Next Article in Journal
Oxidative Power: Tools for Assessing LPMO Activity on Cellulose
Previous Article in Journal
Recent Advances in the Molecular Effects of Biostimulants in Plants: An Overview
Previous Article in Special Issue
Endothelial Dysfunction Driven by Hypoxia—The Influence of Oxygen Deficiency on NO Bioavailability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nitric Oxide-Dependent Pathways as Critical Factors in the Consequences and Recovery after Brain Ischemic Hypoxia

by
Joanna M Wierońska
1,
Paulina Cieślik
1 and
Leszek Kalinowski
2,3,4,*
1
Maj Institute of Pharmacology, Polish Academy of Sciences, Smętna Street 12, 31-343 Kraków, Poland
2
Department of Medical Laboratory Diagnostics—Biobank Fahrenheit BBMRI.pl, Medical University of Gdansk, Debinki Street 7, 80-211 Gdansk, Poland
3
Biobanking and Biomolecular Resources Research Infrastructure Poland (BBMRI.PL), Debinki Street 7, 80-211 Gdansk, Poland
4
BioTechMed Center/Department of Mechanics of Materials and Structures, Gdansk University of Technology, Narutowicza 11/12, 80-223 Gdansk, Poland
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(8), 1097; https://doi.org/10.3390/biom11081097
Submission received: 31 May 2021 / Revised: 17 July 2021 / Accepted: 20 July 2021 / Published: 26 July 2021
(This article belongs to the Special Issue Hypoxia and Hypoxia-Inducible Factors in Human Endothelium)

Abstract

:
Brain ischemia is one of the leading causes of disability and mortality worldwide. Nitric oxide (NO), a molecule that is involved in the regulation of proper blood flow, vasodilation, neuronal and glial activity constitutes the crucial factor that contributes to the development of pathological changes after stroke. One of the early consequences of a sudden interruption in the cerebral blood flow is the massive production of reactive oxygen and nitrogen species (ROS/RNS) in neurons due to NO synthase uncoupling, which leads to neurotoxicity. Progression of apoptotic or necrotic neuronal damage activates reactive astrocytes and attracts microglia or lymphocytes to migrate to place of inflammation. Those inflammatory cells start to produce large amounts of inflammatory proteins, including pathological, inducible form of NOS (iNOS), which generates nitrosative stress that further contributes to brain tissue damage, forming vicious circle of detrimental processes in the late stage of ischemia. S-nitrosylation, hypoxia-inducible factor 1α (HIF-1α) and HIF-1α-dependent genes activated in reactive astrocytes play essential roles in this process. The review summarizes the roles of NO-dependent pathways in the early and late aftermath of stroke and treatments based on the stimulation or inhibition of particular NO synthases and the stabilization of HIF-1α activity.

1. Brain Ischemic Stroke and the Role of NO Its Pathology

Insufficient blood flow to a tissue results in hypoxia (the lack of an adequate oxygen supply at the tissue level) or anoxia (the absence of oxygen). Acute arterial thrombus formation, chronic narrowing of a supply artery and arterial vasospasm are the most critical factors contributing to the local or generalized deprivation of oxygen, which results in ischemia. In the 50-year-old and older age group, brain stroke is the second leading cause of disability [1] and, after coronary artery disease, constitutes the second most common cause of death worldwide [2].
As shown in vivo in transient forebrain ischemia model in rats, the mechanisms of neuronal death following ischemia include both apoptosis, i.e., a form of programmed cell death, highly regulated and controlled process that involves events such as cell shrinkage, nuclear fragmentation, chromatin condensation, chromosomal DNA fragmentation [3] and necrosis, a form of traumatic cell death that results from the disruption of membrane integrity, which attracts leukocytes to accumulate around the necrotic cells and to release cytokines [3], inducing collateral damage thus enabling the healing of the tissue.
Among the variety of factors that contribute to the cellular events leading to ischemic neuronal death, the fundamental factors are N-methyl-D-aspartate receptor (NMDA)-induced excitotoxicity and NO-dependent pathways which are functionally linked [4,5,6].
NO is formed by the oxidation of nitrogen and is biosynthesized endogenously from L-arginine, which is converted first to N-hydroxyl-arginine, then to L-citrulline and finally to NO in the presence of NADPH and tetrahydrobiopterin (BH4) as cofactors [7]. The main enzymatic target of NO is activation of guanylyl cyclase (GC) [8,9], which leads to cGMP production and the subsequent activation of a variety of proteins essential for a number of critical processes in the brain [9] amplifying the excitatory cell responses modulated by NMDA-dependent signaling [10].
At least three NOS isoforms are responsible for NO synthesis, including neuronal NO synthase (nNOS), which mediates the production of NO in neurons; the endothelial (eNOS) isoform, which is found on the inner surface of blood vessels. The activity of eNOS and nNOS, which are called constitutive NOS isoforms (cNOS), is dependent on Ca2+ and calmodulin complex, and the production of NO is initiated in response to elevations in intracellular Ca2+ levels triggered by both mechanical forces and substances circulating in the blood (e.g., glutamate, acetylcholine, ATP) [11,12].
In contrast to cNOS, the activity of inducible NO synthase (iNOS) is Ca2+-independent [12,13]. The enzyme is activated mostly in glial cells or leukocytes by certain cytokines, such as interferon-γ, tumor necrosis factor-α or interleukin-1β, in response to inflammation [12,13].
In the physiological state, glutamate activates the synaptic pool of NMDA receptors and is rapidly taken up from the synaptic space by presynaptic mechanisms and astrocytes. Opening of NMDA ion channels and subsequent Ca2+ influx into the neurons activates constitutive nitric oxide synthases (cNOS) to produce nitric oxide (NO) [4,5,6].
Under pathological conditions, due to the “uncoupling” of NOS activity from electron donation by NADPH and/or reduced availability of L-arginine or BH4, electrons are transferred from NADPH to flavins via the reductase domain, forming superoxide (O2), and other reactive oxygen species (ROS) [11,14]. NO reacts with O2 to form reactive nitrogen species (RNS) including the highly toxic peroxynitrite (ONOO), leading to the production of other secondary components of nitroxidative stress, such as NO2+, NO2 and OH, which initiate a cascade of redox reactions [11,14].
iNOS produces much greater amounts of NO than both eNOS and nNOS combined. Therefore, iNOS is often referred to as the “pathological” form of NO synthase, as it may promote the production of ONOO and subsequently highly reactive hydroxyl radicals [14,15].
The key role of NO in the brain is its impact on the components of the neurovascular unit. Thus, the vasodilation or vasoconstriction of blood vessels, neuronal excitability and glial cell functioning are dependent on NO signaling mediated by constitutive NOS isoforms (nNOS and eNOS) (Figure 1) [11,16].
The distribution and concentration of NO in brain tissue change after ischemic episode and differs both temporally and spatially postinjury. Time-dependent changes induced by ischemia are related in particular to the activity of individual NOS isoforms, and in general, NO produced by nNOS or iNOS plays detrimental role [14]; while eNOS is neuroprotective [17]. The expression of nNOS increases rapidly, in parallel with membrane depolarization and Ca2+ elevation in the cells, while massive production of iNOS is observed several hours after ischemic episode [14]. Such correlations were observed predominantly in the animal model of cerebral ischemia-reperfusion based on middle cerebral artery occlusion (MCAO), which is usually used and is believed to reproduce the pattern of ischemic brain damage observed in many humans ischemic stroke patients [18,19].
Based on these pre-clinical observations the early and late stage of the ischemia, alternatively called early and late neuronal post-ischemic damage, has been described in several research papers [14,20,21,22,23], and in reviews [24]. Clinical studies concerning the progression of ischemia damage and the rsole of NO can be found elsewhere [25,26,27,28,29].

1.1. Early Stage of the Ischemia (Early Neuronal Damage)

The obstruction of blood flow by the clot dramatically reduces glucose and oxygen supply in ischemic brain region and triggers ischemic cascades, which include accumulation of lactate and malfunction of ion pumps (Na+/K+-ATPase and Ca2+/H-ATPase), subsequently inducing membrane depolarization and calcium ion (Ca2+) overload. Membrane depolarization causes the release of excitotoxic amino acid glutamate and its translocation into the extracellular compartment. The reversal of the activity of glutamate transport proteins prevents glutamate reuptake which results in robust increase in glutamate level and activation of extrasynaptic NMDA receptors [4,16,30,31].
In vivo studies show, that in the MCAO model in rats, immediate early genes (c-fos, c-jun) are activated by protein kinases and other second messengers shortly after the onset of ischemia.
The synthesis of NO through nNOS is mainly related to calcium overload induced by glutamate in ischemic neurons [32]. The exacerbated activation of nNOS and the excess NO production due to uncoupled state, contributes to neurotoxicity and neurodegeneration by free radical formation [33,34,35].
Neurons are particularly sensitive to stress caused by ROS/RNS overproduction because of the relatively low levels of antioxidants compared with other cells. Physiologically, basal ROS/RNS production in neurons is constitutively generated by mitochondrial metabolism, which is higher than that in other cells due to the necessity to maintain neuronal circuit activity and synaptic transmission. Excessive nNOS activation under hyperactivation of NMDA receptors results in massive and uncontrolled generation of NO followed by the downstream formation of ROS/RNS, resulting in neuroinflammation and nitro oxidative burst. The prolonged destabilization of nNOS contributes to neurotoxicity and the potentiation of ischemic damage (Figure 2) [36,37]. Genetically engineered mice that overexpress the free radical scavenger, superoxide dismutase, have less edema formation than wild-type littermates confirming the detrimental role of free radicals in ischemia pathology [38,39,40].
A second glutamate-independent rise in Ca2+ is observed 2–3 h after an ischemic episode, further contributing to the activation of Ca2+-dependent enzymes (except NOS also proteases, phospholipases, cyclooxygenases, endonucleases) and leading to irreversible changes that promote the apoptotic and necrotic death of neurons [41,42,43,44,45].
In animal studies with the use of nNOS−/− knockout (KO) mice, in the model of cerebral ischemia-reperfusion, the smaller infarct size comparing to their wild-type littermates was observed. Relative cerebral blood flow after reperfusion was also higher and the levels of nitrates was significantly decreased in nNOS−/− mice [46,47]. No 3-nitrotyrosine immunoreactivity, a marker for ONOO formation associated with cell death, was reported in mutant mice deficient in nNOS activity, which were subjected to reversible MCAO [16].
In contrast to nNOS, NO generated by eNOS expressed in the endothelial cells has been suggested to have beneficial effects [17]. Here, eNOS not only promotes vascular dilation but also increases vascular smooth muscle cell proliferation and migration, and thereby enhances arteriogenesis after stroke [48].
In this case, eNOS deficient mice suffered from more severe ischemia-reperfusion injury in MCAO model, significantly reduced cerebral blood flow which subsequently resulted in greater infarct size [46,47,49].
The results of in vivo studies clearly suggest that the generation of NO and its neurotoxicity after reperfusion is closely related with the activity of nNOS. Additionally, nNOS-derived NO may play a major role in early blood-brain barrier BBB disruption following transient focal cerebral ischemia [50].

1.2. Late Stage of the Ischemia (Delayed Neuronal Damage)

Immediate early gene products activate transcription sites on the cytokine genes such as tumor necrosis factor-α (TNF-α) or interleukin-1β (IL-1β) that appear within hours after stroke. Appearance of cytokines attracts inflammatory cells such as neutrophils, that appear 4 to 6 h after a stroke [51,52]. P-selectin, E-selectin, vascular cell adhesion molecule (VCAM) and intercellular adhesion molecule (ICAM) appear on the vessels with different time courses, but the ultimate effect is that neutrophils rolling along the vessel attach and cross into the brain. At later times macrophages become the predominate cell type in the injured site [15,53].
The role of glia cells (astrocytes and microglia) is pivotal in progressing cerebral ischemia. Astrocytes play critical role BBB integrity and the maintenance of extracellular ion homeostasis by buffering excitatory transmitters released by neurons and producing trophic factors that support neuronal growth and survival. In response to pathological situations in surrounding tissue such as ischemic injury astrocytes undergo morphological, molecular and functional changes and become reactive astrocytes [54] The changes associated with the reactive state can directly impact synaptic transmission and neuronal circuit activity, thereby potentially contributing to the pathological changes observed after ischemia [55].
On the other hand, resting microglia in the mature brain is a powerful weapon under pathological activation. Following ischemia, as a result of rapid neuronal death, microglia migrates to damaged tissue to exert a neuroprotective effect by clearing dead tissue, inhibiting cytotoxic neuronal damage and releasing neuroprotective growth factors. However, microglial activation triggers iNOS production, which, together with the parallel appearance of reactive astrocytes, results in massive iNOS expression observed relatively late after an ischemic episode (after approximately 24 h) and generation of ROS/RNS [51,55,56,57,58]. In some brain areas, astrocytes also express nNOS, which may be tonically active after chain reactions initiated by ischemia. However, some contrary reports indicate the minor role of microglial iNOS in mediating brain injury after stroke [59].
In contrast to nNOS, the induction of iNOS expression begins 12 h after induction of ischemia, increases progressively over time and reaches maximal levels after approximately 24–48 h as shown in transient MCAO model in animals [60]. The other in vivo studies confirm that NO derived from iNOS appears to contribute to neurotoxicity after ischemic stroke. The infarct and the motor deficits produced by MCAO were smaller in iNOS knockouts than in wild-type mice confirming that iNOS-derived NO is one of the factors contributing to the expansion of the brain damage that occurs in the late stage of post-ischemic period [61]. Furthermore, the reduction of infarct size and improvement of neurological deficits was not observed up to 24 h after MCAO, indicating that iNOS does not participate in the initiation of ischemic brain damage [61].
Pathological glial activation escapes endogenous control and turns into an autoaggressive pathomechanism that contributes to secondary neuronal damage occurring after brain ischemia [58]. Leukocytes accumulating at the site of neuronal necrosis act as an additional barrier to this cascade of destructive events by secreting cytokines, including interferon-γ and interleukins (Figure 3).
Overall increased activity of nNOS/iNOS isoforms results in the subsequent production of NO in neurons, glia, neutrophils and the rest inflammatory cells inducing the release of inflammatory factors that promote cytotoxicity, including the increased activation of cell adhesion molecules, cytokines, TNFα and matrix metalloproteinases (MMPs) [62]. TNFα, caspases and MMPs are particularly important in triggering apoptotic signals in cells and contribute to BBB damage, neurotoxic substance release, free radical generation, oxidative stress and brain edema [63,64]. As a result, after hours and days the infarct size expands due to excitatory amino acid release, loss of ion homeostasis, decreased pH, inflammation and edema, causing additional damage and apoptosis in the surrounding areas [65]. The strategy for post-ischemic neuroprotective therapies is to target the peri-infarct or penumbra region to prevent or rescue the spreading damage of the initial infarct.
Reassuming, ischemic brain injury involves a complex interaction between leukocytes, glia, neurons and the endothelium, which form a vicious chain of events that results in massive and continuous production of iNOS resulting is subsequent generation of ROS/RNS [59,60]. The pathological events associated with ischemic brain injury involve energy failure, oxidative stress, acidosis, disruption of ion homeostasis, neuronal cell excitotoxicity or inflammation and evolve progressively over time. The early and late stages of ischemia described above concern not only sequential NO production but also immunological responses (which are in part related to NO generation) described elsewhere in more details (for review see: [66,67]).

2. NO-Dependent Factors Aggravating Ischemic Cascade

In addition to guanylyl cyclase activation and subsequent initialization of cGMP production (or pathological generation of ROS/RNS), NO-mediated processes control the functioning of many proteins and genes expression. Among them, s-nitrosylation and HIF-1α stabilization seem to be of importance in pathophysiology of brain ischemia.

2.1. S-Nitrosylation

The NO-mediated S-nitrosylation process is a redox-based posttranslational modification that modulates protein function and activity. S-nitrosylation is the chemical reaction of an NO moiety with the sulfhydryl groups of target proteins, which leads to the formation of S-nitrosothiols (R-SNO), producing S-nitrosylated proteins (SNO-proteins) [68].
S-nitrosylation can occur both extracellularly and intracellularly [69]. The S-nitrosylation of the regulatory binding partners of transcription factors (TFs) (for example, HIF-1α) may impose an extranuclear influence on their activation, stability and nuclear targeting [69,70]. On the other hand, the S-nitrosylation of critical redox-sensitive Cys residues in the DNA-binding or allosteric sites of TFs invokes alterations in gene expression [71,72]. Finally, S-nitrosylation may regulate protein function by the covalent addition of an NO group to a cysteine thiol/sulfhydryl group (RSH or, more properly, thiolate anion, RS) to form S-nitrosothiol derivatives (RS-NO), which is the most important aspect of the S-nitrosylation process [70]. The reaction is mediated by NO-related species including NO, NO (nitroxyl anion, which is NO with one additional electron) and NO+ (nitrosonium ion, which has one fewer electron than NO) [70].
In addition to the process of S-nitrosylation, functional equivalents of NO+ can be transferred from one nitrosothiol to another in a process called transnitrosylation, whereby an NO moiety is transferred from a SNO-protein to a free thiol on another protein, which occurs when two proteins interact directly and possess appropriate redox potentials to allow electron transfer [73].
Aberrant S-nitrosylation occurs as a consequence of exacerbated nitrosative stress via the excessive production of NO, which nitrosylates cysteine thiols with only partial SNO motifs or located more distant from the NO source [74]. These aberrantly S-nitrosylated proteins may contribute to pathological changes by triggering protein misfolding, mitochondrial dysfunction, transcriptional dysregulation, synaptic damage and neuronal injury [68]. In contrast, some SNO-proteins lose NO groups from their Cys thiols and undergo denitrosylation, which contributes to the regulation of the SNO signaling cascade [73]. Therefore, both S-nitrosylation and denitrosylation regulate protein activity and may be involved in pathological processes.
Under physiological conditions, nNOS is S-nitrosylated by NO. The studies were performed both in vitro in HEK223 cell lines, cultured primary cortical neurons treated with OGD/reoxygenation and in vivo in rat hippocampus during cerebral ischemia-reperfusion. In all experimental schedules the enzyme is S-nitrosylated in resting or physiological state and undergoes significant denitrosylation under oxygen deprivation, which is coupled with its increased activity [75]. The subsequent increase in NO production mediates the S-nitrosylation of proteins that initiate apoptotic signals in neurons, such as GluR6, c-Jun N-terminal kinase 3, phosphatases or tensin homolog [76].
In contrast to nNOS, the process of S-nitrosylation inhibits the activity of eNOS by inducing dimer disruption, dephosphorylation and changes in the subcellular targeting status as shown in vitro [77,78].
Two of the most important proteins that undergo aberrant S-nitrosylation in response to ischemic injury are GAPDH [79,80] and matrix metalloproteinase 9 (MMP9) [81]. GAPDH has been implicated in neurotoxicity and neurodegeneration and regulates transcriptional activation, apoptosis initiation, ER to Golgi vesicle shuttling and fast axonal or axoplasmic transport [82]. Following an ischemic episode, GAPDH accumulates rapidly both in the ischemic core and in penumbral apoptotic neurons as shown in vivo using MCAO model in rats [83]. Aberrant GAPDH s-nitrosylation, translation to nucleus, concomitant neuronal death occur during the early stages of reperfusion as shown, e.g., in the rat four-vessel occlusion ischemic model [80].
MMP9 belongs to the zinc metalloproteinase family, which is involved in the degradation of the extracellular matrix. The enzyme is acutely activated during ischemia and is selectively S-nitrosylated by NO during cerebral ischemia in vivo [81]. In in vitro studies s-nitrosylation of the cysteine switch at the active site of the enzyme together with ROS-mediated oxidation of MMP9 to sulfinic or sulfonic acid derivatives triggered the apoptotic form of cell death [81], suggesting a potential extracellular proteolysis pathway to neuronal cell death in which s-nitrosylation activates MM9.

2.2. Hypoxia-Inducible Factor 1α (HIF-1α)

HIF-1α is the primary mammalian transcription factor specifically regulated by hypoxia and plays an essential role in cellular and systemic O2 homeostasis by regulating the expression of genes important in tissue survival, that regulate glycolysis, erythropoiesis, angiogenesis or catecholamine metabolism [84].
HIF is primarily regulated by changes in protein stability and transcriptional activity in oxygen-dependent manner. Under physiological conditions HIF-1α is rapidly degraded and α subunit is hydroxylated by asparagine and proline residues by the family of prolyl-4-hydroxylase domain (PHD) proteins and factor inhibiting HIF (FIH), whose activity is dependent on molecular oxygen, ferrous iron, 2-oxoglutarate [85,86,87]. In response to low tissue oxygenation during ischemic stroke, the activity of PHD and FIH declines resulting in stabilization of HIF-1α, its accumulation within the cell and translocation into the nucleus. It results in the transcriptional activation of several dozen hypoxia-responsive genes through binding the hypoxia-responsive element (HRE) in their promoter region [88,89,90,91].
The activity of HIF-1α is regulated by NO via the mechanism of S-nitrosylation. NO metabolites such as S-nitrosoglutathione (GSNO) and peroxynitrite stabilizes or destabilizes HIF-1α, respectively [92,93,94].
The impact of HIF-1α on neuronal survival upon stroke is controversial, as it mediates both anti- and pro-survival genes [85]. Up regulation of vascular endothelial growth factor (VEGF) or erythropoietin (EPO) promotes adaptation to hypoxic/ischemic stress [95] and results in reduced infarct size after cerebral ischemia in MCAO model in rodents [58,85]. The HIF-1α dependent activation of eNOS transcription and subsequent NO release in the endothelium also contribute to reduction of ischemia infarct volume also shown in MCAO model [96]. Exogenously administered NO metabolite S-nitrosoglutathione (GSNO) was found not only to stabilize HIF-1α and to induce HIF-1α-dependent genes but also to stimulate the regeneration process and to aid in functional recovery in traumatic brain injury animal model [93].
On the other hand HIF-1α regulates iNOS transcription preserving iNOS translation during ischemia, thus contributing to progressive ischemia-induced inflammation [97,98].
Tissue-specific knockouts have also given conflicting results on the role of HIF-1α during ischemia. The studies of Helton et al. showed that the brains from neuron-specific HIF-1α deficient mice were protected from hypoxia-induced cell death, when subjected to normobaric chamber or bilateral carotid artery occlusion (BCAO) model, suggesting that decreasing HIF-1α level can be neuroprotective [84]. Additionally, in HIF-1α KO mice genes involved in apoptotic pathway were uniquely downregulated when compared with WT animals. Endothelial-specific HIF-1α knock-out reduced BBB permeability and brain infarction in diabetic mice subjected to MCAO procedure [99]. In in vitro studies, in primary cortical neuron cultures, silencing HIF-1α attenuated the accumulation of iNOS [100].
On the other hand in the studies of Baranova et al. neuron-specific inactivation of HIF-1α increased brain injury in mice MCAO model [101]. Additionally, neuron-specific PHD inactivation which resulted in up-regulation of HIF-1α lead to smaller infarct size and reduced edema formation in transient MCAO model in mice [102].
In the models of permanent MCAO or acute phase of ischemic stroke in mice the expression of HIF-1α was enhanced in the place of injury, causing the massive production of iNOS [103], indicating that activation of HIF-1α might be involved in the mechanisms through which iNOS promotes cell death or survival constituting a critical factor in widespread inflammation and subsequent pathological events. Another studies showed that intermittent hypoxic conditions after brain ischemia displayed a neuroprotective effect, and despite relatively high expression of HIF-1α, lower expression of iNOS in the border between infarcts and normal tissue was observed, suggesting that overactivation of HIF-1α may suppress the activation of microglia in ischemic mice [104]. Enhanced HIF-1α activation was responsible for triggering the transcription of HIF-regulated genes (VEGF, erythropoietin, eNOS), reduced infarct size and caspase-3 activation in MCAO or common carotid arteries occlusion models in mice [96]. Increase HIF-1α was accompanied by increased iNOS expression as shown in MCAO-reperfusion injury model in rats [105].
In in vitro studies the upregulation of HIF-1α signaling was shown to improve cGMP production following ischemia through the maintenance of cGMP protein kinase activity [106], thus preventing the NO-mediated production of ROS/RNS instead of cGMP.
Overall, the role of HIF-1α in ischemia remains inconclusive, but despite controversial results the role of HIF-1α was shown to be of special importance in ischemia preconditioning and may either promote or prevent neuronal survival [107]. Partially the pro-survival and pro-inflammatory roles in the ischemic brain might depend on the injury model, time point or cell type assessed [101,108].
The contradictory results indicate that phenotype and transcriptional response to hypoxia in vivo is much more complex that would have been supposed. The brain has multiple ways of inducing HIF-1α-dependent genes involved in the response to hypoxia (that promote erythropoiesis, angiogenesis or vasodilation), and one of the alternatives is HIF-2α [109,110,111,112]. HIF-1α and HIF-2α combined KO mice exhibited reduced expression of the anti-survival genes in MCAO model in mice [85]. Even though the mice initially performed better, became more impaired 72 h after reperfusion, accompanied by increased apoptosis and reduced angiogenesis. HIF-1α and HIF-2α can partially compensate for each other, although specific target genes are differentially regulated after ischemia. Combined loss of neuronal HIF-1α and HIF-2α impairs functional recovery after cerebral ischemia, which may be beneficial predominantly in the early phase after stroke, indicating a timely regulated activation-inhibition of hypoxia-regulated cytoprotective and damaging factors which may be important for the functional outcome after stroke [85].
Another example of mutuality that may promote proapoptotic genes in response to hypoxia is p53/HIF-1α interaction [113,114,115]. Specific disruption of this interplay (attributable to the lack of HIF-1α) leads to downregulation and loss of expression of genes that promote cell death [84,98].

3. Treatment Strategies

As discussed above, a variety of factors contribute to the severity of stroke and its long-term consequences. Treatment strategies based on NO signaling may lower the risk of severe and irreversible complications after stroke, but they may also have detrimental effects and must therefore be provided with caution.
The advantage of NO-based therapies is the possibility of modulating endogenous mechanisms activated after cerebral ischemia with exogenously applied compounds. The objective is to promote neuroprotective outcomes and integrate cellular signaling pathways at different stages of brain damage. The intervention time, pharmacokinetics, pharmacodynamics and activities of the compounds are critical to successfully counteract the consequences of stroke.
The present knowledge indicates that the therapeutic window to reduce the pathological consequences of stroke, essentially neuronal damage, is estimated as 0–6 h for primary interventions [116,117] and may extend up to 24 h poststroke.
As discussed above, NO-mediated actions clearly indicate that the activation of eNOS contributes to proper vasodilation, exerts antioxidant, anti-inflammatory and anti-atherogenic effects, and regulates glucose uptake and insulin sensitivity [118,119], thus exerting protective effects in stroke. In contrast, the inhibition of nNOS and iNOS, the main generators of free radicals, alternatively elimination of free radicals, may counteract neurodegeneration.

3.1. Free Radical Scavengers

Reperfusion of ischemic areas can exacerbate ischemic brain damage through the generation of ROS/RNS (e.g., O2, hydroxyl radicals and ONOO) by excessive production of both NO and O2 by nNOS and iNOS. It has been reported that excess of NO immediately reacts with O2 to form ONOO [120,121], which is responsible for the nitration of both free and protein bound tyrosine residues which are known to disrupt cell signaling cascades leading to tissue injury [120,121]. Therefore, enhanced degradation of ROS/RNS with pharmacological agents has been found to limit the extent of brain damage following stroke-induced excess of NO generation. Elimination of free radicals attenuates cytokine formation that drives up regulation of inflammatory proteins including iNOS.
On the other hand ROS provides a redox signal for hypoxic HIF-1α activation [122]. Assuming, that under specific pathological conditions, increased activity of HIF-1α is responsible for ischemia-induced detrimental effects, application of anti-oxidants eliminates ROS and consequently reduces HIF-1α levels by its destabilization and loss of transcriptional activity [123].
Under normal conditions, endogenous protective enzyme systems, such as superoxide dismutase (SOD) and reduced glutathione (GSH), limit the overproduction of free radicals. However, their capacities may be overwhelmed by pathological changes when free radical generation is uncontrolled. Removal of pathologically produced free radicals may be regarded as a viable approach to neuroprotection and may be achieved by scavenging or trapping free radicals. Among the compounds that possess free radical scavenging properties are nitrones, thiols, iron chelators, phenols and catechols. Tirilazad, ebselen and edaravone are examples of compounds with scavenging activity, and compounds with free radical trapping properties include NXY-059 and NSP-116.
Edaravone (Radicut) is a free radical scavenger marketed in Japan to treat acute ischemic stroke [124]. Edaravone scavenged •HO, NO and ONOO in a concentration-dependent manner [125,126,127]. The efficacy of the drug ranges from large clinical improvements to modest effects measured with standard stroke scales when administered up to 72 h following ischemic stroke [128,129].
Tirilazad mesylate, a lazaroid, has been investigated as a neuroprotective agent in patients after acute ischemic stroke. The compound was ineffective in treating acute ischemic stroke [130,131]. Another seleno-organic compound, ebselen, was shown to reduce delayed ischemic neurological deficits after SAH [132] and to improve outcomes after stroke [133,134].

3.2. Enhancement of NO Production

L-arginine, a substrate for NO biosynthesis, is an obvious candidate of choice to improve NO bioavailability. The administration of L-arginine increased regional blood flow and prevented tissue damage in the rat MCAO model [135,136,137]. However, contradictory results with L-arginine administration were obtained that showed a beneficial effect [135,136,137], no effect [138] and the potentiation of pathological changes [139,140]. These effects may result from the fact that L-arginine enhances NO synthesis via all three isoforms [141]; therefore, synthesis of potentially detrimental NO from iNOS or nNOS may counteract the neuroprotective effects of eNOS activation. The ability of L-arginine to stimulate the release of hormones such as insulin [142], glucagon [143], growth hormone [144] or catecholamines [145] and its ability to transform into toxic byproducts such as polyamines or agmatine [146,147,148,149] are also limitations of its use.
NO donors, such as organic nitrates, sodium nitroprusside (SNP), sydnonimines, S-nitrosothiols, NONOates and hybrid donors potentiate NO production and thus may potentially be regarded as treatment strategies.
Administration of DETA/NONOate [150], SNP [151,152], 3-morpholinosydnonimine [153,154], ZJM-289 [155] or LA-419 [156,157] significantly increased cell proliferation and/or reversed ischemia-induced tissue damage in the selected structures of the rat brain. Stimulation of eNOS activity with concomitant suppression of nNOS and iNOS function was observed [155,156,157]. Beyond this point, neurovascular toxicity instead of vasodilation or infarct size reduction was observed [152].
These results from preclinical studies were confirmed in clinical observations in which SNP reduced the mean arterial blood pressure and led to improvements in CBF in patients after acute stroke [158,159].
The mechanism of the neuroprotective action of NO donors includes their abilities to reduce oxidative stress both in the brain and blood, inhibit the expression and activities of MMPs, scavenge NO or quench ROS, reduce inflammation, exert antiplatelet effects and attenuate I/R-mediated increases in ICAM-1 and E-selectin mRNA expression [160,161].
Suppression of the hydrolysis of L-arginine into ornithine and urea by arginases increased the production of NO and prevented the development of endothelial dysfunction [162], similar to the effects of eliminating asymmetric dimethylarginine (ADMA)—endogenous NOS inhibitor, through the stimulation of ADMA-hydrolyzing enzyme (DDAH II) [163].
The available data concerning the efficacy of the enhancement of NO production-based therapies are collected in Table 1.

3.3. NOS Inhibitors

Selective inhibition of nNOS and iNOS activities is thought to counteract the aftermath of stroke.
The neuroprotective abilities of N(ω)-nitro-L-arginine methyl ester (L-NAME) and 7-nitroindazole (7-NI), widely used NOS/nNOS inhibitors, are time-dependent and are observed in the early stages of ischemic insult [164]. The choice of dose is important, as the compounds prevent at low doses but potentiate ischemia-induced neurodegeneration at high doses [165]. Long-term inhibition of NOS might be too risky because of the off-target effects on eNOS, particularly in patients with cardiovascular risk or metabolic diseases.
Direct or indirect inhibition of iNOS through selective inhibitors, such as aminoguanidine (AG) or the BH4 rate-limiting enzyme GTP cyclohydrolase I, attenuated cerebral infarction, ischemia-induced pathologies and prevented the progression of cerebral aneurysms [166,167,168].
The available data concerning the efficacy of NOS inhibitor-based therapies are collected in Table 1. For review please see also: [164].
Table 1. Enhancement of NO production and NOS inhibitor-based therapies. Pretreatment indicates that the compound was administered before experimental ischemia; posttreatment indicates that the compound was administered after the episode. MCAO—middle cerebral artery occlusion, ODG—oxygen and glucose deprivation.
Table 1. Enhancement of NO production and NOS inhibitor-based therapies. Pretreatment indicates that the compound was administered before experimental ischemia; posttreatment indicates that the compound was administered after the episode. MCAO—middle cerebral artery occlusion, ODG—oxygen and glucose deprivation.
CompoundEffect
NO Donors
PretreatmentLA-419in vivo:
reduced iNOS, nNOS, nitrotyrosine expression and increased apparent diffusion coefficient in endothelin-1-induced focal cerebral ischemia or global cerebral ischemia model in rats induced by oxygen and glucose deprivation
[156,157]
GSNOin vivo:
reduced caspase-3, -8, -9, tBID cleavage in global ischemia models in rats
increased Bid, pro-caspase-3 and pro-caspase-9 expression in global ischemia model in rats
reduced Fas, CaMKII, MKK4 S-nitrosylation, increased nNOS S-nitrosylation and phosphorylation, increased CaMKII phosphorylation in global ischemia models in rats
increased cell density in CA1 in global ischemia models in rats
reduced infarct size and edema in MCAO model in mice
[169,170,171,172]
ZJM-289in vitro:
increased cell viability in primary cortical neuron in OGD model
reduced mitochondrial dysfunction in primary cortical neuron in OGD model
decreased Ca2+ release and ROS production in primary cortical neuron in OGD model
in vivo:
reduced infarct size and edema and improved neurologic deficit in MCAO model in rats
[155,173]
SIN-1in vivo:
no influence or reduced infarct size in MCAO model in rats
[174,175]
DETA NONOatein vivo:
no influence on infarct size in MCAO model in rats
[174]
NBPin vitro:
increased cell viability in primary cortical neuron in OGD model
reduced mitochondrial dysfunction in primary cortical neuron in OGD model
decreased Ca2+ release and ROS production in primary cortical neuron in OGD model
in vivo:
reduced infarct size and improved neurologic deficit in MCAO model in rats
[173,176]
Spermine NONOatein vivo:
reduced infarct size in MCAO model in rats
increased cortical perfusion in MCAO model in rats
[152]
sodium nitroprussidein vivo:
reduced infarct size in MCAO model in rats
[152]
PosttreatmentGSNOin vivo:
reduced infarct size
[171]
DETA NONOatein vivo:
no influence on infarct size in MCAO model in rats
improved neurologic deficit and increased cGMP level in MCAO model in rats
increased cell proliferation in subventricular zone, olfactory bulb and dentate gyrus in MCAO model in rats
[150]
SIN-1in vivo:
reduced infarct size in MCAO model in rats
[153,154]
sodium nitroprussidein vivo:
reduced infarct size in MCAO model in rats
[153]
NOS or nNOS inhibitors
Pretreatment7-NIin vivo:
increased nNOS S-nitrosylation and phosphorylation, decreased CaMKII and MKK4 S-nitrosylation and increased CaMKII phosphorylation in global ischemia model or MCAO in rats
decreased caspase-3 cleavage in MCAO model in rats
increased cell density in CA1 in global ischemia model or MCAO model in rats
reduced maximal NO concentration in bilateral common carotid artery occlusion in rats
[170,172,177]
L-NAMEin vivo:
reduced infarct size and improved neurological deficit in MCAO model in rats
reduced glutamate, aspartate, glutamine synthetase and nitrate/nitrite level in MCAO model in rats
increased ATP and NAD level in MCAO model in rats
reduced TNF-α expression and increased IL-10 expression in MCAO model in rats
[174,178]
Posttreatment7-NIin vivo:
reduced neuronal damage in global cerebra ischemia in rats
[179]
L-NAMEin vivo:
reduced infarct volume in MCAO model in rats and mice
improved neurological deficit in MCAO model in rats and mice
reduced level of tissue nitric oxide end products in MCAO model in mice
reduced nitrate/nitrite level and increased NAD level in MCAO model in rats
[178,180]
iNOS inhibitors
Pretreatmentaminoguanidinein vivo:
reduced infarct volume, edema, neurological deficits, necrotic cell death in penumbra and core and reduced apoptosis in penumbra in MCAO model in rats
[181]
Posttreatmentaminoguanidinein vivo:
reduced infarct volume in permanent MCAO model in mice
reduced ischemia-induced neurogenesis in dentate gyrus in MCAO model in rats
[167,182,183]
1400 Win vivo:
reduced infarct size neurological deficit in MCAO model in rats
inhibited delayed increase in glutamate level in MCAO model in rats
[184]
S-methylisothioreain vivo:
reduced neurological deficit, mortality, infarct volume ratio in MCAO model in rats
attenuated morphological changes in cortical neurons in MCAO model in rats
[185]

3.4. HIF-1α

HIF-1α is an essential component in changing the transcriptional repertoire of tissues during oxygen deprivation and plays a pivotal role in the regulation of iNOS activity. Thus, HIF-1α and the genes regulated by it have been the center of intense research. A growing number of pre-clinical studies in rodents suggests that the activation of HIF-1α signaling pathway prior or shortly after ischemic stroke reduces tissue damage and increases functional recovery from ischemic stroke [87,96,186,187,188,189]. One example of an agent that stabilizes the transcriptional activator HIF-1α and activates target genes involved in compensation for ischemia are inhibitors of HIF-1α prolyl hydrolases (PHD1, PHD2 and PHD3) [190]. The beneficial effects of PHD inhibition after ischemia require the activity of HIF-1α as shown in in vitro in oxygen glucose deprivation model of ischemia and in vivo in MCAO model in mice [96]. Dimethyloxalylglycine (DMOG) enhanced the activation of HIF-1α and enhanced transcription of the HIF-regulated genes. In vivo the infarct size, activation of pro-apoptotic proteins and behavioral deficits after stroke were reduced. The effect of DMOG was decreased after inhibition of HIF-1α with digoxin [96]. Similar effects were observed with small molecule hypoxia mimics, such as deferoxamine, cobalt chloride or GSNO. Additionally, these agents increased the expression of HIF-1α target genes [92,93,191]. Cocaine, andrographolide or vitamin E activated the HIF-VEGF pathway, thus increasing microvascular density, restoring local blood flow and protecting the brain from ischemic insults [192,193,194]. However, available data indicate that sustained and prolonged activation of the HIF-1α pathway may lead to a transition from neuroprotection to neurodegeneration, reflecting the dual features of the HIF system [195,196,197], which should be taken into consideration when considering therapy to modify HIF-1α. Some contradictory results indicated that inhibition of HIF-1α improved brain function in ischemia-reperfusion brain injury-related disorders [105,198]. Inhibition of free radical formation, followed by inhibition of HIF-1α activation, apoptosis formation, neutrophil activation and iNOS expression resulted in reduction in the infarct volume in ischemia-reperfusion brain injury (MCAO model) in rats [105]. In rat model of transient MCAO the early inhibition of HIF-1α by application of inhibitors or small interfering RNA reduced infarct size and BBB hyperpermeability, decreased mortality and improved neurological deficits through inhibition of HIF-1α activity [198,199]. The studies of Hsiao et al. showed that the administration of PMC [105].
The therapeutic strategies might lead to pleiotropic activation of HIF signaling through all cell types in CNS, making it difficult to draw conclusions about the significance of the HIF signaling pathway in the treatment of ischemia-reperfusion injury. However, the stimulation of the non-ischemic penumbra regions to initiate HIF-1α (e.g., by inhibition of PHD) and subsequent downstream induction of HIF-1α mediated antiapoptotic, vascular and glycolytic metabolic changes before the area is enveloped by the spreading ischemia may be one of the proposed therapeutic HIF-dependent effect. Additionally, inhibition of endothelial HIF-1α warrants further investigation as a therapeutic target for the treatment of stroke patients with diabetes [99].
The available data concerning the efficacy of HIF-1α-based therapies are collected in Table 2.

3.5. Combination Therapies

At present, it seems that combination therapies are more effective for treating stroke than monotherapies. Scavenging reactive oxygen species and concomitantly inhibiting NO synthesis by administering statins with resveratrol, an approved antioxidant or with nifedipine, a calcium channel blocker, is just one example of combination therapy for stroke [226]. Administration of resveratrol upregulated antioxidant enzyme activities, decreased O2 production by downregulating NADPH oxidase activity and attenuated oxidative stress-mediated eNOS uncoupling. Some hope may be placed in combination therapy, which could allow minimization of the dose of NMDA antagonists by administering a compound enhancing its effect, such as caffeine [227] or cordycepin (an adenosine analog) [228]. The use of different agents at the same time enables avoidance of detrimental consequences induced by undesired side effects and reinforcement of beneficial impacts by mutual action. The examples of effective combined treatment therapies in clinical trials are collected in Table 3.

4. Conclusions

Stroke is one of the most common indications with unmet medical needs in medicine, and it is a major challenge to develop effective treatments, as stroke treatments need to reduce cell death and infarct size, stabilize the blood-brain barrier, reduce reoxygenation-induced leakage and preserve neuromotor function in a supra-additive manner. The NO-dependent pathway plays a crucial role in all processes involved in subsequent events after ischemia [49,154,235]. However, NO-based treatments can be burdened with a variety of adverse effects. The risk of exacerbating pathological changes with NO-regulating agents is also high. The objective is to restore the imbalance L-arginine/ADMA-NOS-NO [236] and protect against hypoxic/ischemic-derived damage.
The use of combined treatment therapies seems to be the best alternative. The number of plausible combinations is almost unlimited. The only limitation is the satisfactory effectiveness and the lack of the induction of adverse effects. The challenge is to find satisfactory solution. NO-related pathways and the agents that inhibit excess of NO production especially of that which comes from nNOS or iNOS, free radicals’ scavengers or compounds targeting HIF-1α give a number of possibilities that still have not been fully investigated so far. The vast range of studies discussed in the present review and by others indicates, that the area is still open for investigation.

Author Contributions

All the authors equally contributed to manuscript preparation and writing. All authors have read and agreed to the published version of the manuscript.

Funding

Grants no 2019/33/B/NZ7/02699 (OPUS 17, National Science Center) and no DIR/WK/2017/01 (Ministry of Science and Higher Education Poland).

Acknowledgments

The authors would like to thank Studio Grafiki i Komiksu studiokomiks.pl (Krzysztof Czachura) Design for graphical support in preparation of the figures.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. GBD 2019 Diseases and Injuries Collaborators. Global burden of 369 diseases and injuries in 204 countries and territories, 1990–2019: A systematic analysis for the Global Burden of Disease Study 2019. Lancet 2020, 396, 1204–1222. [Google Scholar] [CrossRef]
  2. GBD 2016 Causes of Death Collaborators. Global, regional, and national age-sex specific mortality for 264 causes of death, 1980–2016: A systematic analysis for the Global Burden of Disease Study 2016. Lancet 2017, 390, 1151–1210. [Google Scholar] [CrossRef] [Green Version]
  3. Müller, G.J.; Stadelmann, C.; Bastholm, L.; Elling, F.; Lassmann, H.; Johansen, F.F. Ischemia leads to apoptosis--and necrosis-like neuron death in the ischemic rat hippocampus. Brain Pathol. 2004, 14, 415–424. [Google Scholar] [CrossRef]
  4. Aarts, M.; Liu, Y.; Liu, L.; Besshoh, S.; Arundine, M.; Gurd, J.W.; Wang, Y.-T.; Salter, M.W.; Tymianski, M. Treatment of ischemic brain damage by perturbing NMDA receptor- PSD-95 protein interactions. Science 2002, 298, 846–850. [Google Scholar] [CrossRef]
  5. Zhang, S.J.; Steijaert, M.N.; Lau, D.; Schutz, G.; Delucinge-Vivier, C.; Descombes, P.; Bading, H. Decoding NMDA receptor signaling: Identification of genomic programs specifying neuronal survival and death. Neuron 2007, 53, 549–562. [Google Scholar] [CrossRef] [Green Version]
  6. Christopherson, K.S.; Hillier, B.J.; Lim, W.A.; Bredt, D.S. PSD-95 assembles a ternary complex with the N-methyl-D-aspartic acid receptor and a bivalent neuronal NO synthase PDZ domain. J. Biol. Chem. 1999, 274, 27467–27473. [Google Scholar] [CrossRef] [Green Version]
  7. Palmer, R.M.; Rees, D.D.; Ashton, D.S.; Moncada, S. L-arginine is the physiological precursor for the formation of nitric oxide in endothelium-dependent relaxation. Biochem. Biophys. Res. Commun. 1988, 153, 1251–1256. [Google Scholar] [CrossRef]
  8. Katsuki, S.; Arnold, W.; Mittal, C.; Murad, F. Stimulation of guanylate cyclase by sodium nitroprusside, nitroglycerin and nitric oxide in various tissue preparations and comparison to the effects of sodium azide and hydroxylamine. J. Cyclic Nucleotide Res. 1977, 3, 23–35. [Google Scholar] [PubMed]
  9. Miki, N.; Kawabe, Y.; Kuriyama, K. Activation of cerebral guanylate cyclase by nitric oxide. Biochem. Biophys. Res. Commun. 1977, 75, 851–856. [Google Scholar] [CrossRef]
  10. Bon, C.L.M.; Garthwaite, J. On the role of nitric oxide in hippocampal long-term potentiation. J. Neurosci. 2003, 23, 1941–1948. [Google Scholar] [CrossRef]
  11. Kalinowski, L.; Malinski, T. Endothelial NADH/NADPH-dependent enzymatic sources of superoxide production: Relationship to endothelial dysfunction. Acta Biochim. Pol. 2004, 51, 459–469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Leone, A.M.; Palmer, R.M.; Knowles, R.G.; Francis, P.L.; Ashton, D.S.; Moncada, S. Constitutive and inducible nitric oxide synthases incorporate molecular oxygen into both nitric oxide and citrulline. J. Biol. Chem. 1991, 266, 23790–23795. [Google Scholar] [CrossRef]
  13. Lelchuk, R.; Radomski, M.W.; Martin, J.F.; Moncada, S. Constitutive and inducible nitric oxide synthases in human megakaryoblastic cells. J. Pharmacol. Exp. Ther. 1992, 262, 1220–1224. [Google Scholar]
  14. Eliasson, M.J.; Huang, Z.; Ferrante, R.J.; Sasamata, M.; Molliver, M.E.; Snyder, S.H.; Moskowitz, M.A. Neuronal nitric oxide synthase activation and peroxynitriteformation in ischemic stroke linked to neural damage. J. Neurosci. 1999, 19, 5910–5918. [Google Scholar] [CrossRef] [PubMed]
  15. Xia, Y.; Zweier, J.L. Superoxide and peroxynitrite generation from inducible nitric oxide synthase in macrophages. Proc. Natl. Acad. Sci. USA 1997, 94, 6954–6958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pellegrini-Giampietro, D.E.; Cherici, G.; Alesiani, M.; Carlà, V.; Moroni, F. Excitatory amino acid release from rat hippocampal slices as a consequence of free-radical formation. J. Neurochem. 1988, 51, 1960–1963. [Google Scholar] [CrossRef] [PubMed]
  17. Raghavan, S.; Dikshit, M. Vascular regulation by the L-arginine metabolites, nitric oxide and agmatine. Pharmacol. Res. 2004, 49, 397–414. [Google Scholar] [CrossRef]
  18. Dereski, M.O.; Chopp, M.; Knight, R.A.; Rodolosi, L.C.; Garcia, J.H. The heterogeneous temporal evolution of focal ischemic neuronal damage in the rat. Acta Neuropathol. 1993, 85, 327–333. [Google Scholar] [CrossRef]
  19. Ginsberg, M.D.; Busto, R. Rodent models of cerebral ischemia. Stroke 1989, 20, 1627–1642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Samdani, A.F.; Dawson, T.M.; Dawson, V.L. Nitric oxide synthase in models of focal ischemia. Stroke 1997, 28, 1283–1288. [Google Scholar] [CrossRef]
  21. Garry, P.S.; Ezra, M.; Rowland, M.J.; Westbrook, J.; Pattinson, K.T.S. The role of the nitric oxide pathway in brain injury and its treatment—From bench to bedside. Exp. Neurol. 2015, 263, 235–243. [Google Scholar] [CrossRef] [Green Version]
  22. Bolaños, J.P.; Almeida, A. Roles of nitric oxide in brain hypoxia-ischemia. Biochim. Biophys. Acta 1999, 1411, 415–436. [Google Scholar] [CrossRef] [Green Version]
  23. Dirnagl, U.; Iadecola, C.; Moskowitz, M.A. Pathobiology of ischaemic stroke: An integrated view. Trends Neurosci. 1999, 22, 391–397. [Google Scholar] [CrossRef]
  24. Chen, Z.-Q.; Mou, R.-T.; Feng, D.-X.; Wang, Z.; Chen, G. The role of nitric oxide in stroke. Med. Gas Res. 2017, 7, 194–203. [Google Scholar] [CrossRef] [Green Version]
  25. Heiss, W.D. Experimental evidence of ischemic thresholds and functional recovery. Stroke 1992, 23, 1668–1672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Furlan, M.; Marchal, G.; Viader, F.; Derlon, J.M.; Baron, J.C. Spontaneous neurological recovery after stroke and the fate of the ischemic penumbra. Ann. Neurol. 1996, 40, 216–226. [Google Scholar] [CrossRef]
  27. Baird, A.E.; Benfield, A.; Schlaug, G.; Siewert, B.; Lövblad, K.O.; Edelman, R.R.; Warach, S. Enlargement of human cerebral ischemic lesion volumes measured by diffusion-weighted magnetic resonance imaging. Ann. Neurol. 1997, 41, 581–589. [Google Scholar] [CrossRef]
  28. Castillo, J.; Rama, R.; Dávalos, A. Nitric oxide-related brain damage in acute ischemic stroke. Stroke 2000, 31, 852–857. [Google Scholar] [CrossRef] [Green Version]
  29. Forster, C.; Clark, H.B.; Ross, M.E.; Iadecola, C. Inducible nitric oxide synthase expression in human cerebral infarcts. Acta Neuropathol. 1999, 97, 215–220. [Google Scholar] [CrossRef]
  30. Ketheeswaranathan, P.; Turner, N.A.; Spary, E.J.; Batten, T.F.C.; McColl, B.W.; Saha, S. Changes in glutamate transporter expression in mouse forebrain areas following focal ischemia. Brain Res. 2011, 1418, 93–103. [Google Scholar] [CrossRef]
  31. Mayhan, W.G.; Didion, S.P. Glutamate-induced disruption of the blood-brain barrier in rats. Role of nitric oxide. Stroke 1996, 27, 965–970. [Google Scholar] [CrossRef] [PubMed]
  32. Zhou, L.; Zhu, D.Y. Neuronal nitric oxide synthase: Structure, subcellular localization, regulation, and clinical implications. Nitric Oxide Biol. Chem. 2009, 20, 223–230. [Google Scholar] [CrossRef] [PubMed]
  33. Fabian, R.H.; Perez-Polo, J.R.; Kent, T.A. Perivascular nitric oxide and superoxide in neonatal cerebral hypoxia-ischemia. Am. J. Physiol. Heart Circ. Physiol. 2008, 295, H1809–H1814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Onodera, H.; Kogure, K.; Ono, Y.; Igarashi, K.; Kiyota, Y.; Nagaoka, A. Proto-oncogene c-fos is transiently induced in the rat cerebral cortex after forebrain ischemia. Neurosci. Lett. 1989, 98, 101–104. [Google Scholar] [CrossRef]
  35. Wessel, T.C.; Joh, T.H.; Volpe, B.T. In situ hybridization analysis of c-fos and c-jun expression in the rat brain following transient forebrain ischemia. Brain Res. 1991, 567, 231–240. [Google Scholar] [CrossRef]
  36. Jiang, X.; Mu, D.; Manabat, C.; Koshy, A.A.; Christen, S.; Täuber, M.G.; Vexler, Z.S.; Ferriero, D.M. Differential vulnerability of immature murinae neurons to oxygen-glucose deprivation. Exp. Neurol. 2004, 190, 224–232. [Google Scholar] [CrossRef]
  37. Grima, G.; Benz, B.; Do, K.Q. Glial-derived arginine, the nitric oxide precursor, protects neurons from NMDA-induced excitotoxicity. Eur. J. Neurosci. 2001, 14, 1762–1770. [Google Scholar] [CrossRef]
  38. Kondo, T.; Reaume, A.G.; Huang, T.T.; Carlson, E.; Murakami, K.; Chen, S.F.; Hoffman, E.K.; Scott, R.W.; Epstein, C.J.; Chan, P.H. Reduction of CuZn-superoxide dismutase activity exacerbates neuronal cell injury and edema formation after transient focal cerebral ischemia. J. Neurosci. 1997, 17, 4180–4189. [Google Scholar] [CrossRef]
  39. Kondo, T.; Reaume, A.G.; Huang, T.T.; Murakami, K.; Carlson, E.; Chen, S.; Scott, R.W.; Epstein, C.J.; Chan, P.H. Edema formation exacerbates neurological and histological outcomes after focal cerebral ischemia in CuZn-superoxide dismutase gene knockout mutant mice. Acta Neurochir. Suppl. 1997, 70, 62–64. [Google Scholar]
  40. Chan, P.H.; Epstein, C.J.; Li, Y.; Huang, T.T.; Carlson, E.; Kinouchi, H.; Yang, G.; Kamii, H.; Mikawa, S.; Kondo, T.; et al. Transgenic mice and knockout mutants in the study of oxidative stress in brain injury. J. Neurotrauma 1995, 12, 815–824. [Google Scholar] [CrossRef]
  41. Léveillé, F.; Gaamouch, F.E.; Gouix, E.; Lecocq, M.; Lobner, D.; Nicole, O.; Buisson, A. Neuronal viability is controlled by a functional relation between synaptic and extrasynaptic NMDA receptors. FASEB J. 2008, 22, 4258–4271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Singh, P.; Doshi, S.; Spaethling, J.M.; Hockenberry, A.J.; Patel, T.P.; Geddes-Klein, D.M.; Lync, D.R.; Meaney, D.F. N-methyl-D-aspartate receptor mechanosensitivity is governed by C terminus of NR2B subunit. J. Biol. Chem. 2012, 287, 4348–4359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Maneshi, M.M.; Maki, B.; Gnanasambandam, R.; Belin, S.; Popescu, G.K.; Sachs, F.; Hua, S.Z. Mechanical stress activates NMDA receptors in the absence of agonists. Sci. Rep. 2017, 7, 39610. [Google Scholar] [CrossRef] [Green Version]
  44. Amorini, A.M.; Lazzarino, G.; Pietro, V.D.; Signoretti, S.; Lazzarino, G.; Belli, A.; Tavazzi, B. Severity of experimental traumatic brain injury modulates changes in concentrations of cerebral free amino acids. J. Cell. Mol. Med. 2017, 21, 530–542. [Google Scholar] [CrossRef] [PubMed]
  45. Palmer, A.M.; Marion, D.W.; Botscheller, M.L.; Bowen, D.M.; DeKosky, S.T. Increased transmitter amino acid concentration in human ventricular CSF after brain trauma. Neuroreport 1994, 6, 153–156. [Google Scholar] [CrossRef]
  46. Ito, Y.; Ohkubo, T.; Asano, Y.; Hattori, K.; Shimazu, T.; Yamazato, M.; Nagoya, H.; Kato, Y.; Araki, N. Nitric oxide production during cerebral ischemia and reperfusion in eNOS- and nNOS-knockout mice. Curr. Neurovasc. Res. 2010, 7, 23–31. [Google Scholar] [CrossRef]
  47. Wei, G.; Dawson, V.L.; Zweier, J.L. Role of neuronal and endothelial nitric oxide synthase in nitric oxide generation in the brain following cerebral ischemia. Biochim. Biophys. Acta 1999, 1455, 23–34. [Google Scholar] [CrossRef] [Green Version]
  48. Cui, X.; Chopp, M.; Zacharek, A.; Zhang, C.; Roberts, C.; Chen, J. Role of endothelial nitric oxide synthetase in arteriogenesis after stroke in mice. Neuroscience 2009, 159, 744–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Huang, Z.; Huang, P.L.; Ma, J.; Meng, W.; Ayata, C.; Fishman, M.C.; Moskowitz, M.A. Enlarged infarcts in endothelial nitric oxide synthase knockout mice are attenuated by nitro-L-arginine. J. Cereb. Blood Flow Metab. 1996, 16, 981–987. [Google Scholar] [CrossRef]
  50. Jiang, Z.; Li, C.; Arrick, D.M.; Yang, S.; Baluna, A.E.; Sun, H. Role of nitric oxide synthases in early blood-brain barrier disruption following transient focal cerebral ischemia. PLoS ONE 2014, 9, e93134. [Google Scholar] [CrossRef] [PubMed]
  51. Zarruk, J.G.; Greenhalgh, A.D.; David, S. Microglia and macrophages differ in their inflammatory profile after permanent brain ischemia. Exp. Neurol. 2018, 301, 120–132. [Google Scholar] [CrossRef]
  52. Pei, X.; Li, Y.; Zhu, L.; Zhou, Z. Astrocyte-derived exosomes suppress autophagy and ameliorate neuronal damage in experimental ischemic stroke. Exp. Cell Res. 2019, 382, 111474. [Google Scholar] [CrossRef] [PubMed]
  53. Stanimirovic, D.B.; Wong, J.; Shapiro, A.; Durkin, J.P. Increase in surface expression of ICAM-1, VCAM-1 and E-selectin in human cerebromicrovascular endothelial cells subjected to ischemia-like insults. Acta Neurochir. Suppl. 1997, 70, 12–16. [Google Scholar] [CrossRef]
  54. Wilhelmsson, U.; Bushong, E.A.; Price, D.L.; Smarr, B.L.; Phung, V.; Terada, M.; Ellisman, M.H.; Pekny, M. Redefining the concept of reactive astrocytes as cells that remain within their unique do-mains upon reaction to injury. Proc. Natl. Acad. Sci. USA 2006, 103, 17513–17518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.-S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef]
  56. Buskila, Y.; Farkash, S.; Hershfinkel, M.; Amitai, Y. Rapid and reactive nitric oxide production by astrocytes in mouse neocortical slices. Glia 2005, 52, 169–176. [Google Scholar] [CrossRef]
  57. Zhao, S.-C.; Ma, L.-S.; Chu, Z.-H.; Xu, H.; Wu, W.-Q.; Liu, F. Regulation of microglial activation in stroke. Acta Pharmacol. Sin. 2017, 38, 445–458. [Google Scholar] [CrossRef]
  58. Hewett, S.I.; Csernansky, C.A.; Choi, D.W. Selective potentiation of NMDA-induced neuronal injury following induction of astrocytic iNOS. Neuron 1994, 13, 487–494. [Google Scholar] [CrossRef]
  59. Garcia-Bonilla, L.; Moore, J.M.; Racchumi, G.; Zhou, P.; Butler, J.M.; Iadecola, C.; Anrather, J. Inducible Nitric Oxide Synthase in Neutrophils and Endothelium Contributes to Ischemic Brain Injury in Mice. J. Immunol. 2014, 193, 2531–2537. [Google Scholar] [CrossRef] [Green Version]
  60. Iadecola, C.; Zhang, F.; Xu, S.; Casey, R.; Ross, M.E. Inducible nitric oxide synthase gene expression in brain following cerebral ischemia. J. Cereb. Blood Flow Metab. 1995, 15, 378–384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Iadecola, C.; Zhang, F.; Casey, R.; Nagayama, M.; Ross, M.E. Delayed reduction of ischemic brain injury and neurological deficits in mice lacking the inducible nitric oxide synthase gene. J. Neurosci. 1997, 17, 9157–9164. [Google Scholar] [CrossRef] [PubMed]
  62. Chan, P.H.; Schmidley, J.W.; Fishman, R.A.; Longar, S.M. Brain injury, edema, and vascular permeability changes induced by oxygen-derived free radicals. Neurology 1984, 34, 315–320. [Google Scholar] [CrossRef] [PubMed]
  63. Gasche, Y.; Fujimura, M.; Morita-Fujimura, Y.; Copin, J.C.; Kawase, M.; Massengale, J.; Chan, P.H. Early appearance of activated matrix metalloproteinase-9 after focal cerebral ischemia in mice: A possible role in blood-brain barrier dysfunction. Int. Soc. Cereb. Blood Flow Metab. 1999, 19, 1020–1028. [Google Scholar] [CrossRef] [Green Version]
  64. Guo, M.; Cox, B.; Mahale, S.; Davis, W.; Carranza, A.; Hayes, K.; Sprague, S.; Jimenez, D.; Ding, Y. Pre-ischemic exercise reduces matrix metalloproteinase-9 expression and ameliorates blood-brain barrier dysfunction in stroke. Neuroscience 2008, 151, 340–351. [Google Scholar] [CrossRef]
  65. Broughton, B.R.S.; Reutens, D.C.; Sobey, C.G. Apoptotic mechanisms after cerebral ischemia. Stroke 2009, 40, e331–e339. [Google Scholar] [CrossRef] [Green Version]
  66. Iadecola, C.; Alexander, M. Cerebral ischemia and inflammation. Curr. Opin. Neurol. 2001, 14, 89–94. [Google Scholar] [CrossRef]
  67. Iadecola, C.; Anrather, J. The immunology of stroke: From mechanisms to translation. Nat. Med. 2011, 17, 796–808. [Google Scholar] [CrossRef]
  68. Lipton, S.A.; Choi, Y.B.; Pan, Z.H.; Lei, S.Z.; Chen, H.S.; Sucher, N.J.; Loscalzo, J.; Singel, D.J.; Stamler, J.S. A redox-based mechanism for the neuroprotective and neurodestructive effects of nitric oxide and related nitroso-compounds. Nature 1993, 364, 626–632. [Google Scholar] [CrossRef] [PubMed]
  69. Marozkina, N.V.; Gaston, B. S-Nitrosylation signaling regulates cellular protein interactions. Biochim. Biophys. Acta Gen. Subj. 2012, 1820, 722–729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Lipton, S.A. Neuronal protection and destruction by NO. Cell Death Differ. 1999, 6, 943–951. [Google Scholar] [CrossRef] [PubMed]
  71. Sha, Y.; Marshall, H.E. S-nitrosylation in the regulation of gene transcription. Biochim. Biophys. Acta Gen. Subj. 2012, 1820, 701–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Chung, H.S.; Murray, C.I.; Venkatraman, V.; Crowgey, E.L.; Rainer, P.P.; Cole, R.N.; Bomgarden, R.D.; Rogers, J.C.; Balkan, W.; Hare, J.M.; et al. Dual Labeling Biotin Switch Assay to Reduce Bias Derived From Different Cysteine Subpopulations: A Method to Maximize S-Nitrosylation Detection. Circ. Res. 2015, 117, 846–857. [Google Scholar] [CrossRef] [Green Version]
  73. Doulias, P.-T.; Greene, J.L.; Greco, T.M.; Tenopoulou, M.; Seeholzer, S.H.; Dunbrack, R.L.; Ischiropoulos, H. Structural profiling of endogenous S-nitrosocysteine residues reveals unique features that accommodate diverse mechanisms for protein S-nitrosylation. Proc. Natl. Acad. Sci. USA 2010, 107, 16958–16963. [Google Scholar] [CrossRef] [Green Version]
  74. Yao, D.; Gu, Z.; Nakamura, T.; Shi, Z.-Q.; Ma, Y.; Gaston, B.; Palmer, L.A.; Rockenstein, E.M.; Zhang, Z.; Masliah, E.; et al. Nitrosative stress linked to sporadic Parkinson’s disease: S-nitrosylation of parkin regulates its E3 ubiquitin ligase activity. Proc. Natl. Acad. Sci. USA 2004, 101, 10810–10814. [Google Scholar] [CrossRef] [Green Version]
  75. Chen, T.; Cao, L.; Dong, W.; Luo, P.; Liu, W.; Qu, Y.; Fei, Z. Protective effects of mGluR5 positive modulators against traumatic neuronal injury through PKC-dependent activation of MEK/ERK pathway. Neurochem. Res. 2012, 37, 983–990. [Google Scholar] [CrossRef]
  76. Qu, Z.W.; Miao, W.Y.; Hu, S.Q.; Li, C.; Zhuo, X.L.; Zong, Y.Y.; Wu, Y.P.; Zhang, G.Y. N-Methyl-D-Aspartate Receptor-Dependent Denitrosylation of Neuronal Nitric Oxide Synthase Increase the Enzyme Activity. PLoS ONE 2012, 7, e52788. [Google Scholar] [CrossRef] [Green Version]
  77. Ravi, K.; Brennan, L.A.; Levic, S.; Ross, P.A.; Black, S.M. S-nitrosylation of endothelial nitric oxide synthase is associated with monomerization and decreased enzyme activity. Proc. Natl. Acad. Sci. USA 2004, 101, 2619–2624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Erwin, P.A.; Lin, A.J.; Golan, D.E.; Michel, T. Receptor-regulated dynamic S-nitrosylation of endothelial nitric-oxide synthase in vascular endothelial cells. J. Biol. Chem. 2005, 280, 19888–19894. [Google Scholar] [CrossRef] [Green Version]
  79. Hara, M.R.; Agrawal, N.; Kim, S.F.; Cascio, M.B.; Fujimuro, M.; Ozeki, Y.; Takahashi, M.; Cheah, J.H.; Tankou, S.K.; Hester, L.D.; et al. S-nitrosylated GAPDH initiates apoptotic cell death by nuclear translocation following Siah1 binding. Nat. Cell Biol. 2005, 7, 665–674. [Google Scholar] [CrossRef] [PubMed]
  80. Li, C.; Feng, J.J.; Wu, Y.P.; Zhang, G.Y. Cerebral ischemia-reperfusion induces GAPDH S-nitrosylation and nuclear translocation. Biochemistry (Moscow) 2012, 77, 671–678. [Google Scholar] [CrossRef] [PubMed]
  81. Gu, Z.; Kaul, M.; Yan, B.; Kridel, S.J.; Cui, J.; Strongin, A.; Smith, J.W.; Liddington, R.C.; Lipton, S.A. S-nitrosylation of matrix metalloproteinases: Signaling pathway to neuronal cell death. Science 2002, 297, 1186–1190. [Google Scholar] [CrossRef] [PubMed]
  82. Tristan, C.; Shahani, N.; Sedlak, T.W.; Sawa, A. The diverse functions of GAPDH: Views from different subcellular compartments. Cell. Signal. 2011, 23, 317–323. [Google Scholar] [CrossRef] [Green Version]
  83. Tanaka, R.; Mochizuki, H.; Suzuki, A.; Katsube, N.; Ishitani, R.; Mizuno, Y.; Urabe, T. Induction of glyceraldehyde-3-phosphate dehydrogenase (GAPDH) expression in rat brain after focal ischemia/reperfusion. J. Cereb. Blood Flow Metab. 2002, 22, 280–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Helton, R.; Cui, J.; Scheel, J.R.; Ellison, J.A.; Ames, C.; Gibson, C.; Blouw, B.; Ouyang, L.; Dragatsis, I.; Zeitlin, S.; et al. Brain-specific knock-out of hypoxia-inducible factor-1alpha reduces rather than increases hypoxic-ischemic damage. J. Neurosci. 2005, 25, 4099–4107. [Google Scholar] [CrossRef] [PubMed]
  85. Barteczek, P.; Li, L.; Ernst, A.-S.; Böhler, L.-I.; Marti, H.H.; Kunze, R. Neuronal HIF-1α and HIF-2α deficiency improves neuronal survival and sensorimotor function in the early acute phase after ischemic stroke. J. Cereb. Blood Flow Metab. 2017, 37, 291–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Lando, D.; Peet, D.J.; Gorman, J.J.; Whelan, D.A.; Whitelaw, M.L.; Bruick, R.K. FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor. Genes Dev. 2002, 16, 1466–1471. [Google Scholar] [CrossRef] [Green Version]
  87. Chen, R.L.; Ogunshola, O.O.; Yeoh, K.K.; Jani, A.; Papadakis, M.; Nagel, S.; Schofield, C.J.I.; Buchan, A.M. HIF prolyl hydroxylase inhibition prior to transient focal cerebral ischaemia is neuroprotective in mice. J. Neurochem. 2014, 131, 177–189. [Google Scholar] [CrossRef] [Green Version]
  88. Semenza, G.L. Hypoxia-inducible factor 1: Master regulator of O2 homeostasis. Curr. Opin. Genet. Dev. 1998, 8, 588–594. [Google Scholar] [CrossRef]
  89. Semenza, G.L.; Wang, G.L. A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol. Cell. Biol. 1992, 12, 5447–5454. [Google Scholar] [CrossRef] [Green Version]
  90. Manalo, D.J.; Rowan, A.; Lavoie, T.; Natarajan, L.; Kelly, B.D.; Ye, S.Q.; Garcia, J.G.N.; Semenza, G.L. Transcriptional regulation of vascular endothelial cell responses to hypoxia by HIF-1. Blood 2005, 105, 659–669. [Google Scholar] [CrossRef]
  91. Semenza, G.L.; Nejfelt, M.K.; Chi, S.M.; Antonarakis, S.E. Hypoxia-inducible nuclear factors bind to an enhancer element located 3′ to the human erythropoietin gene. Proc. Natl. Acad. Sci. USA 1991, 88, 5680–5684. [Google Scholar] [CrossRef] [Green Version]
  92. Khan, M.; Dhammu, T.S.; Dhaindsa, T.S. An NO/GSNO-based Neuroregeneration Strategy for Stroke Therapy. J. Neurol. Neurosci. 2015, 6, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Khan, M.; Khan, H.; Singh, I.; Singh, A.K. Hypoxia inducible factor-1 alpha stabilization for regenerative therapy in traumatic brain injury. Neural Regen. Res. 2017, 12, 696–701. [Google Scholar] [CrossRef]
  94. Khan, M.; Dhammu, T.S.; Baarine, M.; Kim, J.; Paintlia, M.K.; Singh, I.; Singh, A.K. GSNO promotes functional recovery in experimental TBI by stabilizing HIF-1α. Behav. Brain Res. 2016, 340, 63–70. [Google Scholar] [CrossRef]
  95. Kaelin, W.G.; Ratcliffe, P.J. Oxygen sensing by metazoans: The central role of the HIF hydroxylase pathway. Mol. Cell 2008, 30, 393–402. [Google Scholar] [CrossRef] [PubMed]
  96. Ogle, M.E.; Gu, X.; Espinera, A.R.; Wei, L. Inhibition of prolyl hydroxylases by dimethyloxaloylglycine after stroke reduces ischemic brain injury and requires hypoxia inducible factor-1α. Neurobiol. Dis. 2012, 45, 733–742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Moro, M.A.; Alba, J.D.; Leza, J.C.; Lorenzo, P.; Fernández, A.P.; Bentura, M.L.; Boscá, L.; Rodrigo, J.; Lizasoain, I. Neuronal expression of inducible nitric oxide synthase after oxygen and glucose deprivation in rat forebrain slices. Eur. J. Neurosci. 1998, 10, 445–456. [Google Scholar] [CrossRef]
  98. Matrone, C.; Pignataro, G.; Molinaro, P.; Irace, C.; Scorziello, A.; Di Renzo, G.F.; Annunziato, L.; Renzo, G.F.D.; Annunziato, L. HIF-1alpha reveals a binding activity to the promoter of iNOS gene after permanent middle cerebral artery occlusion. J. Neurochem. 2004, 90, 368–378. [Google Scholar] [CrossRef] [PubMed]
  99. Zhang, Z.; Yan, J.; Shi, H. Role of Hypoxia Inducible Factor 1 in Hyperglycemia-Exacerbated Blood-Brain Barrier Disruption in Ischemic Stroke. Neurobiol. Dis. 2016, 95, 82–92. [Google Scholar] [CrossRef] [Green Version]
  100. Li, Q.F.; Xu, H.; Sun, Y.; Hu, R.; Jiang, H. Induction of inducible nitric oxide synthase by isoflurane post-conditioning via hypoxia inducible factor-1α during tolerance against ischemic neuronal injury. Brain Res. 2012, 1451, 1–9. [Google Scholar] [CrossRef] [PubMed]
  101. Baranova, O.; Miranda, L.F.; Pichiule, P.; Dragatsis, I.; Johnson, R.S.; Chavez, J.C. Neuron-specific inactivation of the hypoxia inducible factor 1α increases brain injury in a mouse model of transient focal cerebral ischemia. J. Neurosci. 2007, 27, 6320–6332. [Google Scholar] [CrossRef]
  102. Kunze, R.; Zhou, W.; Veltkamp, R.; Wielockx, B.; Breier, G.; Marti, H.H. Neuron-specific prolyl-4-hydroxylase domain 2 knockout reduces brain injury after transient cerebral ischemia. Stroke 2012, 43, 2748–2756. [Google Scholar] [CrossRef] [Green Version]
  103. Bok, S.; Kim, Y.E.; Woo, Y.; Kim, S.; Kang, S.J.; Lee, Y.; Park, S.K.; Weissman, I.L.; Ahn, G.O. Hypoxia-inducible factor-1a regulates microglial functions affecting neuronal survival in the acute phase of ischemic stroke in mice. Oncotarget 2017, 8, 111508–111521. [Google Scholar] [CrossRef] [Green Version]
  104. Huang, T.; Huang, W.; Zhang, Z.; Yu, L.; Xie, C.; Zhu, D.; Peng, Z.; Chen, J. Hypoxia-inducible factor-1α upregulation in microglia following hypoxia protects against ischemia-induced cerebral infarction. Neuroreport 2014, 25, 1122–1128. [Google Scholar] [CrossRef]
  105. Hsiao, G.; Lee, J.-J.J.; Chen, Y.-C.C.; Lin, J.-H.H.; Shen, M.-Y.Y.; Lin, K.-H.H.; Chou, D.-S.S.; Sheu, J.-R.R. Neuroprotective effects of PMC, a potent α-tocopherol derivative, in brain ischemia-reperfusion: Reduced neutrophil activation and anti-oxidant actions. Biochem. Pharmacol. 2007, 73, 682–693. [Google Scholar] [CrossRef] [PubMed]
  106. Luciano, J.A.; Tan, T.; Zhang, Q.; Huang, E.; Scholz, P.; Weiss, H.R. Hypoxia inducible factor-1 improves the actions of nitric oxide and natriuretic peptides after simulated ischemia-reperfusion. Cell. Physiol. Biochem. 2008, 21, 421–428. [Google Scholar] [CrossRef] [PubMed]
  107. Shi, H. Hypoxia Inducible Factor 1 as a Therapeutic Target in Ischemic Stroke. Curr. Med. Chem. 2009, 16, 4593. [Google Scholar] [CrossRef] [Green Version]
  108. Aminova, L.R.; Chavez, J.C.; Lee, J.; Ryu, H.; Kung, A.; Lamanna, J.C.; Ratan, R.R. Prosurvival and prodeath effects of hypoxia-inducible factor-1alpha stabilization in a murine hippocampal cell line. J. Biol. Chem. 2005, 280, 3996–4003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Wiesener, M.S.; Turley, H.; Allen, W.E.; Willam, C.; Eckardt, K.U.; Talks, K.L.; Wood, S.M.; Gatter, K.C.; Harris, A.L.; Pugh, C.W.; et al. Induction of endothelial PAS domain protein-1 by hypoxia: Characterization and comparison with hypoxia-inducible factor-1alpha. Blood 1998, 92, 2260–2268. [Google Scholar] [CrossRef] [PubMed]
  110. Wiesener, M.S.; Jürgensen, J.S.; Rosenberger, C.; Scholze, C.K.; Hörstrup, J.H.; Warnecke, C.; Mandriota, S.; Bechmann, I.; Frei, U.A.; Pugh, C.W.; et al. Widespread hypoxia-inducible expression of HIF-2alpha in distinct cell populations of different organs. FASEB J. 2003, 17, 271–273. [Google Scholar] [CrossRef] [Green Version]
  111. Brusselmans, K.; Compernolle, V.; Tjwa, M.; Wiesener, M.S.; Maxwell, P.H.; Collen, D.; Carmeliet, P. Heterozygous deficiency of hypoxia-inducible factor-2alpha protects mice against pulmonary hypertension and right ventricular dysfunction during prolonged hypoxia. J. Clin. Investig. 2003, 111, 1519–1527. [Google Scholar] [CrossRef] [PubMed]
  112. Hu, C.-J.; Wang, L.-Y.; Chodosh, L.A.; Keith, B.; Simon, M.C. Differential roles of hypoxia-inducible factor 1alpha (HIF-1alpha) and HIF-2alpha in hypoxic gene regulation. Mol. Cell. Biol. 2003, 23, 9361–9374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Halterman, M.W.; Miller, C.C.; Federoff, H.J. Hypoxia-Inducible Factor-1α Mediates Hypoxia-Induced Delayed Neuronal Death That Involves p53. J. Neurosci. 1999, 19, 6818–6824. [Google Scholar] [CrossRef] [Green Version]
  114. Goda, N.; Ryan, H.E.; Khadivi, B.; McNulty, W.; Rickert, R.C.; Johnson, R.S. Hypoxia-inducible factor 1alpha is essential for cell cycle arrest during hypoxia. Mol. Cell. Biol. 2003, 23, 359–369. [Google Scholar] [CrossRef] [Green Version]
  115. Renton, A.; Llanos, S.; Lu, X. Hypoxia induces p53 through a pathway distinct from most DNA-damaging and stress-inducing agents. Carcinogenesis 2003, 24, 1177–1182. [Google Scholar] [CrossRef]
  116. Zivin, J.A. Factors determining the therapeutic window for stroke. Neurology 1998, 50, 599–603. [Google Scholar] [CrossRef]
  117. Jaffer, H.; Morris, V.B.; Stewart, D.; Labhasetwar, V. Advances in stroke therapy. Drug Deliv. Transl. Res. 2011. [Google Scholar] [CrossRef] [Green Version]
  118. Wung, B.S.; Cheng, J.J.; Shyue, S.K.; Wang, D.L. NO modulates monocyte chemotactic prootein-1 expression in endothelial cells under cyclic strain. Arterioscler. Thromb. Vasc. Biol. 2001, 21, 1941–1947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Begum, N.; Sandu, O.A.; Duddy, N. Negative regulation of Rho signaling by insulin and its impact on actin cytoskeleton organization in vascular smooth muscle cells: Role of nitric oxide and cyclic guanosine monophosphate signaling pathways. Diabetes 2002, 51, 2256–2263. [Google Scholar] [CrossRef] [Green Version]
  120. Hirabayashi, H.; Takizawa, S.; Fukuyama, N.; Nakazawa, H.; Shinohara, Y. Nitrotyrosine generation via inducible nitric oxide synthase in vascular wall in focal ischemia-reperfusion. Brain Res. 2000, 852, 319–325. [Google Scholar] [CrossRef]
  121. Gow, A.; Duran, D.; Thom, S.R.; Ischiropoulos, H. Carbon Dioxide Enhancement of Peroxynitrite-Mediated Protein Tyrosine Nitration. Arch. Biochem. Biophys. 1996, 333, 42–48. [Google Scholar] [CrossRef] [PubMed]
  122. Chandel, N.S.; McClintock, D.S.; Feliciano, C.E.; Wood, T.M.; Melendez, J.A.; Rodriguez, A.M.; Schumacker, P.T. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: A mechanism of O2 sensing. J. Biol. Chem. 2000, 275, 25130–25138. [Google Scholar] [CrossRef] [Green Version]
  123. Sanjuán-Pla, A.; Cervera, A.M.; Apostolova, N.; Garcia-Bou, R.; Víctor, V.M.; Murphy, M.P.; McCreath, K.J. A targeted antioxidant reveals the importance of mitochondrial reactive oxygen species in the hypoxic signaling of HIF-1alpha. FEBS Lett. 2005, 579, 2669–26674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Kikuchi, K.; Tancharoen, S.; Takeshige, N.; Yoshitomi, M.; Morioka, M.; Murai, Y.; Tanaka, E. The efficacy of edaravone (radicut), a free radical scavenger, for cardiovascular disease. Int. J. Mol. Sci. 2013, 14, 13909–13930. [Google Scholar] [CrossRef] [Green Version]
  125. Satoh, K.; Ikeda, Y.; Shioda, S.; Tobe, T.; Yoshikawa, T. Edarabone scavenges nitric oxide. Redox Rep. 2002, 7, 219–222. [Google Scholar] [CrossRef] [PubMed]
  126. Banno, M.; Mizuno, T.; Kato, H.; Zhang, G.; Kawanokuchi, J.; Wang, J.; Kuno, R.; Jin, S.; Takeuchi, H.; Suzumura, A. The radical scavenger edaravone prevents oxidative neurotoxicity induced by peroxynitrite and activated microglia. Neuropharmacology 2005, 48, 283–290. [Google Scholar] [CrossRef]
  127. Watanabe, T.; Yuki, S.; Egawa, M.; Nishi, H. Protective effects of MCI-186 on cerebral ischemia: Possible involvement of free radical scavenging and antioxidant actions. J. Pharmacol. Exp. Ther. 1994, 268, 1597–1604. [Google Scholar]
  128. Feng, S.; Yang, Q.; Liu, M.; Li, W.; Yuan, W.; Zhang, S.; Wu, B.; Li, J. Edaravone for acute ischaemic stroke. Cochrane Database Syst. Rev. 2011, 12, CD007230. [Google Scholar] [CrossRef]
  129. Otomo, E.; Tohgi, H.; Kogure, K.; Hirai, S.; Takakura, K.; Terashi, A.; Gotoh, F.; Maruyama, S.; Tazaki, Y.; Shinohara, Y.; et al. Effect of a novel free radical scavenger, edaravone (MCI-186), on acute brain infarction: Randomized, placebo-controlled, double-blind study at multicenters. Cerebrovasc. Dis. 2003, 15, 222–229. [Google Scholar]
  130. The RANTTAS Investigators. A randomized trial of tirilazad mesylate in patients with acute stroke (RANTTAS). Stroke 1996, 27, 1453–1458. [Google Scholar] [CrossRef]
  131. The Tirilazad International Steering Committee. Tirilazad for acute ischaemic stroke. Cochrane Database Syst. Rev. 2001, CD002087. [Google Scholar] [CrossRef]
  132. Saito, I.; Asano, T.; Sano, K.; Takakura, K.; Abe, H.; Yoshimoto, T.; Kikuchi, H.; Ohta, T.; Ishibashi, S. Neuroprotective effect of an antioxidant, ebselen, in patients with delayed neurological deficits after aneurysmal subarachnoid hemorrhage. Neurosurgery 1998, 42, 269–277. [Google Scholar] [CrossRef] [Green Version]
  133. Ogawa, A.; Yoshimoto, T.; Kikuchi, H.; Sano, K.; Saito, I.; Yamaguchi, T.; Yasuhara, H. Ebselen in acute middle cerebral artery occlusion: A placebo-controlled, double-blind clinical trial. Cerebrovasc. Dis. 1999, 9, 112–118. [Google Scholar] [CrossRef]
  134. Yamaguchi, T.; Sano, K.; Takakura, K.; Saito, I.; Shinohara, Y.; Asano, T.; Yasuhara, H. Ebselen in acute ischemic stroke: A placebo-controlled, double-blind clinical trial. Stroke 1998, 29, 12–17. [Google Scholar] [CrossRef] [PubMed]
  135. Morikawa, E.; Rosenblatt, S.; Moskowitz, M.A. l-Arginine dilates rat pial arterioles by nitric oxide-dependent mechanisms and increases blood flow during focal cerebral ischaemia. Br. J. Pharmacol. 1992, 107, 905–907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Morikawa, E.; Huang, Z.; Moskowitz, M.A. L-Arginine decreases infarct size caused by middle cerebral arterial occlusion in SHR. Am. J. Physiol. Hear. Circ. Physiol. 1992, 263, H1632–H1635. [Google Scholar] [CrossRef] [PubMed]
  137. Morikawa, E.; Moskowitz, M.A.; Huang, Z.; Yoshida, T.; Irikura, K.; Dalkara, T. L-arginine infusion promotes nitric oxide-dependent vasodilation, increases regional cerebral blood flow, and reduces infarction volume in the rat. Stroke 1994, 25, 429–435. [Google Scholar] [CrossRef] [Green Version]
  138. Zhao, H.; Asai, S.; Ishikawa, K. Neither L-NAME nor L-arginine changes extracellular glutamate elevation and anoxic depolarization during global ischemia and reperfusion in rat. Neuroreport 1999, 10, 313–318. [Google Scholar] [CrossRef]
  139. Kurt, T.; Oǧuzhanoǧlu, A.; Ortaç, R.; Turman, B.; Adigüzel, E. Effects of L-arginine on the brain ischaemia-reperfusion damage in rats: An investigation by somatosensory evoked potentials and histopathology. Neurosci. Res. Commun. 2002, 31, 175–182. [Google Scholar] [CrossRef]
  140. Zhao, X.; Ross, M.E.; Iadecola, C. L-arginine increases ischemic injury in wild-type mice but not in iNOS-deficient mice. Brain Res. 2003, 966, 308–311. [Google Scholar] [CrossRef]
  141. Alderton, W.K.; Cooper, C.E.; Knowles, R.G. Nitric oxide synthases: Structure, function and inhibition. Biochem. J. 2001, 357, 593–615. [Google Scholar] [CrossRef] [PubMed]
  142. Schmidt, H.H.; Warner, T.D.; Ishii, K.; Sheng, H.; Murad, F. Insulin secretion from pancreatic B cells caused by L-arginine-derived nitrogen oxides. Science 1992, 255, 721–723. [Google Scholar] [CrossRef]
  143. Takahashi, K.; Yamatani, K.; Hara, M.; Sasaki, H. Gliclazide directly suppresses arginine-induced glucagon secretion. Diabetes Res. Clin. Pract. 1994, 24, 143–151. [Google Scholar] [CrossRef]
  144. Alba-Roth, J.; Müller, O.A.; Schopohl, J.; Von Werder, K. Arginine stimulates growth hormone secretion by suppressing endogenous somatostatin secretion. J. Clin. Endocrinol. Metab. 1988, 67, 1186–1189. [Google Scholar] [CrossRef] [Green Version]
  145. Strasser, A.; McCarron, R.M.; Ishii, H.; Stanimirovic, D.; SpatZ, M. L-arginine induces dopamine release from the striatum in vivo. Neuroreport 1994, 5, 2298–2300. [Google Scholar] [CrossRef] [PubMed]
  146. Li, G.; Regunathan, S.; Barrow, C.J.; Eshraghi, J.; Cooper, R.; Reis, D.J. Agmatine: An endogenous clonidine-displacing substance in the brain. Science 1994, 263, 966–969. [Google Scholar] [CrossRef]
  147. Lortie, M.J.; Novotny, W.F.; Peterson, O.W.; Vallon, V.; Malvey, K.; Mendonca, M.; Satriano, J.; Insel, P.; Thomson, S.C.; Blantz, R.C. Agmatine, a bioactive metabolite of arginine. Production, degradation, and functional effects in the kidney of the rat. J. Clin. Investig. 1996, 97, 413–420. [Google Scholar] [CrossRef] [PubMed]
  148. Sastre, M.; Galea, E.; Feinstein, D.; Reis, D.J.; Regunathan, S. Metabolism of agmatine in macrophages: Modulation by lipopolysaccharide and inhibitory cytokines. Biochem. J. 1998, 330, 1405–1409. [Google Scholar] [CrossRef] [Green Version]
  149. Galea, E.; Regunathan, S.; Eliopoulos, V.; Feinstein, D.L.; Reis, D.J. Inhibition of mammalian nitric oxide synthases by agmatine, an endogenous polyamine formed by decarboxylation of arginine. Biochem. J. 1996, 316, 247–249. [Google Scholar] [CrossRef] [Green Version]
  150. Zhang, R.; Zhang, L.; Zhang, Z.; Wang, Y.; Lu, M.; LaPointe, M.; Chopp, M. A nitric oxide donor induces neurogenesis and reduces functional deficits after stroke in rats. Ann. Neurol. 2001, 50, 602–611. [Google Scholar] [CrossRef]
  151. Zhang, F.; Ladecola, C. Nitroprusside improves blood flow and reduces brain damage after focal ischemia. Neuroreport 1993, 4, 559–562. [Google Scholar] [CrossRef]
  152. Salom, J.B.; Ortí, M.; Centeno, J.M.; Torregrosa, G.; Alborch, E. Reduction of infarct size by the NO donors sodium nitroprusside and spermine/NO after transient focal cerebral ischemia in rats. Brain Res. 2000, 865, 149–156. [Google Scholar] [CrossRef]
  153. Zhang, F.; White, J.G.; Iadecola, C. Nitric oxide donors increase blood flow and reduce brain damage in focal ischemia: Evidence that nitric oxide is beneficial in the early stages of cerebral ischemia. J. Cereb. Blood Flow Metab. 1994, 14, 217–226. [Google Scholar] [CrossRef] [Green Version]
  154. Zhang, F.; Iadecola, C. Reduction of focal cerebral ischemic damage by delayed treatment with nitric oxide donors. J. Cereb. Blood Flow Metab. 1994, 14, 574–580. [Google Scholar] [CrossRef] [Green Version]
  155. Zhuang, P.; Ji, H.; Zhang, Y.H.; Min, Z.L.; Ni, Q.G.; You, R. ZJM-289, a novel nitric oxide donor, alleviates the cerebral ischaemic-reperfusion injury in rats. Clin. Exp. Pharmacol. Physiol. 2010, 37, e121–e127. [Google Scholar] [CrossRef]
  156. Martínez-Murillo, R.; Fernández, A.P.; Serrano, J.; Rodrigo, J.; Salas, E.; Mourelle, M.; Martínez, A. The nitric oxide donor LA 419 decreases brain damage in a focal ischemia model. Neurosci. Lett. 2007, 415, 149–153. [Google Scholar] [CrossRef] [PubMed]
  157. Serrano, J.; Fernández, A.P.; Martínez-Murillo, R.; Alonso, D.; Rodrigo, J.; Salas, E.; Mourelle, M.; Martínez, A. The nitric oxide donor LA 419 decreases ischemic brain damage. Int. J. Mol. Med. 2007, 19, 229–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Butterworth, R.J.; Cluckie, A.; Jackson, S.H.D.; Buxton-Thomas, M.; Bath, P.M.W. Pathophysiological assessment of nitric oxide (given as sodium nitroprusside) in acute ischaemic stroke. Cerebrovasc. Dis. 1998, 8, 158–165. [Google Scholar] [CrossRef] [PubMed]
  159. Thomas, J.E.; Rosenwasser, R.H.; Armonda, R.A.; Harrop, J.; Mitchell, W.; Galaria, I. Safety of intrathecal sodium nitroprusside for the treatment and prevention of refractory cerebral vasospasm and ischemia in humans. Stroke 1999, 30, 1409–1416. [Google Scholar] [CrossRef]
  160. Khan, M.; Jatana, M.; Elango, C.; Singh Paintlia, A.; Singh, A.K.; Singh, I. Cerebrovascular protection by various nitric oxide donors in rats after experimental stroke. Nitric Oxide 2006, 15, 114–124. [Google Scholar] [CrossRef]
  161. Khan, M.; Sekhon, B.; Giri, S.; Jatana, M.; Gilg, A.G.; Ayasolla, K.; Elango, C.; Singh, A.K.; Singh, I. S-Nitrosoglutathione reduces inflammation and protects brain against focal cerebral ischemia in a rat model of experimental stroke. J. Cereb. Blood Flow Metab. 2005, 25, 177–192. [Google Scholar] [CrossRef] [Green Version]
  162. Barakat, W.; Fahmy, A.; Askar, M.; El-Kannishy, S. Effectiveness of arginase inhibitors against experimentally induced stroke. Naunyn. Schmiedebergs. Arch. Pharmacol. 2018, 391, 603–612. [Google Scholar] [CrossRef]
  163. Jung, C.S.; Iuliano, B.A.; Harvey-White, J.; Espey, M.G.; Oldfield, E.H.; Pluta, R.M. Association between cerebrospinal fluid levels of asymmetric dimethyl-L-arginine, an endogenous inhibitor of endothelial nitric oxide synthase, and cerebral vasospasm in a primate model of subarachnoid hemorrhage. J. Neurosurg. 2004, 101, 836–842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Willmot, M.; Gibson, C.; Gray, L.; Murphya, S.; Batha, P.; Murphy, S.; Bath, P. Nitric oxide synthase inhibitors in experimental ischemic stroke and their effects on infarct size and cerebral blood flow: A systematic review. Free Radic. Biol. Med. 2005, 39, 412–425. [Google Scholar] [CrossRef] [PubMed]
  165. Margaill, I.; Allix, M.; Boulu, R.G.; Plotkine, M. Dose- and time-dependence of L-NAME neuroprotection in transient focal cerebral ischaemia in rats. Br. J. Pharmacol. 1997, 120, 160–163. [Google Scholar] [CrossRef] [Green Version]
  166. Fukuda, S.; Hashimoto, N.; Naritomi, H.; Nagata, I.; Nozaki, K.; Kondo, S.; Kurino, M.; Kikuchi, H. Prevention of rat cerebral aneurysm formation by inhibition of nitric oxide synthase. Circulation 2000, 101, 2532–2538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Zhang, F.; Casey, R.M.; Ross, M.E.; Iadecola, C. Aminoguanidine ameliorates and L-arginine worsens brain damage from intraluminal middle cerebral artery occlusion. Stroke 1996, 27, 317–323. [Google Scholar] [CrossRef] [PubMed]
  168. Kidd, G.A.; Hong, H.; Majid, A.; Kaufman, D.I.; Chen, A.F. Inhibition of brain GTP cyclohydrolase I and tetrahydrobiopterin attenuates cerebral infarction via reducing inducible NO synthase and peroxynitrite in ischemic stroke. Stroke 2005, 36, 2705–2711. [Google Scholar] [CrossRef] [Green Version]
  169. Yin, X.; Yan, J.; Hou, X.; Wu, S.; Zgang, G. Neuroprotection of S-nitrosoglutathione against ischemic injury by down-regulating Fas S-nitrosylation and downstream signaling. Neuroscience 2013, 248, 209–298. [Google Scholar] [CrossRef]
  170. Yu, L.-M.; Zhang, T.-Y.; Yin, X.-H.; Yang, Q.; Lu, F.; Yan, J.-Z.; Li, C. Denitrosylation of nNOS induced by cerebral ischemia-reperfusion contributes to nitrosylation of CaMKII and its inhibition of autophosphorylation in hippocampal CA1. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 7674–7683. [Google Scholar]
  171. Khan, M.; Dhammu, T.S.; Qiao, F.; Kumar, P.; Singh, A.K.; Singh, I. S-Nitrosoglutathione Mimics the Beneficial Activity of Endothelial Nitric Oxide Synthase-Derived Nitric Oxide in a Mouse Model of Stroke. J. Stroke Cerebrovasc. Dis. 2019, 28, 104470. [Google Scholar] [CrossRef] [PubMed]
  172. Wei, X.W.; Hao, L.Y.; Qia, S.H. Inhibition on the S-nitrosylation of MKK4 can protect hippocampal CA1 neurons in rat cerebral ischemia/reperfusion. Brain Res. Bull. 2016, 124, 123–128. [Google Scholar] [CrossRef]
  173. Zhao, Q.; Zhang, C.; Wang, X.; Chen, L.; Ji, H.; Zhang, Y. (S)-ZJM-289, a nitric oxide-releasing derivative of 3-n-butylphthalide, protects against ischemic neuronal injury by attenuating mitochondrial dysfunction and associated cell death. Neurochem. Int. 2012, 60, 134–144. [Google Scholar] [CrossRef]
  174. Coert, B.A.; Anderson, R.E.; Meyer, F.B. A comparative study of the effects of two nitric oxide synthase inhibitors and two nitric oxide donors on temporary focal cerebral ischemia in the Wistar rat. J. Neurosurg. 1999, 90, 332–338. [Google Scholar] [CrossRef]
  175. Coert, B.A.; Anderson, R.E.; Meyer, F.B. Effects of the nitric oxide donor 3-morpholinosydnonimine (SIN-1) in focal cerebral ischemia dependent on intracellular brain pH. J. Neurosurg. 2002, 97, 914–921. [Google Scholar] [CrossRef]
  176. Zhang, P.; Guo, Z.; Xu, Y.; Li, Y.; Song, J. N-Butylphthalide (NBP) ameliorated cerebral ischemia reperfusion-induced brain injury via HGF-regulated TLR4/NF-κB signaling pathway. Biomed. Pharmacother. 2016, 83, 658–666. [Google Scholar] [CrossRef] [PubMed]
  177. Jiang, M.H.; Kaku, T.; Hada, J.; Hayashi, Y. 7-Nitroindazole reduces nitric oxide concentration in rat hippocampus after transient forebrain ischemia. Eur. J. Pharmacol. 1999, 380, 117–121. [Google Scholar] [CrossRef]
  178. Pramila, B.; Kalaivani, P.; Anita, A.; Saravana, C. l-NAME combats excitotoxicity and recuperates neurological deficits in MCAO/R rats. Pharmacol. Biochem. Behav. 2015, 134, 246–253. [Google Scholar] [CrossRef]
  179. Nanri, K.; Montécot, C.; Springhetti, V.; Seylaz, J.; Pinard, E. The Selective Inhibitor of Neuronal Nitric Oxide Synthase, 7-Nitroindazole, Reduces the Delayed Neuronal Damage Due to Forebrain Ischemia in Rats. Stroke 1998, 29, 1248–1254. [Google Scholar] [CrossRef] [Green Version]
  180. Ding-Zhou, L.; Marchand-Verrecchia, C.; Croci, N.; Plotkine, M.; Margaill, I. L-NAME reduces infarction, neurological deficit and blood-brain barrier disruption following cerebral ischemia in mice. Eur. J. Pharmacol. 2002, 457, 137–146. [Google Scholar] [CrossRef]
  181. Sun, M.; Zhao, Y.; Gu, Y.; Xu, C. Neuroprotective actions of aminoguanidine involve reduced the activation of calpain and caspase-3 in a rat model of stroke. Neurochem. Int. 2010, 56, 634–641. [Google Scholar] [CrossRef]
  182. Zhu, D.Y.; Liu, S.H.; Sun, H.S.; Lu, Y.M. Expression of inducible nitric oxide synthase after focal cerebral ischemia stimulates neurogenesis in the adult rodent dentate gyrus. J. Neurosci. 2003, 23, 233–239. [Google Scholar] [CrossRef]
  183. Sugimoto, K.; Iadecola, C. Effects of aminoguanidine on cerebral ischemia in mice: Comparison between mice with and without inducible nitric oxide synthase gene. Neurosci. Lett. 2002, 331, 25–28. [Google Scholar] [CrossRef]
  184. Pérez-Asensio, F.J.; Hurtado, O.; Burguete, M.C.; Moro, M.A.; Salom, J.B.; Lizasoain, I.; Torregrosa, G.; Leza, J.C.; Alborch, E.; Castillo, J.; et al. Inhibition of iNOS activity by 1400W decreases glutamate release and ameliorates stroke outcome after experimental ischemia. Neurobiol. Dis. 2005, 18, 375–384. [Google Scholar] [CrossRef]
  185. Zheng, L.; Ding, J.; Wang, J.; Zhou, C.; Zhang, W. Effects and Mechanism of Action of Inducible Nitric Oxide Synthase on Apoptosis in a Rat Model of Cerebral Ischemia-Reperfusion Injury. Anat. Rec. Adv. Integr. Anat. Evol. Biol. 2016, 299, 246–255. [Google Scholar] [CrossRef] [Green Version]
  186. Nagel, S.; Papadakis, M.; Chen, R.; Hoyte, L.C.; Brooks, K.J.; Gallichan, D.; Sibson, N.R.; Pugh, C.; Buchan, A.M. Neuroprotection by dimethyloxalylglycine following permanent and transient focal cerebral ischemia in rats. J. Cereb. Blood Flow Metab. 2011, 31, 132–143. [Google Scholar] [CrossRef] [Green Version]
  187. Prass, K.; Ruscher, K.; Karsch, M.; Isaev, N.; Megow, D.; Priller, J.; Scharff, A.; Dirnagl, U.; Meisel, A. Desferrioxamine induces delayed tolerance against cerebral ischemia in vivo and in vitro. J. Cereb. Blood Flow Metab. 2002, 22, 520–525. [Google Scholar] [CrossRef] [Green Version]
  188. Reischl, S.; Li, L.; Walkinshaw, G.; Flippin, L.A.; Marti, H.H.; Kunze, R. Inhibition of HIF prolyl-4-hydroxylases by FG-4497 reduces brain tissue injury and edema formation during ischemic stroke. PLoS ONE 2014, 9, e84767. [Google Scholar] [CrossRef] [Green Version]
  189. Siddiq, A.; Ayoub, I.A.; Chavez, J.C.; Aminova, L.; Shah, S.; LaManna, J.C.; Patton, S.M.; Connor, J.R.; Cherny, R.A.; Volitakis, I.; et al. Hypoxia-inducible factor prolyl 4-hydroxylase inhibition: A target for neuroprotection in the central nervous system. J. Biol. Chem. 2005, 280, 41732–41743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Harten, S.K.; Ashcroft, M.; Maxwell, P.H. Prolyl hydroxylase domain inhibitors: A route to HIF activation and neuroprotection. Antioxid. Redox Signal. 2010, 12, 459–480. [Google Scholar] [CrossRef]
  191. Zaman, K.; Ryu, H.; Hall, D.; O’Donovan, K.; Lin, K.I.; Miller, M.P.; Marquis, J.C.; Baraban, J.M.; Semenza, G.L.; Ratan, R.R. Protection from oxidative stress-induced apoptosis in cortical neuronal cultures by iron chelators is associated with enhanced DNA binding of hypoxia-inducible factor-1 and ATF-1/CREB and increased expression of glycolytic enzymes, p21(waf1/cip1), and ery. J. Neurosci. 1999, 19, 9821–9830. [Google Scholar] [CrossRef] [Green Version]
  192. Yin, W.; Clare, K.; Zhang, Q.; Volkow, N.D.; Du, C. Chronic cocaine induces HIF-VEGF pathway activation along with angiogenesis in the brain. PLoS ONE 2017, 12, e0175499. [Google Scholar] [CrossRef]
  193. Zhang, B.; Tanaka, J.; Yang, L.; Yang, L.; Sakanaka, M.; Hata, R.; Maeda, N.; Mitsuda, N. Protective effect of vitamin E against focal brain ischemia and neuronal death through induction of target genes of hypoxia-inducible factor-1. Neuroscience 2004, 126, 433–440. [Google Scholar] [CrossRef]
  194. Chern, C.M.; Liou, K.T.; Wang, Y.H.; Liao, J.F.; Yen, J.C.; Shen, Y.C. Andrographolide inhibits PI3K/AKT-dependent NOX2 and iNOS expression protecting mice against hypoxia/ischemia-induced oxidative brain injury. Planta Med. 2011, 77, 1669–1679. [Google Scholar] [CrossRef] [Green Version]
  195. Huang, L.E.; Arany, Z.; Livingston, D.M.; Bunn, H.F. Activation of hypoxia-inducible transcription factor depends primarily upon redox-sensitive stabilization of its alpha subunit. J. Biol. Chem. 1996, 271, 32253–32259. [Google Scholar] [CrossRef] [Green Version]
  196. Jeong, J.W.; Bae, M.K.; Ahn, M.Y.; Kim, S.H.; Sohn, T.K.; Bae, M.H.; Yoo, M.A.; Song, E.J.; Lee, K.J.; Kim, K.W. Regulationand destabilization of HIF-1alpha by ARD1-mediated acetylation. Cell 2002, 111, 709–720. [Google Scholar] [CrossRef] [Green Version]
  197. Piret, J.-P.; Mottet, D.; Raes, M.; Michiels, C. Is HIF-1alpha apro- or an anti-apoptotic protein? Biochem. Pharmacol. 2002, 64, 889–892. [Google Scholar] [CrossRef]
  198. Chang, Y.; Hsieh, C.Y.; Peng, Z.A.; Yen, T.L.; Hsiao, G.; Chou, D.S.; Chen, C.M.; Sheu, J.R. Neuroprotective mechanisms of puerarin in middle cerebral artery occlusion-induced brain infarction in rats. J. Biomed. Sci. 2009, 16, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Chen, C.; Hu, Q.; Yan, J.; Lei, J.; Qin, L.; Shi, X.; Luan, L.; Yang, L.; Wang, K.; Han, J.; et al. Multiple effects of 2ME2 and D609 on the cortical expression of HIF-1α and apoptotic genes in a middle cerebral artery occlusion-induced focal ischemia rat model. J. Neurochem. 2007, 102, 1831–1841. [Google Scholar] [CrossRef] [PubMed]
  200. Bergeron, M.; Gidday, J.M.; Aimee, Y.Y.; Semenza, G.L.; Ferriero, D.M.; Sharp, F.R. Role of hypoxia-inducible factor-1 in hypoxia-induced ischemic tolerance in neonatal rat brain. Ann. Neurol. 2000, 48, 285–296. [Google Scholar] [CrossRef]
  201. Freret, T.; Valable, S.; Chazalviel, L.; Saulnier, R.; Mackenzie, E.T.; Petit, E.; Bernaudin, M.; Boulouard, M.; Schumann-Bard, P. Delayed administration of deferoxamine reduces brain damage and promotes functional recovery after transient focal cerebral ischemia in the rat. Eur. J. Neurosci. 2006, 23, 1757–1765. [Google Scholar] [CrossRef]
  202. Zhao, Y.; Rempe, D.A. Prophylactic neuroprotection against stroke: Low-dose, prolonged treatment with deferoxamine or deferasirox establishes prolonged neuroprotection independent of HIF-1 function. J. Cereb. Blood Flow Metab. 2011, 31, 1412–1423. [Google Scholar] [CrossRef]
  203. Sharp, F.R.; Bergeron, M.; Bernaudin, M. Hypoxia-inducible factor in brain. In Hypoxia. Advances in Experimental Medicine and Biology; Roach, R.C., Wagner, P.D., Hackett, P., Eds.; Springer: Boston, MA, USA, 2001; Volume 502, pp. 273–291. [Google Scholar]
  204. Jones, N.M.; Bergeron, M. Hypoxic preconditioning induces changes in HIF-1 target genes in neonatal rat brain. J. Cereb. Blood Flow Metab. 2001, 21, 1105–1114. [Google Scholar] [CrossRef] [Green Version]
  205. Kovalenko, T.N.; Ushakova, G.A.; Osadchenko, I.; Skibo, G.G.; Pierzynowski, S.G. The neuroprotective effect of 2-oxoglutarate in the experimental ischemia of hippocampus. J. Physiol. Pharmacol. 2011, 62, 239–246. [Google Scholar] [PubMed]
  206. Zhou, J.; Li, J.; Rosenbaum, D.M.; Zhuang, J.; Poon, C.; Qin, P.; Rivera, K.; Lepore, J.; Willette, R.N.; Hu, E.; et al. The prolyl 4-hydroxylase inhibitor GSK360A decreases post-stroke brain injury and sensory, motor, and cognitive behavioral deficits. PLoS ONE 2017, 12, e0184049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Singh, A.; Wilson, J.W.; Schofield, C.J.; Chen, R. Hypoxia-inducible factor (HIF) prolyl hydroxylase inhibitors induce autophagy and have a protective effect in an in-vitro ischaemia model. Sci. Rep. 2020, 10, 1597. [Google Scholar] [CrossRef] [Green Version]
  208. Sakanaka, M.; Wen, T.C.; Matsuda, S.; Masuda, S.; Morishita, E.; Nagao, M.; Sasaki, R. In vivo evidence that erythropoietin protects neurons from ischemic damage. Proc. Natl. Acad. Sci. USA 1998, 95, 4635–4640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Bernaudin, M.; Marti, H.H.; Roussel, S.; Divoux, D.; Nouvelot, A.; MacKenzie, E.T.; Petit, E. A potential role for erythropoietin in focal permanent cerebral ischemia in mice. J. Cereb. Blood Flow Metab. 1999, 19, 643–651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Sirén, A.L.; Fratelli, M.; Brines, M.; Goemans, C.; Casagrande, S.; Lewczuk, P.; Keenan, S.; Gleiter, C.; Pasquali, C.; Capobianco, A.; et al. Erythropoietin prevents neuronal apoptosis after cerebral ischemia and metabolic stress. Proc. Natl. Acad. Sci. USA 2001, 98, 4044–4049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Brines, M.L.; Ghezzi, P.; Keenan, S.; Agnello, D.; De Lanerolle, N.C.; Cerami, C.; Itri, L.M.; Cerami, A. Erythropoietin crosses the blood-brain barrier to protect against experimental brain injury. Proc. Natl. Acad. Sci. USA 2000, 97, 10526–10531. [Google Scholar] [CrossRef] [Green Version]
  212. Sadamoto, Y.; Igase, K.; Sakanaka, M.; Sato, K.; Otsuka, H.; Sakaki, S.; Masuda, S.; Sasaki, R. Erythropoietin prevents place navigation disability and cortical infarction in rats with permanent occlusion of the middle cerebral artery. Biochem. Biophys. Res. Commun. 1998, 253, 26–32. [Google Scholar] [CrossRef]
  213. Gan, Y.; Xing, J.; Jing, Z.; Anne Stetler, R.; Zhang, F.; Luo, Y.; Ji, X.; Gao, Y.; Cao, G. Mutant erythropoietin without erythropoietic activity is neuroprotective against ischemic brain injury. Stroke 2012, 43, 3071–3077. [Google Scholar] [CrossRef] [Green Version]
  214. Zhu, L.; Bai, X.; Wang, S.; Hu, Y.; Wang, T.; Qian, L.; Jiang, L. Recombinant human erythropoietin augments angiogenic responses in a neonatal rat model of cerebral unilateral hypoxia-ischemia. Neonatology 2014, 106, 143–148. [Google Scholar] [CrossRef]
  215. Lee, S.T.; Chu, K.; Sinn, D.I.; Jung, K.H.; Kim, E.H.; Kim, S.J.; Kim, J.M.; Ko, S.Y.; Kim, M.; Roh, J.K. Erythropoietin reduces perihematomal inflammation and cell death with eNOS and STAT3 activations in experimental intracerebral hemorrhage. J. Neurochem. 2006, 96, 1728–1739. [Google Scholar] [CrossRef]
  216. Sun, Y.; Jin, K.; Xie, L.; Childs, J.; Mao, X.O.; Logvinova, A.; Greenberg, D.A. VEGF-induced neuroprotection, neurogenesis, and angiogenesis after focal cerebral ischemia. J. Clin. Investig. 2003, 111, 1843–1851. [Google Scholar] [CrossRef]
  217. Zhang, Z.G.; Zhang, L.; Jiang, Q.; Zhang, R.; Davies, K.; Powers, C.; Van Bruggen, N.; Chopp, M. VEGF enhances angiogenesis and promotes blood-brain barrier leakage in the ischemic brain. J. Clin. Investig. 2000, 106, 829–838. [Google Scholar] [CrossRef] [Green Version]
  218. Amin, N.; Chen, S.; Ren, Q.; Tan, X.; Botchway, B.O.A.; Hu, Z.; Chen, F.; Ye, S.; Du, X.; Chen, Z.; et al. Hypoxia Inducible Factor-1α Attenuates Ischemic Brain Damage by Modulating Inflammatory Response and Glial Activity. Cells 2021, 10, 1359. [Google Scholar] [CrossRef] [PubMed]
  219. Yan, J.; Zhou, B.; Taheri, S.; Shi, H. Differential Effects of HIF-1 Inhibition by YC-1 on the Overall Outcome and Blood-Brain Barrier Damage in a Rat Model of Ischemic Stroke. PLoS ONE 2011, 6, e27798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. El Khashab, I.H.; Abdelsalam, R.M.; Elbrairy, A.I.; Attia, A.S. Chrysin attenuates global cerebral ischemic reperfusion injury via suppression of oxidative stress, inflammation and apoptosis. Biomed. Pharmacother. 2019, 112, 108619. [Google Scholar] [CrossRef]
  221. Yao, Y.; Chen, L.; Xiao, J.; Wang, C.; Jiang, W.; Zhang, R.; Hao, J. Chrysin Protects against Focal Cerebral Ischemia/Reperfusion Injury in Mice through Attenuation of Oxidative Stress and Inflammation. Int. J. Mol. Sci. 2014, 15, 20913–201926. [Google Scholar] [CrossRef] [Green Version]
  222. Chen, W.; Jadhav, V.; Tang, J.; Zhang, J.H. HIF-1α inhibition ameliorates neonatal brain injury in a rat pup hypoxic-ischemic model. Neurobiol. Dis. 2008, 31, 433–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Guo, Y.; Zhou, J.; Li, X.; Xiao, Y.; Zhang, J.; Yang, Y.; Feng, L.; Kang, Y.J. The Association of Suppressed Hypoxia-Inducible Factor-1 Transactivation of Angiogenesis With Defective Recovery From Cerebral Ischemic Injury in Aged Rats. Front. Aging Neurosci. 2021, 13, 648115. [Google Scholar] [CrossRef] [PubMed]
  224. Ehrenreich, H.; Hasselblatt, M.; Dembowski, C.; Cepek, L.; Lewczuk, P.; Stiefel, M.; Rustenbeck, H.H.; Breiter, N.; Jacob, S.; Knerlich, F.; et al. Erythropoietin therapy for acute stroke is both safe and beneficial. Mol. Med. 2002, 8, 495–505. [Google Scholar] [CrossRef] [Green Version]
  225. Tsai, T.H.; Lu, C.H.; Wallace, C.G.; Chang, W.N.; Chen, S.F.; Huang, C.R.; Tsai, N.W.; Lan, M.Y.; Sung, P.H.; Liu, C.F.; et al. Erythropoietin improves long-term neurological outcome in acute ischemic stroke patients: A randomized, prospective, placebo-controlled clinical trial. Crit. Care 2015, 19, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Srivastava, K.; Bath, P.M.W.; Bayraktutan, U. Current therapeutic strategies to mitigate the eNOS dysfunction in ischaemic stroke. Cell. Mol. Neurobiol. 2012, 32, 319–336. [Google Scholar] [CrossRef]
  227. Zhao, X.; Strong, R.; Piriyawat, P.; Palusinski, R.; Grotta, J.C.; Aronowski, J. Caffeinol at the receptor level: Anti-ischemic effect of n-methyl-d-aspartate receptor blockade is potentiated by caffeine. Stroke 2010, 41, 363–367. [Google Scholar] [CrossRef] [Green Version]
  228. Dong, Z.S.W.; Cao, Z.P.; Shang, Y.J.; Liu, Q.Y.; Wu, B.Y.; Liu, W.X.; Li, C.H. Neuroprotection of cordycepin in NMDA-induced excitotoxicity by modulating adenosine A1 receptors. Eur. J. Pharmacol. 2019, 853, 325–335. [Google Scholar] [CrossRef]
  229. Imai, K.; Mori, T.; Izumoto, H.; Takabatake, N.; Kunieda, T.; Watanabe, M. Hyperbaric oxygen combined with intravenous edaravone for treatment of acute embolic stroke: A pilot clinical trial. Neurol. Med. Chir. 2006, 46, 373–378. [Google Scholar] [CrossRef] [Green Version]
  230. Kimura, K.; Aoki, J.; Sakamoto, Y.; Kobayashi, K.; Sakai, K.; Inoue, T.; Iguchi, Y.; Shibazaki, K. Administration of edaravone, a free radical scavenger, during t-PA infusion can enhance early recanalization in acute stroke patients—A preliminary study. J. Neurol. Sci. 2012, 313, 132–136. [Google Scholar] [CrossRef]
  231. Aoki, J.; Kimura, K.; Morita, N.; Harada, M.; Metoki, N.; Tateishi, Y.; Todo, K.; Yamagami, H.; Hayashi, K.; Terasawa, Y.; et al. YAMATO Study (Tissue-Type Plasminogen Activator and Edaravone Combination Therapy). Stroke 2017, 48, 712–719. [Google Scholar] [CrossRef]
  232. Grotta, J.; Combination Therapy Stroke Trial Investigators. Combination therapy stroke trial: Recombinant tissue-type plasminogen activator with/without lubeluzole. Cerebrovasc. Dis. 2001, 12, 258–263. [Google Scholar] [CrossRef]
  233. Montaner, J.; Bustamante, A.; García-Matas, S.; Martínez-Zabaleta, M.; Jiménez, C.; De La Torre, J.; Rubio, F.R.; Segura, T.; Masjuán, J.; Cánovas, D.; et al. Combination of thrombolysis and statins in acute stroke is safe: Results of the STARS randomized trial (Stroke Treatment With Acute Reperfusion and Simvastatin). Stroke 2016, 47, 2870–2873. [Google Scholar] [CrossRef] [PubMed]
  234. Zhu, H.; Chandra, A.; Geng, X.; Cheng, Z.; Tong, Y.; Du, H.; Ding, Y. Low dose concomitant treatment with chlorpromazine and promethazine is safe in acute ischemic stroke. J. Neurosurg. Sci. 2019, 63, 265–269. [Google Scholar] [CrossRef] [PubMed]
  235. Vellimana, A.K.; Milner, E.; Azad, T.D.; Harries, M.D.; Zhou, M.-L.L.; Gidday, J.M.; Han, B.H.; Zipfel, G.J. Endothelial nitric oxide synthase mediates endogenous protection against subarachnoid hemorrhage-induced cerebral vasospasm. Stroke 2011, 42, 776–782. [Google Scholar] [CrossRef] [Green Version]
  236. Chen, S.; Li, N.; Deb-Chatterji, M.; Dong, Q.; Kielstein, J.T.; Weissenborn, K.; Worthmann, H. Asymmetric Dimethyarginine as marker and mediator in Ischemic stroke. Int. J. Mol. Sci. 2012, 13, 15983–16004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The activity of nNOS/eNOS dependent pathways in the brain under physiological conditions: glutamate released from the presynaptic neuron is rapidly uptaken by the astrocyte; nNOS and eNOS are kept at the normal level and both enzymes synthesize NO which activates cGMP synthesis.
Figure 1. The activity of nNOS/eNOS dependent pathways in the brain under physiological conditions: glutamate released from the presynaptic neuron is rapidly uptaken by the astrocyte; nNOS and eNOS are kept at the normal level and both enzymes synthesize NO which activates cGMP synthesis.
Biomolecules 11 01097 g001
Figure 2. Early stage of the ischemia (early neuronal damage (0–6 h after ischemic episode): when the clot is formed due to atherosclerotic plaque, under restriction of oxygen supply, the glutamate released by presynaptic neuron accumulates in the synaptic cleft due to reversed activity of GLT1 transporter in the astrocyte; this leads to overactivation of postsynaptic neuron, overactivation of voltage-dependent calcium channels and Ca2+ influx into postsynaptic neuron, that leads to overactivation of nNOS and its uncoupling, and accumulation of ROS/RNS, instead of cGMP.
Figure 2. Early stage of the ischemia (early neuronal damage (0–6 h after ischemic episode): when the clot is formed due to atherosclerotic plaque, under restriction of oxygen supply, the glutamate released by presynaptic neuron accumulates in the synaptic cleft due to reversed activity of GLT1 transporter in the astrocyte; this leads to overactivation of postsynaptic neuron, overactivation of voltage-dependent calcium channels and Ca2+ influx into postsynaptic neuron, that leads to overactivation of nNOS and its uncoupling, and accumulation of ROS/RNS, instead of cGMP.
Biomolecules 11 01097 g002
Figure 3. Late stage of the ischemia (delayed neuronal damage (begins several hours after episode and reaches maximal level after 24–48 h): parts of dead neurons attract leucocytes and microglia, which accumulate in the place of the stroke releasing cytokines that start to induce the production of iNOS in these cells. Additionally, astrocytes undergo morphological, molecular and functional remodeling in response to injury and also becoming a source of proinflammatory cytokines and iNOS. Thus, the cascade of events triggered by ischemia leads to massive production of cytokines and accumulation of proinflammatory cells with high activity of iNOS. Under these circumstances, in the strong deprivation of tetrahydrobiopterin (BH4) and/or substrate L-arginine, all uncoupled NOS isoforms generate simultaneously high amounts of both NO and O2 favoring production of ONOO. The uncoupling of NOS isoforms can be additionally potentiated by O2 overproduction from other enzymatic sources in the cells accumulated in ischemic lesion, such as upregulated NAD(P)H-oxidase. It finally results in the generation of massive ROS/RNS and a vicious molecular circle is formed which aggravates the neurodegeneration process.
Figure 3. Late stage of the ischemia (delayed neuronal damage (begins several hours after episode and reaches maximal level after 24–48 h): parts of dead neurons attract leucocytes and microglia, which accumulate in the place of the stroke releasing cytokines that start to induce the production of iNOS in these cells. Additionally, astrocytes undergo morphological, molecular and functional remodeling in response to injury and also becoming a source of proinflammatory cytokines and iNOS. Thus, the cascade of events triggered by ischemia leads to massive production of cytokines and accumulation of proinflammatory cells with high activity of iNOS. Under these circumstances, in the strong deprivation of tetrahydrobiopterin (BH4) and/or substrate L-arginine, all uncoupled NOS isoforms generate simultaneously high amounts of both NO and O2 favoring production of ONOO. The uncoupling of NOS isoforms can be additionally potentiated by O2 overproduction from other enzymatic sources in the cells accumulated in ischemic lesion, such as upregulated NAD(P)H-oxidase. It finally results in the generation of massive ROS/RNS and a vicious molecular circle is formed which aggravates the neurodegeneration process.
Biomolecules 11 01097 g003
Table 2. HIF-1α-based therapies. Pretreatment indicates that the compound was administered before experimental ischemia; posttreatment indicates that the compound was administered after the episode. MCAO—middle cerebral artery occlusion, ODG—oxygen and glucose deprivation.
Table 2. HIF-1α-based therapies. Pretreatment indicates that the compound was administered before experimental ischemia; posttreatment indicates that the compound was administered after the episode. MCAO—middle cerebral artery occlusion, ODG—oxygen and glucose deprivation.
CompoundEffect
HIF-1α Stabilizing Agents
PretreatmentDeferoxamine (DFO)-iron chelatorin vivo:
reduced infarct volume in animal models of focal and global ischemia
in vitro:
tolerance against oxygen-glucose deprivation in purified cortical neurons
[101,187,200,201,202,203]
Cobalt chloride (CoCl2)in vivo:
reduced brain injury, and infarct size in newborn rat model of ischemia
[200,203,204]
Prolyl hydroxylase (PHD) inhibitors
PHD1
PHD2
PHD3
in vivo:
reduced brain infarct size, reduced post-stroke neurological deficit, reduced cognitive dysfunction and decreased formation of vasogenic edema in transient or permanent MCAO model in rats or mice
attenuated neuronal death and reactive astrogliosis by 20–50% in transient bilateral common carotid artery occlusion ischemia model in gerbils
in vitro:
reduced cell death (OGD, PC12 cell line and primary rat cortical neurons)
reduced number of early apoptotic cells (OGD, PC12 cell line)
[87,186,188,189,205,206,207]
PosttreatmentProlyl hydroxylase inhibitors
PHD1
PHD2
PHD3
in vivo:
reduced brain infarct size in MCAO model in rats
attenuated neuronal death and reactive astrogliosis by 20–50% in transient bilateral common carotid artery occlusion ischemia model in gerbils
in vitro:
stabilized HIF-1α and up-regulated HIF-1 dependent target genes in primary murine astrocytes and murine cerebrovascular endothelial cell line (bEnd.3) in normal conditions
protective effects in oxygen-glucose deprivation in mouse hippocampal neuronal HT-22 cell line
[188,205]
HIF-1α dependent proteins
PretreatmentErythropoietin (EPO)in vivo:
protection against ischemia-induced cell death in bilateral common carotid artery occlusion in gerbils
reduction of infarct volume in permanent MCAO model in mice or transient MCAO model in rats
[208,209,210,211]
PosttreatmentErythropoietin (EPO), or its analogs (MEPO, S104I-EPO)in vivo:
neuroprotective effect in transient MCAO model in rats up to 6 h posttreatment
enhancement of neuronal survival in stroke-prone hypertensive rats after permanent MCAO
enhancement of angiogenic responses in rat model of neonatal ischemia (permanent right CCAO)
reduced perihematomal inflammation and apoptosis, induced functional recovery and upregulation of eNOS, STAT3, ERK in intracerebral hemorrhage model in rats
in vitro:
reduced NMDA-induced excitotoxicity in primary cortical neurons
reduced cell death after hypoxia in rat primary hippocampal neurons
[211,212,213,214,215]
Vascular endothelial growth factor (VEGF)in vivo:
reduced infarct size, enhanced neurogenesis, angiogenesis, cerebral microvascular perfusion and neurological dysfunction in MCAO and focal cerebral embolic ischemia in rats
[216,217]
Direct and indirect HIF-1α inhibitors
PretreatmentAcriflavinein vivo:
increased neurological deficit in endothelin-1-induced focal cerebral ischemia in mice
no influence in infarct volume, number of neurons, number of IL-10-positive cells, iNOS expression and pro-inflammatory cytokines endothelin-1-induced focal cerebral ischemia in mice
decreased expression of HIF-1α and increased expression of NF-κB endothelin-1-induced focal cerebral ischemia in mice
increased number of GFAP-positive cells and GFAP expression endothelin-1-induced focal cerebral ischemia in mice
[218]
2,2,5,7,8-Pentamethyl-6-hydroxychromane (PMC)in vivo:
reduced infarct volume in MCAO model in rats
reduced caspase-3 activation, reduced HIF-1α, iNOS and nitrotyrosine expression in MCAO model in rats
[105]
YC-1in vivo:
reduced HIF-1α, VEGF, EPO and GLUT-3 expression in MCAO model in rats
increased mortality, infarct size and edema in MCAO model in rats
reduced BBB permeability in MCAO model in rats
[219]
Chrysinin vivo:
reduced oxidative stress biomarkers in hippocampus in bilateral carotid artery occlusion in rats
reduced TNF-α, IL-6, BAX and Hsp90 level in hippocampus in bilateral carotid artery occlusion in rats
increased IL-10 and Bcl-2 level in hippocampus in bilateral carotid artery occlusion in rats
increased aspartate and glutamate level in hippocampus in bilateral carotid artery occlusion in rats
reduced infarct volume and neurological deficit in MCAO model in mice
reduced SOD activity and malondialdehyde level in MCAO model in mice
reduced number iNOS-, COX2- and NF-κB-positive cells and their protein expression in MCAO model in mice
reduced number of GFAP- and Iba-1-positive cells in MCAO model in mice
[220,221]
Posttreatment2ME2in vivo:
reduced infarct volume and edema in neonatal model of ischemia or MCAO model in rats
reduced HIF-1α and VEGF expression in neonatal model of ischemia
reduced mortality in MCAO model in rats
increased neurological score in MCAO model in rats
reduced HIF-1α expression in MCAO model in rats
reduced density of HIF-1α-, VEGF-, BNIP3- and caspase-3-positive cells in MCAO model in rats
[199,222]
D609in vivo:
reduced infarct size and mortality in MCAO model in rats
increased neurological score in MCAO model in rats
reduced density of HIF-1α expression in MCAO model in rats
reduced HIF-1α-, VEGF-, BNIP3- and caspase-3 positive cells in MCAO model in rats
[199]
Chetominin vivo:
increased infarct size and neurological deficit in MCAO model in rats
[223]
Clinical trials
rhEPO
(double-blind placebo controlled proof-of-concept trial; i.v.)
better clinical recovery
normalization of circulating marker of injury (S100β)
[224]
EPO
(prospective, randomized, placebo-controlled trial; s.c.)
improved long-term clinical outcome
[225]
Table 3. Combined treatment therapies with the use of NO-related compounds in clinical trials. SOC—standard of care.
Table 3. Combined treatment therapies with the use of NO-related compounds in clinical trials. SOC—standard of care.
Mode of ActionCombined TreatmentEffect
Free radical scavengerEdaravone + Hyperbaric oxygen + Heparin
(pilot trial)
-
reduction of neurological symptoms (NIHSS)
[229]
Edaravone + t-PA
(pilot trial)
-
early recanalization
-
higher percent of remarkable recovery (≥8-point reduction on NIHSS)
[230]
Edaravone + t-PA
(multicenter, prospective, randomized and open-label trial)
-
no effect on the rate of early recanalization, symptomatic intracerebral hemorrhage or favorable outcome after tPA therapy
[231]
NMDA antagonist, VGSC and VGCC blocker, inhibitor of NO synthesisLubeluzole + t-PA
(feasibility, safety and efficacy trial—uncompleted)
-
study was terminated after Lubeluzole’s phase III trial showed no overall improvement
[232]
StatinSimvastatin + t-PA
(phase IV, prospective, randomized, double-blind, placebo-controlled trial)
-
higher proportion of patients undergo major neurological recovery (post hoc analysis)
[233]
Typical antipsychotic
+ First-generation antihistamine
Chlorpromazine + Promethazine + SOC
(pilot trial)
-
no effect on neurological symptoms (NIHSS, mRS)
[234]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wierońska, J.M.; Cieślik, P.; Kalinowski, L. Nitric Oxide-Dependent Pathways as Critical Factors in the Consequences and Recovery after Brain Ischemic Hypoxia. Biomolecules 2021, 11, 1097. https://doi.org/10.3390/biom11081097

AMA Style

Wierońska JM, Cieślik P, Kalinowski L. Nitric Oxide-Dependent Pathways as Critical Factors in the Consequences and Recovery after Brain Ischemic Hypoxia. Biomolecules. 2021; 11(8):1097. https://doi.org/10.3390/biom11081097

Chicago/Turabian Style

Wierońska, Joanna M, Paulina Cieślik, and Leszek Kalinowski. 2021. "Nitric Oxide-Dependent Pathways as Critical Factors in the Consequences and Recovery after Brain Ischemic Hypoxia" Biomolecules 11, no. 8: 1097. https://doi.org/10.3390/biom11081097

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