Open access peer-reviewed chapter

Role of Nitric Oxide Synthase in the Function of the Central Nervous System under Normal and Infectious Conditions

Written By

Patricia Alves Reis, Cassiano Felippe Gonçalves de Albuquerque, Tatiana Maron‐Gutierrez, Adriana Ribeiro Silva and Hugo Caire de Castro Faria Neto

Submitted: 19 May 2016 Reviewed: 10 February 2017 Published: 17 May 2017

DOI: 10.5772/67816

From the Edited Volume

Nitric Oxide Synthase - Simple Enzyme-Complex Roles

Edited by Seyed Soheil Saeedi Saravi

Chapter metrics overview

2,160 Chapter Downloads

View Full Metrics

Abstract

Nitric oxide (NO) was discovered as an endothelium‐derived relaxing factor more than two decades ago. Since then, it has been shown to participate in many pathways. NO has been described as a key mediator of different pathways in the central nervous system (CNS) in both healthy and diseased processes. The three isoforms of nitric oxide synthase differ in their activity patterns and expression in different cells. Neuronal nitric oxide synthase (nNOS) is localized in synaptic spines, astrocytes, and the loose connective tissue surrounding blood vessels in the brain; eNOS is present in both cerebral vascular endothelial cells and motor neurons; and iNOS is induced in astrocytes and microglia under pathological conditions. During physiological processes, NO produced by eNOS/nNOS, respectively, controls blood flow activation, and act as a messenger during long‐term potentiation (LTP). However, under pathological conditions, eNOS appears to be impaired, leading to a reduction in blood flow and, consequently, low oxygen/metabolites delivery, efflux of toxicological agents from the brain tissue and disturbance in the blood‐brain barrier. The NO produced by iNOS in glial cells and nNOS, which triggers the NMDA‐excitotoxic pathway, combines with superoxide anion and results in peroxynitrite synthesis, a potent free radical that contributes to tissue damage in the brain. Here, we intend to show the controversial role of the nitric oxide delivered by the three isoforms of the nitric oxide synthase in the CNS, assess its impact under healthy/pathological conditions and speculate on its possible sequela, particularly in long‐term cognitive decline.

Keywords

  • nitric oxide
  • nitric oxide synthase
  • central nervous system
  • memory
  • infection

1. Introduction

Three isoforms of nitric oxide synthase (NOS) have been identified: two constitutive enzymes, neuronal NOS (nNOS) and endothelial NOS (eNOS), and one inducible enzyme (iNOS). These three isoforms of the enzyme nitric oxide synthase (NOS) that are present in the central nervous system (CNS) can produce nitric oxide (NO). eNOS is expressed in the vascular endothelium and choroid plexus; neuronal NOS is mainly expressed in neuronal cell bodies, especially in the cortex, hippocampus, hypothalamus, olfactory bulb, claustrum, amygdala, and thalamus; and inducible NOS is expressed in macrophages, glial cells, infiltrating neutrophils, and, to some extent, neurons [1]. It has also been reported that eNOS can be found in a subset of neurons and astrocytes and that nNOS can be found at low levels in astrocytes [2]. Because these enzymes may have different sites of expression and activation, they have a pivotal role in the divergent functions of NO [3].

Several studies have demonstrated that NO, a freely diffusible gaseous compound, has an important role in a variety of neurobiological processes [4]. Numerous functions of this regulatory molecule have been identified in the CNS, in the process of endothelium‐dependent vasodilatation [58], in neurotransmission [9, 10], and in host‐defense mechanisms [11, 12].

NO is produced from the oxidation of the terminal guanidine nitrogen of the amino acid arginine. This reaction is catalyzed by the NADPH‐dependent enzyme, nitric oxide synthase (NOS). After its formation, NO diffuses outside the cell [13]. NO derived from eNOS maintains the CNS microcirculation [14] by inhibiting platelet aggregation and leukocyte adhesion and migration [15]. NO derived from nNOS is an important neurotransmitter related to neuronal plasticity, memory formation, regulation of CNS blood flow, and neurotransmitter release [16, 17].

Advertisement

2. Implications of nitric oxide synthase in the physiological central nervous system

Under physiological conditions, the concentration of NO fluctuates within the range of low values [18] and is produced mainly by nNOS and eNOS. Unlike the other two enzymes, iNOS is not expressed unless it is induced by inflammatory mediators, cytokines, and other agents, such as endotoxins [19]. Due to its calcium‐independent activation, iNOS can produce a large amount (100–1000 times greater) of NO in relation to eNOS and nNOS [20]. Until the enzyme is degraded, iNOS constitutively produces NO [21].

NO binds to guanylyl‐cyclase, which is a soluble NO receptor and, through cGMP‐mediated signaling, acts either as a post‐ or a presynaptic messenger [22]. As a neurotransmitter, NO may activate the cGMP‐dependent protein kinase G (PKG) pathway that phosphorylates synaptophysin, which is critical for fusion of presynaptic vesicles, thereby potentiating and facilitating neurotransmission [22] (Figure 1). NO also acts on inhibitory gamma‐aminobutyric acid (GABA)‐ergic synaptic transmission via cGMP‐dependent pathways as well as on ion channels and exchangers [9].

Figure 1.

NO act as an unconventional neurotransmitter that is not stored in synaptic vesicles and not released upon membrane depolarization; it releases as soon it is synthesized and does not bind to any receptors, but diffuses from one neuron to another. NO stimulate soluble guanylyl‐cyclase to form the second messenger molecule, cyclic guanosine monophosphate (cGMP) either as a post‐ or a presynaptic messenger. PKG, protein kinase G; ERK, extracellular signal‐regulated kinases; LTP, long‐term potentiation. Images: https://mindthegraph.com.

The brain relies on a constant and adequate supply of oxygen and glucose that is provided by blood. Cerebral blood flow is altered in response to both neural activity and humoral stimuli (e.g., arterial PO2 and PCO2). Thus, augmented neural activation results in locally increased cerebral blood flow via functional hyperemia, whereas humoral stimuli produce overall increase in cerebral blood flow [23]. The physiological production of endothelium‐derived NO by eNOS is protective in hypoxic or ischemic brain injury.

All NOS isoforms have phosphorylation sites for different protein kinases, including protein kinase A, B, and C, and calcium‐calmodulin kinase [24]. NOS enzymes are very important for the maintenance of physiological mechanisms within an organism, and the genetic ablation of NOS in mice has been informative for establishing the functional roles of NOS‐generated NO in different systems [25]. For instance, nNOS‐KO mice present with intense gastroparesis due to impaired vagal innervation of stomach muscle cells [26], decreased apoptosis induced by striatal N‐methyl‐D‐aspartate (NMDA) microinjections [27], and early dysfunction of hippocampal‐dependent spatial memory [28]. Additionally, disrupting the gene that encodes eNOS in mice‐induced spontaneous systemic and pulmonary hypertension [29] and inhibited growth factor‐mediated angiogenesis [30]. Thus, NO acts through numerous mechanisms of different physiological systems and in living cells.

Shortly after its identification, NO emerged as a possible mediator of neurovascular coupling. Neurovascular coupling is an active mechanism with vessel diameter alterations in response to increased metabolic demands from neuronal activity. Under these conditions, NO acts as a potent vasodilator that is released during enhanced neuronal activity and is well suited to mediate the coupling between neuronal activity and cerebral blood flow [31].

The importance of NO as an intermediary in cell communication in the brain is highlighted by the fact that the excitatory amino acid glutamate is an initiator of the reaction that forms NO. NO can act as a “double‐edged sword”. Whereas NO supports vascular homeostasis in the endothelium‐dependent vasodilatation, its over‐ or underproduction is linked to pathological conditions [4].

Advertisement

3. Implications of nitric oxide synthase in the pathological central nervous system conditions

NO is also an important mediator under pathological conditions. For instance, in brain ischemia‐reperfusion injury, NO formation is initially increased and has a protective function by inducing collateral perfusion as a result of its powerful stimulatory effect on vasodilatation and angiogenesis [32]. NO donors induce neuroprotective effects.

NO can exist in distinct oxidation/reduction states and present dual biological actions as either a neuroprotective or a neurotoxic molecule [4]. Under physiological conditions, nNOS produces hydrogen peroxide (H2O2) and superoxide (O2•−) in addition to NO [33].

The downstream cascade in the breakdown of the BBB appears to be mediated by eNOS activity; the systemic administration of a selective eNOS inhibitor abrogates VEGF‐A‐induced BBB disruption and protects against neurologic damage in models of inflammatory disease [34, 35]. Nevertheless, it was suggested that eNOS‐derived NO is a neuromodulator that participates in the BBB‐mediated control of the cerebellum in experimental models [36].

Curiously, perivascular macrophage‐derived iNOS‐generated NO, that strategically localizes to leukocytes at brain penetration sites, can serve as a negative feedback regulator that prevents the unlimited influx of inflammatory cells by restoring BBB integrity [37].

In the brain, NOS regulates cerebral blood flow and neurotransmitter release, and the proper operative eNOS/NO system accounts for the microenvironment homeostasis that is essential for the normal functioning of neurons and glial cells [38]. Therefore, the NO produced by NOS results in vasodilation and controls vascular resistance, platelet adhesion and aggregation, leukocyte‐to‐endothelium interaction, and the maintenance of BBB integrity [36].

Several reports have indicated significant nervous system morbidity due to viral, bacterial, fungal, and parasitic infections (review [39]). During a systemic response to infections, cytokines, chemokines, and damage‐associated soluble mediators of systemic inflammation can gain access to the CNS via blood flow [40]. These mediators can access the brain tissue after the disruption of the BBB. The presence of proinflammatory mediators leads to a disturbance of neuronal and glial homeostasis, with subsequent cognitive and behavioral manifestations that are common during acute infections (anorexia, malaise, depression, and decreased physical activity) and are collectively known as sickness behavior [40]. Although sickness behavior manifestations are transient and self‐limited, the cognitive and behavioral changes can become permanent or long‐lasting under a persistent systemic inflammatory response. For example, cognitive decline is a common consequence in sepsis and cerebral malaria survivors, as found in both clinical and experimental approaches.

Under healthy conditions, the main cell types that are present in the brain are neurons, oligodendrocytes, astrocytes, and microglia. Neurons connect to each other through long axonal processes with synapses to transmit electrical/chemical signals, thereby generating memory and emotions that are associated with learning, and to control organ and systemic functions. Oligodendrocytes support axons with myelin sheaths. Astrocytes interact with blood vessels to form the BBB and maintain neuronal synapses. Microglia form long processes to phagocyte apoptotic cells and prune inactive synapses without inducing inflammation while maintaining a type of surveillance of neurons. Systemic inflammatory conditions result in the disruption of the BBB and the efflux of proinflammatory cytokines/chemokines as well as pathogen‐associated molecular patterns (PAMPs) and danger‐associated molecular patterns (DAMPS), which together activate glial cells. This results in an inflammatory environment due to the release of cytokines/chemokines and reactive oxygen/nitrogen species that exert direct and indirect neuronal cytotoxic effects [40, 41]. Oligodendroglial myelin sheaths can be affected, thereby leading to axonal degeneration. Astrocytosis leads to reduced BBB and synaptic maintenance. Microgliosis results in a proinflammatory microglial phenotype with reduced tissue maintenance functions [41, 42]. Neuronal cells can develop an excitotoxicity process due to excessive glutamate in the synaptic cleft and subsequent extra‐synaptic NMDA receptor activation that results in a subsequent increase of Ca2+ efflux in neuronal cells and the activation of proteins calpain 1 and neuronal nitric oxide. This results in mitochondrial dysfunction and oxidative damage by reactive oxygen and nitrogen species [43]. Together, these mechanisms lead to neuronal death, thereby contributing to long‐term cognitive decline, which has been shown to be a consequence of several infectious diseases.

As shown in Figure 2, the gaseous signaling molecule NO has a variety of cellular functions, including neurotransmission, regulation of blood‐vessel tone, and immunity. Under pathological conditions, free radicals may deplete NO produced by eNOS through the formation of ONOO, thus decreasing the vascular bioavailability of NO, which results in BBB dysfunction. This ultimately results in endothelial damage, edema development, and hypoxia. Furthermore, the NO produced by iNOS in glial cells or by nNOS under excitotoxic process can form with free radicals (particularly O2) ONOO and produce several deleterious effects on tissue, such as through tyrosine nitration and cysteine oxidation in various proteins. These free radicals can further decompose into highly toxic‐free radicals, such as NO2 and OH (as reviewed by [44]).

Figure 2.

Different steps in the NO signaling cascade under physiological/pathological conditions in the brain. During long‐term potentiation, NOS1 or neuronal NOS (nNOS) catalyze the NO synthesis after the activation of the NMDA receptor by Ca2+. Under excitotoxic conditions, excessive Ca2+ leads to nNOS hyperactivity, whereas excessive NO production can combine with superoxide to form peroxynitrite, which is responsible for tissue damage due to several biological effects, including blockage of the eNOS pathway and BBB impairment. NO is synthesized following the transcriptional expression of a Ca2+‐independent NOS2 or iNOS isoform in glial cells (astrocytes and microglia) after cytokine exposure, thereby contributing to neuroinflammation and tissue damage in the brain. Intracellular Ca2+ activates NOS3 or eNOS to release NO from brain microvessels. This NO binds to soluble guanylyl‐cyclase (sGC) receptors, which triggers a cGMP‐dependent pathway and interacts with its downstream mediators of the physiological regulation of vasodilation and vascular resistance, platelet adhesion and aggregation, leukocyte‐endothelial interaction, and BBB integrity maintenance. Images: https://mindthegraph.com.

Nitric oxide, as described above, is a key molecule in the regulation of physiological brain homeostasis. Nitric oxide is synthetized by neuronal vessels (eNOS), by neurons (nNOS) under physiological/pathological stimuli, and by iNOS during inflammatory conditions. It is well known that high levels of nitric oxide (NO) released by the inducible NO synthase (iNOS) are critical for defense against parasites and mediate inflammatory tissue damage. However, the suppression or lack of NO production results in the impaired clearance of some types of bacteria by the host [45].

Our group and others have shown that eNOS is impaired during systemic infection, which leads to brain microcirculation dysfunction [4648]. We showed that during different infectious diseases, there is a functional capillary rarefaction (Figure 3) that can be caused by the impairment of endothelial NO production by a reduction of eNOS activity or by reduced levels of the enzyme cofactor tetrahydrobiopterin—BH4 [49].

Figure 3.

Photomicrography from intravital microscopy showing capillary impairment (red arrows) under healthy (A) and infectious conditions (B) malaria, (C) sepsis, and (D) Chagas’ disease [47, 48].

Vascular function can be evaluated by the vasodilator response of cerebral arterioles to acetylcholine (Ach). Our group used intravital microscopy to evaluate the response to Ach in sepsis and Chagas’ disease model. The vasodilation‐to‐Ach test is directly related to the availability of endothelium‐derived NO generated from l‐arginine by the action of endothelial NO synthase (eNOS). The diffusion of NO to vascular smooth muscle cells and the activation of guanylyl‐cyclase resulted in cGMP‐mediated vasodilation. However, the vasoconstrictive response to Ach results from a direct muscarinic smooth muscle effect [50]. This vasoconstrictive effect can lead to the slow delivery of oxygen to brain tissue, generating hypoxia and increased glycose consume by glycolytic pathway to generate ATP for neuronal functioning. Consequently, mitochondrial function is compromised by low O2 concentration [51].

In both sepsis and Chagas’ disease models, we observed a vasoconstrictive response to Ach. One mechanism of eNOS “uncoupling” from the reduction of NO production and ROS generation involves the oxidative degradation of the cofactor BH4 (tetrahydrobiopterin), especially by peroxynitrite (ONOO), which is produced in association with superoxide and nitric oxide. Several studies have shown that there is an increase in impaired eNOS‐derived oxidative stress in brain tissue exposed to systemic infection, NO consumption and capillary dysfunction [46, 47, 52, 53].

Interestingly, in experimental models of malaria [48] and sepsis (Reis et al., submitted) blood flow was recovered by treatment with statins. As reviewed by Giannopoulos et al. [54], statins are shown to enhance eNOS expression and can contribute to the restoration of brain capillary function.

Resident glial cells in the CNS (i.e., astroglia and microglia) express inducible nitric oxide synthase (iNOS) and produce high levels of NO in response to a wide variety of proinflammatory and degenerative stimuli. Excessive NO production by glial cells evoked by inflammatory signals contributes to the pathogenesis of several neurodegenerative diseases, such as multiple sclerosis, HIV dementia, brain ischemia, trauma, Parkinson's disease, and Alzheimer's disease. NO can also be released in glial cells in response to infectious agent. The inducible form of nitric oxide synthase can be strongly associated with tissue damage, despite its activity as an antimicrobial agent. As reviewed by Saha and Pahan [55], astrocytes and microglia can express iNOS and synthetize NO. According to Radi [56], nitric oxide can combine with superoxide (produced mainly by the NADPH oxidase system and partially by mitochondria) under conditions of oxidative stress. It is common under neuroinflammatory conditions to form peroxynitrite, which binds to tyrosine to produce 3‐nitrotyrosine, and a marker of reactive nitrogen species production. The reactions of peroxynitrite with biomolecules can lead to cytotoxic events and may result in apoptotic or necrotic cell death. Peroxynitrite can act via antioxidant enzyme inhibition; antioxidant depletion; protein aggregation (e.g., α‐synuclein and microtubule‐associated tau protein modifications in the CNS); activation of specific enzymes (e.g., matrix metalloproteinases (MMPs), cytochrome c, glutathione‐S‐transferase, protein kinase C‐ε (PKCε), and fibrinogen); impairment of enzyme cofactors (e.g., BH4); modification of mediator pathways, receptors, and cellular signaling molecules; calcium deregulation; DNA injury; and mitochondrial dysfunction, among other processes (reviewed by [57]).

Data from our group show that the absence of the iNOS enzyme (knockout) or treatment with iNOS‐specific inhibitors (e.g., aminoguanidine) has a beneficial effect on the prevention of cognitive impairment (unpublished data) in experimental cerebral malaria. This could be due to the reduced production of the peroxynitrite radical and the subsequent reduction of tissue damage. Other studies are being conducted to clarify the impact of the synthesis of nitric oxide by iNOS in infectious diseases and in the development of cognitive decline. Weberpals et al. also observed that iNOS gene deficiency prevents cognitive decline in addition to promoting a reduction in gliosis (astrocytes and microglia), proinflammatory cytokines TNF‐α and IL‐1β release, and reduction of synaptic dysfunction in sepsis model [58].

Advertisement

4. NMDA‐nNOS pathway and excitotoxicity development

The NMDA receptor is a key regulator during glutamatergic long‐term potentiation (LTP) response. During physiological conditions, glutamatergic ionotropic (e.g., kainate, NMDA, and AMPA) and metabotropic receptors (mGlut) are activated, delivering Ca2+ into neuronal cells and resulting in depolarization and the triggering of mitogen‐activated protein kinase (MEK)/extracellular signal‐regulated kinase (ERK) and phosphatidylinositol 3‐kinase (PI3K) signaling pathways. This then activates phosphoinositide‐dependent kinase 1 or 2 (PDK1/2), Akt, and mammalian target of rapamycin (mTOR) downstream signaling [59]. Additionally, calcium‐dependent kinase II (CAMKII) is rapidly activated and all events together lead to the activation of transcription factors (e.g., CREB) and an increase in the synthesis of neurotrophic factors (e.g., BDNF) and proteins associated with the long‐term potentiation (LTP) process.

Despite the major activity of NMDA receptors in LTP, several reports have related this receptor with excitotoxicity, which is a neurodegenerative process. This process has been associated with a different class of NMDA receptor: the extra‐synaptic receptor. As reviewed by Parsons and Raymond, synaptic NMDA (which expresses the subunit GlutN2A) is associated with LTP, whereas extra‐synaptic receptors (which express the subunit GlutN2B) is associated with excitotoxicity and cell death [42]. NMDARs recruit the calcium‐dependent enzyme nNOS via PSD95 (postsynaptic density: membrane‐associated guanylate kinase (MAGUK) scaffolding protein located in neural postsynaptic densities), which is also associated with the LTP process. This is considered a key contributor to excitotoxicity lesions in both stroke and neurodegenerative diseases [60]. nNOS is activated by calcium/calmodulin signaling and is PSD95 protein‐dependent [60]. During excitotoxicity, the activation of NMDA enhances intracellular calcium, leading to nNOS activation and NO production. Additionally, excessive intracellular calcium activates calpain 1, which then disrupts mitochondrial function, thereby triggering the intrinsic apoptotic pathway by releasing cytochrome C and activating apoptosomal protein complex [61]. NO can combine with superoxide, which results in peroxynitrite formation and cellular damage. Peroxynitrite can disturb cellular function by the nitration of proteins, which reduces or eliminates protein function, as described earlier, and by DNA damage via the activation of poly (ADP‐ribose) polymerase 1 (PARP‐1). The impact of PARP‐1 on intracellular concentration of its nicotinamideadenine dinucleotide substrate (NAD), creates a bioenergetic imbalance that culminates with ATP depletion, thereby triggering necrotic neuronal death [62, 63]. In addition, NO can drive the retraction of the synaptic button via the activation of small GTPase RhoA/ROCK signaling through a paracrine/retrograde‐signaling pathway [64]. Taken together, these events could contribute to neuronal dysfunction and death associated with cognitive decline.

The roles of NO in neuronal damage following insults, such as hypoxia, traumatic brain injury, and ischemia, have been well established. Recent evidence has implicated an imbalance of ROS and NO signaling in neurodegenerative disorders, such as Alzheimer's disease and Parkinson's disease, and in cognitive impairments associated with normal physiological aging [6567]. Whereas mild oxidative/nitrosative (nitric oxide (NO)‐related) stress mediates normal neuronal signaling, the accumulation of free radicals is associated with neuronal cell injury or death.

As described above, NO from eNOS modulates blood flow in the brain, and its impairment could be associated with hypoxic events in the brain. Several degenerative and infectious diseases relate hypoxia to neuronal dysfunction and cognitive decline. Using partial eNOS knockout mice model, Tan et al. [68] showed that the development of spontaneous thrombotic cerebral infarction is followed by amyloid protein deposit and cognitive decline. Cognitive decline is a major cause of disability in stroke survivors [69]. We have shown that microvascular impairment in malaria, sepsis, and Chagas’ diseases may occur in response to other infectious agents [4648]. Cerebral metabolism is dependent upon the glucose and oxygen that are delivered by blood, and it is clear that alterations in endothelial function can disrupt neuronal functions. Additionally, activation of endothelial cells by systemic cytokines, PAMP/DAMP and prooxidant molecules can contribute to eNOS dysfunction [70], blood flow disturbance and neuronal dysfunction, which subsequently results in cognitive decline.

The activation of glial cells that may be related with systemic inflammation associated to host response to pathogens can activate inducible isoforms of iNOS and subsequently generate cellular damage via the generation of peroxynitrite by the combination of NO and superoxide radicals. In this way, the induction of iNOS may result in the development of cognitive impairment [71]. Experimental models of sepsis [58] and malaria (unpublished data) have shown that the inhibition of iNOS has a beneficial effect on the central nervous system, particularly by abolishing cognitive decline.

The main enzyme target in the central nervous system seems to be nNOS. The NO generated by nNOS controls the release of neurotransmitters and is involved in synaptogenesis, synaptic plasticity, memory function, and neuroendocrine secretion. However, the overproduction of NO during NMDA‐excitotoxic events can lead to neuronal death and directly impact the cognitive function.

Death pathways are activated in mouse brains during experimental model of sepsis and malaria [72, 73]. Additionally, reduced levels of neurotrophic factors, impairment of neurogenesis, and synaptic dysfunction were observed [7478]. However, the role of excitotoxicity and nNOS delivery during infectious diseases and long‐term cognitive impairment is not yet clear. Neuroinflammation can also induce cell damage/death, and the modulation of the activation of glial cells has been suggested to prevent neuronal damage and cognitive decline. Cognitive impairment is prevented by antioxidants and statin treatment [48, 53, 79], which suggests that the control of oxidative or inflammatory damage is also efficient to avoid cognitive decline.

Advertisement

5. Conclusion

NO is a key molecule involved in the regulation of CNS function in health and disease (Table 1). The impairment of enzymatic activity or the overproduction of NO under inflammatory/excitotoxic conditions can contribute to neurological sequela during systemic‐infectious diseases (Figure 3). The NO synthase complex can be considered a target of pharmacological intervention focusing on the prevention of cognitive sequela; however, this field requires further studies.

NOS isoformPhysiological effect on CNSPathological effect on CNS
nNOSLearning and memory↑—Neuronal death by excitotoxicity
eNOSVasodilation and increased blood flow↓—Hypoxia
iNOSSleep [33]↑—Tissue damage

Table 1.

Nitric oxide synthase enzymes functions on health/pathological conditions.

References

  1. 1. Gunther M, Al Nimer F, Gahm C, Piehl F, Mathiesen T. iNOS‐mediated secondary inflammatory response differs between rat strains following experimental brain contusion. Acta Neurochirurgica (Wien). 2012;154:689–697
  2. 2. Brown GC. Nitric oxide and neuronal death. Nitric Oxide. 2010;23:153–165
  3. 3. Garry PS, Ezra M, Rowland MJ, Westbrook J, Pattinson KT. The role of the nitric oxide pathway in brain injury and its treatment—from bench to bedside. Experimental Neurology. 2015;263:235–243
  4. 4. Dzoljic E, Grbatinic I, Kostic V. Why is nitric oxide important for our brain? Functional Neurology. 2015;30:159–163
  5. 5. Duncan AJ, Heales SJ. Nitric oxide and neurological disorders. Molecular Aspects of Medicine. 2005;26:67–96
  6. 6. Michell AW, Lewis SJ, Foltynie T, Barker RA. Biomarkers and Parkinson’s disease. Brain. 2004;127:1693–1705
  7. 7. Stojanovic R, Todorovic Z, Nesic Z, Vuckovic S, Cerovac‐Cosic N, et al. NG‐nitro‐L‐arginine methyl ester‐induced potentiaton of the effect of aminophylline on rat diaphragm: The role of extracellular calcium. Journal of Pharmacological Sciences. 2004;96:493–498
  8. 8. Stojanovic R, Todorovic Z, Vuckovic S, Nesic Z, Prostran M. NG‐nitro‐L‐arginine methyl ester potentiates the effect of aminophylline on the isolated rat hemidiaphragm. Journal of Pharmacological Sciences. 2003;92:157–162
  9. 9. Yamamoto K, Takei H, Koyanagi Y, Koshikawa N, Kobayashi M. Presynaptic cell type‐dependent regulation of GABAergic synaptic transmission by nitric oxide in rat insular cortex. Neuroscience. 2015;284:65–77
  10. 10. Yassin L, Radtke‐Schuller S, Asraf H, Grothe B, Hershfinkel M, et al. Nitric oxide signaling modulates synaptic inhibition in the superior paraolivary nucleus (SPN) via cGMP‐dependent suppression of KCC2. Frontiers in Neural Circuits. 2014;8:65
  11. 11. Akyol O, Zoroglu SS, Armutcu F, Sahin S, Gurel A. Nitric oxide as a physiopathological factor in neuropsychiatric disorders. In Vivo. 2004;18:377–390
  12. 12. Boje KM. Nitric oxide neurotoxicity in neurodegenerative diseases. Frontiers in Bioscience. 2004;9:763–776
  13. 13. Virarkar M, Alappat L, Bradford PG, Awad AB. L‐Arginine and nitric oxide in CNS function and neurodegenerative diseases. Critical Reviews in Food Science and Nutrition. 2013;53:1157–1167
  14. 14. Toda N, Ayajiki K, Okamura T. Cerebral blood flow regulation by nitric oxide in neurological disorders. Canadian Journal of Physiology and Pharmacology. 2009;87:581–594
  15. 15. Broos K, Feys HB, De Meyer SF, Vanhoorelbeke K, Deckmyn H. Platelets at work in primary hemostasis. Blood Reviews. 2011;25:155–167
  16. 16. Toda N, Okamura T. The pharmacology of nitric oxide in the peripheral nervous system of blood vessels. Pharmacological Reviews. 2003;55:271–324
  17. 17. Schuman EM, Madison DV. Nitric oxide and synaptic function. Annual Review of Neuroscience. 1994;17:153–183
  18. 18. Tieu K, Ischiropoulos H, Przedborski S. Nitric oxide and reactive oxygen species in Parkinson’s disease. International Union of Biochemistry and Molecular Biology Life. 2003;55:329–335
  19. 19. Zhu H, Vishwamitra D, Curry CV, Manshouri R, Diao L, et al. NPM‐ALK up‐regulates iNOS expression through a STAT3/microRNA‐26a‐dependent mechanism. Journal of Pathology. 2013;230:82–94
  20. 20. Pautz A, Art J, Hahn S, Nowag S, Voss C, et al. Regulation of the expression of inducible nitric oxide synthase. Nitric Oxide. 2010;23:75–93
  21. 21. MacMicking JD, Willenborg DO, Weidemann MJ, Rockett KA, Cowden WB. Elevated secretion of reactive nitrogen and oxygen intermediates by inflammatory leukocytes in hyperacute experimental autoimmune encephalomyelitis: Enhancement by the soluble products of encephalitogenic T cells. The Journal of Experimental Medicine. 1992;176:303–307
  22. 22. Wang HG, Lu FM, Jin I, Udo H, Kandel ER, et al. Presynaptic and postsynaptic roles of NO, cGK, and RhoA in long‐lasting potentiation and aggregation of synaptic proteins. Neuron. 2005;45:389–403
  23. 23. Steinback CD, Poulin MJ. Influence of hypoxia on cerebral blood flow regulation in humans. Advances in Experimental Medicine and Biology. 2016;903:131–144
  24. 24. Boehning D, Snyder SH. Novel neural modulators. Annual Review of Neuroscience. 2003;26:105–131
  25. 25. Cossenza M, Socodato R, Portugal CC, Domith IC, Gladulich LF, et al. Nitric oxide in the nervous system: Biochemical, developmental, and neurobiological aspects. Vitamins and Hormones. 2014;96:79–125
  26. 26. Mashimo H, Kjellin A, Goyal RK. Gastric stasis in neuronal nitric oxide synthase‐deficient knockout mice. Gastroenterology. 2000;119:766–773
  27. 27. Ayata C, Ayata G, Hara H, Matthews RT, Beal MF, et al. Mechanisms of reduced striatal NMDA excitotoxicity in type I nitric oxide synthase knock‐out mice. Journal of Neuroscience. 1997;17:6908–6917
  28. 28. Kirchner L, Weitzdoerfer R, Hoeger H, Url A, Schmidt P, et al. Impaired cognitive performance in neuronal nitric oxide synthase knockout mice is associated with hippocampal protein derangements. Nitric Oxide. 2004;11:316–330
  29. 29. Huang PL, Huang Z, Mashimo H, Bloch KD, Moskowitz MA, et al. Hypertension in mice lacking the gene for endothelial nitric oxide synthase. Nature. 1995;377:239–242
  30. 30. Lee PC, Salyapongse AN, Bragdon GA, Shears LL, 2nd, Watkins SC, et al. Impaired wound healing and angiogenesis in eNOS‐deficient mice. The American Journal of Physiology. 1999;277:H1600–1608
  31. 31. Girouard H, Iadecola C. Neurovascular coupling in the normal brain and in hypertension, stroke, and Alzheimer disease. Journal of Applied Physiology (1985). 2006;100:328–335
  32. 32. Su KH, Lin SJ, Wei J, Lee KI, Zhao JF, et al. The essential role of transient receptor potential vanilloid 1 in simvastatin‐induced activation of endothelial nitric oxide synthase and angiogenesis. Acta Physiologica (Oxf). 2014;212:191–204
  33. 33. Costa ED, Rezende BA, Cortes SF, Lemos VS. Neuronal nitric oxide synthase in vascular physiology and diseases. Frontiers in Physiology. 2016;7:206
  34. 34. Argaw AT, Asp L, Zhang J, Navrazhina K, Pham T, et al. Astrocyte‐derived VEGF‐A drives blood‐brain barrier disruption in CNS inflammatory disease. The Journal of Clinical Investigation. 2012;122:2454–2468
  35. 35. Park HJ, Shin JY, Kim HN, Oh SH, Song SK, et al. Mesenchymal stem cells stabilize the blood‐brain barrier through regulation of astrocytes. Stem Cell Research & Therapy. 2015;6:187
  36. 36. Soares ES, Mendonca MC, da Cruz‐Hofling MA. eNOS uncoupling in the cerebellum after BBB disruption by exposure to phoneutria nigriventer spider venom. Toxicon. 2015;104:7–13
  37. 37. Olivera GC, Ren X, Vodnala SK, Lu J, Coppo L, et al. Nitric oxide protects against infection‐induced neuroinflammation by preserving the stability of the blood‐brain barrier. PLoS Pathogens. 2016;12:e1005442
  38. 38. Northington FJ, Tobin JR, Harris AP, Traystman RJ, Koehler RC. Developmental and regional differences in nitric oxide synthase activity and blood flow in the sheep brain. Journal of Cerebral Blood Flow and Metabolism. 1997;17:109–115
  39. 39. John CC, Carabin H, Montano SM, Bangirana P, Zunt JR, et al. Global research priorities for infections that affect the nervous system. Nature. 2015;527:S178–186
  40. 40. Sankowski R, Mader S, Valdes‐Ferrer SI. Systemic inflammation and the brain: Novel roles of genetic, molecular, and environmental cues as drivers of neurodegeneration. Frontiers in Cellular Neuroscience. 2015;9:28
  41. 41. Kopitar‐Jerala N. Innate immune response in brain, NF‐Kappa B signaling and cystatins. Frontiers in Molecular Neuroscience. 2015;8:73
  42. 42. Parsons MP, Raymond LA. Extrasynaptic NMDA receptor involvement in central nervous system disorders. Neuron. 2014;82:279–293
  43. 43. Fujikawa DG. The role of excitotoxic programmed necrosis in acute brain injury. Computational and Structural Biotechnology Journal. 2015;13:212–221
  44. 44. Berg RM, Moller K, Bailey DM. Neuro‐oxidative‐nitrosative stress in sepsis. Journal of Cerebral Blood Flow and Metabolism. 2011;31:1532–1544
  45. 45. Nathan C. Inducible nitric oxide synthase: What difference does it make? The Journal of Clinical Investigation. 1997;100:2417–2423
  46. 46. Araujo CV, Estato V, Tibirica E, Bozza PT, Castro‐Faria‐Neto HC, et al. PPAR gamma activation protects the brain against microvascular dysfunction in sepsis. Microvascular Research. 2012;84:218–221
  47. 47. Nisimura LM, Estato V, de Souza EM, Reis PA, Lessa MA, et al. Acute Chagas disease induces cerebral microvasculopathy in mice. PLoS Neglected Tropical Diseases. 2014;8:e2998
  48. 48. Reis PA, Estato V, da Silva TI, d’Avila JC, Siqueira LD, et al. Statins decrease neuroinflammation and prevent cognitive impairment after cerebral malaria. PLoS Pathogens. 2012;8:e1003099
  49. 49. Ong PK, Melchior B, Martins YC, Hofer A, Orjuela‐Sanchez P, et al. Nitric oxide synthase dysfunction contributes to impaired cerebroarteriolar reactivity in experimental cerebral malaria. PLoS Pathogens. 2013;9:e1003444
  50. 50. Pesic S, Grbovic L, Jovanovic A. Acetylcholine‐induced contractions in the perforating branch of the human internal mammary artery: Protective role of the vascular endothelium. Pharmacology. 2002;64:182–188
  51. 51. Bozza FA, D’Avila JC, Ritter C, Sonneville R, Sharshar T, et al. Bioenergetics, mitochondrial dysfunction, and oxidative stress in the pathophysiology of septic encephalopathy. Shock. 2013;39 (Suppl 1):10–16
  52. 52. Barichello T, Fortunato JJ, Vitali AM, Feier G, Reinke A, et al. Oxidative variables in the rat brain after sepsis induced by cecal ligation and perforation. Critical Care Medicine. 2006;34:886–889
  53. 53. Reis PA, Comim CM, Hermani F, Silva B, Barichello T, et al. Cognitive dysfunction is sustained after rescue therapy in experimental cerebral malaria, and is reduced by additive antioxidant therapy. PLoS Pathogens. 2010;6:e1000963
  54. 54. Giannopoulos S, Katsanos AH, Tsivgoulis G, Marshall RS. Statins and cerebral hemodynamics. Journal of Cerebral Blood Flow and Metabolism. 2012;32:1973–1976
  55. 55. Saha RN, Pahan K. Regulation of inducible nitric oxide synthase gene in glial cells. Antioxidants & Redox Signaling. 2006;8:929–947
  56. 56. Radi R. Peroxynitrite, a stealthy biological oxidant. The Journal of Biological Chemistry. 2013;288:26464–26472
  57. 57. Szabo C, Ischiropoulos H, Radi R. Peroxynitrite: Biochemistry, pathophysiology and development of therapeutics. Nature Reviews Drug Discovery. 2007;6:662–680
  58. 58. Weberpals M, Hermes M, Hermann S, Kummer MP, Terwel D, et al. NOS2 gene deficiency protects from sepsis‐induced long‐term cognitive deficits. The Journal of Neuroscience. 2009;29:14177–14184
  59. 59. Klann E, Dever TE. Biochemical mechanisms for translational regulation in synaptic plasticity. Nature Reviews Neuroscience. 2004;5:931–942
  60. 60. Courtney MJ, Li LL, Lai YY. Mechanisms of NOS1AP action on NMDA receptor‐nNOS signaling. Frontiers in Cellular Neuroscience. 2014;8:252
  61. 61. Lipton SA. NMDA receptor activity regulates transcription of antioxidant pathways. Nature Neuroscience. 2008;11:381–382
  62. 62. Lai TW, Zhang S, Wang YT. Excitotoxicity and stroke: Identifying novel targets for neuroprotection. Progress in Neurobiology. 2014;115:157–188
  63. 63. Ha HC, Snyder SH. Poly(ADP‐ribose) polymerase is a mediator of necrotic cell death by ATP depletion. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:13978–13982
  64. 64. Sunico CR, Gonzalez‐Forero D, Dominguez G, Garcia‐Verdugo JM, Moreno‐Lopez B. Nitric oxide induces pathological synapse loss by a protein kinase G‐, Rho kinase‐dependent mechanism preceded by myosin light chain phosphorylation. The Journal of Neuroscience. 2010;30:973–984
  65. 65. Halliwell B. Reactive oxygen species and the central nervous system. Journal of Neurochemistry. 1992;59:1609–1623
  66. 66. Finkel T, Holbrook NJ. Oxidants, oxidative stress and the biology of ageing. Nature. 2000;408:239–247
  67. 67. Nakamura T, Cho DH, Lipton SA. Redox regulation of protein misfolding, mitochondrial dysfunction, synaptic damage, and cell death in neurodegenerative diseases. Experimental Neurology. 2012;238:12–21
  68. 68. Tan XL, Xue YQ, Ma T, Wang X, Li JJ, et al. Partial eNOS deficiency causes spontaneous thrombotic cerebral infarction, amyloid angiopathy and cognitive impairment. Molecular Neurodegeneration. 2015;10:24
  69. 69. Levine DA, Galecki AT, Langa KM, Unverzagt FW, Kabeto MU, et al. Trajectory of cognitive decline after incident stroke. The Journal of the American Medical Association. 2015;314:41–51
  70. 70. Weiss N, Miller F, Cazaubon S, Couraud PO. The blood‐brain barrier in brain homeostasis and neurological diseases. Biochimica et Biophysica Acta. 2009;1788:842–857
  71. 71. Eckel B, Ohl F, Bogdanski R, Kochs EF, Blobner M. Cognitive deficits after systemic induction of inducible nitric oxide synthase: A randomised trial in rats. European Journal of Anaesthesiology. 2011;28:655–663
  72. 72. Lackner P, Burger C, Pfaller K, Heussler V, Helbok R, et al. Apoptosis in experimental cerebral malaria: Spatial profile of cleaved caspase‐3 and ultrastructural alterations in different disease stages. Neuropathology and Applied Neurobiology. 2007;33:560–571
  73. 73. Comim CM, Barichello T, Grandgirard D, Dal‐Pizzol F, Quevedo J, et al. Caspase‐3 mediates in part hippocampal apoptosis in sepsis. Molecular Neurobiology. 2013;47:394–398
  74. 74. Anand SS, Babu PP. c‐Jun N terminal kinases (JNK) are activated in the brain during the pathology of experimental cerebral malaria. Neuroscience Letters. 2011;488:118–122
  75. 75. de Miranda AS, Brant F, Campos AC, Vieira LB, Rocha NP, et al. Evidence for the contribution of adult neurogenesis and hippocampal cell death in experimental cerebral malaria cognitive outcome. Neuroscience. 2015;284:920–933
  76. 76. Moraes CA, Santos G, de Sampaio e Spohr TC, D’Avila JC, Lima FR, et al. Activated microglia‐induced deficits in excitatory synapses through IL‐1beta: Implications for cognitive impairment in sepsis. Molecular Neurobiology. 2015;52:653–663
  77. 77. Comim CM, Reis PA, Frutuoso VS, Fries GR, Fraga DB, et al. Effects of experimental cerebral malaria in memory, brain‐derived neurotrophic factor and acetylcholinesterase activity [correction for acitivity] in the hippocampus of survivor mice. Neuroscience Letters. 2012;523:104–107
  78. 78. Comim CM, Silva NC, Mina F, Dominguini D, Scaini G, et al. Evaluation of NCS‐1, DARPP‐32, and neurotrophins in hippocampus and prefrontal cortex in rats submitted to sepsis. Synapse. 2014;68:474–479
  79. 79. Barichello T, Machado RA, Constantino L, Valvassori SS, Reus GZ, et al. Antioxidant treatment prevented late memory impairment in an animal model of sepsis. Critical Care Medicine. 2007;35:2186–2190

Written By

Patricia Alves Reis, Cassiano Felippe Gonçalves de Albuquerque, Tatiana Maron‐Gutierrez, Adriana Ribeiro Silva and Hugo Caire de Castro Faria Neto

Submitted: 19 May 2016 Reviewed: 10 February 2017 Published: 17 May 2017