements, components of the replication machinery, nucleic acids, metabolic enzymes, or with virulence-associated molecules

Indirect anmicrobial effects of NO, e.g., Host cell apoptosis Autophagosomal degradaon of bacteria Fpn1-mediated iron deprivaon Phagolysosomal fusion Inhibion of the expression of bacterial effectors/toxins/adhesins Dispersion of bacterial biofilms Immunoregulatory effects

Macrophage

Nrf2 Fe2+ Phagolysosome

NO

NOS2

L-arg

Direct anmicrobial effects of NO, e.g., Inhibion of pathogen proliferaon DNA mutagenesis Disrupon of [FeS] clusters Metabolic blockade (e.g., Krebs cycle) Inacvaon of virulence factors

Fpn1

Fe2+

Key: Indirect anmicrobial effects of NO

Direct anmicrobial effects of NO

Intracellular pathogen (e.g., S. enterica Typhimurium) TRENDS in Immunology

Figure 2. Examples for direct and indirect antimicrobial activities of NO. NOS2-derived NO exerts numerous direct and indirect antimicrobial effects on intracellular bacterial and protozoan pathogens. Amongst these, NO activates the transcription factor Nrf2, which leads to upregulation of the iron exporter ferroportin-1 (Fpn1) and subsequent iron deprivation of intracellular bacteria such as S. enterica Typhimurium. For further details see text.

6

TREIMM-1157; No. of Pages 18

Feature review of infectious pathogens, which is the basis for their direct antiviral or antimicrobial effects [19,152] (Figure 2). Although NO is likely to attack multiple structures of an infectious pathogen at the same time, the disruption of a single microbial target structure (e.g., [Fe–S] clusters) can be sufficient for a strong antimicrobial effect [158]. New microbial targets of RNS continue to be discovered. In S. enterica serovar Typhimurium, NO inactivated the dihydrolipoyl (or lipoamide) dehydrogenase component (LpdA) of the a-ketoglutarate dehydrogenase (a-KDH), an enzymatic complex of the tricarboxylic acid cycle (Krebs cycle). This led to reduced synthesis of methionine and lysine, whose synthesis is dependent on succinyl-CoA, the product of the a-KDH reaction [159]. In Clostridium difficile infections, NO derived from NOS2 (or NOS3) S-nitrosylated cysteine residues of the two major exotoxins of the bacterium (TcdA and TcdB), which blocked their cysteine protease activity and self-cleavage by catalytic site-inhibition and by displacement of the allosteric activator inositol hexakisphosphate [160]. Control of pathogens by NOS2-derived NO does not necessarily imply their killing and elimination. In fact, NO-mediated reduction of microbial metabolic activity caused sufficient impairment of pathogen proliferation in vivo and allowed clinical resolution of the disease by the immune response, as was shown for the protozoan parasite L. major [161]. Although an enhanced translocation of NOS2 towards a pathogen-containing vacuole has been associated with improved infection control in vitro [19,162], localization of NOS2 in the immediate vicinity to an intracellular microbe is not a conditio sine qua non for executing its antimicrobial activity. Transcellular diffusion of NO from uninfected NOS2-expressing macrophages to Leishmania-infected NOS2-negative host cells was also effective in parasite killing in vitro, and might even represent the prevailing scenario in vivo during acute and latent infections [163,164]. The detection of NOS2 in granulomas of humans and macaques infected with M. tuberculosis is also indicative for an antimicrobial role of NO during acute and latent tuberculosis [79]. Indirect effects against infectious pathogens Whereas the direct toxicity of RNS, especially against extracellular infectious pathogens, is well documented, the activity of NOS2 against intracellular microbes appears to be far more complex and can also entail indirect effects (Figure 2). In mouse bone marrow-derived macrophages, NO generated by NOS2 was found to elicit host cell apoptosis and thereby restricted the growth of intracellular M. tuberculosis [165], whereas in L. monocytogenesinfected macrophages NOS2-derived NO caused host cell necrosis without affecting bacterial survival [166]. NOS2derived NO also led to the formation of 8-nitro-cGMP, which caused S-guanylation and subsequent autophagosomal degradation of cytosolic bacteria such as group A streptococci [167]. Autophagy, in turn, might support NOS2 expression, as was seen in Pseudomonas aeruginosa-infected macrophages [168]. In S. typhimurium infection NOS2 expressed by macrophages was required for the Nrf2-dependent transcriptional induction and upregulation of ferroportin-1 (Fpn1), an iron exporter expressed in

Trends in Immunology xxx xxxx, Vol. xxx, No. x

the phagosomal and the cell membrane. The resulting iron deprivation partially accounted for the NOS2-mediated control of intracellular Salmonella by macrophages in vitro and in vivo [169]. Other indirect antimicrobial activities of NOS2-derived NO include (i) the stimulation of phagosomal maturation and phagolysosomal fusion [170]; (ii) the inhibition of the expression of bacterial secretion systems, effector molecules, toxins, or adhesins [171,172]; (iii) the dispersion of bacterial biofilms [173]; and (iv) various immunomodulatory effects on different types of immune cells that help to convey protection against infectious pathogens (reviewed in [14,15,17,19]). In Anopheles mosquitoes, tyrosine nitration of Plasmodium falciparum parasites, which resulted from the cooperation of the invertebrate NOS2 with heme peroxidase 2 and NADPH oxidase 5, was reported to facilitate the recognition of the pathogen by the complement-like innate defense system in the hemolymph of the vector ([174] and references therein). Future research will need to tackle the question how these multiple direct and indirect antimicrobial activities of NOS2 interact and cooperate with NOS2-independent effector mechanisms (e.g., NADPH oxidase, myeloperoxidase) in vivo to achieve pathogen control in various cell types and organs during the different phases of infection. Single cell and in situ kinetic analyses have provided the first insights [175]. NO and microbial escape of host defenses Microorganisms have developed various constitutive or inducible mechanisms to resist oxidative and nitrosative stress, and also to evade killing by activated phagocytes. Classically, these include (i) the production of scavenger molecules (e.g., thiols), (ii) detoxifying enzymes (e.g., mycobacterial peroxiredoxins; truncated hemoglobin HbN of M. tuberculosis, which converts host-derived NO into nitrite; flavohemoglobin Hmp of e.g., Salmonella or Escherichia coli, which oxidizes – ‘denitrosylates’ – host-derived NO to nitrate; nitric oxide reductase NorB of Neisseria meningitidis, which deprives the host cells of S-nitrosothiols; assimilatory reductases of Salmonella which reduce nitrite to nitrogen or ammonium), (iii) mechanisms of repair (e.g., degradation of nitrosylated proteins by mycobacterial proteasomes), and (iv) strategies of avoidance (e.g., inhibition of phagosomal recruitment of NOS2 by M. tuberculosis or Salmonella effector proteins) (reviewed in [19,152]; [133,176]). Several recent discoveries illustrate that this list is far from being complete. Host- or pathogen-derived RNS support pathogen survival and contribute to disease pathogenesis The observation that NOS2-derived NO itself might facilitate the survival of intracellular pathogens has added an unexpected facet to the scenario of microbial escape of the antimicrobial machinery of phagocytes. In infections with the food-borne pathogen Listeria (L.) monocytogenes, Cole et al. showed that host cell derived NO promoted the listeriolysin-dependent escape of the bacterium from the phagosomal vacuole into the cytosol by inhibiting the proton-pumping activity of V-ATPase and delaying phagolysosomal fusion [177,178]. An even more intricate case of 7

TREIMM-1157; No. of Pages 18

Feature review NO function was observed with bacterial pathogens that express their own bacterial NO synthase (bNOS). In Bacillus (B.) anthracis, bNOS-derived NO activated the bacterial expression of catalase and antagonized the generation of hydroxyl radicals in the Fenton reaction [179]. bNOS of S. aureus conferred resistance to ROS, cathelicidin, and particular antibiotics, and caused enhanced virulence in vivo, presumably also due to the induction of antioxidative mechanisms (e.g., superoxide dismutase – SOD) [180]. These results need to be reconciled with previous studies in which NO generated by NO donors or host cells strongly inhibited catalase activity and exerted synergistic antibacterial effects together with ROS [19,152,181,182]. Presumably, the amount of NO generated by bNOS is sufficient to elicit signaling events, but too low to inactivate catalase or to damage the bacteria. Bacteria such as S. enterica Typhimurium, which are devoid of bNOS, express ROS- and/or RNS-sensitive transcription factors (e.g., OxyR, SoxR) or transcriptional repressors (e.g., NsrR, Fur), which become activated or deactivated, respectively, during oxidative or nitrosative stress, and which regulate the expression of detoxification mechanisms (e.g., Hmp, NorB) [152,183,184]. S. aureus lacks the NO-sensor NsrR, but instead uses the SrrAB twocomponent system to control the expression its NO resistance genes [185]. A special form of evasion and adaptation to NO is seen in bacteria that are capable of utilizing nitrate for respiration, especially in hypoxic tissues such as the gut or intracellularly in granulomata when oxygen is limiting. In mouse models of colitis, NOS2-derived NO3 promoted the outgrowth of facultative anaerobic S. enterica Typhimurium or Escherichia (E.) coli (which express nitrate, nitrite, and nitric oxide reductases) in the inflamed colon [186–188]. This raises the possibility that the expression of NOS2 during intestinal inflammations catalyzes the dysbiosis frequently seen in colitis. Importantly, anaerobic respiration of E. coli on nitrate was accompanied by extensive S-nitrosylation of bacterial proteins – including the transcription factor OxyR, which switched on a specific set of genes that protected E. coli from endogenous nitrosative stress [189]. Similarly, NO produced by human macrophages supported the intracellular growth of M. tuberculosis, a phenotype that required bacterial nitrate reductase [190]. The mycobacterial respiration of nitrate and the generation of nitrite protected the bacteria against various forms of stress typically encountered in infected tissues (acidic pH, NO, and hypoxia) [191,192]. The pro-survival effect of nitrite resulted from bacterial growth retardation, reduced consumption of ATP, and the expression of resistance-mediating genes [193]. Another respiratory tract pathogen, Moraxella (M.) catarrhalis, generated NO by reducing nitrite. The nitrite-derived NO not only activated bronchial epithelial cells for the production of proinflammatory cytokines (TNF and IL-1a), but also inhibited host cell division and caused host cell apoptosis by caspasedependent and -independent mechanisms [194]. Depletion of host cell arginine The consumption of arginine by extra- and intracellular microorganisms can deprive host cell NOS2 of its substrate. 8

Trends in Immunology xxx xxxx, Vol. xxx, No. x

Metabolic pathways that were found to impair the NOS2dependent production of NO by myeloid cells include the arginase activity (e.g., Helicobacter pylori; reviewed in [195]) and the arginine deiminase activity of bacteria (e.g., Streptococcus pyogenes, S. enterica Typhimurium) or protozoa (e.g., Giardia intestinalis) [147,196,197] (Figure 1B). Mutant bacteria lacking the arginine deiminase gene showed attenuated virulence in wild type, but not in NOS2-deficient mice [196]. Depletion of host cell arginine can presumably also result from upregulation of microbial arginine transporters and subsequent sequestration of arginine within the pathogen [198]. Evasion mechanisms that enable microbes to at least partially or temporarily circumvent the antimicrobial effects of NOS2 reflect the coevolution of host organisms and infectious pathogens. It is important to bear in mind that pathogen killing and escape is frequently not a blackand-white situation. Evasion of host defenses can lead to pathogen persistence while still allowing the control and clinical resolution of the infection [199]. As outlined above, NOS2 activity is one of the players in this scenario that can serve both the host (pathogen killing, immunoregulation) and the pathogen (growth on NOS2 metabolites; transcriptional induction of antioxidant defenses by NO). NOS and cell differentiation, function, and survival NOS and myeloid cells The expression of NOS2 by myeloid cells is associated with two major functional consequences: the acquisition and/or alteration of cell-intrinsic capabilities and phenotypes, and regulatory effects on neighboring immune cells. The first category is reflected not only by the NOS2-mediated antimicrobial activity of macrophages and other myeloid cells (as discussed above), but also by their cell-autonomous modulation of phagocytosis, expression of MHC class II and costimulatory molecules, antigen presentation, cytokine production, survival, and apoptosis of myeloid cells [15,200]. The second category is best exemplified by the NOS2-dependent inhibition of T cell responses (see below). Myeloid cells specialized in this latter function have been described as ‘suppressor macrophages’, ‘myeloid-derived suppressor cells’ (MDSC), ‘regulatory macrophages’, or ‘immunosuppressive monocytes’. In recent years, several new facets of NO function in myeloid cells have emerged. First, NO was shown to drive the differentiation of myeloid cells. In experimental myocarditis, macrophage colony-stimulating factor (M-CSF) converted CD133+ progenitor cells into CD133+F4/80+ macrophages and prevented the development of myofibroblasts, provided that the mice and macrophages expressed NOS2 [201]. In human DCs, inflammatory cytokines induced NOS1 (but not NOS2) which was required for DC maturation [38]. Second, the cNOS isoforms (NOS1 and NOS3) were found to contribute to the inflammatory functions of macrophages. In mouse bone marrow-derived macrophages, LPS-induced NOS3 supported the expression of NOS2 in vitro and in vivo [35]; engagement of Fcgreceptors triggered NOS1 (and NOS3) activity and lowoutput NO production, which facilitated autocrine and paracrine phagocytosis [44]. Third, tissue regeneration is dependent on NOS2-expression by myeloid cells even after

TREIMM-1157; No. of Pages 18

Feature review sterile injury. In a muscle-injury mouse model, NO released by infiltrating NOS2-positive macrophages was necessary for the proliferation and differentiation of myogenic precursor cells [202]. Finally, NO, derived from any of the three NOS isoforms, can either upregulate or inhibit autophagy in distinct contexts, thereby supporting antimicrobial defense (see above) or contributing to the development of neurodegenerative or malignant diseases [167,203]. NOS and T cells The response of T lymphocytes to NO generated by NOS2positive myeloid cells was one of its first discovered immunomodulatory activities. Exogenous NO (i.e., NO not generated by T cells themselves) inhibited the proliferation or even caused the death of T cells (reviewed in [15,17]). Mice lacking an important antioxidant mechanism (i.e., GSNOR) exhibited a significant deficiency of T and B cells in the periphery as a result of excessive S-nitrosylation and lymphocyte apoptosis [204]. On the other hand, smaller quantities of NO supported the survival and differentiation of T cell subpopulations, notably Th1 cells and a FoxP3-negative regulatory T cell population termed NOTreg [15,17], which potently inhibited Th17 cell differentiation [205]. Furthermore, recent studies have shown that exogenous NO also regulates Th9 and Th17 cells. Under in vitro Th9 stimulatory conditions (i.e., antiCD3 plus anti-CD28 plus IL-4 and TGF-b), the presence of a NO donor (50 mM) promoted the generation of mouse and human Th9 cells. NOS2-deficient mice showed a reduced frequency of Th9 cells during allergic airway inflammation using the ovalbumin asthma model. The mechanism underlying the induction of IL-9 by NO was identified by Niedbala et al. as enhanced expression of p53 (presumably via S-nitrosylation and inactivation of its negative regulator MDM2) followed by increased IL-2 production, STAT5phosphorylation, and IRF4 activity, and also by upregulation of IL-4Ra and TGF-bR2. Evidence for Th9 cell-intrinsic expression and function of NOS2 was not obtained [206]. The expression of NOS isoforms by T cells has been a matter of debate. The important methodological issues of cell purity, cell origin (primary T cells vs cell lines), and NOS detection method (mRNA vs protein vs enzyme activity; analysis at a population vs single cell level), as well as the results of previous studies on NOS1, NOS2, or NOS3 in T cells, have been discussed in detail elsewhere [17]. Two groups recently provided convincing evidence for the expression of NOS2 in mouse or human Th17 cells, but reached different conclusions as to its functional role. Yang et al. analyzed mouse CD4+ T cells at the single cell level and found that stimulation with plate-bound anti-CD3 plus anti-CD28 under neutral (Th0) or Th17-polarizing conditions (IL-6 plus TGF-b) induced NOS2 protein. In the absence of NOS2, the differentiation of Th0 cells into Th17 cells was strikingly enhanced, whereas the development of Th1 and Th2 cells was unimpaired; NOS1- or NOS3-deficiency had no detectable effect [207]. Exogenous NO (10–200 mM) inhibited Th17 differentiation, confirming earlier data by others [208]. The negative regulatory effect of T cell NOS2 resulted from the nitration of defined

Trends in Immunology xxx xxxx, Vol. xxx, No. x

tyrosine residues within the transcription factor RORgt, and this strongly impaired its ability to activate the IL-17 promoter [207]. The in vivo relevance of the NOS2-dependent downregulation of Th17 responses was demonstrated in T cell transfer colitis and in experimental allergic encephalomyelitis (elicited by immunization with myelin oligodendrocyte glycoprotein), where the absence of NOS2 led to more severe disease driven by higher rates of Th17 cells [207]. Obermajer et al. observed that low doses of exogenous NO (10–25 mM), delivered by NO donors or myeloid-derived suppressor cells (MDSC), strongly supported the development of RORgt+IL-23R+IL-17+ Th17 cells from naive human CD4+ T cells stimulated with anti-CD3/ CD28 beads in the presence of a Th17-inducing cytokine cocktail (IL-1b, IL-6, IL-23, TGF-b). More importantly, human Th17 cells, especially memory Th cells, expressed NOS2 mRNA and protein, and this was confirmed at the single cell level. Endogenous NOS2 and NO-induced cGMP signaling were required for maintaining the Th17 cell phenotype (IL-17- and IL23R-expression) [209]. To date, there is no molecular explanation for the differential effects of NOS2-derived NO in mouse versus human Th17 cells. The effects of exogenous or endogenous NO on T cell differentiation are illustrated in Figure 3. NOS, B cells, and plasma cells Mouse or human peritoneal B1 cells, IgD+ B cells, and B220lowCD138+ plasma cells were found to express NOS2 or NOS3 [14,15,210,211]. However, to date, we do not have a coherent picture of the function of NOS2 in B cells. The serum of influenza A virus-infected NOS2-deficient mice contained higher levels of virus-specific IgG2a compared to wild type controls [212]. These findings contrasted with another report where naive Nos2/ mice exhibited a striking lack of mucosal IgA and IgG2b in the serum. In naive IgD+ B cells from the spleen, mesenteric lymph node, or intestinal lamina propria of NOS2-deficient mice, both T cell dependent (simulated by anti-CD40 plus TGF-b) and T cell independent (mediated by the B cell stimulatory cytokines BAFF and APRIL) IgA class-switch recombination and IgA production were severely impaired. Mechanistic analyses revealed that the expression of TGF-bRII on B cells and TLR-induced production of APRIL and BAFF by dendritic cells strictly required NOS2-derived NO [213]. In a recent study, the antibody responses towards a type 2 Tindependent (TI-2) antigen (NP-Ficoll) were enhanced in NOS2-deficient mice, which correlated with a raised production of BAFF by monocytes and DCs in the absence of NOS2 [214]. Thus, NOS2-derived NO, either expressed by B cells themselves or delivered by myeloid cells, appeared to be essential for homeostatic mucosal IgA production, but was inhibitory to antibody responses driven by viral pathogens or type 2 TI-2 B cell antigens. In a study by Saini et al., the yield of antibody-secreting B220lowCD138+ plasma cells [following polyclonal B cell stimulation with LPS or after immunization of mice with a TI-2 antigen (NP-Ficoll) or a T cell dependent antigen (ovalbumin)] was strikingly lower in the absence of NOS2. Unlike to their wild type counterparts, Nos2/ plasma cells failed to show prolonged survival or increased 9

TREIMM-1157; No. of Pages 18

Feature review

Trends in Immunology xxx xxxx, Vol. xxx, No. x

Mouse (A)

Exogenous NO

NO donor

NO donor

Th1↑ (B)

Human

Endogenous NO

Th0

–/–

Nos2

NO

NO

Th0

Th0

Th9↑

Treg↑ Th17↓

IL-17↓

Th0

Th9↑

Th17↑

IL- 23R↑

RORγt ↓

Th1

++

++

Th2

++

++

NO

++

cGMP ↑ NO

Th17 NOS2

Th17

IL-17↑

+/+

Nos2

Th17 NOS2

(+) TRENDS in Immunology

Figure 3. Effect of exogenous or endogenous NOS2-derived NO on mouse and human T cell differentiation. (A) Under defined cytokine conditions (not shown), exogenous NO (derived from chemical compounds donating NO or from NOS2-expressing myeloid cells) facilitated the differentiation of mouse Th0 cells into Th1 cells [12], Th9 cells [206], or into a subset of regulatory T cells (termed NO-Tregs) [205], but inhibited the development of Th17 cells [207,208]. NO-Tregs also blocked Th17 cell differentiation [205]. In human T cells, exogenous NO supported the development of Th9 [206] and Th17 cells [209]. (B) Expression of endogenous NOS2 in mouse Th0 cells had no effect on Th1 or Th2 development, but hindered the differentiation of Th17 cells by inactivating the transcription factor RORgt [207]. In human (memory) Th17 cells expression of NOS2 and production of NO activated guanylate cyclase. The resulting cGMP helped to sustain the expression of IL-17 and of the IL-23 receptor which signify the Th17 phenotype [209].

antibody secretion in response to IL-6 or APRIL. In addition, Nos2/ plasma cells exhibited reduced activity of the NO/cGMP-dependent protein kinase G, an impaired ER stress response, and enhanced activation of initiator and executioner caspases [211]. Notably, NOS2 did not affect the activation (expression of CD69, MHC class II, and CD44) or proliferation of B cells in response to LPS (a TI-1 antigen) or anti-IgM stimulation [211]. These data strongly argue for a cell-intrinsic prosurvival effect of NOS2 specifically in plasma cells. NOS2 and mesenchymal cells Non-hematopoietic mesenchyme, such as connective, adipose, or muscle tissue, is derived from mesenchymal stem cells (MSCs), also termed multipotent mesenchymal stroma cells. MSCs prove to be important modulators of immune responses. In response to supernatants from activated T cells or recombinant proinflammatory cytokines (IFN-g together with TNF, IL-1a or IL-1b) mouse MSCs not only strongly expressed a plethora of chemokines, but also produced high amounts of NO derived from NOS2. Several of these chemokines (notably CXCL-9, -10, and -11 acting via the CXCR3 receptor) attracted T cells towards the MSCs, whereas the released NO potently suppressed T cell proliferation [215]. Human MSCs, by contrast, caused T cell suppression mainly by producing prostaglandin E2 and by upregulating indoleamine 2,3dioxygenase (IDO), which led to local tryptophan depletion 10

(reviewed in [216]). Similarly to MSCs, differentiated stromal cells in mouse lymphoid organs [i.e., fibroblastic reticular cells (FRCs) and lymphoid endothelial cells (LECs)] also inhibited T cell proliferation in a NOS2-dependent manner [217,218]. The findings summarized above support the concept that cell-intrinsic NO or low doses of exogenous NO (10– 50 mM) primarily regulate the differentiation and survival of lymphocytes, whereas high amounts of exogenous NO (>100 mM) delivered by neighboring myeloid or mesenchymal stromal cells exert an anti-proliferative effect. NOS-mediated induction versus resolution of inflammation Owing to its diverse sources, complex regulation, and multiple cellular and molecular targets and interaction partners, it is not surprising that NO has both stimulatory and suppressive properties in the immune system (see above and [15,17]). Which of these two categories will dominate the outcome of an immune response in a given in vivo situation, is hard to predict and difficult to analyze, especially because transgenic mice for cell type-specific and inducible NOS2 or NOS3 deletion are still unavailable. Nevertheless, available in vitro systems and mouse models have revealed fascinating new insights into the pro- and anti-inflammatory activities of NO. Although the available data are not yet consistent, NOS2 and NOS3 (expressed by endothelial and myeloid cells) are clearly

TREIMM-1157; No. of Pages 18

Feature review involved in the homeostatic and inflammatory regulation of the frequency and strength of lymphatic vessel contractions, which are essential for the transport of dendritic cells and antigen into the draining lymph node [40,219]. During infection with herpes simplex virus type 2, NOS3 was indispensable for the remodeling (caliper increase) of arterioles feeding the draining lymph node [45]. NOS3 was also required for the adhesion and transendothelial migration of neutrophils [43]. Several findings support the concept that NOS2, and to some extent also the other NOS isoforms, can help to limit immune responses and thereby contribute to the resolution of inflammation. These include (i) the downregulation of leukocyte recruitment by NOS2-dependent functional inactivation of chemokines via nitration of crucial tyrosine residues or suppression of chemokine production [202,220,221]; (ii) the NOS1-, NOS2-, or NOS3-dependent inhibition of leukocyte adhesion and transendothelial migration [15,23,29,39,222]; (iii) the blockade of T cell expansion in lymph nodes by NOS2-expressing FRCs and LECs [217,218]; (iv) the restriction of Th1 responses by NOS2positive Ly6C+CD86+PDCA1+ DCs, which express less T cell stimulatory cytokines compared to their NOS2-deficient counterparts [223], confirming earlier data obtained with macrophages [15]; (v) the nitration of STAT1 by NOS2, which switched off an IFN-g or IFN-a/b response [118]; (vi) the S-nitrosylation of NLRP3 by IFN-g-, IFN-a/ b-, or LPS-induced NOS2, which led to the suppression of caspase 1 activation and IL-1b production, and thereby protected tuberculous or endotoxemic mice from excessive immunopathology [224–226]; (vii) the shedding of the TNF receptor 1 by the protease TACE (TNF converting enzyme), which became activated via the NOS2/NO/cGMP pathway [257]; and (viii) the induction of apototic cell death in myeloid cells and lymphoid cells (e.g., effector memory T cells) [15,200,227]. NO in disease and treatment The striking antimicrobial and immunoregulatory effects of NO in vitro, and the expression of NOS2 and other NOS isoforms during infectious, autoimmune, chronic inflammatory, malignant, and degenerative diseases in humans and in animal models [14,15,42,152,228,229], have stimulated immunologists and clinicians to investigate the functional role of NOS2 in vivo (e.g., by using transgenic mouse models) and to test the therapeutic application of NOS inhibitors or NO donors [230]. Because comprehensive discussion of this topic is beyond the scope of this review, I will briefly highlight some results recently obtained in three disease models (sepsis, tumor, neurodegeneration). In endotoxemia the earlier use of NOS2 inhibitors for treatment has yielded disappointing results in preclinical models [231]. Retrospectively, this is not surprising given the potent anti-inflammatory effects of NOS2-derived NO in addition to its sepsis-promoting proinflammatory activities. A recent study by Fletcher et al. now documents an unexpected therapeutic potency of the transfer of ex vivo expanded FRCs into mice with endotoxemia or cecal-ligation-and-puncture sepsis, which was completely dependent on the NOS2 activity expressed by the FRCs [232]. In a similar sepsis model, NOS2-dependent upregulation of

Trends in Immunology xxx xxxx, Vol. xxx, No. x

Box 3. Overview of the principal effects of NOS2-derived NO in tumor biology and antitumor defensea Tumor-promoting effects  Increased growth and metastasis of tumors due to (i) enhanced blood flow (vasodilatation), (ii) improved vascular permeability, (iii) increased tumor vessel formation [expression of angiogenic growth factors such as vascular endothelial growth factor (VEGF) and basic fibroblast growth factor (bFGF)], (iv) activation of tumorigenic signaling cascades by S-nitrosylation [e.g., Wnt/bcatenin, epidermal growth factor (EGF) receptor, inhibitor of differentiation 4 (ID4)–JAGGED–NOTCH signaling, Ets-1 transcription factor), (v) suppression of anti-tumorigenic signaling cascades (e.g., cell division autoantigen-1 – CDA1), and (vi) invasion of host cell tissue [e.g., upregulation of metalloproteinases (MMPs) and, downregulation of tissue inhibitor of metalloproteinase-1 (TIMP-1)].  Inhibition of immune cell chemotaxis, adhesion, and infiltration (downregulation of VCAM-1, ICAM-1, or P-selectin; inactivation of T cell chemokines).  Direct inhibition of T cell proliferation, induction of T cell death.  Indirect blockade of T cell proliferation via increased infiltration of MDSC. Anti-tumor effects  Killing of tumor cells by induction of apoptosis.  Normalization of tumor vasculature (improved expression of adhesion molecules on tumor endothelial cells).  Suppression of angiogenic and tumor growth factors.  Upregulation of chemokines with subsequent improved recruitment of tumor-specific cytotoxic T cells.  Inhibition of metastasis by downregulation of matrix metalloproteases and upregulation of TIMP-1 and E-cadherin. a Further details can be found in [14,15,221,229,234–236,250–255].

cGMP and subsequent activation of TACE protected against organ injury [257]. The role of NOS2 in tumor immunology is controversial because both beneficial (i.e., tumoricidal and tumor immunity-inducing) and detrimental (i.e., immunosuppressive and tumor growth- and metastasis-promoting) effects have been described (Box 3). Accordingly, NOS2 expression in human tumors has been positively or negatively correlated with tumor progression [229]. In addition to the use of different in vitro and in vivo models, key factors for the discrepancies between different published results appear to be (i) the spatiotemporal and quantitative relationship between NOS2-positive tumor cells, NOS2-expressing infiltrating myeloid cells, and anti-tumor effector T cells [233], (ii) the balance between the actual effects exerted by tumor-expressed NOS2 or by myeloid cell-expressed NOS2, and (iii) the amount of NO generated and available within the tumor tissue. Although high amounts of NO (e.g., produced by tumor cells expressing NOS2 or by NOS2-positive tumor-infiltrating myeloid cells) have been associated with tumor cell death [234], smaller amounts of NO produced by hypoxic tumor tissue or by infiltrating myeloid cells were found to have regulatory effects on angiogenesis, metastasis, or on the adhesion and infiltration of immune cells with either suppressive (e.g., MDSC) or (in-)direct tumoricidal activities (e.g., NOS2+Arg1 macrophages, cytotoxic T cells) [229,235]. The expression of NOS2 by tumor cells is subject to regulation by cytokines, miRNAs, hypoxia, and the availability of arginine [99,100,236–238], which in addition to its role as a NOS2 substrate is thought to function as a tumor-promoting amino acid. In tumors that lack Ass1, arginine-degrading 11

TREIMM-1157; No. of Pages 18

Feature review

Trends in Immunology xxx xxxx, Vol. xxx, No. x

Table 1. Selection of recent studies demonstrating the effects of immune cell- or tumor cell-derived NOS2 on tumor cell survival and/or anti-tumor immune responsesa Tumor model

Source of NOS2

Induction of NOS2 protein by

Effect of NOS2-derived NO

Refs

Tumor-infiltrating macrophage

Low-dose local g- irradiation

[252]

Mouse renal adenocarcinoma (RENCA)

Tumor cell

Mouse renal adenocarcinoma (RENCA)

Tumor-infiltrating macrophage

IFN-g/LPS plus miR-146a inhibitor IL-2/anti-CD40 immunotherapy; IFN-g

Suppression of angiogenic VEGF and MDSCrecruiting GM-CSF Endothelial cell activation (VCAM-1") Induction of T cell-recruiting chemokines (RANTES) and improved infiltration of T cells Promotes macrophage-mediated tumor cell death in vitro and in vivo Inhibition of metastasis by downregulation of matrix metalloproteinases and upregulation of TIMP-1

Tumor cell

N.a.

Anti-tumorigenic effects of NO Spontaneous mouse pancreatic island carcinoma (RIP1-Tag5-mice); tumor control by low-dose g-irradiation and adoptive T cell transfer

Protumorigenic effects of NO Human glioma stem cells

NPM-ALK+ human T cell lymphoma

Tumor cell

Mouse melanoma MT-RET-1

Tumor cell

Mouse colon adenocarcinoma C26

Gr1+ myeloid cells (MDSC) Tumor cells and myeloid cells (MDSC)

Mouse colon adenocarcinoma C26; mouse thymoma EG7-OVA; mouse prostate carcinoma (TRAMP mouse); human prostate, colon, or nasopharyngeal carcinoma Mouse xenograft model with human breast cancer cells (e.g., MDA-MB-231)

Tumor cells (and myeloid cells?)

Lack of miR-26a expression N.a.

N.a. N.a.

IFN-g, hypoxia, serum withdrawal

Downregulation of cell cycle inhibitor cell division autoantigen-1 (CDA1) Upregulation of inhibitor of differentiation 4 (ID4) Tumor cell survival, adhesion, and migration Secretion of VEGF by tumor cells, which in turn causes recruitment, maturation, and accumulation of MDSC in the tumor and T cell suppression Inhibition of splenocyte IFN-a and IFN-g response due to nitration of STAT1 Nitration of tyrosine- and tryptophan-residues in mouse and human CCL2; nitrated CCL2 still attracts myeloid cells, but no longer effector T cells Upregulation of proinflammatory markers associated with poor tumor survival (e.g., S100A8, IL-6, IL-8) Increased tumor cell migration and brain metastasis

[100] [254]

[253] [256] [99] [251]

[118] [221]

[236]

a

GM-CSF, granulocyte-macrophage colony stimulating factor; MDSC, myeloid-derived suppressor cells; N.a., not analyzed; NPM-ALK, nucleophosmin-anaplastic lymphoma kinase; RANTES, regulated on activation, normal T cell expressed and secreted; TIMP-1, tissue inhibitor of metalloproteinase 1; VCAM-1, vascular cellular adhesion molecule-1; VEGF, vascular endothelial growth factor.

enzymes (e.g., pegylated forms of bacterial ADI or human arginase) are currently being evaluated as novel therapeutics [239,240]. Table 1 summarizes the results of some recent mouse and human studies, which illustrate the complexity of the role of NOS2 in tumor biology and anti-tumor immune defenses. Novel tumor treatment strategies are likely to arise from the molecular targets of NOS2-derived NO. In mouse models of Alzheimer’s dementia, NOS3 activity protected against loss of spatial learning and memory [241]. By contrast, the expression of NOS2 was associated with the deposition of extracellular amyloid b and cognitive dysfunction [242]. The disease-promoting effect of NOS2, which is upregulated by the NLRP3 inflammasome, appeared to result from the NO-dependent inhibition of amyloid b-degrading enzymes and from the accelerated aggregation and plaque formation of amyloid b following nitration of defined tyrosine residues ([243,244] and references therein). In humans, progressive Alzheimer’s disease was paralleled by loss of NOS1 and NOS3 expression and activity in various regions of the brain versus normal agematched controls, whereas conflicting results were obtained for NOS2 ([245] and references therein). These data argue for an isoform-dependent effect of NOS in 12

Alzheimer’s dementia, which might reflect the canonical cellular distribution and function of NOS1, NOS2, and NOS3. Concluding remarks In recent years our understanding of the role of NO in immunity has significantly changed. Originally conceived as a key effector mechanism of myeloid cells and the innate immune response, NOS2 must now also be considered as an integral component of the development, differentiation, and function of B and T lymphocytes, and also of non-hematopoietic cells. In addition to NOS2, NOS1 and NOS3 have to be taken into account as sources of immunomodulatory NO. Whereas in the past NO was selectively recognized as a product of host cells for defense against pathogens, novel research has revealed that NO generated by bacterial NOS or nitrite reductases clearly helps bacterial pathogens to resist the antimicrobial activities of myeloid cells or even modulates host cell immune responses. All these insights into the activities of NOS and NO provide the basis for novel therapeutic approaches of infectious, malignant, autoimmune, and chronic inflammatory or neurodegenerative diseases. On the other hand, the increasing knowledge of the multiple functions and targets of NO also raises our

TREIMM-1157; No. of Pages 18

Feature review

Box 4. Outstanding questions  What is the chemical basis for stimulatory versus inhibitory effects of H2S/HS on the expression of NOS isoforms?  How does CO, the product of the heme oxygenase-1, induce or inhibit NOS2 expression?  What is the phenotype of mice with a cell type (e.g., macrophage, dendritic cells, CD4+ T cell, B cell)-specific deletion of NOS2 in models of infection, autoimmunity, or tumor?  What are the inducers of human NOS2 in vivo? Are microenvironmental factors (e.g., hypoxia, tonicity, microbial ligands) key signals in humans?  Does NOS2 during infection of humans fulfill an antimicrobial (direct? indirect?) or rather a T cell regulatory function?  Does the tissue microenvironment determine whether NOmediated pathogen killing out-competes the NO-triggered antioxidant defense mechanisms?  Do molecular mechanisms other than impaired protein synthesis underlie the suppression of T cell proliferation following arginine depletion?  How can NO generated by NOS2 or NOS3 support or inhibit the endothelial adhesion and transmigration of immune cells?  Is the putatively therapeutic effect of arginine degradation in particular malignancies due to nutrient deprivation of the tumor cells or to impaired generation of pro-tumorigenic NO?  Can NO donors be designed that are only taken up by specific cell types and only release NO inside a defined compartment of the target cell?

awareness of the complexity of events that might ensue from the activation or inhibition of NOS isoforms. Outstanding questions are listed in Box 4. Clearly, without tissue- or cellspecific strategies for the deletion of NOS or for the delivery of NO it currently appears unlikely that systemically applied NOS-based therapies will readily become clinical practice. However, this skepticism continues to be dispelled by the truly encouraging results from preclinical mouse models as outlined above. Acknowledgements The preparation of this manuscript and some of the work reviewed was supported by the Deutsche Forschungsgemeinschaft (SFB 643, project grant A6; GRK 1660), the Interdisciplinary Center for Clinical Research (IZKF) of the Universita¨tsklinikum Erlangen (project grants A49 and A61), the Emerging Field Initiative of FAU Erlangen-Nu¨rnberg (project grant within the ‘Metal Redox Inorganic Chemistry’ consortium) and by the Dr Robert Pfleger Stiftung. I apologize to all authors, whose work could only be acknowledged by referring to previous review articles because of space limitations.

References 1 Stuehr, D.J. and Marletta, M.A. (1985) Mammalian nitrite biosynthesis: mouse macrophages produce nitrite and nitrate in response to Escherichia coli lipopolysaccharide. Proc. Natl. Acad. Sci. U.S.A. 82, 7738–7742 2 Stuehr, D.J. and Marletta, M.A. (1987) Induction of nitrite/nitrate synthesis in murine macrophages by BCG infection, lymphokines or interferon-g. J. Immunol. 139, 518–525 3 Hibbs, J.B. et al. (1987) L-arginine is required for expression of the activated macrophage effector mechanism causing selective metabolic inhibition in target cells. J. Immunol. 138, 550–565 4 Hibbs, J.B. et al. (1987) Macrophage cytotoxicity: role of L-arginine deiminase and imino nitrogen oxidation to nitrite. Science 235, 473– 476 5 Xie, Q-W. et al. (1992) Cloning and characterization of inducible nitric oxide synthase from mouse macrophages. Science 256, 225–228 6 Lowenstein, C.J. et al. (1992) Cloned and expressed macrophage nitric oxide synthase contrasts with the brain synthase. Proc. Natl. Acad. Sci. U.S.A. 89, 6711–6715

Trends in Immunology xxx xxxx, Vol. xxx, No. x

7 Connor, J.R. et al. (1995) Suppression of adjuvant-induced arthritis by selective inhibition of inducible nitric oxide synthase. Eur. J. Pharmacol. 273, 15–24 8 Stenger, S. et al. (1995) L-N6-(1-iminoethyl)lysine potently inhibits inducible nitric oxide synthase and is superior to NG-monomethylarginine in vitro and in vivo. Eur. J. Pharmacol. 294, 703–712 9 MacMicking, J.D. et al. (1995) Altered responses to bacterial infection and endotoxic shock in mice lacking inducible nitric oxide synthase. Cell 81, 641–650 10 Laubach, V.E. et al. (1995) Mice lacking inducible nitric oxide synthase are not resistant to lipopolysaccharide-induced death. Proc. Natl. Acad. Sci. U.S.A. 92, 10688–10692 11 Wei, X-Q. et al. (1995) Altered immune responses in mice lacking inducible nitric oxide synthase. Nature 375, 408–411 12 Niedbala, W. et al. (1999) Effects of nitric oxide on the induction and differentiation of Th1 cells. Eur. J. Immunol. 29, 2498–2505 13 Nathan, C. (1992) Nitric oxide as a secretory product of mammalian cells. FASEB J. 6, 3051–3064 14 Bogdan, C. (2000) The function of nitric oxide in the immune system. In Handbook of Experimental Pharmacology: Nitric Oxide (Mayer, B., ed.), pp. 443–492, Springer 15 Bogdan, C. (2001) Nitric oxide and the immune response. Nat. Immunol. 2, 907–916 16 Nathan, C. (2003) Specificity of a third kind: reactive oxygen and nitrogen intermediates in cell signaling. J. Clin. Invest. 111, 769–778 17 Bogdan, C. (2011) Regulation of lymphocytes by nitric oxide. Methods Mol. Biol. 677, 375–393 18 Wink, D.A. et al. (2011) Nitric oxide and redox mechanisms in the immune response. J. Leukoc. Biol. 89, 873–891 19 Bogdan, C. (2011) Reactive oxygen and reactive nitrogen intermediates in the immune system. In The Immune Response to Infection (Kaufmann, S.H.E. et al., eds), pp. 69–84, ASM Press 20 Vodovotz, Y. et al. (1995) Vesicle membrane association of nitric oxide synthase in primary mouuse macrophages. J. Immunol. 154, 2914–2925 21 Webb, J.L. et al. (2001) Macrophage nitric oxide synthase associates with cortical actin but is not recruited to phagosomes. Infect. Immun. 69, 6391–6400 22 Rahat, M.A. et al. (2011) Molecular mechanisms regulating macrophage response to hypoxia. Front. Immunol. 2, 45 23 Thom, S.R. et al. (2013) Nitric-oxide synthase-2 linkage to focal adhesion kinase in neutrophils influences enzyme activity and beta2 integrin function. J. Biol. Chem. 288, 4810–4818 24 Glynne, P.A. et al. (2002) Epithelial inducible nitric-oxide synthase is an apical EBP50-binding protein that directs vectorial nitric oxide output. J. Biol. Chem. 277, 33132–33138 25 Jyoti, A. et al. (2014) Interaction of inducible nitric oxide synthase with rac2 regulates reactive oxygen and nitrogen species generation in the human neutrophil phagosomes: implication in microbial killing. Antioxid. Redox Signal. 20, 417–431 26 Zhang, Y. et al. (2011) Kinase suppressor of Ras-1 protects against pulmonary Pseudomonas aeruginosa infections. Nat. Med. 17, 341–346 27 Pandit, L. et al. (2009) The physiologic aggresome mediates cellular inactivation of iNOS. Proc. Natl. Acad. Sci. U.S.A. 106, 1211–1215 28 Saini, R. et al. (2006) Nitric oxide synthase localization in the rat neutrophils: immunocytochemical, molecular, and biochemical studies. J. Leukoc. Biol. 79, 519–528 29 Chakrabarti, S. et al. (2012) Neuronal nitric oxide synthase regulates endothelial inflammation. J. Leukoc. Biol. 91, 947–956 30 Fo¨rstermann, U. et al. (1998) Expressional control of the ‘constitutive’ isoforms of nitric oxide synthase (NOSI and NOSIII). FASEB J. 12, 773–790 31 Iwase, K. et al. (2000) Induction of endothelial nitric oxide synthase in rat brain astrocytes by systemic lipopolysaccharide treatment. J. Biol. Chem. 275, 11929–11933 32 Boissel, J.P. et al. (2004) The neuronal nitric oxide synthase is upregulated in mouse skin repair and in response to epidermal growth factor in human HaCaT keratinocytes. J. Invest. Dermatol. 123, 132–139 33 Dudzinski, D.M. et al. (2006) The regulation and pharmacology of endothelial nitric oxide synthase. Annu. Rev. Pharmacol. Toxicol. 46, 235–276 13

TREIMM-1157; No. of Pages 18

Feature review 34 Wahl, S.M. et al. (2003) Nitric oxide in experimental joint inflammation. Benefit or detriment? Cells Tissues Organs 174, 26–33 35 Connelly, L. et al. (2005) Resistance to endotoxic shock in endothelial nitric-oxide synthase (eNOS) knock-out mice: a pro-inflammatory role for eNOS-derived no in vivo. J. Biol. Chem. 280, 10040–10046 36 Li, E. et al. (2006) Neuronal nitric oxide synthase is necessary for elimination of Giardia lamblia infections in mice. J. Immunol. 176, 516–521 37 Martinelli, R. et al. (2009) ICAM-1-mediated endothelial nitric oxide synthase activation via calcium and AMP-activated protein kinase is required for transendothelial lymphocyte migration. Mol. Biol. Cell 20, 995–1005 38 Adler, H.S. et al. (2010) Neuronal nitric oxide synthase modulates maturation of human dendritic cells. J. Immunol. 184, 6025–6034 39 Fritzsche, C. et al. (2010) Endothelial nitric oxide synthase limits the inflammatory response in mouse cutaneous leishmaniasis. Immunobiology 215, 826–832 40 Liao, S. et al. (2011) Impaired lymphatic contraction associated with immunosuppression. Proc. Natl. Acad. Sci. U.S.A. 108, 18784–18789 41 Duma, D. et al. (2011) NOS-1-derived NO is an essential triggering signal for the development of systemic inflammatory responses. Eur. J. Pharmacol. 668, 285–292 42 Forstermann, U. and Sessa, W.C. (2012) Nitric oxide synthases: regulation and function. Eur. Heart J. 33, 829–837 43 Liu, G. et al. (2012) ICAM-1-activated Src and eNOS signaling increase endothelial cell surface PECAM-1 adhesivity and neutrophil transmigration. Blood 120, 1942–1952 44 Huang, Z. et al. (2012) Stimulation of unprimed macrophages with immune complexes triggers a low output of nitric oxide by calciumdependent neuronal nitric-oxide synthase. J. Biol. Chem. 287, 4492–4502 45 Sellers, S.L. et al. (2013) Nitric oxide and TNFalpha are critical regulators of reversible lymph node vascular remodeling and adaptive immune response. PLoS ONE 8, e60741 46 Lundberg, J.O. et al. (2008) The nitrate–nitrite–nitric oxide pathway in physiology and therapeutics. Nat. Rev. Drug Discov. 7, 156–167 47 Doyle, M.P. et al. (1981) Kinetics and mechanism of the oxidation of human deoxyhemoglobin by nitrites. J. Biol. Chem. 256, 12393–12398 48 Li, H. et al. (2008) Nitric oxide production from nitrite occurs primarily in tissues not in the blood: critical role of xanthine oxidase and aldehyde oxidase. J. Biol. Chem. 283, 17855–17863 49 Miljkovic, J. et al. (2013) Generation of HNO and HSNO from nitrite by heme–iron-catalyzed metabolism with H2S. Angew. Chem. Int. Ed. Engl. 52, 12061–12064 50 Oplander, C. et al. (2013) Mechanism and biological relevance of bluelight (420–453 nm)-induced nonenzymatic nitric oxide generation from photolabile nitric oxide derivates in human skin in vitro and in vivo. Free Radic. Biol. Med. 65, 1363–1377 51 Oplander, C. and Suschek, C.V. (2012) The role of photolabile dermal nitric oxide derivates in ultraviolet radiation (UVR)-induced cell death. Int. J. Mol. Sci. 14, 191–204 52 Liu, D. et al. (2014) UVA irradiation of human skin vasodilates arterial vasculature and lowers blood pressure independently of nitric oxide synthase. J. Invest. Dermatol. 134, 1839–1846 53 Kraft, B. et al. (2011) Microbial nitrate respiration – genes, enzymes and environmental distribution. J. Biotechnol. 155, 104–117 54 Crane, B.R. et al. (2010) Bacterial nitric oxide synthases. Annu. Rev. Biochem. 79, 445–470 55 Anand, P. and Stamler, J.S. (2012) Enzymatic mechanisms regulating protein S-nitrosylation: implications in health and disease. J. Mol. Med. (Berl.) 90, 233–244 56 Radi, R. (2013) Peroxynitrite, a stealthy biological oxidant. J. Biol. Chem. 288, 26464–26472 57 Vanin, A.F. (2009) Dinitrosyl iron complexes with thiolate ligands: physico-chemistry, biochemistry and physiology. Nitric Oxide 21, 1–13 58 Kroncke, K.D. and Klotz, L.O. (2009) Zinc fingers as biologic redox switches? Antioxid. Redox Signal. 11, 1015–1027 59 Sawa, T. and Ohshima, H. (2006) Nitrative DNA damage in inflammation and its possible role in carcinogenesis. Nitric Oxide 14, 91–100 60 Jones, L.H. (2012) Chemistry and biology of biomolecule nitration. Chem. Biol. 19, 1086–1092 14

Trends in Immunology xxx xxxx, Vol. xxx, No. x

61 Bogdan, C. (2001) Nitric oxide and the regulation of gene expression. Trends Cell Biol. 11, 66–75 62 Gould, N. et al. (2013) Regulation of protein function and signaling by reversible cysteine S-nitrosylation. J. Biol. Chem. 288, 26473–26479 63 Hernansanz-Agustin, P. et al. (2013) Nitrosothiols in the immune system: signaling and protection. Antioxid. Redox Signal. 18, 288–308 64 Lee, Y.I. et al. (2014) Protein microarray characterization of the Snitrosoproteome. Mol. Cell. Proteomics 13, 63–72 65 Doulias, P.T. et al. (2013) Nitric oxide regulates mitochondrial fatty acid metabolism through reversible protein S-nitrosylation. Sci. Signal. 6, rs1 66 Filipovic, M.R. et al. (2012) Chemical characterization of the smallest S-nitrosothiol, HSNO; cellular cross-talk of H2S and S-nitrosothiols. J. Am. Chem. Soc. 134, 12016–12027 67 King, S.B. (2013) Potential biological chemistry of hydrogen sulfide (H2S) with the nitrogen oxides. Free Radic. Biol. Med. 55, 1–7 68 Wang, R. (2012) Physiological implications of hydrogen sulfide: a whiff exploration that blossomed. Physiol. Rev. 92, 791–896 69 Kleinert, H. et al. (2004) Regulation of the expression of inducible nitric oxide synthase. Eur. J. Pharmacol. 500, 255–266 70 Gao, J.J. et al. (1998) Autocrine/paracrine IFN-a/b mediates the lipopolysaccharide-induced activation of transcription factor Stat1a in mouse macrophages: pivotal role of Stat1a in induction of the inducible nitric oxide synthase gene. J. Immunol. 161, 4803–4810 71 Toshchakov, V. et al. (2002) TLR4, but not TLR2, mediates IFN-binduced STAT1ab-dependent gene expression in macrophages. Nat. Immunol. 3, 392–398 72 Farlik, M. et al. (2010) Nonconventional initiation complex assembly by STAT and NF-kappaB transcription factors regulates nitric oxide synthase expression. Immunity 33, 25–34 73 Wienerroither, S. et al. (2014) Regulation of NO synthesis, local inflammation, and innate immunity to pathogens by BET family proteins. Mol. Cell. Biol. 34, 415–427 74 Bogdan, C. (2004) Reactive oxygen and reactive nitrogen metabolites as effector molecules against infectious pathogens. In The Innate Immune Response to Infection (Kaufmann, S.H.E. et al., eds), pp. 357–396, ASM Press 75 Fang, F.C. and Nathan, C.F. (2007) Man is not a mouse: reply. J. Leukoc. Biol. 81, 580 76 Ralph, A.P. et al. (2008) L-arginine and vitamin D: novel adjunctive immunotherapies in tuberculosis. Trends Microbiol. 16, 336–344 77 Thomas, A.C. and Mattila, J.T. (2014) ‘Of mice and men’: arginine metabolism in macrophages. Front. Immunol. 5, 479 78 Ralph, A.P. et al. (2013) Impaired pulmonary nitric oxide bioavailability in pulmonary tuberculosis: association with disease severity and delayed mycobacterial clearance with treatment. J. Infect. Dis. 208, 616–626 79 Mattila, J.T. et al. (2013) Microenvironments in tuberculous granulomas are delineated by distinct populations of macrophage subsets and expression of nitric oxide synthase and arginase isoforms. J. Immunol. 191, 773–784 80 Wilsmann-Theis, D. et al. (2013) Generation and functional analysis of human TNF-alpha/iNOS-producing dendritic cells (Tip-DC). Allergy 68, 890–898 81 Zhang, X. et al. (1996) Transcriptional basis for hyporesponsiveness of the human inducible nitric oxide synthase gene to lipopolysaccharide/ interferon-g. J. Leukoc. Biol. 59, 575–585 82 Pautz, A. et al. (2010) Regulation of the expression of inducible nitric oxide synthase. Nitric Oxide 23, 75–93 83 Gross, T.J. et al. (2014) Epigenetic silencing of the human NOS2 gene: rethinking the role of nitric oxide in human macrophage inflammatory responses. J. Immunol. 192, 2326–2338 84 Saldarriaga, O.A. et al. (2012) Identification of hamster inducible nitric oxide synthase (iNOS) promoter sequences that influence basal and inducible iNOS expression. J. Leukoc. Biol. 92, 205–218 85 Lima-Junior, D.S. et al. (2013) Inflammasome-derived IL-1beta production induces nitric oxide-mediated resistance to Leishmania. Nat. Med. 19, 909–915 86 Guo, Z. et al. (2012) miRNA-939 regulates human inducible nitric oxide synthase posttranscriptional gene expression in human hepatocytes. Proc. Natl. Acad. Sci. U.S.A. 109, 5826–5831

TREIMM-1157; No. of Pages 18

Feature review 87 Matsui, K. et al. (2008) Natural antisense transcript stabilizes inducible nitric oxide synthase messenger RNA in rat hepatocytes. Hepatology 47, 686–697 88 Li, C. et al. (2014) Interleukin-33 increases antibacterial defense by activation of inducible nitric oxide synthase in skin. PLoS Pathog. 10, e1003918 89 Brines, M. and Cerami, A. (2012) The receptor that tames the innate immune response. Mol. Med. 18, 486–496 90 Nairz, M. et al. (2011) Erythropoietin contrastingly affects bacterial infection and experimental colitis by inhibiting nuclear factorkappaB-inducible immune pathways. Immunity 34, 61–74 91 Aoshiba, K. et al. (2009) Therapeutic effects of erythropoietin in murine models of endotoxin shock. Crit. Care Med. 37, 889–898 92 Banerjee, S. et al. (2013) miR-125a-5p regulates differential activation of macrophages and inflammation. J. Biol. Chem. 288, 35428–35436 93 Wang, X. et al. (2009) Inducible nitric-oxide synthase expression is regulated by mitogen-activated protein kinase phosphatase-1. J. Biol. Chem. 284, 27123–27134 94 Yang, K. et al. (2014) miR-155 suppresses bacterial clearance in Pseudomonas aeruginosa-induced keratitis by targeting Rheb. J. Infect. Dis. 210, 89–98 95 Huang, L. et al. (2013) Down-regulation of miR-301a suppresses proinflammatory cytokines in Toll-like receptor-triggered macrophages. Immunology 140, 314–322 96 Dai, R. et al. (2008) Suppression of LPS-induced Interferon-gamma and nitric oxide in splenic lymphocytes by select estrogen-regulated microRNAs: a novel mechanism of immune modulation. Blood 112, 4591–4597 97 Xu, C. et al. (2013) miR-155 regulates immune modulatory properties of mesenchymal stem cells by targeting TAK1-binding protein 2. J. Biol. Chem. 288, 11074–11079 98 Palmieri, D. et al. (2014) TNFalpha induces the expression of genes associated with endothelial dysfunction through p38MAPK-mediated down-regulation of miR-149. Biochem. Biophys. Res. Commun. 443, 246–251 99 Zhu, H. et al. (2013) NPM-ALK up-regulates iNOS expression through a STAT3/microRNA-26a-dependent mechanism. J. Pathol. 230, 82–94 100 Perske, C. et al. (2010) Loss of inducible nitric oxide synthase expression in the mouse renal cell carcinoma cell line RENCA is mediated by microRNA miR-146a. Am. J. Pathol. 177, 2046–2054 101 Khorkova, O. et al. (2014) Natural antisense transcripts. Hum. Mol. Genet. 23, R54–R63 102 Yoshigai, E. et al. (2013) Natural antisense transcript-targeted regulation of inducible nitric oxide synthase mRNA levels. Nitric Oxide 30, 9–16 103 Ho, J.J. et al. (2013) Active stabilization of human endothelial nitric oxide synthase mRNA by hnRNP E1 protects against antisense RNA and microRNAs. Mol. Cell. Biol. 33, 2029–2046 104 Korneev, S.A. et al. (2008) Novel noncoding antisense RNA transcribed from human anti-NOS2A locus is differentially regulated during neuronal differentiation of embryonic stem cells. RNA 14, 2030–2037 105 Olson, N. and van der Vliet, A. (2011) Interactions between nitric oxide and hypoxia-inducible factor signaling pathways in inflammatory disease. Nitric Oxide 25, 125–137 106 Campbell, E.L. et al. (2014) Transmigrating neutrophils shape the mucosal microenvironment through localized oxygen depletion to influence resolution of inflammation. Immunity 40, 66–77 107 Mahnke, A. et al. (2014) Hypoxia in Leishmania major skin lesions impairs the NO-dependent leishmanicidal activity of macrophages. J. Invest. Dermatol. 134, 2339–2346 108 Robinson, M.A. et al. (2011) Oxygen-dependent regulation of nitric oxide production by inducible nitric oxide synthase. Free Radic. Biol. Med. 51, 1952–1965 109 Jantsch, J. et al. (2011) Toll-like receptor activation and hypoxia use distinct signaling pathways to stabilize hypoxia-inducible factor 1alpha (HIF1A) and result in differential HIF1A-dependent gene expression. J. Leukoc. Biol. 90, 551–562 110 Elks, P.M. et al. (2013) Hypoxia inducible factor signaling modulates susceptibility to mycobacterial infection via a nitric oxide dependent mechanism. PLoS Pathog. 9, e1003789

Trends in Immunology xxx xxxx, Vol. xxx, No. x

111 Go, W.Y. et al. (2004) NFAT5/TonEBP mutant mice define osmotic stress as a critical feature of the lymphoid microenvironment. Proc. Natl. Acad. Sci. U.S.A. 101, 10673–10678 112 Wiig, H. et al. (2013) Immune cells control skin lymphatic electrolyte homeostasis and blood pressure. J. Clin. Invest. 123, 2803–2815 113 Shapiro, L. and Dinarello, C.A. (1995) Osmotic regulation of cytokine synthesis in vitro. Proc. Natl. Acad. Sci. U.S.A. 92, 12230–12234 114 Jantsch, J. et al. (2015) Cutaneous Na+ storage strengthens the antimicrobial barrier function of the skin and boosts macrophagedriven host defense. Cell Metab. (in press) 115 Neuhofer, W. (2010) Role of NFAT5 in inflammatory disorders associated with osmotic stress. Curr. Genomics 11, 584–590 116 Buxade, M. et al. (2012) Gene expression induced by Toll-like receptors in macrophages requires the transcription factor NFAT5. J. Exp. Med. 209, 379–393 117 Lee, M. and Choy, J.C. (2013) Positive feedback regulation of human inducible nitric-oxide synthase expression by Ras protein Snitrosylation. J. Biol. Chem. 288, 15677–15686 118 Mundy-Bosse, B.L. et al. (2011) Myeloid-derived suppressor cell inhibition of the IFN response in tumor-bearing mice. Cancer Res. 71, 5101–5110 119 Kolluru, G.K. et al. (2013) Hydrogen sulfide chemical biology: pathophysiological roles and detection. Nitric Oxide 35, 5–20 120 Altaany, Z. et al. (2014) The coordination of S-sulfhydration, Snitrosylation, and phosphorylation of endothelial nitric oxide synthase by hydrogen sulfide. Sci. Signal. 7, ra87 121 Oh, G.S. et al. (2006) Hydrogen sulfide inhibits nitric oxide production and nuclear factor-kappaB via heme oxygenase-1 expression in RAW264.7 macrophages stimulated with lipopolysaccharide. Free Radic. Biol. Med. 41, 106–119 122 Badiei, A. et al. (2013) Inhibition of hydrogen sulfide production by gene silencing attenuates inflammatory activity of LPS-activated RAW264.7 cells. Appl. Microbiol. Biotechnol. 97, 7845–7852 123 Gozzelino, R. et al. (2010) Mechanisms of cell protection by heme oxygenase-1. Annu. Rev. Pharmacol. Toxicol. 50, 323–354 124 Fourquet, S. et al. (2010) Activation of NRF2 by nitrosative agents and H2O2 involves KEAP1 disulfide formation. J. Biol. Chem. 285, 8463– 8471 125 Sarady, J.K. et al. (2004) Carbon monoxide protection against endotoxic shock involves reciprocal effects on iNOS in the lung and liver. FASEB J. 18, 854–856 126 Vareille, M. et al. (2008) Heme oxygenase-1 is a critical regulator of nitric oxide production in enterohemorrhagic Escherichia coliinfected human enterocytes. J. Immunol. 180, 5720–5726 127 Nairz, M. et al. (2014) Iron at the interface of immunity and infection. Front. Pharmacol. 5, 152 128 Tzima, S. et al. (2009) Myeloid heme oxygenase-1 regulates innate immunity and autoimmunity by modulating IFN-beta production. J. Exp. Med. 206, 1167–1179 129 Jais, A. et al. (2014) Heme oxygenase-1 drives metaflammation and insulin resistance in mouse and man. Cell 158, 25–40 130 Daniliuc, S. et al. (2003) Hypoxia inactivates inducible nitric oxide synthase in mouse macrophages by disrupting its interaction with alpha-actinin 4. J. Immunol. 171, 3225–3232 131 Zhang, W. et al. (2003) Protein–protein interactions involving inducible nitric oxide synthase. Acta Physiol. Scand. 179, 137–142 132 Yoshida, M. and Xia, Y. (2003) Heat shock protein 90 as an endogenous protein enhancer of inducible nitric-oxide synthase. J. Biol. Chem. 278, 36953–36958 133 Davis, A.S. et al. (2007) Mechanism of inducible nitric oxide synthase exclusion from mycobacterial phagosomes. PLOS Pathog. 3, 1887–1894 134 Mazumdar, T. et al. (2010) Regulation of NF-kappaB activity and inducible nitric oxide synthase by regulatory particle non-ATPase subunit 13 (Rpn13). Proc. Natl. Acad. Sci. U.S.A. 107, 13854–13859 135 Hadkar, V. et al. (2004) Carboxypeptidase-mediated enhancement of nitric oxide production in rat lungs and microvascular endothelial cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 287, L35–L45 136 Goto, Y. et al. (2015) Substrate-dependent nitric oxide synthesis by secreted endoplasmic reticulum aminopeptidase 1 in macrophages. J. Biochem. Published online January 9, 2015. http://dx.doi.org/10.1093/ jb/mvv001

15

TREIMM-1157; No. of Pages 18

Feature review 137 Morris, S.M., Jr (2009) Recent advances in arginine metabolism: roles and regulation of the arginases. Br. J. Pharmacol. 157, 922–930 138 Michel, T. (2013) R is for arginine: metabolism of arginine takes off again, in new directions. Circulation 128, 1400–1404 139 El Kasmi, K.C. et al. (2008) Toll-like receptor-induced arginase 1 in macrophages thwarts effective immunity against intracellular pathogens. Nat. Immunol. 9, 1399–1406 140 Ko¨nig, T. et al. (2009) Translational repression of inducible NO synthase in macrophages by L-arginine depletion is not associated with an increased phosphorylation of eIF2alpha. Immunobiology 214, 822–827 141 Duque-Correa, M.A. et al. (2014) Macrophage arginase-1 controls bacterial growth and pathology in hypoxic tuberculosis granulomas. Proc. Natl. Acad. Sci. U.S.A. 111, E4024–E4032 142 Elms, S. et al. (2013) Insights into the arginine paradox: evidence against the importance of subcellular location of arginase and eNOS. Am. J. Physiol. Heart Circ. Physiol. 305, H651–H666 143 Qualls, J.E. et al. (2012) Sustained generation of nitric oxide and control of mycobacterial infection requires argininosuccinate synthase 1. Cell Host Microbe 12, 313–323 144 Wijnands, K.A. et al. (2012) Citrulline a more suitable substrate than arginine to restore NO production and the microcirculation during endotoxemia. PLoS ONE 7, e37439 145 Grady, S.L. et al. (2013) Argininosuccinate synthetase 1 depletion produces a metabolic state conducive to herpes simplex virus 1 infection. Proc. Natl. Acad. Sci. U.S.A. 110, E5006–E5015 146 Wu, G. et al. (2009) Arginine metabolism and nutrition in growth, health and disease. Amino Acids 37, 153–168 147 Stadelmann, B. et al. (2013) The role of arginine and argininemetabolizing enzymes during Giardia–host cell interactions in vitro. BMC Microbiol. 13, 256 148 Regunathan, S. and Piletz, J.E. (2003) Regulation of inducible nitric oxide synthase and agmatine synthesis in macrophages and astrocytes. Ann. N. Y. Acad. Sci. 1009, 20–29 149 Molderings, G.J. and Haenisch, B. (2012) Agmatine (decarboxylated L-arginine): physiological role and therapeutic potential. Pharmacol. Ther. 133, 351–365 150 Deeb, R.S. et al. (2013) Characterization of a cellular denitrase activity that reverses nitration of cyclooxygenase. Am. J. Physiol. Heart Circ. Physiol. 305, H687–H698 151 Nathan, C. and Shiloh, M.U. (2000) Reactive oxygen and nitrogen intermediates in the relationship between mammalian hosts and microbial pathogens. Proc. Natl. Acad. Sci. U.S.A. 97, 8841–8848 152 Fang, F.C. (2004) Antimicrobial reactive oxygen and nitrogen species: concepts and controversies. Nat. Rev. Immunol. 2, 820–832 153 Feng, H.M. and Walker, D.H. (2000) Mechanisms of intracellular killing of Rickettsia conorii in infected human endothelial cells, hepatocytes, and macrophages. Infect. Immun. 68, 6729–6736 154 Serbina, N.V. et al. (2003) TNF/iNOS-producing dendritic cells mediate innate immune defense against bacterial infection. Immunity 19, 59–70 155 Copin, R. et al. (2012) In situ microscopy analysis reveals local innate immune response developed around Brucella infected cells in resistant and susceptible mice. PLoS Pathog. 8, e1002575 156 Gebreselassie, N.G. et al. (2012) Eosinophils preserve parasitic nematode larvae by regulating local immunity. J. Immunol. 188, 417–425 157 Su, Y.C. et al. (2015) Dual pro-inflammatory and antiviral properties of pulmonary eosinophils in respiratory syncytial virus (RSV) vaccineenhanced disease. J. Virol. 89, 1564–1578 158 Jones-Carson, J. et al. (2008) Inactivation of [Fe–S] metalloproteins mediates nitric oxide-dependent killing of Burkholderia mallei. PLoS ONE 3, e1976 159 Richardson, A.R. et al. (2011) Multiple targets of nitric oxide in the tricarboxylic acid cycle of Salmonella enterica serovar typhimurium. Cell Host Microbe 10, 33–43 160 Savidge, T.C. et al. (2011) Host S-nitrosylation inhibits clostridial small molecule-activated glucosylating toxins. Nat. Med. 17, 1136– 1141 161 Muller, A.J. et al. (2013) Photoconvertible pathogen labeling reveals nitric oxide control of Leishmania major infection in vivo via dampening of parasite metabolism. Cell Host Microbe 14, 460–467

16

Trends in Immunology xxx xxxx, Vol. xxx, No. x

162 Burkholder, K.M. et al. (2011) A small molecule deubiquitinase inhibitor increases localization of inducible nitric oxide synthase to the macrophage phagosome and enhances bacterial killing. Infect. Immun. 79, 4850–4857 163 Bogdan, C. et al. (2000) Fibroblasts as host cells in latent leishmaniosis. J. Exp. Med. 191, 2121–2129 164 Olekhnovitch, R. et al. (2014) Collective nitric oxide production provides tissue-wide immunity during Leishmania infection. J. Clin. Invest. 124, 1711–1722 165 Herbst, S. et al. (2011) Interferon gamma activated macrophages kill mycobacteria by nitric oxide induced apoptosis. PLoS ONE 6, e19105 166 Zwaferink, H. et al. (2008) Stimulation of inducible nitric oxide synthase expression by beta interferon increases necrotic death of macrophages upon Listeria monocytogenes infection. Infect. Immun. 76, 1649–1656 167 Ito, C. et al. (2013) Endogenous nitrated nucleotide is a key mediator of autophagy and innate defense against bacteria. Mol. Cell 52, 794– 804 168 Li, X. et al. (2015) Atg7 enhances host defense against infection via downregulation of superoxide but upregulation of nitric oxide. J. Immunol. 194, 1112–1121 169 Nairz, M. et al. (2013) Nitric oxide-mediated regulation of ferroportin1 controls macrophage iron homeostasis and immune function in Salmonella infection. J. Exp. Med. 210, 855–873 170 Axelrod, S. et al. (2008) Delay of phagosome maturation by a mycobacterial lipid is reversed by nitric oxide. Cell Microbiol. 10, 1530–1545 171 Vareille, M. et al. (2007) Nitric oxide inhibits Shiga-toxin synthesis by enterohemorrhagic Escherichia coli. Proc. Natl. Acad. Sci. U.S.A. 104, 10199–10204 172 Branchu, P. et al. (2014) NsrR, GadE, and GadX interplay in repressing expression of the Escherichia coli O157:H7 LEE pathogenicity island in response to nitric oxide. PLoS Pathog. 10, e1003874 173 Barraud, N. et al. (2014) Nitric oxide: a key mediator of biofilm dispersal with applications in infectious diseases. Curr. Pharm. Des. 21, 21–42 174 Ramphul, U.N. et al. (2014) Plasmodium falciparum evades mosquito immunity by disrupting JNK-mediated apoptosis of invaded midgut cells. Proc. Natl. Acad. Sci. U.S.A. Published online December 31, 2014. http://dx.doi.org/10.1073/pnas.1423586112 175 Burton, N.A. et al. (2014) Disparate impact of oxidative host defenses determines the fate of Salmonella during systemic infection in mice. Cell Host Microbe 15, 72–83 176 Lu, W. et al. (2012) Structural and functional characterization of the nitrite channel NirC from Salmonella typhimurium. Proc. Natl. Acad. Sci. U.S.A. 109, 18395–18400 177 Cole, C. et al. (2012) Nitric oxide increases susceptibility of Toll-like receptor-activated macrophages to spreading Listeria monocytogenes. Immunity 36, 807–820 178 Bogdan, C. (2012) Listeria monocytogenes: no spreading without NO. Immunity 36, 697–699 179 Shatalin, K. et al. (2008) Bacillus anthracis-derived nitric oxide is essential for pathogen virulence and survival in macrophages. Proc. Natl. Acad. Sci. U.S.A. 105, 1009–1013 180 van Sorge, N.M. et al. (2013) Methicillin-resistant Staphylococcus aureus bacterial nitric-oxide synthase affects antibiotic sensitivity and skin abscess development. J. Biol. Chem. 288, 6417–6426 181 Brown, G.C. (1995) Reversible binding and inhibition of catalase by nitric oxide. Eur. J. Biochem. 232, 188–191 182 Sigfrid, L.A. et al. (2003) Cytokines and nitric oxide inhibit the enzyme activity of catalase but not its protein or mRNA expression in insulin-producing cells. J. Mol. Endocrinol. 31, 509–518 183 Karlinsey, J.E. et al. (2012) The NsrR regulon in nitrosative stress resistance of Salmonella enterica serovar Typhimurium. Mol. Microbiol. 85, 1179–1193 184 Husain, M. et al. (2014) Ferric uptake regulator-dependent antinitrosative defenses in Salmonella enterica serovar Typhimurium pathogenesis. Infect. Immun. 82, 333–340 185 Kinkel, T.L. et al. (2013) The Staphylococcus aureus SrrAB twocomponent system promotes resistance to nitrosative stress and hypoxia. MBio 4, e00696–00613

TREIMM-1157; No. of Pages 18

Feature review 186 Lopez, C.A. et al. (2012) Phage-mediated acquisition of a type III secreted effector protein boosts growth of salmonella by nitrate respiration. MBio 3, e00143–e212 187 Winter, S.E. et al. (2013) Host-derived nitrate boosts growth of E. coli in the inflamed gut. Science 339, 708–711 188 Spees, A.M. et al. (2013) Streptomycin-induced inflammation enhances Escherichia coli gut colonization through nitrate respiration. MBio 4, e00143–e212 189 Seth, D. et al. (2012) Endogenous protein S-mitrosylation in E. coli: regulation by OxyR. Science 336, 470–473 190 Jung, J.Y. et al. (2013) The intracellular environment of human macrophages that produce nitric oxide promotes growth of mycobacteria. Infect. Immun. 81, 3198–3209 191 Tan, M.P. et al. (2010) Nitrate respiration protects hypoxic Mycobacterium tuberculosis against acid- and reactive nitrogen species stresses. PLoS ONE 5, e13356 192 Sohaskey, C.D. (2008) Nitrate enhances the survival of Mycobacterium tuberculosis during inhibition of respiration. J. Bacteriol. 190, 2981–2986 193 Cunningham-Bussel, A. et al. (2013) Nitrite produced by Mycobacterium tuberculosis in human macrophages in physiologic oxygen impacts bacterial ATP consumption and gene expression. Proc. Natl. Acad. Sci. U.S.A. 110, E4256–E4265 194 Mocca, B. and Wang, W. (2012) Bacterium-generated nitric oxide hijacks host tumor necrosis factor alpha signaling and modulates the host cell cycle in vitro. J. Bacteriol. 194, 4059–4068 195 Das, P. et al. (2010) Modulation of the arginase pathway in the context of microbial pathogenesis: a metabolic enzyme moonlighting as an immune modulator. PLoS Pathog. 6, e1000899 196 Cusumano, Z.T. et al. (2014) Streptococcus pyogenes arginine and citrulline catabolism promotes infection and modulates innate immunity. Infect. Immun. 82, 233–242 197 Choi, Y. et al. (2012) Expression of STM4467-encoded arginine deiminase controlled by the STM4463 regulator contributes to Salmonella enterica serovar Typhimurium virulence. Infect. Immun. 80, 4291–4297 198 Darlyuk, I. et al. (2009) Arginine homeostasis and transport in the human pathogen Leishmania donovani. J. Biol. Chem. 284, 19800– 19807 199 Bogdan, C. (2008) Mechanisms and consequences of persistence of intracellular pathogens: leishmaniasis as an example. Cell. Microbiol. 10, 1221–1234 200 Brune, B. et al. (2013) Redox control of inflammation in macrophages. Antioxid. Redox Signal. 19, 595–637 201 Blyszczuk, P. et al. (2013) Nitric oxide synthase 2 is required for conversion of pro-fibrogenic inflammatory CD133+ progenitors into F4/80+ macrophages in experimental autoimmune myocarditis. Cardiovasc. Res. 97, 219–229 202 Rigamonti, E. et al. (2013) Requirement of inducible nitric oxide synthase for skeletal muscle regeneration after acute damage. J. Immunol. 190, 1767–1777 203 Sarkar, S. et al. (2011) Complex inhibitory effects of nitric oxide on autophagy. Mol. Cell 43, 19–32 204 Yang, Z. et al. (2010) Lymphocyte development requires Snitrosoglutathione reductase. J. Immunol. 185, 6664–6669 205 Niedbala, W. et al. (2013) Nitric oxide-induced regulatory T cells inhibit Th17 but not Th1 cell differentiation and function. J. Immunol. 191, 164–170 206 Niedbala, W. et al. (2014) Nitric oxide enhances Th9 cell differentiation and airway inflammation. Nat. Commun. 5, 4575 207 Yang, J. et al. (2013) T cell-derived inducible nitric oxide synthase switches off Th17 cell differentiation. J. Exp. Med. 210, 1447–1462 208 Lee, S.W. et al. (2011) Nitric oxide modulates TGF-beta-directive signals to suppress Foxp3+ regulatory T cell differentiation and potentiate Th1 development. J. Immunol. 186, 6972–6980 209 Obermajer, N. et al. (2013) Induction and stability of human Th17 cells require endogenous NOS2 and cGMP-dependent NO signaling. J. Exp. Med. 210, 1433–1445 210 Tumurkhuu, G. et al. (2010) B1 cells produce nitric oxide in response to a series of toll-like receptor ligands. Cell. Immunol. 261, 122–127

Trends in Immunology xxx xxxx, Vol. xxx, No. x

211 Saini, A.S. et al. (2014) Inducible nitric oxide synthase is a major intermediate in signaling pathways for the survival of plasma cells. Nat. Immunol. 15, 275–282 212 Jayasekera, J.P. et al. (2006) Enhanced antiviral antibody secretion and attenuated immunopathology during influenza virus infection in nitric oxide synthase-2-deficient mice. J. Gen. Virol. 87, 3361–3371 213 Tezuka, H. et al. (2007) Regulation of IgA production by naturally occurring TNF/iNOS-producing dendritic cells. Nature 448, 929–933 214 Giordano, D. et al. (2014) Nitric oxide regulates BAFF expression and T cell-independent antibody responses. J. Immunol. 193, 1110–1120 215 Ren, G. et al. (2008) Mesenchymal stem cell-mediated immunosuppression occurs via concerted action of chemokines and nitric oxide. Cell Stem Cell 2, 141–150 216 Shi, Y. et al. (2012) How mesenchymal stem cells interact with tissue immune responses. Trends Immunol. 33, 136–143 217 Lukacs-Kornek, V. et al. (2011) Regulated release of nitric oxide by nonhematopoietic stroma controls expansion of the activated T cell pool in lymph nodes. Nat. Immunol. 12, 1096–1104 218 Siegert, S. et al. (2011) Fibroblastic reticular cells from lymph nodes attenuate T cell expansion by producing nitric oxide. PLoS ONE 6, e27618 219 Scallan, J.P. and Davis, M.J. (2013) Genetic removal of basal nitric oxide enhances contractile activity in isolated murine collecting lymphatic vessels. J. Physiol. 591, 2139–2156 220 Chandrasekar, B. et al. (2013) Regulation of chemokines, CCL3 and CCL4, by interferon gamma and nitric oxide synthase 2 in mouse macrophages and during Salmonella enterica serovar typhimurium infection. J. Infect. Dis. 207, 1556–1568 221 Molon, B. et al. (2011) Chemokine nitration prevents intratumoral infiltration of antigen-specific T cells. J. Exp. Med. 208, 1949–1962 222 Norman, M.U. et al. (2008) Interferon-gamma limits Th1 lymphocyte adhesion to inflamed endothelium: a nitric oxide regulatory feedback mechanism. Eur. J. Immunol. 38, 1368–1380 223 Giordano, D. et al. (2011) Nitric oxide controls an inflammatory-like Ly6ChiPDCA1+ DC subset that regulates Th1 immune responses. J. Leukoc. Biol. 89, 443–455 224 Hernandez-Cuellar, E. et al. (2012) Cutting edge: nitric oxide inhibits the NLRP3 inflammasome. J. Immunol. 189, 5113–5117 225 Mishra, B.B. et al. (2013) Nitric oxide controls the immunopathology of tuberculosis by inhibiting NLRP3 inflammasome-dependent processing of IL-1beta. Nat. Immunol. 14, 52–60 226 Mao, K. et al. (2013) Nitric oxide suppresses NLRP3 inflammasome activation and protects against LPS-induced septic shock. Cell Res. 23, 201–212 227 Purushothaman, D. et al. (2013) Apoptotic programs are determined during lineage commitment of CD4+ T effectors: selective regulation of T effector-memory apoptosis by inducible nitric oxide synthase. J. Immunol. 190, 97–105 228 Bogdan, C. (1998) The multiplex function of nitric oxide in (auto)immunity. J. Exp. Med. 187, 1361–1365 229 Burke, A.J. et al. (2013) The yin and yang of nitric oxide in cancer progression. Carcinogenesis 34, 503–512 230 Schairer, D.O. et al. (2012) The potential of nitric oxide releasing therapies as antimicrobial agents. Virulence 3, 271–279 231 Hauser, B. et al. (2005) Nitric oxide synthase inhibition in sepsis? Lessons learned from large-animal studies. Anesth. Analg. 101, 488– 498 232 Fletcher, A.L. et al. (2014) Lymph node fibroblastic reticular cell transplants show robust therapeutic efficacy in high-mortality murine sepsis. Sci. Transl. Med. 6, 249ra109 233 Beury, D.W. et al. (2014) Cross-talk among myeloid-derived suppressor cells, macrophages, and tumor cells impacts the inflammatory milieu of solid tumors. J. Leukoc. Biol. 96, 1109–1118 234 Weigert, A. and Brune, B. (2008) Nitric oxide, apoptosis and macrophage polarization during tumor progression. Nitric Oxide 19, 95–102 235 Gabrilovich, D.I. et al. (2012) Coordinated regulation of myeloid cells by tumours. Nat. Rev. Immunol. 12, 253–268 236 Heinecke, J.L. et al. (2014) Tumor microenvironment-based feedforward regulation of NOS2 in breast cancer progression. Proc. Natl. Acad. Sci. U.S.A. 111, 6323–6328

17

TREIMM-1157; No. of Pages 18

Feature review 237 Long, Y. et al. (2013) Arginine deiminase resistance in melanoma cells is associated with metabolic reprogramming, glucose dependence, and glutamine addiction. Mol. Cancer Ther. 12, 2581–2590 238 Huang, H.L. et al. (2013) Attenuation of argininosuccinate lyase inhibits cancer growth via cyclin A2 and nitric oxide. Mol. Cancer Ther. 12, 2505–2516 239 Phillips, M.M. et al. (2013) Targeting arginine-dependent cancers with arginine-degrading enzymes: opportunities and challenges. Cancer Res. Treat. 45, 251–262 240 Feun, L.G. et al. (2015) Arginine deprivation in cancer therapy. Curr. Opin. Clin. Nutr. Metab. Care 18, 78–82 241 Austin, S.A. et al. (2013) Endothelial nitric oxide deficiency promotes Alzheimer’s disease pathology. J. Neurochem. 127, 691–700 242 Nathan, C. et al. (2005) Protection from Alzheimer’s-like disease in the mouse by genetic ablation of inducible nitric oxide synthase. J. Exp. Med. 202, 1163–1169 243 Heneka, M.T. et al. (2013) NLRP3 is activated in Alzheimer’s disease and contributes to pathology in APP/PS1 mice. Nature 493, 674–678 244 Kummer, M.P. et al. (2012) Nitric oxide decreases the enzymatic activity of insulin degrading enzyme in APP/PS1 mice. J. Neuroimmune Pharmacol. 7, 165–172 245 Liu, P. et al. (2014) Altered arginine metabolism in Alzheimer’s disease brains. Neurobiol. Aging 35, 1992–2003 246 Stuehr, D.J. et al. (2009) Structural and mechanistic aspects of flavoproteins: electron transfer through the nitric oxide synthase flavoprotein domain. FEBS J. 276, 3959–3974 247 Campbell, M.G. et al. (2014) Molecular architecture of mammalian nitric oxide synthases. Proc. Natl. Acad. Sci. U.S.A. 111, E3614–E3623

18

Trends in Immunology xxx xxxx, Vol. xxx, No. x

248 Smith, B.C. and Marletta, M.A. (2012) Mechanisms of S-nitrosothiol formation and selectivity in nitric oxide signaling. Curr. Opin. Chem. Biol. 16, 498–506 249 Heinrich, T.A. et al. (2013) Biological nitric oxide signalling: chemistry and terminology. Br. J. Pharmacol. 169, 1417–1429 250 Motz, G.T. and Coukos, G. (2013) Deciphering and reversing tumor immune suppression. Immunity 39, 61–73 251 Jayaraman, P. et al. (2012) Tumor-expressed inducible nitric oxide synthase controls induction of functional myeloid-derived suppressor cells through modulation of vascular endothelial growth factor release. J. Immunol. 188, 5365–5376 252 Klug, F. et al. (2013) Low-dose irradiation programs macrophage differentiation to an iNOS+/M1 phenotype that orchestrates effective T cell immunotherapy. Cancer Cell 24, 589–602 253 Eyler, C.E. et al. (2011) Glioma stem cell proliferation and tumor growth are promoted by nitric oxide synthase-2. Cell 146, 53–66 254 Weiss, J.M. et al. (2010) Macrophage-dependent nitric oxide expression regulates tumor cell detachment and metastasis after IL-2/anti-CD40 immunotherapy. J. Exp. Med. 207, 2455–2467 255 Switzer, C.H. et al. (2012) S-nitrosylation of EGFR and Src activates an oncogenic signaling network in human basal-like breast cancer. Mol. Cancer Res. 10, 1203–1215 256 Jeon, H.M. et al. (2014) Crosstalk between glioma-initiating cells and endothelial cells drives tumor progression. Cancer Res. 74, 4482–4492 257 Deng, M. et al. (2015) Shedding of the tumor necrosis factor (TNF) receptor from the surface of hepatocytes during sepsis limits inflammation through cGMP signaling. Sci. Signal. http:// stke.sciencemag.org/content/8/361/ra11.long

Nitric oxide synthase in innate and adaptive immunity: an update.

Thirty years after the discovery of its production by activated macrophages, our appreciation of the diverse roles of nitric oxide (NO) continues to g...
1MB Sizes 0 Downloads 15 Views