Next Article in Journal
Ameliorative Effect of Tocotrienols on Perimenopausal-Associated Osteoporosis—A Review
Next Article in Special Issue
Nitric Oxide Sensing by a Blue Fluorescent Protein
Previous Article in Journal
Effect of Tomato, Beetroot and Carrot Juice Addition on Physicochemical, Antioxidant and Texture Properties of Wheat Bread
Previous Article in Special Issue
The Breast Cancer Protooncogenes HER2, BRCA1 and BRCA2 and Their Regulation by the iNOS/NOS2 Axis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Strategies of Pathogens to Escape from NO-Based Host Defense

by
Giovanna De Simone
1,*,
Alessandra di Masi
1 and
Paolo Ascenzi
2
1
Dipartimento di Scienze, Università Roma Tre, 00146 Roma, Italy
2
Laboratorio Interdipartimentale di Microscopia Elettronica, Via della Vasca Navale 79, 00146 Roma, Italy
*
Author to whom correspondence should be addressed.
Antioxidants 2022, 11(11), 2176; https://doi.org/10.3390/antiox11112176
Submission received: 20 October 2022 / Accepted: 27 October 2022 / Published: 3 November 2022
(This article belongs to the Special Issue NO Role in Evolution: Significance and Signaling)

Abstract

:
Nitric oxide (NO) is an essential signaling molecule present in most living organisms including bacteria, fungi, plants, and animals. NO participates in a wide range of biological processes including vasomotor tone, neurotransmission, and immune response. However, NO is highly reactive and can give rise to reactive nitrogen and oxygen species that, in turn, can modify a broad range of biomolecules. Much evidence supports the critical role of NO in the virulence and replication of viruses, bacteria, protozoan, metazoan, and fungi, thus representing a general mechanism of host defense. However, pathogens have developed different mechanisms to elude the host NO and to protect themselves against oxidative and nitrosative stress. Here, the strategies evolved by viruses, bacteria, protozoan, metazoan, and fungi to escape from the NO-based host defense are overviewed.

1. Introduction

Nitric oxide (NO) is a pivotal messenger molecule able to control, via the induction of cyclic guanosine monophosphate (cGMP) production, a wide range of biological processes including vasomotor tone, neurotransmission, and immune response [1,2,3,4,5,6,7,8,9,10,11]. Furthermore, NO controls gene transcription and mRNA translation by modulating the transcriptional activity of the iron-responsive elements [12,13,14,15].
NO is generated in a variety of cell types by the concomitant conversion of L-arginine to L-citrulline through a reaction catalyzed by at least three distinct isoforms of NO synthase (NOS) [16,17,18,19,20,21,22,23,24]. In mammals, NOS-I and NOS-III are constitutively expressed calcium-dependent enzymes also known as neuronal NOS (nNOS) and endothelial NOS (eNOS), respectively, since they have been originally isolated in neuronal and vascular endothelial cells. NOS-II is a calcium-independent inducible form (iNOS) whose expression is induced by pathological conditions (e.g., inflammation and infection). NO produced by NOS-II contributes to the antimicrobial activity of macrophages and represents the fraction of exhaled NO, which is considered an inflammatory marker [22,23,25,26]. As NOS-II acts as a cytotoxic effector and immune modulator [22,23,25,26], inducible NO production needs a tight control as it can be detrimental to the host [25,27,28,29,30,31,32] (Figure 1). Indeed, free radical NO (NO): (i) forms the nitrosonium cations (NO+); (ii) generates the nitroxyl anion (NO); (iii) auto-oxidizes into dinitrogen trioxide (N2O3), which rapidly is converted to nitrite (NO2); and (iv) reacts with the superoxide radical anion (O2●−) and H2O2-producing reactive nitrogen species (RNS) [30]. Indeed, the reaction of NO with O2●− leads to the powerful oxidant peroxynitrite (ONOO/HOONO), which is more reactive than its precursors NO and O2●− [33,34,35,36]. Peroxynitrite reacts with carbon dioxide (CO2) to form 1-carboxylato-2-nitrosodioxidane (ONOOC[O]O), which decays by homolysis of the O–O bond to yield the reactive species nitrogen dioxide (NO2) and trioxocarbonate (CO3●−); of note, CO3●− is a stronger oxidant than NO2 and ONOO [34,35,37] (Figure 1). Most reactions of CO3●− are one-electron oxidations with preference for Tyr and Trp residues. In addition, NO auto-oxidation leads to the formation of N2O3 that induces the nitrosation of Cys residues (Figure 2). In turn, S-nitrosothiols (RSNOs) are involved in cell signaling and regulatory processes (e.g., bronchodilation and neuroprotection). Furthermore, NO binds to transition metals (e.g., the heme-Fe atom). In the vasculature, NO is rapidly converted to NO3 by oxyhemoglobin (Hb(II)O2), which contributes to the very short half-life of NO (0.1–2.0 s) and in turn acts as a local modulator of vasodilation. Moreover, NO participates to O2 transport by delivery from hemoglobin (Hb) [22,30,38,39,40,41,42,43,44] (Figure 2).
NO, nitrite, nitrate, and peroxynitrite are bactericidal molecules that play a central role in the ability of activated macrophages to kill pathogens [45]. Macrophages respond to cytokines and recognize several molecules exhibiting pathogen-associated molecular patterns (PAMPs) through specific receptors (e.g., primarily Toll-like receptors (TLRs) and Nod-like receptors (NLRs)) [45].
Much evidence supports the critical role of NO in the virulence and replication of many viruses, bacteria, and parasites, thus representing a general mechanism of host defense [26,31,46,47,48]. However, pathogens have developed several mechanisms to elude the host barriers to protect themselves against oxidative and nitrosative stress [32,49]. Under anaerobic conditions, many environmental commensal (e.g., Escherichia coli) and pathogenic bacteria (e.g., Mycobacterium tuberculosis, Salmonella typhimurium) reduce the NOS-II-derived NO and dietary nitrate or nitrite to NO via nitrate respiration, nitrate dissimilation, or denitrification [45]. On the contrary, Listeria monocytogenes uses the host-derived NO as a gateway to enter the host [45,50,51]. Here, the strategies evolved by viruses, bacteria, protozoan, metazoan, and fungi to escape from the NO-based host defense are reviewed.
Figure 2. Effects of NO under physiological and pathological conditions. The activity of the constitutive NOS-I and NOS-III enzymes depends on the intracellular calcium levels. On the contrary, NOS-II expression is calcium-independent and is induced under inflammatory conditions. At low physiological concentrations (green box), NO produced by constitutive NOS-I and NOS-III acts as a physiological modulator of protein activities (e.g., reacting with heme groups, iron–sulfur cluster, and thiol groups). At high pathological NO concentration (red box) (e.g., during inflammation and infections) and in the presence of oxidative agents (i.e., O2●−), NO can trigger irreversible damages to biological macromolecules (i.e., DNA oxidation, protein nitrosylation, oxidation of thiol groups, and lipid peroxidation) and can induce cell overall damage (e.g., inhibition of cytochrome c oxidase (CcO) activity, inactivation of mitochondrial respiration, cell membrane alteration, and induction of cell death and differentiation). The three-dimensional structures of Hb (PDB ID: 1NQP) [52], guanylate cyclase (PDB ID: 4NI2) [53], aconitase (PDB ID: 2B3Y) [54], and NMDA receptor (PDB ID: 6IRA) [55] were drawn using UCSF-Chimera [56].
Figure 2. Effects of NO under physiological and pathological conditions. The activity of the constitutive NOS-I and NOS-III enzymes depends on the intracellular calcium levels. On the contrary, NOS-II expression is calcium-independent and is induced under inflammatory conditions. At low physiological concentrations (green box), NO produced by constitutive NOS-I and NOS-III acts as a physiological modulator of protein activities (e.g., reacting with heme groups, iron–sulfur cluster, and thiol groups). At high pathological NO concentration (red box) (e.g., during inflammation and infections) and in the presence of oxidative agents (i.e., O2●−), NO can trigger irreversible damages to biological macromolecules (i.e., DNA oxidation, protein nitrosylation, oxidation of thiol groups, and lipid peroxidation) and can induce cell overall damage (e.g., inhibition of cytochrome c oxidase (CcO) activity, inactivation of mitochondrial respiration, cell membrane alteration, and induction of cell death and differentiation). The three-dimensional structures of Hb (PDB ID: 1NQP) [52], guanylate cyclase (PDB ID: 4NI2) [53], aconitase (PDB ID: 2B3Y) [54], and NMDA receptor (PDB ID: 6IRA) [55] were drawn using UCSF-Chimera [56].
Antioxidants 11 02176 g002

2. Differences in Macrophage NO Production and Regulation in Mice and Humans

In macrophages, an important component of the host immune defense is represented by NOS-II-mediated NO production [57]. Murine macrophages produce large amounts of NO via NOS-II, consuming most of the L-arginine generated by arginase, an enzyme inactive in human macrophages [58,59,60,61,62]. Moreover, in murine macrophages the transcription of NOS-II is induced by IFN-γ, IFN-α/β, and microbial products (e.g., lipopolysaccharide (LPS)). Furthermore, murine macrophages synthetize the tetrahydrobiopterin (BH4) cofactor that is necessary to stabilize and promote NOS-II activity [45,59,61,63]. In contrast, human, rabbit, Syrian hamster, and goat macrophages do not synthesize BH4 [61,64,65].
In murine macrophages, LPS and INFs induce NOS-II transcription by the activation of the nuclear factor NF-kB and the ISGF3-complex (composed of STAT1, STAT2, and IRF9) [45,66,67]. In human cells, the expression of NOS-II depends on: (i) the cell type, (ii) the activation NF-kB and STAT1 by LPS, INFs, and cytokines, (iii) structural changes in the promoter of NOS-II, and (iv) post-translational histone modifications [62,68]. For example, human alveolar macrophages stimulated with IFN-γ and LPS are not able to express NOS-II, as the NOS-II promoter is subjected to epigenetic gene silencing via CpG methylation, histone modifications, and chromatin compaction [45,57,69,70,71].
Differences in the regulation of the NO pathway have been reported in rats and humans [57]. Arteries from rats subjected to bacterial peritonitis or infused with LPS show a contractile hypo-responsiveness independent from NO production; in contrast, in vitro activation of the aortic smooth muscle cells with a variety of molecules (e.g., LPS, IL-1β, and TNF-α) induces NO production [57]. Similarly, human vessels treated with cytokines do not display any change in NOS mRNA levels, while isolated human aortic smooth muscle cells stimulated with LPS plus cytokines increase NOS-II mRNA expression [57].
Interestingly, both in murine and human cells the mRNA and protein expression levels of NOS-II are affected by microRNAs (miRNAs) [45]. Most of the identified miRNAs (including miR-125a-5p, miR-146a, miR-149, and miR-155) indirectly regulate the expression of NOS-II by blocking the expression of transcription factors (e.g., IRAK-1, NF-kB repressing factor, and SOCS-1) [45,72]. In contrast, miR-939, miR-26a, and miR-146a expressed in human hepatocytes, T cells, lymphoma cells, and mouse renal carcinoma cells, respectively, cause a downregulation of NOS-II expression and NO production by interacting with the 3′-UTR of NOS-II mRNA [72,73,74].

3. Antiviral Action of NO

The antiviral action of NO has been reported for several DNA and RNA virus families including Picornaviridae, Flaviviridae, Coronaviridae, Rhabdoviridae, Reoviridae, Retroviridae, Parvoviridae, Herpesviridae, and Poxviridae [30,31,75,76,77,78,79]. In fact, NO affects: (i) the transcription process, (ii) the post-transcriptional enzyme activity, and (iii) the protein assembly [75,78,80,81,82,83,84]. NO can prevent virus replication by the S-nitrosylation of Cys residues present in viral proteases, reductases, and reverse transcriptases [30,31,75,79] and/or by modulating early transcript levels through the introduction of strand breakages and deamination of adenine and guanine bases [85].
During viral infection, the effect of NO may depend on both the time of infection and the pathogenic mechanism. Once viruses infect the host cells, the increased NO production can upregulate NOS-II through three pathways [30] (Figure 3). The first pathway is dependent on TLRs, which are expressed on the membrane of macrophages, T and B lymphocytes, and non-immune epithelial cells. TLRs recognize the viral DNA or RNA and activate both NF-κB and AP1 [78]. The second pathway involves IFN-γ, which is produced by lymphocytes and activates STAT-1 [30]. The third pathway involves viral double-stranded RNAs that induce IFNs and R-protein kinases (PKR) expression [86] (Figure 2 and Figure 3).
In non-respiratory infections (e.g., hepatitis C, dengue fever, and herpes simplex), NO production is typically associated with pathogenic effects [79,87,88,89,90]. In fact, a positive correlation between NOS-II activity and liver damage has been found in patients chronically infected by hepatitis C virus [88,89]. Moreover, the serum of humans affected by dengue fever shows increased NO levels that, by inducing dilatation of the blood vessels, could be relevant in the evolution of the infection into a severe hemorrhagic form [90]. NO also displays a role in the pathogenesis of herpes simplex virus type 1 (HSV-1) as demonstrated by the fact that mice infected with HSV-1 and treated with the NOS-II inhibitor L-NMMA show a higher survival rate and a better lung compliance compared to the untreated control group [79,87,91].
In the respiratory infections (e.g., common cold caused by human rhinovirus (HRV) and influenza A virus (IAV), MERS and COVID-19 severe acute respiratory syndrome coronavirus (SARS-CoV-2) infections), the NO production may provide a first line of defense against viruses, acting as an innate immune response component [79]. Patients infected with HRV display NOS-II upregulation that results in high levels of exhaled NO (3- to 5-fold higher compared to the baseline), which in turn mitigates HRV-induced symptoms and contributes to the disease resolution [79,92,93]. Similarly, high NO levels have been also observed in patients affected by IAV infection; however, NO does not influence IAV patient symptoms [94] (Figure 3).
Viruses might avoid the NO-based antiviral effect though different strategies: (i) by inhibiting NOS-II activity during the early stages of the viral infection [95], (ii) by lacking the nitrosylation consensus sequence in the viral protease [96,97], and (iii) by infecting hosts with pre-existing endothelial dysfunction(s) and reduced NO bioavailability [32,98,99,100]. Viruses that can interfere with NO synthesis seem to replicate more rapidly [101]. In this regard, the adenovirus E1A reduces NOS-II transcription thus eluding the host immune responses. Similarly, in vitro studies suggest that HSV-1 inhibits the activity of constitutive NOS in the brain during the early stages of infection and in the presence of circulating glucocorticoids. Brain NOS inhibition by HSV-1 may play a role in the neuronal viral invasion and in the activation of the adrenocortical system [95]. In vitro evidence shows that Coxsackievirus has a selective advantage when the 3Cpro viral protease lacks the consensus nitrosylation motif (i.e., Xxx(Lys,Arg,His)Cys(Asp,Glu)) [96,97].While NO does not inhibit genome replication of viruses encoding serine proteases (e.g., Alphavirus), it blocks the replication of viruses encoding cysteine proteases (e.g., Picornavirus and Coronavirus) [84,97,102]. Indeed, in silico and in vitro studies suggest that S-nitrosylation of cysteine proteases (e.g., TMPRSS2 and 3CL) can play a key role in the inhibition of the SARS-CoV-2 viral life cycle [103,104,105,106,107,108]. Indeed, S-nitrosylation of Cys281 and/or Cys348 residue(s) of TMPRSS2 impairs SARS-CoV-2 entry into lung epithelial cells [108]. In addition, S-nitrosylation of the Cys145 residue of the 3CL protease interrupts the maturation of the viral polyproteins that are necessary for replication [105,108]. Of note, in vitro studies suggest that the NO donor sodium nitroprusside can nitrosylate the Cys289 residue of the Ang II type 1 receptor (AT1R), thus impairing AT1R:Ang II recognition [109,110]. Furthermore, in vivo studies show that SARS-CoV-2-infected rats treated with NO can restore the ACE2-angiotensin (1–7)-Mas system, outweighing the deleterious effects of COVID-19 progression by: (i) increasing circulating S-nitrosothiol levels: (ii) reducing vasoconstriction; and (iii) increasing aortic PKC nitrosylation [111,112]. COVID-19 mortality has been correlated to NO production in the endothelium and to NO bioavailability in different groups (e.g., age and gender) and/or in the presence of comorbidities. SARS-CoV-2 induces an inflammatory response mediated by various chemokines, cytokines, and other immune-related factors that in turn leads to endothelial dysfunction(s) such as a proliferative and prothrombotic status causing the disruption of the vascular integrity [113,114]. These conditions are associated with aging and cause the decrease in NO availability in the elderly [115,116,117]. Interestingly, these SARS-CoV-2-dependent inflammatory responses are attenuated in females, where estrogen generates a protective environment by increasing the expression and activity of the NOS-II enzyme and consequently of NO [118]. Therapeutic strategies based on NO donors and NOS inhibitors [31], as well as lifestyle factors (e.g., a diet rich in nitrate and physical exercise), have been proposed to restore the NO bioavailability in the host thus improving the health conditions of COVID-19 patients [32,98,99,100].

4. Effects of NO on Pathogenic Bacteria

Physical and chemical barriers of the innate immune system generally protect the host from invading pathogens by activating macrophages that produce reactive oxygen species (ROS) (e.g., O2●−) and RNS, including NO. ROS and RNS can alter proteins, lipids, and nucleic acids of pathogenic microorganisms, leading to pathogen killing [119]. Interestingly, NO generated in the stomach participates in the protection against pathogens in addition to the stomach acidity [49,120]. However, some bacteria including E. coli, S. typhimurium, Campylobacter jejuni, Mycobacterium leprae, and Mycobacterium tuberculosis have evolved detoxification systems to protect themselves against host-induced oxidative and nitrosative stresses [48,49,121,122]. Interestingly, bacteria produce NO as a by-product of their own metabolism during anaerobic nitrate respiration [49] (Figure 4).

4.1. NO Detoxification in Enteric Bacteria

E. coli and S. typhimurium defend themselves against NO produced by the host immune system by expressing: (i) soluble Hmp flavohemoglobin (flavoHb); (ii) truncated hemoglobins (trHbs); (iii) di-iron-centered flavorubredoxin NorV and NADH-dependent oxidoreductase NorW (NorVW); and (iv) cytochrome c nitrite reductase NrfA [49,123,124,125,126]. FlavoHb expression and activity is NO-dependent [127,128,129,130]. Indeed, flavoHb displays an NO-dioxygenase activity, catalyzing the conversion of NO to NO3 in the presence of O2 and NADH [131,132]. Moreover, flavoHb displays an NADH-dependent alkylhydroperoxide reductase activity, reducing several alkyl-hydroperoxides into their corresponding alcohols [130]. Furthermore, flavoHb has been suggested to repair lipid membrane oxidative damage generated during oxidative and nitrosative stress [126,130]. Under anoxic conditions, a flavoHb-dependent mechanism facilitating NO scavenging and reduction into N2O has been reported [127,128,129,130]. To date, flavoHbs have never been found in higher organisms [126,130].
While flavoHb can act both under oxic and anoxic conditions, leading to the production of NO3 and N2O, respectively, NorVW and NrfA are active only under anaerobic or microoxic conditions, thus resulting in the most important enzymes acting in anaerobic NO detoxification [124,125,133]. While NorVW reduces NO to N2O, NrfA utilizes either NO or NO2 to synthesize ammonia in O2-limited environments [49,124,125,133].
E. coli and S. typhimurium can survive in several different environments thanks to the expression of some transcriptional regulators, including NorR, NsrR, fumarate, nitrate reductase regulator (FNR), ferric-uptake regulator (Fur), and methionine repressor (MetR). The expression of these regulators ensures the response to NO [134,135,136,137,138,139]. FNR acts in the transition between aerobic and anaerobic growth and mediates the upregulation of several operons in response to nitrate and nitrite [136]. Fur controls the expression of genes implicated in the iron uptake, being especially crucial when iron availability is limited. Of note, Fur activity is also impaired by the presence of NO [49,139].
The enteric pathogen C. jejuni is exposed to a range of ROS and RNS produced by the host [140,141]. The ability of C. jejuni to detoxify RNS and ROS has been associated to the expression of a classic 3-on-3 globin (Cgb) and of a group III trHb (trHbP) [142,143,144,145,146]. Cgb helps the micro-aerophilic enteric microorganism to catalyze NO detoxification through a NO deoxygenase or denitrosylase mechanism that implies the transient formation of the heme-Fe(III)-ONOO complex [145,146]. Instead, TrHbP promotes primarily microaerobic growth and moderate respiration; moreover, it participates secondarily to NO metabolism [145,146]. Interestingly, ferrous trHbP (trHbP(II)) binds reversibly NO and displays nitrite reduction activity, leading to heme-Fe(II)-NO [147]. Moreover, ferric trHbP (trHbP(III)) binds reversibly to NO at low pH, whereas it undergoes reductive nitrosylation at alkaline pH. Furthermore, trHbP(III) cooperates with Cgb in the isomerization of peroxynitrite [148]. Of note, under high aeration conditions, C. jejuni strains defective for trHbP are disadvantaged compared to the wild-type ones, achieving lower growth yields and consuming O2 at approximately half the rate displayed by wild-type cells. Remarkably, trHbP mutants do not show increased sensitivity to NO or oxidative stress, suggesting that trHbP may play a role in cell respiration [142,143,144,147]. Overall, while Cgb plays a major role in the resistance to nitrosative stress and aerobically converts NO to nitrate [149], the contribution of trHbP is less prominent. Both Cgb and trHbP are devoid of NO-protective activity under O2-limited conditions that normally exist in vivo [150]. Therefore, the role of trHbP is distinct from that of the Cgb that is involved in O2 metabolism [142,144], likely performing a peroxidase- or P450-like oxygen chemistry [151].

4.2. NO Detoxification in M. leprae and M. tuberculosis

M. leprae and M. tuberculosis represent two of the most dangerous infective pathogens for humans [48,152,153,154]. The ability of mycobacteria to persist in vivo in the presence of RNS produced by activated macrophages [155,156] implies the existence of pseudo-enzymatic detoxification systems, including trHbs [47,48,121,122,157,158,159,160,161,162,163].
The intracellular pathogen M. tuberculosis expresses genes encoding for trHbN (belonging to group I) and trHbO (belonging to group II). TrHbN has primarily been linked to NO detoxification, while trHbO has been proposed to be involved in O2 uptake/transport and/or redox sensing [122,157,164,165,166,167]. M. leprae expresses only trHbO (i.e., GlbO), which shows both O2 uptake/transport and NO detoxification properties [47,121,161].
The involvement of M. tuberculosis trHbN in the protection against RNS has been demonstrated in vivo using both reverse genetic approaches and homologous or heterologous expression systems [158,159,163]. Indeed, the Mycobacterium bovis mutant lacking trHbN does not oxidize NO to NO3 and shows decreased respiration upon exposure to NO [158]. A similar behavior has been predicted for M. tuberculosis given the close phylogenetic relationship between M. bovis and M. tuberculosis and the high identity of trHbNs expressed in these two neighbor species [162]. Moreover, the heterologous expression of M. tuberculosis trHbN significantly protects both M. smegmatis and a flavoHb-deprived mutant of E. coli from NO damage through an O2-sustained detoxification mechanism [159]. A similar protective effect has also been reported for M. smegmatis trHbN [163]. Lastly, the overexpression of M. leprae trHbO alleviates the growth inhibition caused by NO donors in the E. coli hmp mutant, partially complementing the defect in flavoHb synthesis [48,121]. Both M. tuberculosis trHbN and M. leprae GlbO catalyze peroxynitrite scavenging, allowing mycobacteria to survive also in the adverse host macrophagic environment [47,48]. The structural and functional characterization of nitrobindin (Nb) from M. tuberculosis highlighted its ability to scavenge peroxynitrite. This supports the notion that Nb can be part of the pool of proteins required to scavenge RNS produced by the host during the immune response. In this framework, M. tuberculosis Nb may become a novel therapeutic target for the treatment of M. tuberculosis infections as reported also for mycobacterial trHbN [47,48,168,169,170,171,172,173,174,175,176].

4.3. Lysteria Monocytogenes Escapes from NO

The Gram-positive Lysteria monocytogenes is an intracellular pathogen implicated in several outbreaks of foodborne diseases (e.g., gastroenteritis, meningitis, and abortion in susceptible individuals) [51,177]. Upon ingestion, L. monocytogenes survives in TLR-activated macrophages by escaping its internalization via the action of the pore-forming toxin listeriolysin O (LLO) [178]. The maturation of the phagosome requires the activation of the vacuolar H+-ATPase (V-ATPase) that leads to phagosomal acidification through the loss of the early endosomal marker Rab5 and by acquisition of the lysosomal membrane protein-1 (LAMP-1) [50,51]. Once in the cytosol, L. monocytogenes induces actin polymerization via the surface protein ActA to form pseudopod projections that propel bacteria from a primary infected donor cell to a secondary uninfected recipient cell via a process known as cell–cell spread [51]. This process allows the pathogens to remain intracellular, thus avoiding extracellular defense mechanisms and humoral immune factors [177].
L. monocytogenes takes advantage of the NO produced in macrophages by NOS-II in response to TLR to enhance the cell–cell spread during systemic infection [50,51]. In detail, NO delays the maturation of secondary vacuole-containing membrane-encapsulated bacteria, thus increasing both the percentage of infected recipients and the number of bacteria per recipient cell. Although the NO-based mechanism that selectively attenuates secondary vacuole maturation is still unclear, NO produced by NOS-II inhibits the V-ATPase, which could delay maturation of secondary vacuoles [179,180]. However, the inhibition of V-ATPase by bafilomycin and concanamycin A does not reverse NO effects. For this reason, it has been speculated that NO reduces the phagosome acidification and degrades secondary vacuoles by direct modification of Rab4 and LAMP-1 proteins or indirect activation of protein kinase G [51,181]. Of note, during cell–cell spread the effects of TLR stimulation with LPS prevail on the antibacterial action of IFN-γ. Indeed, even if the amount of NO produced in LPS- and IFN-γ-stimulated macrophages is comparable, the TLR stimulation of “recipient” macrophages with IFN-γ alone is insufficient to reduce the L. monocytogenes spreading. TLR signals induced by IFN-αβ, IL-6, or TNF-α inhibit the induction of IFN-γ antimicrobial effectors (e.g., p65 GTPases) impairing L. monocytogenes killing in both primary and secondary vacuoles [50,51]. Furthermore, L. monocytogenes spread is organ specific. Indeed, NOS-II inhibitors partially prevent bacterial spread in the liver but not in the spleen, possibly because NOS-II is significantly less expressed in the spleen compared to the liver [50,51].

5. Antiparasitic Effects of NO on Protozoa and Metazoa

NO exerts antiparasitic effects on Protozoa (i.e., Leishmania, Trypanosoma, Giardia, Trichomonas, Naegleria, Entamoeba, Plasmodium, Toxoplasma, and Babesia) and Metazoa (i.e., Schistosoma, Fasciola, Dicrocoelium, Opisthorchis, Taenia, Echinococcus, Trichinella, Ascaris, and Onchocerca). Interestingly, a role of host heme-proteins (i.e., myoglobin (Mb), neuroglobin (Ngb), Hb, and hemocyanin (Hc)) has been postulated in Trypanosoma, Toxoplasma, Plasmodium, and Schistosoma protection from the parasiticidal effect of NO [26,46,182,183,184] (Figure 4).

5.1. Trypanosoma cruzi

T. cruzi is the protozoan parasite that causes the Chagas disease. T. cruzi shows a complex life cycle, involving the triatomine hematophagous vector and a mammalian host. In the host, trypomastigotes penetrate phagocytic cell lines, transform into the amastigote stage, which replicate, and emerge from ruptured cells as trypomastigotes. Then, while some of the parasites penetrate again in cells continuing with the intracellular division, other trypomastigotes circulate in the blood to be picked up by triatomine bugs during the blood meal [185,186].
NO blocks the T. cruzi life cycle both in vitro and in vivo [187,188]. Macrophages from mice infected by T. cruzi kill trypomastigotes by producing high levels of NO [189,190,191]. Furthermore, cardiomyocytes invaded by T. cruzi express NOS isoforms, thus increasing NO levels and metabolites [192,193,194]. Accordingly, mice deficient in NOS-II [195] or treated with NOS-II inhibitors show an increased susceptibility to T. cruzi. In addition, mice infected with T. cruzi and exposed to NO donors are more protected against trypomastigotes compared to the control group [196,197]. Lastly, NO promotes splenocyte apoptosis during the acute phase of T. cruzi infection in mice [198,199] and modulates parasite cell entry, thus contributing to the pathogenesis of Chagas cardiomyopathy [188,200,201,202].
In mammals, T. cruzi preferentially invades heart as well as skeletal and smooth muscle [185,186] where Mb acts as a NO scavenger [182]. In fact, oxygenated Mb (Mb(II)O2) reacts rapidly and irreversibly with NO leading to the harmless NO3 and the ferric Mb derivative (Mb(III)) [203]. In turn, Mb(III) is reduced to the physiologically active form by Mb reductase [204,205]. In heart and skeletal muscle cells, the high concentration of Mb (~3 × 10−4 M) could capture NO thus reducing NO-related adverse effects (e.g., inhibition of the respiratory chain of the parasite) towards T. cruzi [182]. Of note, it has been also speculated that Trypanosoma brucei preferentially localizes in the brain areas expressing Ngb, a hemeprotein involved in the NO/O2 metabolism [182,206].

5.2. Toxoplasma gondii

The preferential localization in retina, heart, and skeletal muscle cells of T. gondii may reflect the NO scavenging activity of Ngb and Mb [207], whose concentrations are ~1.0 × 10−3 M [203] and ~3.0 × 10−4 M, respectively [204]. Ngb and Mb inactivate efficiently NO, reducing its toxoplasmacidal effect and favoring T. gondii colonization in retina and muscle [207,208,209,210]. In contrast to mycobacterial pathogens, T. gondii does not express trHbs as a protection mechanism against NO [207,211]. However, T. gondii encodes superoxide dismutase, catalase, glutathione/thioredoxin-like peroxidases, and peroxiredoxins that all participate to NO metabolism [212].

5.3. Plasmodium falciparum

Oxygenated Hb (Hb(II)O2) has been postulated to protect intraerythrocytic Plasmodia from the parasiticidal effect of NO [213]. P. falciparum is the etiological agent of malaria, the most important cause of death due to a vector-borne infectious disease. Its life cycle involves two hosts (i.e., humans and the Anopheles stephensi mosquito) and several developmental stages in each one. Human infection with P. falciparum starts when, during the blood meal, a female Anopheles mosquito injects sporozoites in the host peripheral circulation. The parasite reproduces asexually in the liver cells (exoerythrocytic schizogony) and in red blood cells (RBCs) (erythrocytic schizogony) and then develops into sexual precursors (gametocytes), which can be taken up by mosquitoes in a blood meal to complete the life cycle [214]. In RBCs, the parasite is surrounded by Hb which acts as a potent scavenger of NO by blocking its antiparasitic effects [215,216]. In vivo, Hb may operate a dual role, both as a scavenger and as an NO donor. Thus, when RBCs are saturated with oxygen, Hb(II)O2 reacts with NO, preventing the antiparasitic effect. In contrast, at low oxygen tension, Hb(II) readily releases NO that exerts cytotoxic effects towards P. falciparum [213,217,218].

5.4. Schistosoma

Schistosomiasis is a devastating parasitic disease diffused in tropical and subtropical countries [219]. Eggs deposited in water release a free-swimming ciliated larva (miracidium) able to penetrate a freshwater snail host. In the snail tissues, miracidia transform into mother sporocysts that generate daughter sporocysts. The latter, after migration into the digestive gland, give rise to infectious fork-tailed larvae called cercariae. These larvae are shed from the snail and penetrate the human skin. During skin penetration, larvae lose tails and become schistosomulae that migrate via the venous circulation to the lungs and the heart. Then, they develop in the liver, exiting the liver via the portal vein system when mature. Male and female adult worms migrate to mesenteric venules of the bowel/rectum and venous plexus of the bladder, depending on the species. Females deposit eggs that move progressively toward the lumen of intestine and of both bladder and ureters. Lastly, eggs are eliminated with feces or urine, the life cycle starting again [219,220].
RNS affect the Schistosoma life cycle at different stages [28,46,183,219,220,221]. Hemocytes circulating in the hemolymph of resistant strains of the freshwater snail host Biomphalaria glabrata eliminate sporocysts of Schistosoma mansoni through inducible NO. Accordingly, NO synthase inhibitors (e.g., Nω-nitro-L-arginine methylester) and the NO scavenger 2-(4-carboxyphenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide) reduce hemocyte-mediated killing of S. mansoni sporocysts [222]. Although NO-mediated cytotoxicity of macrophages is known to eliminate the schistosomula of S. mansoni [221,223,224], the skin-stage schistosomula can penetrate the host skin [225]. Thus, skin-stage schistosomula of Schistosoma japonicum can evade the macrophage NO-mediated cytotoxicity by Sj-Ca8, a calcium-binding protein expressed in cercariae, skin-stage schistosomula, lung-stage schistosomula, and adult worms [225]. Sj-Ca8 impairs skin penetration by cercariae, suppresses macrophage migration, and impairs NO release [225]. These findings render Sj-Ca8 a potential vaccine candidate and a chemotherapeutic target for the prevention and treatment of schistosomiasis [225].
The survival of schistosomula and of adult worms in the definitive host (e.g., human) and of sporocysts in the intermediate freshwater snail host has been proposed depend on the role of Hb and Hc, respectively. Both proteins impair the antiparasitic effects of NO [183]. In the case of Hc, the binuclear oxygenated metal center of Hc (Cu(II)-O22−-Cu(II)) reacting with NO, generates harmless NO oxidation products (i.e., NO2) and the Cu(I)-Cu(II) complex [183]. In the mammalian host, although anemia occurs during schistosomiasis, the high concentration of Hb(II)O2 intercepts NO and protects the parasite [183].

5.5. Ascaris lumbricoides and Ascaris suum

A. lumbricoides and A. suum are nematodes able to parasitize the intestine of humans and pigs, respectively [226,227,228,229]. Ascariasis occurs when embryonated eggs that contaminate hands, utensils, or food are swallowed. In the small intestine the eggs hatch, releasing the larvae that go through the intestinal wall and migrate through the liver and heart, up to lungs. In the lung, larvae are expectorated and ingested, thus arriving to the small intestine, where they mature into adult worms and produce new eggs which are expelled with feces, contaminating the environment [229].
Ascaris worms express an octameric Hb that acts as an NO-dependent deoxygenase by using the endogenous NO as a substrate to detoxify O2 (Figure 5) [227,228]. Although the primary function of Ascaris Hb is to remove O2 from the perienteric fluids, it may also protect the nematode against the NO present either in the host gut or generated by the host innate immune response [227,228]. The CysE15 residue, located in the distal side of the A. lumbricoides Hb heme-pocket, plays a key role in NO destiny, allowing the NO-mediated enzymatic consumption of O2 [228]. On the other hand, the CysF9 residue, located in the proximal side of the heme-pocket of most tetrameric mammalian Hbs, allows the NO-mediated control of O2 delivery [230,231,232,233]. Therefore, A. lumbricoides Hb has been postulated to be evolutionary positioned between the primordial bacterial flavoHbs catalyzing NO/SNO detoxification and cooperative mammalian tetrameric Hbs, which display an NO-mediated O2 delivery mechanism [227,228] (Figure 5).

6. Effects of NO on Fungal Infection

The mushroom kingdom comprises about 1.5 million species, but only 400 fungal species are pathogenic to humans [234]. Fungal pathogens, including Candida albicans and Cryptococcus neoformans, developed several mechanisms to elude the host immune defenses [234,235,236]. Since many stress-protective enzymes use iron as a cofactor, pathogenic fungi induce the expression of genes involved in iron acquisition in response to RNS [234].
C. albicans is part of the healthy human microbiota and colonizes several niches in the body (e.g., oral cavity, gastrointestinal tract, female reproductive tract, and skin) [237]. In healthy individuals, C. albicans represents a harmless commensal that coexists in harmony with other members of the microbiota. However, alterations in the host immune system, pH variations in the local environment, and/or antibiotic therapy can favor C. albicans proliferation and infection [234]. C. albicans can tolerate high levels of NO produced by the host immune response through the expression of the inducible flavoHb-related protein YHB1, which plays a critical role in NO metabolism and in RNS detoxification [235]. YHB1 acts as an NO dioxygenase by converting NO to harmless nitrate [234]. The induction of YHB1 expression is mediated by Cta4p, a Zn(II)2-Cys6-DNA-binding protein belonging to a family of fungal transcription factors [238].
The virulence of C. neoformans is due to the FHB1 flavoHb, which is involved in the detoxification of NO produced by the host-inducible NOS-II [238].Under nitrosative stress, C. neoformans activates the synthesis of proteins involved in detoxification mechanisms including chaperones, oxidoreductase, thioredoxin reductase, and dehydrogenase [232,234].

7. Conclusions

RNS exert positive antiviral effects. This is supported by the success of NO-based therapies as well as by lifestyle factors that restore the physiological NO levels and consequently improve the clinical settings of patients affected by respiratory infections. Indeed, boosting NO, which is depleted by psychological stress and viral assault, provides protection against viral proliferation. Notably, the FDA approved inhaled NO for the treatment of pulmonary hypertension, thrombocytopenia, and respiratory infections, including COVID-19.
In bacteria, protozoa, metazoa, and fungi, the expression of hemeproteins is at the root of several reactions involving NO. This suggests that the evolution of hemeproteins has been greatly influenced by NO-related functions and, in some cases, by mechanisms required to escape from NO-related toxic effects. Indeed, NO is part of denitrifying bacteria metabolism as an intermediate of the nitrogen cycle. However, NO is toxic and harmful for those microorganisms that have not developed mechanisms to counteract host-produced NO effects. Several trHbs found in bacteria are implicated in the tolerance to nitrosative stress; moreover flavoHb has been shown to protect pathogens against NO both aerobically and anaerobically (Figure 6). Since NO can exert beneficial or detrimental functions, further studies are required to understand the evolutionary adaptations of pathogens to counteract host-produced nitrogen and oxygen reactive species.

Author Contributions

G.D.S., A.d.M. and P.A.: conceptualized, wrote, reviewed, and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rapoport, R.M.; Draznin, M.B.; Murad, F. Endothelium-dependent relaxation in rat aorta may be mediated through cyclic GMP-dependent protein phosphorylation. Nature 1983, 306, 174–176. [Google Scholar] [CrossRef] [PubMed]
  2. O’Dell, T.J.; Hawkins, R.D.; Kandel, E.R.; Arancio, O. Tests of the roles of two diffusible substances in long-term potentiation: Evidence for nitric oxide as a possible early retrograde messenger. Proc. Natl. Acad. Sci. USA 1991, 88, 11285–11289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Schuman, E.M.; Madison, D.V. A requirement for the intercellular messenger nitric oxide in long-term potentiation. Science 1991, 254, 1503–1506. [Google Scholar] [CrossRef] [PubMed]
  4. Cardillo, C.; Kilcoyne, C.M.; Cannon, R.O., 3rd; Panza, J.A. Interactions between nitric oxide and endothelin in the regulation of vascular tone of human resistance vessels in vivo. Hypertension 2000, 35, 1237–1241. [Google Scholar] [CrossRef] [Green Version]
  5. Stuart-Smith, K. Demystified. Nitric oxide. Mol. Pathol. 2002, 55, 360–366. [Google Scholar] [CrossRef] [Green Version]
  6. Guzik, T.J.; Korbut, R.; Adamek-Guzik, T. Nitric oxide and superoxide in inflammation and immune regulation. J. Physiol. Pharmacol. 2003, 54, 469–487. [Google Scholar]
  7. Neuman, R.B.; Hayek, S.S.; Poole, J.C.; Rahman, A.; Menon, V.; Kavtaradze, N.; Polhemus, D.; Veledar, E.; Lefer, D.J.; Quyyumi, A.A. Nitric Oxide Contributes to Vasomotor Tone in Hypertensive African Americans Treated with Nebivolol and Metoprolol. J. Clin. Hypertens. 2016, 18, 223–231. [Google Scholar] [CrossRef] [Green Version]
  8. Ghimire, K.; Altmann, H.M.; Straub, A.C.; Isenberg, J.S. Nitric oxide: What’s new to NO? Am. J. Physiol. Cell Physiol. 2017, 312, C254–C262. [Google Scholar] [CrossRef] [Green Version]
  9. Ghasemi, M. Nitric oxide: Antidepressant mechanisms and inflammation. Adv. Pharmacol. 2019, 86, 121–152. [Google Scholar]
  10. Tewari, D.; Sah, A.N.; Bawari, S.; Nabavi, S.F.; Dehpour, A.R.; Shirooie, S.; Braidy, N.; Fiebich, B.L.; Vacca, R.A.; Nabavi, S.M. Role of Nitric Oxide in Neurodegeneration: Function, Regulation, and Inhibition. Curr. Neuropharmacol. 2021, 19, 114–126. [Google Scholar] [CrossRef]
  11. Kumar, R.; Coggan, A.R.; Ferreira, L.F. Nitric oxide and skeletal muscle contractile function. Nitric Oxide Biol. Chem. 2022, 122–123, 54–61. [Google Scholar] [CrossRef] [PubMed]
  12. Brune, B.; Dimmeler, S.; Molina y Vedia, L.; Lapetina, E.G. Nitric oxide: A signal for ADP-ribosylation of proteins. Life Sci. 1994, 54, 61–70. [Google Scholar] [CrossRef]
  13. Brune, B.; Lapetina, E.G. Protein thiol modification of glyceraldehyde-3-phosphate dehydrogenase as a target for nitric oxide signaling. Genet. Eng. 1995, 17, 149–164. [Google Scholar]
  14. Fillebeen, C.; Pantopoulos, K. Redox control of iron regulatory proteins. Redox Rep. 2002, 7, 15–22. [Google Scholar] [CrossRef] [PubMed]
  15. Muronetz, V.I.; Medvedeva, M.V.; Sevostyanova, I.A.; Schmalhausen, E.V. Modification of Glyceraldehyde-3-Phosphate Dehydrogenase with Nitric Oxide: Role in Signal Transduction and Development of Apoptosis. Biomolecules 2021, 11, 1656. [Google Scholar] [CrossRef]
  16. Nathan, C. Nitric oxide as a secretory product of mammalian cells. FASEB J. 1992, 6, 3051–3064. [Google Scholar] [CrossRef]
  17. Moncada, S.; Higgs, A. The L-arginine-nitric oxide pathway. N. Engl. J. Med. 1993, 329, 2002–2012. [Google Scholar] [PubMed]
  18. Moncada, S.; Palmer, R.M.; Higgs, E.A. Nitric oxide: Physiology, pathophysiology, and pharmacology. Pharmacol. Rev. 1991, 43, 109–142. [Google Scholar]
  19. Stuehr, D.J. Structure-function aspects in the nitric oxide synthases. Annu. Rev. Pharmacol. Toxicol. 1997, 37, 339–359. [Google Scholar] [CrossRef] [Green Version]
  20. Geller, D.A.; Billiar, T.R. Molecular biology of nitric oxide synthases. Cancer Metastasis Rev. 1998, 17, 7–23. [Google Scholar] [CrossRef]
  21. Marletta, M.A.; Hurshman, A.R.; Rusche, K.M. Catalysis by nitric oxide synthase. Curr. Opin. Chem. Biol. 1998, 2, 656–663. [Google Scholar] [CrossRef]
  22. Forstermann, U.; Sessa, W.C. Nitric oxide synthases: Regulation and function. Eur. Heart J. 2012, 33, 829–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Anavi, S.; Tirosh, O. iNOS as a metabolic enzyme under stress conditions. Free Radic. Biol. Med. 2020, 146, 16–35. [Google Scholar] [CrossRef] [PubMed]
  24. Mader, M.M.; Boger, R.; Appel, D.; Schwedhelm, E.; Haddad, M.; Mohme, M.; Lamszus, K.; Westphal, M.; Czorlich, P.; Hannemann, J. Intrathecal and systemic alterations of L-arginine metabolism in patients after intracerebral hemorrhage. J. Cereb. Blood Flow Metab. 2021, 41, 1964–1977. [Google Scholar] [CrossRef]
  25. Moncada, S. Nitric oxide in the vasculature: Physiology and pathophysiology. Ann. N. Y. Acad. Sci. 1997, 811, 60–67. [Google Scholar] [CrossRef]
  26. Ascenzi, P.; Bocedi, A.; Gradoni, L. The anti-parasitic effects of nitric oxide. IUBMB Life 2003, 55, 573–578. [Google Scholar] [CrossRef]
  27. De Groote, M.A.; Fang, F.C. NO inhibitions: Antimicrobial properties of nitric oxide. Clin. Infect. Dis. 1995, 21 (Suppl. 2), S162–S165. [Google Scholar] [CrossRef]
  28. Clark, I.A.; Rockett, K.A. Nitric oxide and parasitic disease. Adv. Parasitol. 1996, 37, 1–56. [Google Scholar]
  29. Wink, D.A.; Mitchell, J.B. Chemical biology of nitric oxide: Insights into regulatory, cytotoxic, and cytoprotective mechanisms of nitric oxide. Free Radic. Biol. Med. 1998, 25, 434–456. [Google Scholar] [CrossRef]
  30. Garren, M.R.; Ashcraft, M.; Qian, Y.; Douglass, M.; Brisbois, E.J.; Handa, H. Nitric oxide and viral infection: Recent developments in antiviral therapies and platforms. Appl. Mater. Today 2021, 22, 100887. [Google Scholar] [CrossRef]
  31. Ignarro, L.J. Inhaled NO and COVID-19. Br. J. Pharmacol. 2020, 177, 3848–3849. [Google Scholar] [CrossRef] [PubMed]
  32. Kobayashi, J. Lifestyle-mediated nitric oxide boost to prevent SARS-CoV-2 infection: A perspective. Nitric Oxide Biol. Chem. 2021, 115, 55–61. [Google Scholar] [CrossRef] [PubMed]
  33. Beckman, J.S.; Beckman, T.W.; Chen, J.; Marshall, P.A.; Freeman, B.A. Apparent hydroxyl radical production by peroxynitrite: Implications for endothelial injury from nitric oxide and superoxide. Proc. Natl. Acad. Sci. USA 1990, 87, 1620–1624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Goldstein, S.; Lind, J.; Merenyi, G. Chemistry of peroxynitrites as compared to peroxynitrates. Chem. Rev. 2005, 105, 2457–2470. [Google Scholar] [CrossRef] [PubMed]
  35. Goldstein, S.; Merenyi, G. The chemistry of peroxynitrite: Implications for biological activity. Methods Enzymol. 2008, 436, 49–61. [Google Scholar]
  36. Ahmad, R.; Hussain, A.; Ahsan, H. Peroxynitrite: Cellular pathology and implications in autoimmunity. J. Immunoass. Immunochem. 2019, 40, 123–138. [Google Scholar] [CrossRef]
  37. Augusto, O.; Goldstein, S.; Hurst, J.K.; Lind, J.; Lymar, S.V.; Merenyi, G.; Radi, R. Carbon dioxide-catalyzed peroxynitrite reactivity—The resilience of the radical mechanism after two decades of research. Free Radic. Biol. Med. 2019, 135, 210–215. [Google Scholar] [CrossRef]
  38. Stamler, J.S. Redox signaling: Nitrosylation and related target interactions of nitric oxide. Cell 1994, 78, 931–936. [Google Scholar] [CrossRef]
  39. Stamler, J.S. S-nitrosothiols and the bioregulatory actions of nitrogen oxides through reactions with thiol groups. Curr. Top. Microbiol. Immunol. 1995, 196, 19–36. [Google Scholar]
  40. Stamler, J.S.; Hausladen, A. Oxidative modifications in nitrosative stress. Nat. Struct. Biol. 1998, 5, 247–249. [Google Scholar] [CrossRef]
  41. Smith, B.C.; Marletta, M.A. Mechanisms of S-nitrosothiol formation and selectivity in nitric oxide signaling. Curr. Opin. Chem. Biol. 2012, 16, 498–506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Heinrich, T.A.; da Silva, R.S.; Miranda, K.M.; Switzer, C.H.; Wink, D.A.; Fukuto, J.M. Biological nitric oxide signalling: Chemistry and terminology. Br. J. Pharmacol. 2013, 169, 1417–1429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Pisoschi, A.M.; Pop, A. The role of antioxidants in the chemistry of oxidative stress: A review. Eur. J. Med. Chem. 2015, 97, 55–74. [Google Scholar] [CrossRef] [PubMed]
  44. Premont, R.T.; Singel, D.J.; Stamler, J.S. The enzymatic function of the honorary enzyme: S-nitrosylation of hemoglobin in physiology and medicine. Mol. Asp. Med. 2022, 84, 101056. [Google Scholar] [CrossRef]
  45. Bogdan, C. Nitric oxide synthase in innate and adaptive immunity: An update. Trends Immunol. 2015, 36, 161–178. [Google Scholar] [CrossRef] [PubMed]
  46. Ascenzi, P.; Gradoni, L. Nitric oxide limits parasite development in vectors and in invertebrate intermediate hosts. IUBMB Life 2002, 53, 121–123. [Google Scholar]
  47. Ascenzi, P.; Bocedi, A.; Bolognesi, M.; Fabozzi, G.; Milani, M.; Visca, P. Nitric oxide scavenging by Mycobacterium leprae GlbO involves the formation of the ferric heme-bound peroxynitrite intermediate. Biochem. Biophys. Res. Commun. 2006, 339, 450–456. [Google Scholar] [CrossRef]
  48. Ascenzi, P.; Visca, P. Scavenging of reactive nitrogen species by mycobacterial truncated hemoglobins. Methods Enzymol. 2008, 436, 317–337. [Google Scholar]
  49. Arkenberg, A.; Runkel, S.; Richardson, D.J.; Rowley, G. The production and detoxification of a potent cytotoxin, nitric oxide, by pathogenic enteric bacteria. Biochem. Soc. Trans. 2011, 39, 1876–1879. [Google Scholar] [CrossRef]
  50. Bogdan, C. Listeria monocytogenes: No spreading without NO. Immunity 2012, 36, 697–699. [Google Scholar] [CrossRef] [Green Version]
  51. Cole, C.; Thomas, S.; Filak, H.; Henson, P.M.; Lenz, L.L. Nitric oxide increases susceptibility of Toll-like receptor-activated macrophages to spreading Listeria monocytogenes. Immunity 2012, 36, 807–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Dasgupta, J.; Sen, U.; Choudhury, D.; Datta, P.; Chakrabarti, A.; Chakrabarty, S.B.; Chakrabarty, A.; Dattagupta, J.K. Crystallization and preliminary X-ray structural studies of hemoglobin A2 and hemoglobin E, isolated from the blood samples of beta-thalassemic patients. Biochem. Biophys. Res. Commun. 2003, 303, 619–623. [Google Scholar] [CrossRef]
  53. Seeger, F.; Quintyn, R.; Tanimoto, A.; Williams, G.J.; Tainer, J.A.; Wysocki, V.H.; Garcin, E.D. Interfacial residues promote an optimal alignment of the catalytic center in human soluble guanylate cyclase: Heterodimerization is required but not sufficient for activity. Biochemistry 2014, 53, 2153–2165. [Google Scholar] [CrossRef] [PubMed]
  54. Dupuy, J.; Volbeda, A.; Carpentier, P.; Darnault, C.; Moulis, J.M.; Fontecilla-Camps, J.C. Crystal structure of human iron regulatory protein 1 as cytosolic aconitase. Structure 2006, 14, 129–139. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, J.B.; Chang, S.; Xu, P.; Miao, M.; Wu, H.; Zhang, Y.; Zhang, T.; Wang, H.; Zhang, J.; Xie, C.; et al. Structural Basis of the Proton Sensitivity of Human GluN1-GluN2A NMDA Receptors. Cell Rep. 2018, 25, 3582–3590.e4. [Google Scholar] [CrossRef] [Green Version]
  56. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
  57. Reade, M.C.; Young, J.D. Of mice and men (and rats): Implications of species and stimulus differences for the interpretation of studies of nitric oxide in sepsis. Br. J. Anaesth. 2003, 90, 115–118. [Google Scholar] [CrossRef] [Green Version]
  58. Schneemann, M.; Schoedon, G.; Hofer, S.; Blau, N.; Guerrero, L.; Schaffner, A. Nitric oxide synthase is not a constituent of the antimicrobial armature of human mononuclear phagocytes. J. Infect. Dis. 1993, 167, 1358–11363. [Google Scholar] [CrossRef]
  59. Schoedon, G.; Schneemann, M.; Hofer, S.; Guerrero, L.; Blau, N.; Schaffner, A. Regulation of the L-arginine-dependent and tetrahydrobiopterin-dependent biosynthesis of nitric oxide in murine macrophages. Eur. J. Biochem. 1993, 213, 833–839. [Google Scholar] [CrossRef]
  60. Albina, J.E.; Reichner, J.S. Oxygen and the regulation of gene expression in wounds. Wound Repair Regen. 2003, 11, 445–451. [Google Scholar] [CrossRef]
  61. Schneemann, M.; Schoeden, G. Macrophage biology and immunology: Man is not a mouse. J. Leukoc. Biol. 2007, 81, 579; discussion 580. [Google Scholar] [CrossRef] [PubMed]
  62. Thomas, A.C.; Mattila, J.T. “Of mice and men”: Arginine metabolism in macrophages. Front. Immunol. 2014, 5, 479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Tayeh, M.A.; Marletta, M.A. Macrophage oxidation of L-arginine to nitric oxide, nitrite, and nitrate. Tetrahydrobiopterin is required as a cofactor. J. Biol. Chem. 1989, 264, 19654–19658. [Google Scholar] [CrossRef]
  64. Pfister, H.; Remer, K.A.; Brcic, M.; Fatzer, R.; Christen, S.; Leib, S.; Jungi, T.W. Inducible nitric oxide synthase and nitrotyrosine in listeric encephalitis: A cross-species study in ruminants. Vet. Pathol. 2002, 39, 190–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Perez, L.E.; Chandrasekar, B.; Saldarriaga, O.A.; Zhao, W.; Arteaga, L.T.; Travi, B.L.; Melby, P.C. Reduced nitric oxide synthase 2 (NOS2) promoter activity in the Syrian hamster renders the animal functionally deficient in NOS2 activity and unable to control an intracellular pathogen. J. Immunol. 2006, 176, 5519–5528. [Google Scholar] [CrossRef] [Green Version]
  66. Kleinert, H.; Pautz, A.; Linker, K.; Schwarz, P.M. Regulation of the expression of inducible nitric oxide synthase. Eur. J. Pharmacol. 2004, 500, 255–266. [Google Scholar] [CrossRef]
  67. Lima-Junior, D.S.; Costa, D.L.; Carregaro, V.; Cunha, L.D.; Silva, A.L.; Mineo, T.W.; Gutierrez, F.R.; Bellio, M.; Bortoluci, K.R.; Flavell, R.A.; et al. Inflammasome-derived IL-1beta production induces nitric oxide-mediated resistance to Leishmania. Nat. Med. 2013, 19, 909–915. [Google Scholar] [CrossRef]
  68. Pautz, A.; Art, J.; Hahn, S.; Nowag, S.; Voss, C.; Kleinert, H. Regulation of the expression of inducible nitric oxide synthase. Nitric Oxide Biol. Chem. 2010, 23, 75–93. [Google Scholar] [CrossRef]
  69. Gross, T.J.; Kremens, K.; Powers, L.S.; Brink, B.; Knutson, T.; Domann, F.E.; Philibert, R.A.; Milhem, M.M.; Monick, M.M. Epigenetic silencing of the human NOS2 gene: Rethinking the role of nitric oxide in human macrophage inflammatory responses. J. Immunol. 2014, 192, 2326–2338. [Google Scholar] [CrossRef]
  70. Chesrown, S.E.; Monnier, J.; Visner, G.; Nick, H.S. Regulation of inducible nitric oxide synthase mRNA levels by LPS, INF-gamma, TGF-beta, and IL-10 in murine macrophage cell lines and rat peritoneal macrophages. Biochem. Biophys. Res. Commun. 1994, 200, 126–134. [Google Scholar] [CrossRef]
  71. Weinberg, J.B.; Misukonis, M.A.; Shami, P.J.; Mason, S.N.; Sauls, D.L.; Dittman, W.A.; Wood, E.R.; Smith, G.K.; McDonald, B.; Bachus, K.E.; et al. Human mononuclear phagocyte inducible nitric oxide synthase (iNOS): Analysis of iNOS mRNA, iNOS protein, biopterin, and nitric oxide production by blood monocytes and peritoneal macrophages. Blood 1995, 86, 1184–1195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Colasanti, M.; Persichini, T.; Venturini, G.; Ascenzi, P. S-nitrosylation of viral proteins: Molecular bases for antiviral effect of nitric oxide. IUBMB Life 1999, 48, 25–31. [Google Scholar] [CrossRef] [PubMed]
  73. Colasanti, M.; Gradoni, L.; Mattu, M.; Persichini, T.; Salvati, L.; Venturini, G.; Ascenzi, P. Molecular bases for the anti-parasitic effect of NO (Review). Int. J. Mol. Med. 2002, 9, 131–134. [Google Scholar] [CrossRef] [PubMed]
  74. Mehta, D.R.; Ashkar, A.A.; Mossman, K.L. The nitric oxide pathway provides innate antiviral protection in conjunction with the type I interferon pathway in fibroblasts. PLoS ONE 2012, 7, e31688. [Google Scholar] [CrossRef] [Green Version]
  75. Abdul-Cader, M.S.; Amarasinghe, A.; Abdul-Careem, M.F. Activation of toll-like receptor signaling pathways leading to nitric oxide-mediated antiviral responses. Arch. Virol. 2016, 161, 2075–2086. [Google Scholar] [CrossRef]
  76. Lisi, F.; Zelikin, A.N.; Chandrawati, R. Nitric Oxide to Fight Viral Infections. Adv. Sci. 2021, 8, 2003895. [Google Scholar] [CrossRef]
  77. Powell, K.L.; Baylis, S.A. The antiviral effects of nitric oxide. Trends Microbiol. 1995, 3, 81–82. [Google Scholar] [CrossRef]
  78. Peterhans, E. Reactive oxygen species and nitric oxide in viral diseases. Biol. Trace Elem. Res. 1997, 56, 107–116. [Google Scholar] [CrossRef]
  79. Maeda, H.; Akaike, T. Nitric oxide and oxygen radicals in infection, inflammation, and cancer. Biochemistry 1998, 63, 854–865. [Google Scholar]
  80. Olcott, M.C.; Andersson, J.; Sjoberg, B.M. Localization and characterization of two nucleotide-binding sites on the anaerobic ribonucleotide reductase from bacteriophage T4. J. Biol. Chem. 1998, 273, 24853–24860. [Google Scholar] [CrossRef]
  81. Reiss, C.S.; Komatsu, T. Does nitric oxide play a critical role in viral infections? J. Virol. 1998, 72, 4547–4551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wei, L.; Gravitt, P.E.; Song, H.; Maldonado, A.M.; Ozbun, M.A. Nitric oxide induces early viral transcription coincident with increased DNA damage and mutation rates in human papillomavirus-infected cells. Cancer Res. 2009, 69, 4878–4884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Xu, W.; Zheng, S.; Dweik, R.A.; Erzurum, S.C. Role of epithelial nitric oxide in airway viral infection. Free Radic. Biol. Med. 2006, 41, 19–28. [Google Scholar] [CrossRef] [PubMed]
  84. Adler, H.; Beland, J.L.; Del-Pan, N.C.; Kobzik, L.; Brewer, J.P.; Martin, T.R.; Rimm, I.J. Suppression of herpes simplex virus type 1 (HSV-1)-induced pneumonia in mice by inhibition of inducible nitric oxide synthase (iNOS, NOS2). J. Exp. Med. 1997, 185, 1533–1540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Mihm, S.; Fayyazi, A.; Ramadori, G. Hepatic expression of inducible nitric oxide synthase transcripts in chronic hepatitis C virus infection: Relation to hepatic viral load and liver injury. Hepatology 1997, 26, 451–458. [Google Scholar] [CrossRef] [PubMed]
  86. Kandemir, O.; Polat, A.; Kaya, A. Inducible nitric oxide synthase expression in chronic viral hepatitis and its relation with histological severity of disease. J. Viral Hepat. 2002, 9, 419–423. [Google Scholar] [CrossRef]
  87. Valero, N.; Espina, L.M.; Anez, G.; Torres, E.; Mosquera, J.A. Short report: Increased level of serum nitric oxide in patients with dengue. Am. J. Trop. Med. Hyg. 2002, 66, 762–764. [Google Scholar] [CrossRef] [Green Version]
  88. Gamba, G.; Cavalieri, H.; Courreges, M.C.; Massouh, E.J.; Benencia, F. Early inhibition of nitric oxide production increases HSV-1 intranasal infection. J. Med. Virol. 2004, 73, 313–322. [Google Scholar] [CrossRef]
  89. Sanders, S.P.; Siekierski, E.S.; Richards, S.M.; Porter, J.D.; Imani, F.; Proud, D. Rhinovirus infection induces expression of type 2 nitric oxide synthase in human respiratory epithelial cells in vitro and in vivo. J. Allergy Clin. Immunol. 2001, 107, 235–243. [Google Scholar] [CrossRef]
  90. Sanders, S.P.; Proud, D.; Permutt, S.; Siekierski, E.S.; Yachechko, R.; Liu, M.C. Role of nasal nitric oxide in the resolution of experimental rhinovirus infection. J. Allergy Clin. Immunol. 2004, 113, 697–702. [Google Scholar] [CrossRef]
  91. McGill, J.; Heusel, J.W.; Legge, K.L. Innate immune control and regulation of influenza virus infections. J. Leukoc. Biol. 2009, 86, 803–812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Hasson, J.; Weidenfeld, J.; Mizrachi-Kol, R.; Ben-Hur, T.; Ovadia, H. The effect of herpes simplex virus-1 on nitric oxide synthase in the rat brain: The role of glucocorticoids. Neuroimmunomodulation 2011, 18, 111–116. [Google Scholar] [CrossRef] [PubMed]
  93. Stamler, J.S.; Toone, E.J.; Lipton, S.A.; Sucher, N.J. (S)NO signals: Translocation, regulation, and a consensus motif. Neuron 1997, 18, 691–696. [Google Scholar] [CrossRef] [Green Version]
  94. Saura, M.; Zaragoza, C.; McMillan, A.; Quick, R.A.; Hohenadl, C.; Lowenstein, J.M.; Lowenstein, C.J. An antiviral mechanism of nitric oxide: Inhibition of a viral protease. Immunity 1999, 10, 21–28. [Google Scholar] [CrossRef] [Green Version]
  95. Banerjee, N.S.; Moore, D.W.; Wang, H.K.; Broker, T.R.; Chow, L.T. NVN1000, a novel nitric oxide-releasing compound, inhibits HPV-18 virus production by interfering with E6 and E7 oncoprotein functions. Antivir. Res. 2019, 170, 104559. [Google Scholar] [CrossRef]
  96. Gandhi, R.T.; Lynch, J.B.; Del Rio, C. Mild or Moderate COVID-19. N. Engl. J. Med. 2020, 383, 1757–1766. [Google Scholar] [CrossRef]
  97. Li, H.; Tian, S.; Chen, T.; Cui, Z.; Shi, N.; Zhong, X.; Qiu, K.; Zhang, J.; Zeng, T.; Chen, L.; et al. Newly diagnosed diabetes is associated with a higher risk of mortality than known diabetes in hospitalized patients with COVID-19. Diabetes Obes. Metab. 2020, 22, 1897–1906. [Google Scholar] [CrossRef]
  98. Cao, W.; Bao, C.; Lowenstein, C.J. Inducible nitric oxide synthase expression inhibition by adenovirus E1A. Proc. Natl. Acad. Sci. USA 2003, 100, 7773–7778. [Google Scholar] [CrossRef] [Green Version]
  99. Mannick, J.B. The antiviral role of nitric oxide. Res. Immunol. 1995, 146, 693–697. [Google Scholar] [CrossRef]
  100. Akerstrom, S.; Gunalan, V.; Keng, C.T.; Tan, Y.J.; Mirazimi, A. Dual effect of nitric oxide on SARS-CoV replication: Viral RNA production and palmitoylation of the S protein are affected. Virology 2009, 395, 1–9. [Google Scholar] [CrossRef]
  101. Kawase, M.; Shirato, K.; van der Hoek, L.; Taguchi, F.; Matsuyama, S. Simultaneous treatment of human bronchial epithelial cells with serine and cysteine protease inhibitors prevents severe acute respiratory syndrome coronavirus entry. J. Virol. 2012, 86, 6537–6545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Akaberi, D.; Krambrich, J.; Ling, J.; Luni, C.; Hedenstierna, G.; Jarhult, J.D.; Lennerstrand, J.; Lundkvist, A. Mitigation of the replication of SARS-CoV-2 by nitric oxide in vitro. Redox Biol. 2020, 37, 101734. [Google Scholar] [CrossRef] [PubMed]
  103. Bestle, D.; Heindl, M.R.; Limburg, H.; Van Lam van, T.; Pilgram, O.; Moulton, H.; Stein, D.A.; Hardes, K.; Eickmann, M.; Dolnik, O.; et al. TMPRSS2 and furin are both essential for proteolytic activation of SARS-CoV-2 in human airway cells. Life Sci. Alliance 2020, 3, e202000786. [Google Scholar] [CrossRef] [PubMed]
  104. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Kruger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.H.; Nitsche, A.; et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell 2020, 181, 271–280.e8. [Google Scholar] [CrossRef]
  105. Gioia, M.; Ciaccio, C.; Calligari, P.; De Simone, G.; Sbardella, D.; Tundo, G.; Fasciglione, G.F.; Di Masi, A.; Di Pierro, D.; Bocedi, A.; et al. Role of proteolytic enzymes in the COVID-19 infection and promising therapeutic approaches. Biochem. Pharmacol. 2020, 182, 114225. [Google Scholar] [CrossRef] [PubMed]
  106. Kidde, J.; Sahebkar, A. From Foe to Friend in COVID-19: Renin-Angiotensin System Inhibitors. J. Infect. Dis. 2021, 223, 174–175. [Google Scholar] [CrossRef] [PubMed]
  107. Leclerc, P.C.; Lanctot, P.M.; Auger-Messier, M.; Escher, E.; Leduc, R.; Guillemette, G. S-nitrosylation of cysteine 289 of the AT1 receptor decreases its binding affinity for angiotensin II. Br. J. Pharmacol. 2006, 148, 306–313. [Google Scholar] [CrossRef] [Green Version]
  108. Hirano, T.; Murakami, M. COVID-19: A New Virus, but a Familiar Receptor and Cytokine Release Syndrome. Immunity 2020, 52, 731–733. [Google Scholar] [CrossRef]
  109. Pinheiro, L.C.; Oliveira-Paula, G.H.; Ferreira, G.C.; Dal-Cin de Paula, T.; Duarte, D.A.; Costa-Neto, C.M.; Tanus-Santos, J.E. Oral nitrite treatment increases S-nitrosylation of vascular protein kinase C and attenuates the responses to angiotensin II. Redox Biol. 2021, 38, 101769. [Google Scholar] [CrossRef]
  110. Forstermann, U. Nitric oxide and oxidative stress in vascular disease. Pflugers Arch. 2010, 459, 923–939. [Google Scholar] [CrossRef]
  111. Rauti, R.; Shahoha, M.; Leichtmann-Bardoogo, Y.; Nasser, R.; Paz, E.; Tamir, R.; Miller, V.; Babich, T.; Shaked, K.; Ehrlich, A.; et al. Effect of SARS-CoV-2 proteins on vascular permeability. Elife 2021, 10, e69314. [Google Scholar] [CrossRef] [PubMed]
  112. Egashira, K.; Inou, T.; Hirooka, Y.; Kai, H.; Sugimachi, M.; Suzuki, S.; Kuga, T.; Urabe, Y.; Takeshita, A. Effects of age on endothelium-dependent vasodilation of resistance coronary artery by acetylcholine in humans. Circulation 1993, 88, 77–81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Taddei, S.; Virdis, A.; Ghiadoni, L.; Salvetti, G.; Bernini, G.; Magagna, A.; Salvetti, A. Age-related reduction of NO availability and oxidative stress in humans. Hypertension 2001, 38, 274–279. [Google Scholar] [CrossRef] [PubMed]
  114. Torregrossa, A.C.; Aranke, M.; Bryan, N.S. Nitric oxide and geriatrics: Implications in diagnostics and treatment of the elderly. J. Geriatr. Cardiol. 2011, 8, 230–242. [Google Scholar] [PubMed] [Green Version]
  115. Nevzati, E.; Shafighi, M.; Bakhtian, K.D.; Treiber, H.; Fandino, J.; Fathi, A.R. Estrogen induces nitric oxide production via nitric oxide synthase activation in endothelial cells. Acta Neurochir. Suppl. 2015, 120, 141–145. [Google Scholar] [PubMed]
  116. Wink, D.A.; Hines, H.B.; Cheng, R.Y.; Switzer, C.H.; Flores-Santana, W.; Vitek, M.P.; Ridnour, L.A.; Colton, C.A. Nitric oxide and redox mechanisms in the immune response. J. Leukoc. Biol. 2011, 89, 873–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Gilchrist, M.; Winyard, P.G.; Benjamin, N. Dietary nitrate-good or bad? Nitric Oxide Biol. Chem. 2010, 22, 104–109. [Google Scholar] [CrossRef]
  118. Fabozzi, G.; Ascenzi, P.; Renzi, S.D.; Visca, P. Truncated hemoglobin GlbO from Mycobacterium leprae alleviates nitric oxide toxicity. Microb. Pathog. 2006, 40, 211–220. [Google Scholar] [CrossRef]
  119. Nardini, M.; Pesce, A.; Bolognesi, M. Truncated (2/2) hemoglobin: Unconventional structures and functional roles in vivo and in human pathogenesis. Mol. Asp. Med. 2022, 84, 101049. [Google Scholar] [CrossRef]
  120. Crawford, M.J.; Goldberg, D.E. Role for the Salmonella flavohemoglobin in protection from nitric oxide. J. Biol. Chem. 1998, 273, 12543–12547. [Google Scholar] [CrossRef]
  121. Mills, P.C.; Rowley, G.; Spiro, S.; Hinton, J.C.D.; Richardson, D.J. A combination of cytochrome c nitrite reductase (NrfA) and flavorubredoxin (NorV) protects Salmonella enterica serovar Typhimurium against killing by NO in anoxic environments. Microbiology 2008, 154, 1218–1228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Bush, M.; Ghosh, T.; Tucker, N.; Zhang, X.; Dixon, R. Transcriptional regulation by the dedicated nitric oxide sensor, NorR: A route towards NO detoxification. Biochem. Soc. Trans. 2011, 39, 289–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Poole, R.K. Flavohaemoglobin: The pre-eminent nitric oxide-detoxifying machine of microorganisms. F1000Research 2020, 9, 7. [Google Scholar] [CrossRef] [Green Version]
  124. Membrillo-Hernandez, J.; Ioannidis, N.; Poole, R.K. The flavohaemoglobin (HMP) of Escherichia coli generates superoxide in vitro and causes oxidative stress in vivo. FEBS Lett. 1996, 382, 141–144. [Google Scholar] [CrossRef] [Green Version]
  125. Membrillo-Hernandez, J.; Coopamah, M.D.; Anjum, M.F.; Stevanin, T.M.; Kelly, A.; Hughes, M.N.; Poole, R.K. The flavohemoglobin of Escherichia coli confers resistance to a nitrosating agent, a “Nitric oxide Releaser,” and paraquat and is essential for transcriptional responses to oxidative stress. J. Biol. Chem. 1999, 274, 748–754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Hausladen, A.; Gow, A.J.; Stamler, J.S. Nitrosative stress: Metabolic pathway involving the flavohemoglobin. Proc. Natl. Acad. Sci. USA 1998, 95, 14100–14105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Bonamore, A.; Boffi, A. Flavohemoglobin: Structure and reactivity. IUBMB Life 2008, 60, 19–28. [Google Scholar] [CrossRef]
  128. Gardner, P.R.; Gardner, A.M.; Martin, L.A.; Salzman, A.L. Nitric oxide dioxygenase: An enzymic function for flavohemoglobin. Proc. Natl. Acad. Sci. USA 1998, 95, 10378–10383. [Google Scholar] [CrossRef] [Green Version]
  129. Gardner, P.R.; Gardner, A.M.; Martin, L.A.; Dou, Y.; Li, T.; Olson, J.S.; Zhu, H.; Riggs, A.F. Nitric-oxide dioxygenase activity and function of flavohemoglobins. sensitivity to nitric oxide and carbon monoxide inhibition. J. Biol. Chem. 2000, 275, 31581–31587. [Google Scholar] [CrossRef] [Green Version]
  130. Gardner, A.M.; Gardner, P.R. Flavohemoglobin detoxifies nitric oxide in aerobic, but not anaerobic, Escherichia coli. Evidence for a novel inducible anaerobic nitric oxide-scavenging activity. J. Biol. Chem. 2002, 277, 8166–8171. [Google Scholar] [CrossRef]
  131. Rodionov, D.A.; Dubchak, I.L.; Arkin, A.P.; Alm, E.J.; Gelfand, M.S. Dissimilatory metabolism of nitrogen oxides in bacteria: Comparative reconstruction of transcriptional networks. PLoS Comput. Biol. 2005, 1, e55. [Google Scholar] [CrossRef] [PubMed]
  132. Spiro, S. Regulators of bacterial responses to nitric oxide. FEMS Microbiol. Rev. 2007, 31, 193–211. [Google Scholar] [CrossRef] [PubMed]
  133. Constantinidou, C.; Hobman, J.L.; Griffiths, L.; Patel, M.D.; Penn, C.W.; Cole, J.A.; Overton, T.W. A reassessment of the FNR regulon and transcriptomic analysis of the effects of nitrate, nitrite, NarXL, and NarQP as Escherichia coli K12 adapts from aerobic to anaerobic growth. J. Biol. Chem. 2006, 281, 4802–4815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Overton, T.W.; Griffiths, L.; Patel, M.D.; Hobman, J.L.; Penn, C.W.; Cole, J.A.; Constantinidou, C. Microarray analysis of gene regulation by oxygen, nitrate, nitrite, FNR, NarL and NarP during anaerobic growth of Escherichia coli: New insights into microbial physiology. Biochem. Soc. Trans. 2006, 34, 104–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Fink, R.C.; Evans, M.R.; Porwollik, S.; Vazquez-Torres, A.; Jones-Carson, J.; Troxell, B.; Libby, S.J.; McClelland, M.; Hassan, H.M. FNR is a global regulator of virulence and anaerobic metabolism in Salmonella enterica serovar Typhimurium (ATCC 14028s). J. Bacteriol. 2007, 189, 2262–2273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Pullan, S.T.; Gidley, M.D.; Jones, R.A.; Barrett, J.; Stevanin, T.M.; Read, R.C.; Green, J.; Poole, R.K. Nitric oxide in chemostat-cultured Escherichia coli is sensed by Fnr and other global regulators: Unaltered methionine biosynthesis indicates lack of S nitrosation. J. Bacteriol. 2007, 189, 1845–1855. [Google Scholar] [CrossRef] [Green Version]
  137. Iovine, N.M.; Pursnani, S.; Voldman, A.; Wasserman, G.; Blaser, M.J.; Weinrauch, Y. Reactive nitrogen species contribute to innate host defense against Campylobacter jejuni. Infect. Immun. 2008, 76, 986–993. [Google Scholar] [CrossRef] [Green Version]
  138. Callahan, S.M.; Dolislager, C.G.; Johnson, J.G. The host cellular immune response to infection by Campylobacter spp. and its role in disease. Infect. Immun. 2021, 89, e0011621. [Google Scholar] [CrossRef]
  139. Wainwright, L.M.; Elvers, K.T.; Park, S.F.; Poole, R.K. A truncated haemoglobin implicated in oxygen metabolism by the microaerophilic food-borne pathogen Campylobacter jejuni. Microbiology 2005, 151, 4079–4091. [Google Scholar] [CrossRef] [Green Version]
  140. Nardini, M.; Pesce, A.; Labarre, M.; Richard, C.; Bolli, A.; Ascenzi, P.; Guertin, M.; Bolognesi, M. Structural determinants in the group III truncated hemoglobin from Campylobacter jejuni. J. Biol. Chem. 2006, 281, 37803–37812. [Google Scholar] [CrossRef]
  141. Wainwright, L.M.; Wang, Y.; Park, S.F.; Yeh, S.R.; Poole, R.K. Purification and spectroscopic characterization of Ctb, a group III truncated hemoglobin implicated in oxygen metabolism in the food-borne pathogen Campylobacter jejuni. Biochemistry 2006, 45, 6003–6011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Tinajero-Trejo, M.; Shepherd, M. The globins of Campylobacter jejuni. Adv. Microb. Physiol. 2013, 63, 97–145. [Google Scholar] [PubMed]
  143. Tinajero-Trejo, M.; Vreugdenhil, A.; Sedelnikova, S.E.; Davidge, K.S.; Poole, R.K. Nitric oxide reactivities of the two globins of the foodborne pathogen Campylobacter jejuni: Roles in protection from nitrosative stress and analysis of potential reductants. Nitric Oxide Biol. Chem. 2013, 34, 65–75. [Google Scholar] [CrossRef]
  144. Ascenzi, P.; di Masi, A.; Tundo, G.R.; Pesce, A.; Visca, P.; Coletta, M. Nitrosylation mechanisms of Mycobacterium tuberculosis and Campylobacter jejuni truncated hemoglobins N, O, and P. PLoS ONE 2014, 9, e102811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Ascenzi, P.; Pesce, A. Peroxynitrite scavenging by Campylobacter jejuni truncated hemoglobin P. J. Biol. Inorg. Chem. 2017, 22, 1141–1150. [Google Scholar] [CrossRef] [PubMed]
  146. Elvers, K.T.; Wu, G.; Gilberthorpe, N.J.; Poole, R.K.; Park, S.F. Role of an inducible single-domain hemoglobin in mediating resistance to nitric oxide and nitrosative stress in Campylobacter jejuni and Campylobacter coli. J. Bacteriol. 2004, 186, 5332–5341. [Google Scholar] [CrossRef] [Green Version]
  147. Avila-Ramirez, C.; Tinajero-Trejo, M.; Davidge, K.S.; Monk, C.E.; Kelly, D.J.; Poole, R.K. Do globins in microaerophilic Campylobacter jejuni confer nitrosative stress tolerance under oxygen limitation? Antioxid. Redox Signal. 2013, 18, 424–431. [Google Scholar] [CrossRef] [Green Version]
  148. Lu, C.; Egawa, T.; Wainwright, L.M.; Poole, R.K.; Yeh, S.R. Structural and functional properties of a truncated hemoglobin from a food-borne pathogen Campylobacter jejuni. J. Biol. Chem. 2007, 282, 13627–13636. [Google Scholar] [CrossRef] [Green Version]
  149. Hussain, T. Leprosy and tuberculosis: An insight-review. Crit. Rev. Microbiol. 2007, 33, 15–66. [Google Scholar] [CrossRef]
  150. Schalcher, T.R.; Vieira, J.L.; Salgado, C.G.; Borges Rdos, S.; Monteiro, M.C. Antioxidant factors, nitric oxide levels, and cellular damage in leprosy patients. Rev. Soc. Bras. Med. Trop. 2013, 46, 645–649. [Google Scholar] [CrossRef]
  151. Keragala, B.; Herath, H.; Janapriya, G.; Vanitha, S.; Balendran, T.; Janani, T.; Keragala, T.S.; Gunasekera, C.N. Coexistence of mycobacterial infections—Mycobacterium tuberculosis and Mycobacterium leprae—In Sri Lanka: A case series. J. Med. Case Rep. 2020, 14, 101. [Google Scholar] [CrossRef] [PubMed]
  152. Schon, T.; Gebre, N.; Sundqvist, T.; Habetmariam, H.S.; Engeda, T.; Britton, S. Increased levels of nitric oxide metabolites in urine from leprosy patients in reversal reaction. Lepr. Rev. 1999, 70, 52–55. [Google Scholar] [CrossRef] [PubMed]
  153. Schon, T.; Hernandez-Pando, R.H.; Negesse, Y.; Leekassa, R.; Sundqvist, T.; Britton, S. Expression of inducible nitric oxide synthase and nitrotyrosine in borderline leprosy lesions. Br. J. Dermatol. 2001, 145, 809–815. [Google Scholar] [CrossRef] [PubMed]
  154. Couture, M.; Yeh, S.R.; Wittenberg, B.A.; Wittenberg, J.B.; Ouellet, Y.; Rousseau, D.L.; Guertin, M. A cooperative oxygen-binding hemoglobin from Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. USA 1999, 96, 11223–11228. [Google Scholar] [CrossRef] [Green Version]
  155. Ouellet, H.; Ouellet, Y.; Richard, C.; Labarre, M.; Wittenberg, B.; Wittenberg, J.; Guertin, M. Truncated hemoglobin HbN protects Mycobacterium bovis from nitric oxide. Proc. Natl. Acad. Sci. USA 2002, 99, 5902–5907. [Google Scholar] [CrossRef] [Green Version]
  156. Pathania, R.; Navani, N.K.; Gardner, A.M.; Gardner, P.R.; Dikshit, K.L. Nitric oxide scavenging and detoxification by the Mycobacterium tuberculosis haemoglobin, HbN in Escherichia coli. Mol. Microbiol. 2002, 45, 1303–1314. [Google Scholar] [CrossRef]
  157. Milani, M.; Pesce, A.; Ouellet, H.; Guertin, M.; Bolognesi, M. Truncated hemoglobins and nitric oxide action. IUBMB Life 2003, 55, 623–627. [Google Scholar] [CrossRef]
  158. Ascenzi, P.; Milani, M.; Visca, P. Peroxynitrite scavenging by ferrous truncated hemoglobin GlbO from Mycobacterium leprae. Biochem. Biophys. Res. Commun. 2006, 351, 528–533. [Google Scholar] [CrossRef]
  159. Ascenzi, P.; Bolognesi, M.; Milani, M.; Guertin, M.; Visca, P. Mycobacterial truncated hemoglobins: From genes to functions. Gene 2007, 398, 42–51. [Google Scholar] [CrossRef]
  160. Lama, A.; Pawaria, S.; Dikshit, K.L. Oxygen binding and NO scavenging properties of truncated hemoglobin, HbN, of Mycobacterium smegmatis. FEBS Lett. 2006, 580, 4031–4041. [Google Scholar] [CrossRef]
  161. Davidge, K.S.; Dikshit, K.L. Haemoglobins of Mycobacteria: Structural features and biological functions. Adv. Microb. Physiol 2013, 63, 147–194. [Google Scholar] [PubMed]
  162. Nardini, M.; Pesce, A.; Milani, M.; Bolognesi, M. Protein fold and structure in the truncated (2/2) globin family. Gene 2007, 398, 2–11. [Google Scholar] [CrossRef] [PubMed]
  163. Vuletich, D.A.; Lecomte, J.T. A phylogenetic and structural analysis of truncated hemoglobins. J. Mol. Evol. 2006, 62, 196–210. [Google Scholar] [CrossRef] [PubMed]
  164. Wittenberg, J.B.; Bolognesi, M.; Wittenberg, B.A.; Guertin, M. Truncated hemoglobins: A new family of hemoglobins widely distributed in bacteria, unicellular eukaryotes, and plants. J. Biol. Chem. 2002, 277, 871–874. [Google Scholar] [CrossRef] [Green Version]
  165. Ascenzi, P.; De Marinis, E.; Visca, P.; Ciaccio, C.; Coletta, M. Peroxynitrite detoxification by ferryl Mycobacterium leprae truncated hemoglobin O. Biochem. Biophys. Res. Commun. 2009, 380, 392–396. [Google Scholar] [CrossRef] [Green Version]
  166. Tripathi, R.P.; Bisht, S.S.; Ajay, A.; Sharma, A.; Misra, M.; Gupt, M.P. Developments in chemical approaches to treat tuberculosis in the last decade. Curr. Med. Chem. 2012, 19, 488–517. [Google Scholar] [CrossRef]
  167. Zumla, A.; Chakaya, J.; Centis, R.; D’Ambrosio, L.; Mwaba, P.; Bates, M.; Kapata, N.; Nyirenda, T.; Chanda, D.; Mfinanga, S.; et al. Tuberculosis treatment and management—An update on treatment regimens, trials, new drugs, and adjunct therapies. Lancet Respir Med. 2015, 3, 220–234. [Google Scholar] [CrossRef]
  168. Ascenzi, P.; Coletta, A.; Cao, Y.; Trezza, V.; Leboffe, L.; Fanali, G.; Fasano, M.; Pesce, A.; Ciaccio, C.; Marini, S.; et al. Isoniazid inhibits the heme-based reactivity of Mycobacterium tuberculosis truncated hemoglobin N. PLoS ONE 2013, 8, e69762. [Google Scholar] [CrossRef] [Green Version]
  169. De Simone, G.; Ascenzi, P.; Polticelli, F. Nitrobindin: An Ubiquitous Family of All beta-Barrel Heme-proteins. IUBMB Life 2016, 68, 423–428. [Google Scholar] [CrossRef] [Green Version]
  170. De Simone, G.; di Masi, A.; Polticelli, F.; Ascenzi, P. Human nitrobindin: The first example of an all-beta-barrel ferric heme-protein that catalyzes peroxynitrite detoxification. FEBS Open Bio 2018, 8, 2002–2010. [Google Scholar] [CrossRef]
  171. De Simone, G.; di Masi, A.; Ciaccio, C.; Coletta, M.; Ascenzi, P. NO Scavenging through reductive nitrosylation of ferric Mycobacterium tuberculosis and Homo sapiens Nitrobindins. Int. J. Mol. Sci. 2020, 21, 9395. [Google Scholar] [CrossRef] [PubMed]
  172. De Simone, G.; di Masi, A.; Vita, G.M.; Polticelli, F.; Pesce, A.; Nardini, M.; Bolognesi, M.; Ciaccio, C.; Coletta, M.; Turilli, E.S.; et al. Mycobacterial and human nitrobindins: Structure and function. Antioxid. Redox Signal. 2020, 33, 229–246. [Google Scholar] [CrossRef] [PubMed]
  173. De Simone, G.; di Masi, A.; Fattibene, P.; Ciaccio, C.; Platas-Iglesias, C.; Coletta, M.; Pesce, A.; Ascenzi, P. Oxygen-mediated oxidation of ferrous nitrosylated nitrobindins. J. Inorg. Biochem. 2021, 224, 111579. [Google Scholar] [CrossRef]
  174. Lecuit, M. Listeria monocytogenes, a model in infection biology. Cell. Microbiol. 2020, 22, e13186. [Google Scholar] [CrossRef] [Green Version]
  175. Cossart, P.; Vicente, M.F.; Mengaud, J.; Baquero, F.; Perez-Diaz, J.C.; Berche, P. Listeriolysin O is essential for virulence of Listeria monocytogenes: Direct evidence obtained by gene complementation. Infect. Immun. 1989, 57, 3629–3636. [Google Scholar] [CrossRef] [Green Version]
  176. Tojo, A.; Guzman, N.J.; Garg, L.C.; Tisher, C.C.; Madsen, K.M. Nitric oxide inhibits bafilomycin-sensitive H(+)-ATPase activity in rat cortical collecting duct. Am. J. Physiol. 1994, 267, F509–F515. [Google Scholar] [CrossRef]
  177. Forgac, M. The vacuolar H+-ATPase of clathrin-coated vesicles is reversibly inhibited by S-nitrosoglutathione. J. Biol. Chem. 1999, 274, 1301–1305. [Google Scholar] [CrossRef] [Green Version]
  178. Akaike, T.; Fujii, S.; Sawa, T.; Ihara, H. Cell signaling mediated by nitrated cyclic guanine nucleotide. Nitric Oxide Biol. Chem. 2010, 23, 166–174. [Google Scholar] [CrossRef]
  179. Ascenzi, P.; Salvati, L.; Brunori, M. Does myoglobin protect Trypanosoma cruzi from the antiparasitic effects of nitric oxide? FEBS Lett. 2001, 501, 103–105. [Google Scholar] [CrossRef] [Green Version]
  180. Ascenzi, P.; Fasano, M.; Gradoni, L. Do hemoglobin and hemocyanin impair schistosoma killing by no? IUBMB Life 2002, 53, 287–288. [Google Scholar] [CrossRef]
  181. Bath, P.M.; Coleman, C.M.; Gordon, A.L.; Lim, W.S.; Webb, A.J. Nitric oxide for the prevention and treatment of viral, bacterial, protozoal and fungal infections. F1000Research 2021, 10, 536. [Google Scholar] [CrossRef] [PubMed]
  182. Gilles, H.M. Protozoal Diseases; Arnold: London, UK, 1999; Volume 94, pp. 113–127. [Google Scholar]
  183. Perez-Molina, J.A.; Molina, I. Chagas disease. Lancet 2018, 391, 82–94. [Google Scholar] [CrossRef]
  184. Burleigh, B.A.; Andrews, N.W. Signaling and host cell invasion by Trypanosoma cruzi. Curr. Opin. Microbiol. 1998, 1, 461–465. [Google Scholar] [CrossRef]
  185. Soeiro Mde, N.; Paiva, M.M.; Barbosa, H.S.; Meirelles Mde, N.; Araujo-Jorge, T.C. A cardiomyocyte mannose receptor system is involved in Trypanosoma cruzi invasion and is down-modulated after infection. Cell Struct. Funct. 1999, 24, 139–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Villalta, F.; Zhang, Y.; Bibb, K.E.; Kappes, J.C.; Lima, M.F. The cysteine-cysteine family of chemokines RANTES, MIP-1α, and MIP-1β induce trypanocidal activity in human macrophages via nitric oxide. Infect. Immun. 1998, 66, 4690–4695. [Google Scholar] [CrossRef] [Green Version]
  187. Aliberti, J.C.; Machado, F.S.; Souto, J.T.; Campanelli, A.P.; Teixeira, M.M.; Gazzinelli, R.T.; Silva, J.S. Chemokines enhance parasite uptake and promote nitric oxide-dependent microbiostatic activity in murine inflammatory macrophages infected with Trypanosoma cruzi. Infect. Immun. 1999, 67, 4819–4826. [Google Scholar] [CrossRef] [Green Version]
  188. Thomson, L.; Gadelha, F.R.; Peluffo, G.; Vercesi, A.E.; Radi, R. Peroxynitrite affects Ca2+ transport in Trypanosoma cruzi. Mol. Biochem. Parasitol. 1999, 98, 81–91. [Google Scholar] [CrossRef]
  189. Eu, J.P.; Xu, L.; Stamler, J.S.; Meissner, G. Regulation of ryanodine receptors by reactive nitrogen species. Biochem. Pharmacol. 1999, 57, 1079–1084. [Google Scholar] [CrossRef]
  190. Chandrasekar, B.; Melby, P.C.; Troyer, D.A.; Freeman, G.L. Differential regulation of nitric oxide synthase isoforms in experimental acute chagasic cardiomyopathy. Clin. Exp. Immunol. 2000, 121, 112–119. [Google Scholar] [CrossRef]
  191. Stamler, J.S.; Meissner, G. Physiology of nitric oxide in skeletal muscle. Physiol. Rev. 2001, 81, 209–237. [Google Scholar] [CrossRef]
  192. Holscher, C.; Kohler, G.; Muller, U.; Mossmann, H.; Schaub, G.A.; Brombacher, F. Defective nitric oxide effector functions lead to extreme susceptibility of Trypanosoma cruzi-infected mice deficient in gamma interferon receptor or inducible nitric oxide synthase. Infect. Immun. 1998, 66, 1208–1215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Vespa, G.N.; Cunha, F.Q.; Silva, J.S. Nitric oxide is involved in control of Trypanosoma cruzi-induced parasitemia and directly kills the parasite in vitro. Infect. Immun. 1994, 62, 5177–5182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Petray, P.; Castanos-Velez, E.; Grinstein, S.; Orn, A.; Rottenberg, M.E. Role of nitric oxide in resistance and histopathology during experimental infection with Trypanosoma cruzi. Immunol. Lett. 1995, 47, 121–126. [Google Scholar] [CrossRef]
  195. Martins, G.A.; Cardoso, M.A.; Aliberti, J.C.; Silva, J.S. Nitric oxide-induced apoptotic cell death in the acute phase of Trypanosoma cruzi infection in mice. Immunol. Lett. 1998, 63, 113–120. [Google Scholar] [CrossRef]
  196. Silva, J.S.; Machado, F.S.; Martins, G.A. The role of nitric oxide in the pathogenesis of Chagas disease. Front. Biosci. 2003, 8, s314–s325. [Google Scholar] [CrossRef] [Green Version]
  197. Huang, H.; Chan, J.; Wittner, M.; Jelicks, L.A.; Morris, S.A.; Factor, S.M.; Weiss, L.M.; Braunstein, V.L.; Bacchi, C.J.; Yarlett, N.; et al. Expression of cardiac cytokines and inducible form of nitric oxide synthase (NOS2) in Trypanosoma cruzi-infected mice. J. Mol. Cell Cardiol. 1999, 31, 75–88. [Google Scholar] [CrossRef]
  198. Machado, F.S.; Martins, G.A.; Aliberti, J.C.; Mestriner, F.L.; Cunha, F.Q.; Silva, J.S. Trypanosoma cruzi-infected cardiomyocytes produce chemokines and cytokines that trigger potent nitric oxide-dependent trypanocidal activity. Circulation 2000, 102, 3003–3008. [Google Scholar] [CrossRef] [Green Version]
  199. Santos, E.; Menezes Falcao, L. Chagas cardiomyopathy and heart failure: From epidemiology to treatment. Rev. Port. Cardiol. (Engl. Ed.) 2020, 39, 279–289. [Google Scholar] [CrossRef]
  200. Eich, R.F.; Li, T.; Lemon, D.D.; Doherty, D.H.; Curry, S.R.; Aitken, J.F.; Mathews, A.J.; Johnson, K.A.; Smith, R.D.; Phillips, G.N., Jr.; et al. Mechanism of NO-induced oxidation of myoglobin and hemoglobin. Biochemistry 1996, 35, 6976–6983. [Google Scholar] [CrossRef]
  201. Livingston, D.J.; McLachlan, S.J.; La Mar, G.N.; Brown, W.D. Myoglobin: Cytochrome b5 interactions and the kinetic mechanism of metmyoglobin reductase. J. Biol. Chem. 1985, 260, 15699–15707. [Google Scholar] [CrossRef]
  202. Ahmad, G.; Chami, B.; El Kazzi, M.; Wang, X.; Moreira, M.T.S.; Hamilton, N.; Maw, A.M.; Hambly, T.W.; Witting, P.K. Catalase-like antioxidant activity is unaltered in hypochlorous acid oxidized horse heart Myoglobin. Antioxidants 2019, 8, 414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Burmester, T.; Weich, B.; Reinhardt, S.; Hankeln, T. A vertebrate globin expressed in the brain. Nature 2000, 407, 520–523. [Google Scholar] [CrossRef] [PubMed]
  204. Ascenzi, P.; Bocedi, A.; Gradoni, L. Do neuroglobin and myoglobin protect Toxoplasma gondii from nitrosative stress? IUBMB Life 2005, 57, 689–691. [Google Scholar] [CrossRef] [PubMed]
  205. Herold, S.; Exner, M.; Nauser, T. Kinetic and mechanistic studies of the NO*-mediated oxidation of oxymyoglobin and oxyhemoglobin. Biochemistry 2001, 40, 3385–3395. [Google Scholar] [CrossRef]
  206. Herold, S.; Fago, A.; Weber, R.E.; Dewilde, S.; Moens, L. Reactivity studies of the Fe(III) and Fe(II)NO forms of human neuroglobin reveal a potential role against oxidative stress. J. Biol. Chem. 2004, 279, 22841–22847. [Google Scholar] [CrossRef] [Green Version]
  207. Brunori, M.; Giuffre, A.; Nienhaus, K.; Nienhaus, G.U.; Scandurra, F.M.; Vallone, B. Neuroglobin, nitric oxide, and oxygen: Functional pathways and conformational changes. Proc. Natl. Acad. Sci. USA 2005, 102, 8483–8488. [Google Scholar] [CrossRef] [Green Version]
  208. Wu, G.; Wainwright, L.M.; Poole, R.K. Microbial globins. Adv. Microb. Physiol. 2003, 47, 255–310. [Google Scholar]
  209. Szewczyk-Golec, K.; Pawlowska, M.; Wesolowski, R.; Wroblewski, M.; Mila-Kierzenkowska, C. Oxidative stress as a possible target in the treatment of toxoplasmosis: Perspectives and ambiguities. Int. J. Mol. Sci. 2021, 22, 5705. [Google Scholar] [CrossRef]
  210. Taylor-Robinson, A.W. Nitric oxide can be released as well as scavenged by haemoglobin: Relevance to its antimalarial activity. Parasite Immunol. 1998, 20, 49–50. [Google Scholar] [CrossRef]
  211. Milner, D.A., Jr. Malaria Pathogenesis. Cold Spring Harb. Perspect. Med. 2018, 8, a025569. [Google Scholar] [CrossRef] [Green Version]
  212. Jones, I.W.; Thomsen, L.L.; Knowles, R.; Gutteridge, W.E.; Butcher, G.A.; Sinden, R.E. Nitric oxide synthase activity in malaria-infected mice. Parasite Immunol. 1996, 18, 535–538. [Google Scholar] [CrossRef] [PubMed]
  213. Lancaster, J.R., Jr. A tutorial on the diffusibility and reactivity of free nitric oxide. Nitric Oxide Biol. Chem. 1997, 1, 18–30. [Google Scholar] [CrossRef] [PubMed]
  214. Seguin, M.C.; Klotz, F.W.; Schneider, I.; Weir, J.P.; Goodbary, M.; Slayter, M.; Raney, J.J.; Aniagolu, J.U.; Green, S.J. Induction of nitric oxide synthase protects against malaria in mice exposed to irradiated Plasmodium berghei infected mosquitoes: Involvement of interferon gamma and CD8+ T cells. J. Exp. Med. 1994, 180, 353–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Nahrevanian, H. Immune effector mechanisms of the nitric oxide pathway in malaria: Cytotoxicity versus cytoprotection. Braz. J. Infect. Dis. 2006, 10, 283–292. [Google Scholar] [CrossRef]
  216. LoVerde, P.T. Schistosomiasis. Adv. Exp. Med. Biol. 2019, 1154, 45–70. [Google Scholar]
  217. Gryseels, B.; Polman, K.; Clerinx, J.; Kestens, L. Human schistosomiasis. Lancet 2006, 368, 1106–1118. [Google Scholar] [CrossRef]
  218. James, S.L.; Glaven, J. Macrophage cytotoxicity against schistosomula of Schistosoma mansoni involves arginine-dependent production of reactive nitrogen intermediates. J. Immunol. 1989, 143, 4208–4212. [Google Scholar]
  219. Hahn, U.K.; Bender, R.C.; Bayne, C.J. Involvement of nitric oxide in killing of Schistosoma mansoni sporocysts by hemocytes from resistant Biomphalaria glabrata. J. Parasitol. 2001, 87, 778–785. [Google Scholar] [CrossRef]
  220. Oswald, I.P.; Eltoum, I.; Wynn, T.A.; Schwartz, B.; Caspar, P.; Paulin, D.; Sher, A.; James, S.L. Endothelial cells are activated by cytokine treatment to kill an intravascular parasite, Schistosoma mansoni, through the production of nitric oxide. Proc. Natl. Acad. Sci. USA 1994, 91, 999–1003. [Google Scholar] [CrossRef] [Green Version]
  221. Wynn, T.A.; Oswald, I.P.; Eltoum, I.A.; Caspar, P.; Lowenstein, C.J.; Lewis, F.A.; James, S.L.; Sher, A. Elevated expression of Th1 cytokines and nitric oxide synthase in the lungs of vaccinated mice after challenge infection with Schistosoma mansoni. J. Immunol. 1994, 153, 5200–5209. [Google Scholar]
  222. Liu, J.; Pan, T.; You, X.; Xu, Y.; Liang, J.; Limpanont, Y.; Sun, X.; Okanurak, K.; Zheng, H.; Wu, Z.; et al. SjCa8, a calcium-binding protein from Schistosoma japonicum, inhibits cell migration and suppresses nitric oxide release of RAW264.7 macrophages. Parasit. Vectors 2015, 8, 513. [Google Scholar] [CrossRef] [Green Version]
  223. Gow, A.J.; Luchsinger, B.P.; Pawloski, J.R.; Singel, D.J.; Stamler, J.S. The oxyhemoglobin reaction of nitric oxide. Proc. Natl. Acad. Sci. USA 1999, 96, 9027–9032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Imai, K. The haemoglobin enzyme. Nature 1999, 401, 437–439. [Google Scholar] [CrossRef] [PubMed]
  225. Minning, D.M.; Gow, A.J.; Bonaventura, J.; Braun, R.; Dewhirst, M.; Goldberg, D.E.; Stamler, J.S. Ascaris haemoglobin is a nitric oxide-activated ‘deoxygenase’. Nature 1999, 401, 497–502. [Google Scholar] [CrossRef]
  226. Dold, C.; Holland, C.V. Ascaris and ascariasis. Microbes Infect. 2011, 13, 632–637. [Google Scholar] [CrossRef]
  227. Singel, D.J.; Stamler, J.S. Chemical physiology of blood flow regulation by red blood cells: The role of nitric oxide and S-nitrosohemoglobin. Annu. Rev. Physiol. 2005, 67, 99–145. [Google Scholar] [CrossRef] [PubMed]
  228. Allen, B.W.; Stamler, J.S.; Piantadosi, C.A. Hemoglobin, nitric oxide and molecular mechanisms of hypoxic vasodilation. Trends Mol. Med. 2009, 15, 452–460. [Google Scholar] [CrossRef] [Green Version]
  229. Palmer, L.A.; Doctor, A.; Gaston, B. SNO-hemoglobin and hypoxic vasodilation. Nat. Med. 2008, 14, 1009, author reply 1009–1010. [Google Scholar] [CrossRef]
  230. Gaston, B.; May, W.J.; Sullivan, S.; Yemen, S.; Marozkina, N.V.; Palmer, L.A.; Bates, J.N.; Lewis, S.J. Essential role of hemoglobin beta-93-cysteine in posthypoxia facilitation of breathing in conscious mice. J. Appl. Physiol. 2014, 116, 1290–1299. [Google Scholar] [CrossRef] [Green Version]
  231. Brown, A.J.; Haynes, K.; Quinn, J. Nitrosative and oxidative stress responses in fungal pathogenicity. Curr. Opin. Microbiol. 2009, 12, 384–391. [Google Scholar] [CrossRef] [Green Version]
  232. Missall, T.A.; Lodge, J.K.; McEwen, J.E. Mechanisms of resistance to oxidative and nitrosative stress: Implications for fungal survival in mammalian hosts. Eukaryot. Cell 2004, 3, 835–846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Reddy, G.K.K.; Padmavathi, A.R.; Nancharaiah, Y.V. Fungal infections: Pathogenesis, antifungals and alternate treatment approaches. Curr. Res. Microbial Sci. 2022, 3, 100137. [Google Scholar] [CrossRef] [PubMed]
  234. Gulati, M.; Nobile, C.J. Candida albicans biofilms: Development, regulation, and molecular mechanisms. Microbes Infect. 2016, 18, 310–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Hromatka, B.S.; Noble, S.M.; Johnson, A.D. Transcriptional response of Candida albicans to nitric oxide and the role of the YHB1 gene in nitrosative stress and virulence. Mol. Biol. Cell 2005, 16, 4814–4826. [Google Scholar] [CrossRef] [PubMed]
  236. Chiranand, W.; McLeod, I.; Zhou, H.; Lynn, J.J.; Vega, L.A.; Myers, H.; Yates, J.R., 3rd; Lorenz, M.C.; Gustin, M.C. CTA4 transcription factor mediates induction of nitrosative stress response in Candida albicans. Eukaryot. Cell 2008, 7, 268–278. [Google Scholar] [CrossRef] [Green Version]
  237. De Jesus-Berrios, M.; Liu, L.; Nussbaum, J.C.; Cox, G.M.; Stamler, J.S.; Heitman, J. Enzymes that counteract nitrosative stress promote fungal virulence. Curr. Biol. 2003, 13, 1963–1968. [Google Scholar] [CrossRef] [Green Version]
  238. Ullmann, B.D.; Myers, H.; Chiranand, W.; Lazzell, A.L.; Zhao, Q.; Vega, L.A.; Lopez-Ribot, J.L.; Gardner, P.R.; Gustin, M.C. Inducible defense mechanism against nitric oxide in Candida albicans. Eukaryot. Cell. 2004, 3, 715–723. [Google Scholar] [CrossRef]
Figure 1. Overview on NO and reactive nitrogen species (RNS) production. NO can be produced from the oxidation of L-arginine in the presence of O2, NADPH, and various co-factors. This reaction is catalyzed by the constitutive NOS-I (nNOS) and NOS-III (eNOS) enzymes, as well as by the inducible NOS-II (iNOS). Once produced, NO can interact with O2, metals, nucleic acids, and proteins as well as with O2●−, generating peroxynitrite (ONOO). ONOO reacts with CO2 to form 1-carboxylato-2-nitrosodioxidane (ONOOC[O]O), which decays by homolysis of the O–O bond to yield the reactive species nitrogen dioxide (NO2) and trioxocarbonate (CO3●−).
Figure 1. Overview on NO and reactive nitrogen species (RNS) production. NO can be produced from the oxidation of L-arginine in the presence of O2, NADPH, and various co-factors. This reaction is catalyzed by the constitutive NOS-I (nNOS) and NOS-III (eNOS) enzymes, as well as by the inducible NOS-II (iNOS). Once produced, NO can interact with O2, metals, nucleic acids, and proteins as well as with O2●−, generating peroxynitrite (ONOO). ONOO reacts with CO2 to form 1-carboxylato-2-nitrosodioxidane (ONOOC[O]O), which decays by homolysis of the O–O bond to yield the reactive species nitrogen dioxide (NO2) and trioxocarbonate (CO3●−).
Antioxidants 11 02176 g001
Figure 3. Role of NO during the viral infection. Constitutive NO exerts an antiviral activity controlling infection whereas the inducible and exogenous NO contributes to inflammation and pro-oxidants effects (e.g., tissue damage and cell death).
Figure 3. Role of NO during the viral infection. Constitutive NO exerts an antiviral activity controlling infection whereas the inducible and exogenous NO contributes to inflammation and pro-oxidants effects (e.g., tissue damage and cell death).
Antioxidants 11 02176 g003
Figure 4. NO detoxification mediated by heme-proteins and dimeric Hmp. Heme-proteins include bacterial truncated hemoglobin, myoglobin, neuroglobin, and hemocyanin.
Figure 4. NO detoxification mediated by heme-proteins and dimeric Hmp. Heme-proteins include bacterial truncated hemoglobin, myoglobin, neuroglobin, and hemocyanin.
Antioxidants 11 02176 g004
Figure 5. NO detoxification mediated by A. lumbricoides Hb.
Figure 5. NO detoxification mediated by A. lumbricoides Hb.
Antioxidants 11 02176 g005
Figure 6. Overview of pathogen strategies to escape from NO-based host defense.
Figure 6. Overview of pathogen strategies to escape from NO-based host defense.
Antioxidants 11 02176 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

De Simone, G.; di Masi, A.; Ascenzi, P. Strategies of Pathogens to Escape from NO-Based Host Defense. Antioxidants 2022, 11, 2176. https://doi.org/10.3390/antiox11112176

AMA Style

De Simone G, di Masi A, Ascenzi P. Strategies of Pathogens to Escape from NO-Based Host Defense. Antioxidants. 2022; 11(11):2176. https://doi.org/10.3390/antiox11112176

Chicago/Turabian Style

De Simone, Giovanna, Alessandra di Masi, and Paolo Ascenzi. 2022. "Strategies of Pathogens to Escape from NO-Based Host Defense" Antioxidants 11, no. 11: 2176. https://doi.org/10.3390/antiox11112176

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop