Next Article in Journal
Slow Evolution toward “Super-Aggregation” of the Oligomers Formed through the Swapping of RNase A N-Termini: A Wish for Amyloidosis?
Next Article in Special Issue
Circulating U13 Small Nucleolar RNA as a Potential Biomarker in Huntington’s Disease: A Pilot Study
Previous Article in Journal
Metal Cluster Triggered-Assembling Heterogeneous Au-Ag Nanoclusters with Highly Loading Performance and Biocompatible Capability
Previous Article in Special Issue
Microglial CD74 Expression Is Regulated by TGFβ Signaling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuro-Inflammation Modulation and Post-Traumatic Brain Injury Lesions: From Bench to Bed-Side

1
Unité de Neuroanesthésie-Réanimation, Hôpital de la Pitié Salpêtrière 43-87, Boulevard de l’Hôpital, F-75013 Paris, France
2
Inserm, Maladies Neurodéveloppementales et Neurovasculaires, Université Paris Cité, F-75019 Paris, France
3
Division of Anaesthesia, Addenbrooke’s Hospital, University of Cambridge, Box 93, Hills Road, Cambridge CB2 2QQ, UK
4
Department of Neuroscience, Istituto di Ricerche Farmacologiche Mario Negri IRCCS, 20156 Milan, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(19), 11193; https://doi.org/10.3390/ijms231911193
Submission received: 25 August 2022 / Revised: 14 September 2022 / Accepted: 15 September 2022 / Published: 23 September 2022

Abstract

:
Head trauma is the most common cause of disability in young adults. Known as a silent epidemic, it can cause a mosaic of symptoms, whether neurological (sensory–motor deficits), psychiatric (depressive and anxiety symptoms), or somatic (vertigo, tinnitus, phosphenes). Furthermore, cranial trauma (CT) in children presents several particularities in terms of epidemiology, mechanism, and physiopathology—notably linked to the attack of an immature organ. As in adults, head trauma in children can have lifelong repercussions and can cause social and family isolation, difficulties at school, and, later, socio-professional adversity. Improving management of the pre-hospital and rehabilitation course of these patients reduces secondary morbidity and mortality, but often not without long-term disability. One hypothesized contributor to this process is chronic neuroinflammation, which could accompany primary lesions and facilitate their development into tertiary lesions. Neuroinflammation is a complex process involving different actors such as glial cells (astrocytes, microglia, oligodendrocytes), the permeability of the blood–brain barrier, excitotoxicity, production of oxygen derivatives, cytokine release, tissue damage, and neuronal death. Several studies have investigated the effect of various treatments on the neuroinflammatory response in traumatic brain injury in vitro and in animal and human models. The aim of this review is to examine the various anti-inflammatory therapies that have been implemented.

1. Introduction

Traumatic brain injury (TBI) is defined as an alteration in brain function caused by an external force [1]. However, disagreements persist regarding this definition, which may account for the literature’s heterogeneity regarding its epidemiology [2,3,4]. TBI is one of the main causes of death and disability in young people, and its incidence is estimated at 69 million per year worldwide [5]. Currently, the incidence of TBI in Europe is estimated at 9.3 million annual cases, increased from 3.7 million in 2004. In France, this figure is around 150 cases per 100.000 inhabitants [6] but it is probably underestimated because not all patients coming to the emergency room are hospitalized [7]. Also in France, the overall incidence of severe TBI has been steadily decreasing for thirty years: 24/100,000 in 1986, 17/100,000 in 1996 and 3/100,000 in 2007 [8]. This decrease is accounted for primarily by patients under 55 years of age; conversely, there is an increase in the incidence of TBI in people over 75 years of age [9]. TBI severity is typically stratified by Glasgow Coma Scale (GCS) on admission, and in turn GCS has a role in predicting outcome, although much smaller than might be expected, suggesting that factors other than the initial injury severity are important determinants in outcome, and may be therapeutic targets. [10,11].
A severe TBI should never be assumed to be isolated; indeed, two descriptive studies demonstrate the ubiquity of associated injuries: 70% with fractures of the extremities, 35% thoracic trauma, 20% abdominal trauma, and 8% cervical spine trauma [12,13]. The neuroinflammatory response to TBI therefore occurs in the context of significant peripheral inflammation, which can in turn exacerbate neuroinflammation [14]. All studies agree that the risk of TBI is higher in men regardless of age. The incidence of TBI varies by age, with a trimodal distribution: young children (0–4 years), adolescents and young adults (15–24 years) and the elderly (>65 years). Many studies have looked at the causes of TBI. It seems evident that the aetiologies depend on sex and age groups and are also correlated with the severity of the trauma. Some studies have focused on a certain type of cause other than TBI associated with sport or military activities. The injury mechanisms are mainly linked to falls (extreme ages) and road traffic accidents (RTA) [15]. Alcohol intake and/or the use of illicit substances is found in 60% of severe TBI.
The pathophysiology of TBI is complex and, in addition to the damage resulting from direct physical injury, pressure-effects and ischaemia, involves a constellation of different intra- or extra-cerebral immunological cellular actors interacting via multiple small molecules (cytokines, or chemokines) through a permeabilized blood-brain barrier (BBB) thus generating an inflammatory storm shown schematically in Figure 1. Historically, two types of post-traumatic injuries have been described: immediate primary injuries related to the impact itself, and secondary injuries that appear within hours and months of trauma. It is only recently that a third type of lesion has been discussed: the so-called late tertiary lesions. Primary lesions result directly from the impact, and in particular from what is classically called “container-content conflict”. There are three physical mechanisms: direct contact trauma, the most frequent; the phenomena of acceleration or deceleration called inertia effect, responsible for disseminated lesions; and finally the static compressive mechanism called crushing, the rarest. The resulting lesions can be focal and multifocal. The focal are mainly represented by cortico-subcortical cerebral contusions, haemorrhages of venous origin most often resulting from the direct impact of the brain against the skull, and different variations of hematoma: epidural (EDH), subdural (SDH), intraparenchymal (IP), or subarachnoid hemorrhage (SAH). These lesions are often very well visualized in the acute phase on computed tomography (CT) scan. Contusions are present under the impact zone (direct contusion by blow) or at a distance (indirect contusion by backlash), and most typically occur in the frontal and temporal lobes. The multifocal or diffuse lesion subtype is also known as diffuse axonal injury (DAI)—little or not at all detectable with conventional neuroradiology techniques, but visualizable in post-mortem histopathology. They are mainly linked to acceleration-deceleration movements leading to shearing and stretching effects, and are found mainly in areas of lower resistance of axons (transition areas between gray matter and white matter) [16], especially in the corpus callosum, the subcortical white matter, and the brainstem [17]. These lesions are frequent in cases of high kinetic trauma in young subjects. In 2010, Skandsen et al. observed DAI in 72% of patients with moderate to severe TBI, associated in 50% of cases with focal lesions. They are a major cause of persistent coma and vegetative state [18,19] and in survivors they are responsible of significant functional and motor sequelae [20].
Secondary lesions can appear within the first few minutes after the TBI and will worsen the initial lesions. They develop in areas of cerebral parenchyma still viable but weakened by primary cerebral insult, and are mediated by factors of systemic or central origin, usually concomitantly. Arterial hypotension, hypoxia, hypercapnia, anemia, as well as some metabolic and fluid electrolyte disturbances are all systemic factors contributing to the development of secondary lesions [21,22,23]. At the cellular level, all of these secondary attacks cause neuronal death via multiple mechanisms with an inflammatory storm linked to both a peripheral and local immune reaction in the central nervous system (CNS) [24,25,26,27]. This inflammatory reaction breaks down into two parts: first, a central intracerebral inflammatory response with in particular glial activation, then an added systemic immune response.
TBI induces immediate neuropathological effects which may be transient in the less severe forms. However, with increasing severity, it exacerbates neuronal damage by degeneration mechanisms. The latter, described in the literature, may occur remotely from the acute phase. The pathophysiological mechanisms of these so-called tertiary lesions are still uncertain but would involve processes common to certain dementias and inflammatory diseases of the CNS leading to neurodegeneration. The two main pathological arguments in favor of this analogy are cerebral atrophy and cerebral inflammation: a postmortem study conducted by Johnson et al. demonstrated a significant decrease in the thickness of the corpus callosum in 22 subjects who had been victims of a TBI more than a year ago, compared to control subjects; microglial reactivity is also observed in 30% of patients after one year following TBI, indicating persistent brain inflammation [28].
Tertiary lesions are thought to influence the long-term prognosis and constitute a clinical entity called chronic post-traumatic encephalopathy (Figure 2).
Severe TBI has long been considered an exclusively acute clinical entity, and the notion of new brain lesions occurring long after the initial impact was not introduced until the 1920s [29]. This hypothesis was already mentioned at the time but it was probably experienced as an inevitability, but over the last thirty years, several studies have focused on better understanding the mechanisms involved in these lesions in order to identify therapeutic targets.
To date, the phenomenon previously stated as pugilistic dementia has been redefined as post-traumatic encephalopathy (PTE), a pathology that has been widely studied [30]. Historically, it was initially described in high performance athletes [31,32,33,34,35,36] but it is now known that it can develop in any patient who has undergone a single TBI. PTE can result in disorders of attention, memory, and concentration with significant impact on patient social functioning and quality of life. PTE can progress to Alzheimer-like dementia in the decades following the trauma. Anatomically, it is characterized by the deposition of hyperphosphorylated TAU protein in neurofibrillary tangles, most often in the perivascular spaces, the depths of the cortical grooves, and the subpial and periventricular areas.
Recently, a new nosological entity called post-concussion syndrome (PCS) has been described [37]. Historically, this entity has been confused with post-traumatic stress disorder (PTSD), though it has clear clinical distinctions. Indeed, complaints reported by TBI patients include not only a constellation of psychiatric symptoms that may overlap with PTSD, such as anxiety, aggressiveness, emotional lability, sleep and eating disorders, memory and attention disorders, and difficulty concentrating, but also somatic symptoms such as headache, dizziness with tinnitus or phosphenes. The Paris-TBI study, which examined the long term outcome of TBI patients treated in Ile-de-France between 2005 and 2007, showed that between 1 and 4 years after the TBI, 39% of the 147 patients had improved clinically while 43% remained stable and 15% had worsened. Thus at 4 years of the trauma only 28% had recovered, while 40% suffered moderate handicap and 32% severe handicap [38]. At present, the literature does not explain why some patients will develop PCS and others will not, or why some worsen later. Even if some genetic susceptibilities may exist, neuroinflammation could be one of the main explanations for this clinical expression.
In addition, the clinical course after TBI does not appear to be linear. Indeed, some studies observe a rapid improvement in the first months, followed by a slower progression and then a plateau phase with, in the most severe cases, an absence of return to the premorbid state [39,40,41]. TBI in children, apart from anatomical differences, also has some clinical specificities with a particularly serious impact in the long term.
The management of the acute phase of head trauma is currently the subject of multiple medical recommendations with both multimodal monitoring tools (intracranial pressure, microdialysis) and therapies to control secondary lesions and more specifically intracranial hypertension (osmotherapies, coma, hypothermia, even surgical treatment) [42,43]. These measures significantly reduce the acute morbidity and mortality linked to head trauma, but unfortunately do not improve post-traumatic disability. The latter results from the accumulation of insults in the acute phase and particularly from neuroinflammation. The main objective below is to present and discuss the various neuro-antiinflammatory treatments that have been studied in the literature over the years (Table 1).

2. Neuro-Anti-Inflammatory Therapeutics

In this first part, we discussed anti-inflammatory therapeutics such as glucocorticoids and non-steroidal anti-inflammatory drugs and also drugs such as melatonin, cyclosporine, oxytocin, statins, and erythropoietin, whose initial mechanism of action is not anti-inflammatory but which can, in certain contexts—similar to TBI—have an anti-inflammatory activity.

2.1. Glucocorticoids

Glucocorticoids (GC) havethe ability to act on various neuroinflammatory mechanisms. They are also known to reduce vasogenic edema but on the other hand worsen cytotoxic edema [44]. In addition, GCs have the potential to act on all three phases of post-traumatic injury (TBI), so their use in TBI seems particularly attractive. Derived from the natural hormone cortisol, synthetic GCs are drugs developed to maximize glucocorticoid effects and minimize mineralocorticoid effects. Synthetic derivatives that have appeared on the market vary in anti-inflammatory efficacy, half-life, and mineralocorticoid action, but all have structural similarity and a common mechanism of action: they circulate bound to transport proteins, with a small fraction of pharmacologically-active unbound form. This free fraction crosses the cell membrane and binds, with high affinity, to specific cytosolic receptors called nuclear glucocorticoid receptors (GR) which can then enter the nucleus.
GCs are steroidal anti-inflammatory drugs. Cortisol, also called hydrocortisone, has glucocorticoid properties (particularly anti-inflammatory) and mineralocorticoid properties (anti-diuretic, anti-natriuretic, and kaliuretic). Their main property is immunomodulation. First, corticosteroids reduce the synthesis of many pro-inflammatory cytokines and chemokines (TNF-α, IL-1ß, IL-2, IL-6, IL-8, IL-4, IL-5, IL-12, IL-17, IL-18, GM-CSF…). They also induce the synthesis of anti-inflammatory lymphocyte cytokines such as IL-10 and TGF-ß [45].
In addition, corticosteroids directly inhibit the synthesis of multiple inflammatory enzymes, such as inducible NO synthetase (iNOS), phospholipases A2 and C, but also cyclooxygenase 2 [3,4]. Many other enzymes involved in the phenomena of cell destruction (proteases, collagenases) or in inflammatory phenomena (C3 convertase) are also inhibited.
Corticosteroids also act on many target cells involved in innate or adaptive immunity [46]: macrophages, dendritic cells, polymorphonuclear cells, and T and B lymphocytes in particular.
They act by controlling their maturation, regulating their activation, modulating their capacity for synthesis (cytokines, chemokines, enzymes, etc.), managing their survival and migration, and by modifying their “learning”, in particular for intrathymic lymphocytes [47]. They can reduce release of lysosomal enzymes and preformed granules containing inflammation mediators (histamine, serotonin). This partly explains the inhibition of cellular activity seen with corticosteroids, especially for immune cells such as lymphocytes. Finally, the effectiveness of corticosteroids is directly related to the cytosolic concentration of the receptor of the GCs (RGCs) available. However, the affinity and cytosolic concentration of GRs are genetically regulated. A particular polymorphism of RGCs could explain an increased sensitivity to corticosteroids in some patients [48].
Finally, corticosteroids inhibit peroxidation and lipid hydrolysis. Effects on the maintenance of aerobic energy metabolism, intracellular accumulation of calcium, and the preservation of cerebral blood flow have also been attributed to them. Experimental data supports neuroprotective action of corticosteroids in models of TBI [49].
Corticosteroids have been used effectively for several years in inflammatory neurological conditions in humans. They are indicated in multiple sclerosis, oncology, and postoperative neurosurgery to reduce peritumoral edema [50]. Concerning TBI, a survey in England carried out in 1996 shows that these drugs were used in 14% of units with TBI patients [51]. Unfortunately, however, the literature is not definitive regarding their effectiveness: meta-analyses show that the work to date is too heterogeneous in terms of patient demographics, doses, corticosteroid type, and timing and duration of treatment [52,53,54,55].

2.2. Nonsteroidal Anti-Inflammatory Drugs

Nonsteroidal anti-inflammatory drugs (NSAIDs) are non-selective COX and selective COX2 inhibitors. Tested on several models of TBI, they reduce neuroinflammation by inhibiting the production of IL-1ß, IL-6 and IL-10, but do not reduce tissue damage and functional consequences of TBI [56].

2.3. Statins

Statins are HMGCoA reductase (hydroxy-methyl-glutaryl-coenzyme A reductase) inhibitors and form a class of lipid-lowering drugs commonly used to control cardiovascular risk factors. In addition to this beneficial effect on lowering cholesterol, statins have an anti-inflammatory effect by reducing oxidative stress through up-regulation of endothelial NO-synthase. However, other signaling pathways also intervene in parallel with the NO-synthase one, such as tissue plasminogen activator and the phosphoinositide 3-kinase (PI3K)/serine-threonine kinase (AKT) pathway.
Statins appear to be a potentially interesting neuroprotective strategy, especially during stroke, subarachnoid hemorrhage, and possibly TBI [57,58,59]. Some authors also show that statin treatment likely not only promotes hippocampal neurogenesis in the dentate gyrus but also improves learning [60]; and possibly reduces apoptosis [59]. Statins can decrease apoptotic cell death and promote neuron survival by suppressing Caspase-3 activity [61,62,63] and by reducing the Bax/Bcl-2 ratio [60]. Furthermore, acute statin treatment attenuates microglial activation and polarization after TBI in rodents [61,64,65]. In addition to protecting pre-existing neurons, statins foster neurogenesis with growth and neuronal differentiation particularly in the hippocampus probably due to an upregulation of neurotrophic factors like brain- derived neurotrophic factors or vascular endothelial growth factor [60,66,67]. Statins also have vascular and endothelial effects. In a murine model of TBI, atorvastatin decreased the level of delayed thrombosis, and this was correlated with a reduction in necrotic brain tissue [68]. Multiple mechanisms are probably associated, but pre-clinical and clinical studies have demonstrated that statin treatment can decrease pro-thrombotic markers like von Willebrand factor [69]. Statins have also been shown to promote angiogenesis in TBI models, with an increase in newly formed vessels and capillary density and VEGF (Vascular Endothelial Growth Factor) levels [66]. The effect of statins could also be anti-inflammatory via decreased expression of pro-inflammatory markers such as TLR (Toll like receptor) 4, NFκB, and IL(interleukin)-1ß or IL-6 [70,71,72]. Statins can also lower microglial activation [64], also probably due to the effect of statin on harmful oxygen free radicals, such as superoxide production [73]. This decrease of pro-inflammatory markers is associated with blood brain barrier (BBB) integrity maintenance [72,74,75]. In a systematic review published in 2021, the authors showed four randomized clinical trials with 296 patients demonstrating that statins can play a neuroprotective role and improve cognitive outcomes by anti-inflammatory effect, for example, in association with lower tumor necrosis factor and c-reactive protein [76]. In summary, statins appear to be a good candidate for further studies on improving cognitive outcome after traumatic brain injury.

2.4. Melatonin

Melatonin (N-acetyl-5methoxytryptamine) is a hormone produced by the pineal gland. It has many functions that are particularly interesting in the context of brain injury [77].
Recently Osier et al. published a review on published evidence of therapeutic mechanisms of melatonin in TBI [77]. It acts as a powerful antioxidant by promoting the elimination of reactive oxygen species (ROS) by trapping and reducing the synthesis of iNOS and nNOS while increasing that of antioxidant enzymes [78]. These antioxidant capacities are confirmed in models of cerebral ischemia, in which melatonin reduced the size of the ischemic lesion and generated an anti-inflammatory response [79,80,81,82]—findings also observed in the context of TBI [83,84,85]. Melatonin also appears to have mitochondrial protective properties by helping to maintain their structure and function [86,87]. One of the other hypothesis is that melatonin could overcome energy depletion via the adenosine 3′5′monophosphate (AMP)-c/p- cAMPc-response element binding (CREB) pathway; in this worked melatonin treatment also decreased apoptotic cell death, lesion volume, and promote post-TBI motor coordination and work memory [88]. Furthermore, melatonin was reported to reduce neuroinflammation and brain edema, decrease late-phase activation of NFkB [89] and attenuate acute microglial and astrocyte activation [90].
Clinically, studies have shown mixed and conflicting results on the effectiveness of melatonin for the treatment of TBI. In pediatric populations, melatonin appears to demonstrate neuroprotective effects [91]: PLAYGAME, a randomized controlled trial, tested different doses of melatonin on children ages 13 to 18 with post-concussion syndrome symptoms [92]: in addition to anti-neuroinflammatory effects, melatonin seems to modulate neurobehavioral and particularly sleep cycle difficulties, and indirectly cognition. However, in a double-blind randomized crossover trial with post-TBI adults suffering from sleep disorders, melatonin supplementation did not improve patients’ sleep or neuropsychiatric well-being [93]. In contrast, another randomized double blind placebo-controlled study found melatonin supplementation over a 4-week period safely and effectively improved subjective sleep quality evaluated by the global Pittsburgh Sleep Quality Index score [94].
Melatonin appears to be a potentially useful anti-inflammatory agent in TBI, particularly due to its anti-oxidant effects. Its absence of toxicity allows for clinical studies. Clinical trials are currently discordant but it could be that melatonin can act on post-TBI sleep disorders. The effects of melatonin appear to be through indirect mechanisms but some studies have looked at other mechanisms that could be specific to melatonin and its receptors [95].

2.5. Minocycline

Minocycline is an antibiotic from the cycline family, the second-generation tetracyclines, and has been demonstrated to protect against neonatal hypoxic-ischemic injury in a rodent model [96]. It also exhibits numerous neuroprotective effects in different animal models of TBI, such as inhibition of microglial activation [97,98] and reduction of cytokine production, including IL-1β and IL-6 as well as chemokines CCL4, CXCL1 and CXCL2 [99]. It reduces the production of nitric oxide [100], and inhibits the excitotoxic N-methyl-D-aspartic acid pathway [101] and ROS, with an associated decrease in cerebral edema and lesion volume.
In a mild blast-induced model of TBI, minocycline treatment normalized tissue levels of inflammatory (CRP), vascular, neuronal (Neuron Specific Enolase [NSE], tau), and glial markers (Glial fibrillary acidic protein [GFAP], S100B) [102]. Others studies have shown negligible effect, however, a 2018 study on neonatal rats found minocycline was ineffective in reducing microglial/macrophage activation and ameliorating post-injury deficit by postnatal day 11 [103]. In another study that similarly found no effect of minocycline, the authors tested two time points for treatment: 1 h or 9 weeks post-injury, with no effect on lesion size or degree of microglial activation after either the early or the late administration of minocycline [104].
Some studies show that minocycline plus N-acetylcysteine (NAC) synergistically improve cognition and memory, modulating neuroinflammation and preventing oligodendrocyte loss [105], preserving myelin, limiting lesion volume [106,107], and promoting remyelination [108]. However, these results seem not to apply in clinical practice as thus far, minocycline has not been shown to benefit neurological outcomes [109]. Recently a clinical study examined the effect of minocycline in the chronic phase of TBI: fifteen patients received either minocycline or placebo at least 6 months after TBI, and the authors observed that while minocycline treatment reduced chronic microglial activation by PET, it increased plasma concentrations of neurofilament light, a marker of neurodegeneration [110]. In another clinical study—the phase IIa open label safety and feasibility study for preliminary data on functional outcomes – Meythaler et al. demonstrated a trend in neurologic improvement for the higher dose of minocycline, but this did not reach statistical significance. These two clinical studies are limited, however, in that they are underpowered. Further higher-powered studies may demonstrate clinical efficacy, but for the moment, minocycline has not been shown to improve neurological outcomes in TBI.

2.6. Cyclosporin

Cyclosporin is a potent immunomodulator whose therapeutic use dates from the early 1980s for prevention of organ transplant rejection; its main mode of action lies in inhibiting the production of cytokines that regulate and activate T lymphocytes, particularly IL-2 [111]. Cyclosporin A acts on mitochondria by inhibiting the permeabilization transition pore—blocking the release of cytochrome c into the cytosol – thus inhibiting apoptosis. This mechanism makes it an interesting candidate treatment for TBI, with a number of animal models suggesting benefit [112,113,114,115,116,117].

2.7. Oxytocin

Oxytocin is a neuropeptide synthesized by paraventricular neurons of the hypothalamus and excreted by the neurohypophysis; it is known to play a fundamental role during pregnancy by ensuring the tone of the uterus then the initiation of contractions and childbirth, and finally promotes lactation for breastfeeding. Numerous works also suggest its role in social interactions and the pleasure felt during these interactions; in fact, some authors show that mice lacking an oxytocin receptor in the nucleus accumbens exhibit disturbances in social interactions [118]. In addition it seems that oxytocin has a neuroprotective effect by a direct action on the microglia in the context of systemic inflammation (injection of lipopolysaccharide [LPS]), post-traumatic stress, and stroke [119,120,121,122]. All of these arguments make it possible to imagine that a treatment modulating the activity of oxytocinergic neurons, by attenuating microglial activation, could improve the quality of life of patients suffering from cerebral pathologies such as post-traumatic encephalopathy and which are manifested by social interaction disorders [123].

2.8. Erythropoietin

Erythropoietin (EPO) is a glycoprotein regulating erythropoiesis in the bone narrow; it is naturally produced by kidney. It has also been found in the brain even though it is on the upper limit of the molecular weight threshold to pass the BBB, and may have a neuroprotective role in TBI [124]. TBI leads to an upregulation to EPO receptor expression particularly in neurons, glial, and endothelial cells. EPO could promote neuroprotection in TBI by activating the antiapoptotic cascade JAK-2/NFkB [125] and PI3K, promoting STAT5 homodimerization [126]. Preclinical studies demonstrated that EPO could be antioxidant, antiedematous and also anti-inflammatory [127,128], and that it could reduce cell loss and promote neurogenesis [127,129]. However, in 2015, a double-blind, placebo-controlled trial undertaken in 29 centers in seven countries did not show a beneficial effect of the EPO treatment [130], but in 2020, Katiyar et al. published a meta-analysis including research studies through December 2019 showing that EPO could reduce 6-month mortality (though not in-hospital mortality), neurological outcome, and risk of deep vein thrombosis [131]. These results suggest the need for other clinical trials.
On the other hand, because of its essential roles during neurodevelopment (genesis, survival and differentiation of neural cells), EPO could also be a very interesting drug for children suffering from TBI. Extended high doses of EPO seem to prevent long term cognitive deficit and white matter loss visible in diffusion tensor imaging in infantile animal controlled cortical impact (CCI) studies [132].
EPO is already in use as a treatment for anemia, particularly in patients with renal insufficiency. It is a very low-toxicity treatment and its anti-apoptotic effects could improve the outcome of TBI patients. It is important to underline the interest of this treatment in the perinatal context, notably because of its essential character during neurodevelopment, which makes it a particularly interesting candidate in children.

2.9. Others

Other anti-inflammatory drugs targeting the selective activity of certain cytokines have also been tested. These include, for example, anti-TNFα [133]: etanercept. In mouse models, etanercept reduced microglial and astrocytic activation while stimulating neurogenesis, and thus improved post-traumatic cognitive performance [134,135,136]. A rat model of lateral fluid percussion showed attenuation of cerebral ischemia, neurological motor deficits, and numbers of microglia-TNFα double positive cells with etanercept therapy. In humans, the administration of anti-TNFα after a stroke promotes regression of pain and chronic deficits [137,138]; these results are also found in a post-traumatic context with a significant improvement in motor, sensory and cognitive functions [139].
Other authors are developing neuroprotective strategies from the IL-1 receptor antagonist (IL-1ra) to which both IL-1β and IL-1α bind. The use of a transgenic mouse hyperexpressing this antagonist makes it possible, by blocking the IL-1 pathway, to decrease the overall production of cytokines and in particular of TNF-α and IL-6 [140]. IL-1ra-treated animals show fewer nitric oxide synthase-2-positive cells in and around the lesion [141] and reduction of oligodendrocyte loss [142]. Anakinra, a recombinant IL1-R antagonist, reduces neuroinflammation and preserves post-TBI cognitive function in mice [143]. Conversely the use of selective antagonist interleukin 1β had no effect on motor recovery [144]. Some authors have studied the effect of anakinra on the mouse eye after blast-mediated traumatic brain injury, where anakinra treatment resulted in a preservation of retinal ganglion cells function and structure compared with saline treated bTBI mice, suggesting that IL-1 blockade also could also prevent axonal damage after blast [145]. Early injection of specific anti-IL-1β antibodies minimizes microglial activation, decreases neutrophilic and T lymphocyte infiltrates and reduce lesion volume [146] and cerebral edema [147]. In humans, a single center phase II randomized therapeutic trial shows the safety of injection of anti-IL-1 as well as its anti-inflammatory action on 20 patients; however, statistically significant clinical improvement was not shown, perhaps in part because the study was underpowered [148].
IL-6 also appears to be a robust marker of both neuroinflammation and intracranial hypertension in the setting of TBI [149,150], and injection of anti-IL-6 antibodies reduced the production of pro-inflammatory cerebral cytokines and improved motor functions in mice [151].
HMGB1, high mobility group box 1, is one of the Damage-Associated Molecular Patterns (DAMPs) proteins; it is a normally intracellular chromatin protein that is released by necrotic cells [152] and macrophages in response to stimulation by LPS or IL-1β [153]. It is believed to act on TLR receptors and, at the microglial level, it activates the NFkB pathway and the production of superoxide [154]. HMGB1, alarmin early involved in post-traumatic neuroinflammation and involved in neurogenesis, thus appears to be a particularly promising therapeutic target by many approaches. Targeting HMGB1 was shown to reduce microglial activation, the production of pro-inflammatory cytokines, and cerebral edema, and to improve the post-traumatic neurological outcome in mouse models [155,156,157]. In another murine model of CCI, HMGB1 antagonism reversed brain damage, and significantly reduced brain edema by protecting BBB integrity [157]. A treatment with an anti-HMGB1 monoclonal antibody improved post-traumatic motor and cognitive functions for fourteen days after the injury, prevented neuronal hippocampal death, and reduced microglial accumulation [158]. HMGB1 appears to be a therapeutic target, however at present the toxicity of using an HMGB1 antagonist has not been studied. Further work is needed to determine the potential utility and use of this treatment in humans.

3. Anesthetic Agents

3.1. Halogenated

Halogenated agents are hydrocarbons some parts of which are substituted by a halogen atom (bromine, chlorine or fluorine), thus explaining their name. They are powerful anesthetics first used by Morton in 1846 with ether, then chloroform [159]. It was not until a century later that methoxyflurane was marketed, which was subsequently discontinued due to its renal toxicity. Similarly, halothane and endoflurane are no longer used because of their cardiovascular toxicity. Isoflurane, desflurane, and sevoflurane, marketed in France in 1984, 1990 and 1996 respectively, are the three halogenated agents most used in clinical practice today. Thanks to PET and MRI imaging techniques, halogenated agents have been shown to modify cerebral metabolism, particularly in specific regions such as the thalamus and the reticulate formation [160], thus modifying global neuronal activity. The neuronal effect of halogens is also apparent in vitro: they are responsible for a decrease in the release of glutamate [161] with inhibition of the transmission of nerve impulses [162,163] as well as a potentiation of the inhibitory effect of gamma-amino-butyric acid (GABA).
Halogens facilitate neuroprotection by decreasing the brain’s electrical activity and its consumption of oxygen and glucose [164,165]. In a model of stroke, isoflurane has been demonstrated to inhibit microglial activation through the Notch pathway [166]. Isoflurane could also decrease the incidence of brain edema by downregulating aquaporin 4 [167,168]. In a study using controlled cortical impact, adult rats pre-teated with isoflurane presented a better cornu amonnis (CA) 3 neuronal survival and better performance in the Morris water maze and beam walking, thus a better motor coordination and a better memory [169]; these results were confirmed in other studies [170,171]. Sevoflurane may also attenuate inflammation [172,173] without modulating microglial activation [172]. However, in a model of cerebral arterial occlusion in rats, Dang et al. demonstrated that sevoflurane treatment impacts microglial/macrophage dynamics, migration, and phagocytosis—and so, indirectly, microglial activation – and promotes brain repair [174]. Moreover, in a model of neonatal ischemia, sevoflurane was shown to promote neuronal survival through the regulation of PI3K/Akt, and to improve neurocognitive performance [175,176]. Isoflurane shows similar results [177,178], but in an in vitro BBB model and controlled cortical impact study in mice, sevoflurane protected from brain edema better than isoflurane [179]. Statler et al. showed that treatment with isoflurane after focal trauma in rats improved the neurological score as well as the size of the lesion via an inhibition of the decrease in cerebral perfusion, an inhibition of the excitotoxicity of glutamate, and stimulation of GABA type A receptors [180].
To conclude, halogenated anesthetics may be neuroprotective from a mechanistic point of view [181], but the literature on its role in traumatic brain injury remains limited, and further studies are needed.

3.2. Inert Gas

Xenon is a colorless, odorless gas, and is the principal inert gas studied for therapeutic purposes in TBI thus far because it is used in anesthesia. Xenon is obtained by a complex air separation process, and its cost of production may unfortunately limit its use clinically. However, it does seem to have neuroprotective properties [182], especially in the context of cerebral ischemia [183]. Its action is not thought to be mediated by GABA receptors but rather by inhibition of NMDA receptors [184], thus reducing trauma-related excitotoxicity. In some models of murine TBI, xenon additionally helps reduce microglial activation and neuronal loss, thereby promoting late neurocognitive development [185,186]. In a study published in 2021, Xenon treatment reduced lesion volume, neuronal loss, microglia, reactive astrocytes, and early locomotor deficits [187]. In 2021, Filev et al. studied the effect of Xenon on gene expression in brain tissue in context of TBI rat model and observed lower expression of inflammatory genes like Irf1 (Interferon Regulatory Factor 1) in the area of damage [188]. Xenon appears to have few side effects: while it causes bradycardia, it has no effect on hemodynamic stability in contrast to other anesthetic agents. To conclude, Xenon appears to act through a variety of pathways, but among the most likely mechanisms—which could explain its protective effects on brain tissue—is the inhibition of NMDA receptors, which become overactive after brain injury. Given the relative safety of xenon and the results of the present study, the researchers hope to be able to quickly study the effectiveness of Xenon in TBI patients.
Argon, the third gas present in air (0.9%), exhibits anesthetic properties in hyperbaric conditions via GABA receptors [189]. Its neuroprotective [190] effect is observed on slices of hippocampal brains subjected to ischemic stress [191] or a trauma [192]. The underlying mechanisms are still poorly understood and the increase in the anti-apoptotic protein B-cell lymphoma 2 (BCL2] has been suggested [192]. In a recent study published in 2021, inhaled Argon reduced brain edema and neuroinflammation, and also accelerated sensorimotor and cognitive recovery. However, these results were not replicated in other rodent studies [193]. Its study in humans has been more limited [194]. To conclude, the evidence for Argon’s neuroprotective effects is less certain than for Xenon in the context of TBI.
Regarding inert gases, the literature on human subjects is very limited. The administration of these gases in intensive care requires a respirator with a closed circuit. Therefore, not only has the therapeutic effect of these gases not yet been proven in humans, but their daily use in intensive care is also inconvenient.

3.3. Propofol

Propofol is a highly hydrophobic intravenous hypnotic in the form of a lipid emulsion. Like many intravenous anesthetics, propofol has the ability to decrease brain oxygen consumption, decrease glutamate release, and modulate GABA-A receptor activity [195,196,197]. In the context of TBI, the literature on propofol is quite limited aside from demonstrating its antioxidant effects [198]. Indeed, propofol decreases ROS production, pro-inflammatory cytokine levels [199]. In vitro, propofol decreases oxidative stress [200]. In a CCI rat model, propofol reduced brain edema by suppressing aquaporin-4 expression, and was associated with a decrease in IL-1β and TNF-α levels [201]. Propofol may also be neuroprotective by blocking microglial activation through the modulation of nicotinamide adenine dinucleotide phosphate (NADPH) oxidation [202]: in a study published in 2013, propofol decreased expression of inducible nitric oxide synthase, nitric oxide, TNF-α, IL1β, ROS, and NADPH oxidase. Propofol has also been demonstrated to improve neurological TBI outcomes [203]. However in one study propofol limited reparative processes at the acute phase [204]. Interestingly propofol had protective effect on the intestine following TBI by lowering inflammatory markers expression [205]. Propofol is among the anesthetic drugs that could indeed improve the management of TBI patients.

3.4. Dexmedetomidine

Demedetomidine is a selective α-2-adrenergic agonist used for sedation during mechanical ventilation. It has already been demonstrated to be neuroprotective in models of ischemia and excitotoxicity. In CCI rat model, dexmedetomidine reduced secondary BBB damages by upregulating tight junction proteins, promoted neurogenesis, and decreased apoptosis, the latter potentially via suppression of NFkB and NOD-like receptor family, pyrin domain containing 3 (NLRP3] inflammasome activation [206]. Other studies show dexmedetomidine protects against apoptosis particularly in hippocampus via upregulation of Hsp70 [207]; it also protects against axonal injury and synaptic degeneration [208]. This neuroprotective effect could be due to the activation of the mTOR pathway [209], or inhibition of microglial activation [210,211,212].
Dexmedetomidine decreases levels of cytokines such as IL-1ß, IL-6, TNF- α, and IL-8 in a mouse model of TBI, but also in patient serum [213]. Recently, Dexmedetomidine has emerged as a potential neuroprotective agent by several mechanisms in TBI. Further work and particularly randomized clinical trials are needed.

3.5. Ketamine

Ketamine is an N-methyl-D-aspartate receptor antagonist shown to present anti-inflammatory effects in models of systemic inflammation [214,215] and brain ischemia [216]. The use of ketamine is very controversial because of its potential role in increasing intracranial pressure (ICP), but in 2013 Chang et al. reviewed the studies that had employed ketamine for sedation, and demonstrated no increase in ICP and eventually a neuroprotective effect [217]. For example, in a model of Maramrou’s weight drop model in mice, ketamine administration ameliorated oxidative stress, induced the expression of NRF2 pathway related factors, and ameliorated secondary brain injury including water content, neuronal apoptosis, and neurological deficit [218]. Ketamine has long been thought to be deleterious in neuroinjured patients, but recently several studies have suggested that it may have neuroprotective effects.

4. Hormonal Treatments

The role of the hypothalamic-pituitary axis appears to be major in TBI. In humans, hypopituitarism is indeed observed in association with TBI [219]; in a cohort study following 116 adults with TBI for up to 6 months, patients showed a decrease in estradiol, follicle stimulating hormone (FSH), and luteinizing hormone (LH) followed by an increase in stress hormones and impaired cognitive function. Female hormones may thus have a protective effect. These data corroborate those of numerous studies reporting that women with TBI not only have a lower mortality rate but also a better functional recovery than men [220,221,222,223].

4.1. Estrogens

In many models of brain damage, early administration of sex hormones may have anti-apoptotic, anti-inflammatory, and antioxidant effects, which can speed up repair processes and reduce long-term sequelae [224].
One of the main hypotheses regarding estrogen is based on the increased production of the protein sonic hedgehog, which regulates neuronal differentiation and may promote neuronal regeneration and protection [225]. Estrogens are also believed to have antioxidant properties [226,227,228,229] and to promote astrocytic production of glutamine, a precursor of glutamate, which will be captured by glutamatergic neurons [230,231]. They also act on microglial pro-inflammatory mediators such as the activation of NFkB or inducible NO synthase [232]. Emerson et al. observed in 1993 that the administration of estrogen before TBI improves the fate of male rats and paradoxically worsens that of females [233]. In addition, Kim et al. showed a decrease in intra-cranial pressure (ICP) and an increase in cerebral perfusion pressure, partial oxygen pressure, and glycolysis at the lesion [234]. Estrogen seems to have powerful effects on acute brain edema. Furthermore, 17 ß-estradiol treatment, the most abundant and potent endogenous vertebrate estrogen, alleviated neurological deficits, neuronal injuries, brain edema, and pro-inflammatory marker expression such as TLR4, NFkB, IL1-β, IL6 or TNFα in TBI [235]. 17 ß-estradiol protects against programmed cell death in the cortical pericontusional zone by decreasing caspase-3 activation [236]. Estrogens decrease free radical production and oxidative stress after TBI [237]. Bazedoxifen, a third generation selective estrogen receptor modulator, promoted neuroprotection by suppressing the activation of the MAPK/NFkB pathway and attenuated cognitive dysfunctions [238]. In order to better understand if the neuroprotective effect of estrogen was due estrogen receptor alpha or beta, some authors compared the effect of ER alpha and beta agonists: estradiol effect on brain edema, BBB permeability and neurological outcome was mediated through both receptors [239,240]. Estrogens could be an interesting therapeutic for both prevention and intervention post-TBI via multiple mechanisms.

4.2. Progesterone

Similar to estrogen, several studies describe a protective role for progesterone [241,242,243]. Three independent studies showed a reduction in post-traumatic brain edema after administration of the hormone in animals [244,245]. One of the explanations lies in the modulation of the expression of aquaporin 4 [246] and the p-glycoprotein. In addition, some authors find a decrease in post-traumatic DAI [247]. Moreover, progesterone presents anti-inflammatory effect reducing cytokine levels like IL1, IL-6, TLR2, TLR4, TNFα, NFkB activity and microglial activation [248,249]. Progesterone protects again lipid peroxidation [250]. Finally, progesterone treatment promotes myelin formation in Schwann cells and increases the number of mature oligodendrocytes [251,252,253].
According to studies from the Wagner laboratory, progesterone plays a crucial role in the development of brain and behavior [254,255]. The progesterone receptor is expressed in the forebrain during brain maturation in the rodent, and may influence neuronal migration, synaptogenesis, and cell death [256,257]. On a pediatric postnatal rat model of weight drop injury, the authors evaluated the effects of progesterone and magnesium separately and administered together: combination therapy was superior to progesterone alone for improving neuronal survival and overall long-term outcome; spatial learning and memory retention at three weeks was improved by each of these treatments alone and in combination [258]. Progesterone decreases anxiety after TBI in the immature brain [259]. Very promising results from clinical trials showed a neurodevelopmental improvement in extremely preterm infants (<1000 g) at 5 years [260,261]. Multicenter randomized trials are needed to confirmed these results.
In adults humans, progesterone promotes neurocognitive recovery [262,263,264,265,266], but unfortunately these data are controversial, and recently two randomized controlled studies investigating the benefit of treatment with intravenous progesterone in the acute phase of trauma did not show any benefit of this treatment compared to placebo [267,268].
To conclude, the effect of a progesterone treatment differs between children and adults. One of the explanations is that children’s brains are more sensitive to brain edema, so therapies aimed specifically at decreasing cerebral edema may have greater effect in children than in adults.

5. Vitamin Supplementation

5.1. Vitamins B

Vitamin B2 (riboflavin) is provided through food (meat and dairy products). Easily absorbed, it is necessary for normal cellular function and has strong antioxidant effects [269]. It delays neuronal death in vitro under excitotoxic conditions and in a dose-dependent manner [270]. Despite its potent antioxidant status, there have been very few studies on neuroprotection with riboflavin. In rat TBI models, it reduces edema, lesion size, and astrocyte activation, and facilitates cognitive and sensorimotor recovery [271,272]. B2 treatment also reduces behavioural troubles in the bilateral tactile removal test and improves reference and working memory [271].
Vitamin B3 (nicotinamide) is the amide form of nicotinic acid. Its neuroprotective action is widely characterized after TBI and stroke [273]. Its effects are manifold and include energy supplementation, scavenging of free radicals, and reduction of cell activation, apoptosis, and DNA damage [274]. The combination of these mechanisms makes nicotinamide an attractive candidate for the treatment of brain damage. In vivo, vitamin B3 administration is effective in several models of traumatic injury: it improves sensory, motor, and cognitive functions after frontal injury [275,276,277] and unilateral injury [278,279,280]. Acutely (<7 days after injury), vitamin B3 treatment reduces apoptosis, neuronal degeneration, edema, and BBB damage; finally, it decreases astrocyte activation and lesion volume [281], the latter even after a substantial latency period [278,279,282].
Preclinical evidence in rats suggests that nicotinamide may be an interesting treatment to explore in a clinical population. However, the putative neuroprotective dose in humans could be responsible for toxicity not yet evaluated in the literature. However, even taking into account possible toxicity problems, nicotinamide could exert protective effects in post-trauma and be particularly interesting to use in combination therapies because it is relatively easy to administer with few negative drug interactions.
Vitamin B6 (pyridoxine) is a water soluble vitamin, easily metabolized and excreted [283]. Pyridoxal 5′-phosphate (PLP) is the active coenzyme of vitamin B6 and is essential for the metabolism, catabolism, and transamination of amino acids [284]; according to some studies, it may have a neuroprotective effect by promoting glycogenolysis and reducing excitotoxicity [285,286]. Vitamin B6 reduces tissue damage post-TBI [287]. However, chronic high doses of vitamin B6 can lead to considerable neuronal toxicity with behavioral disturbances and problems with balance and walking [288,289], which limits the feasibility of long-term treatment at high doses.
Vitamin B9 (folic acid) is well known for its role in neural tube closure but also in the process of cell division, DNA synthesis, and maintenance of DNA methylation [290]. The literature is quite limited concerning TBI and does not permit conclusions regarding a therapeutic effect of vitamin B9: if certain authors conclude folic acid improves cognitive function in a pig model [291], such results were not found in a mouse model [292]. In a piglet pediatric model of TBI, piglet treatment with folic acid presented better exploratory interest and locomotion on day one, but not after. Axonal injury was not affected by the treatment.

5.2. Vitamin C

Ascorbic acid (AA), or vitamin C is widely recognized as one of the most important endogenous free radical scavengers [293]. It has also been suspected to have neuroprotective action by decreasing excitotoxicity-induced damage [294]. As part of the general metabolic dysfunction post-TBI, tissue levels of ascorbic acid have been shown to be immediately greatly reduced [295] and do not return to normal until 72 h after [296]. Polidori et al. published a cohort study of TBI patients in 2001, showing decreased ascorbic acid plasma levels in TBI patients compared with healthy controls, and this decrease was inversely correlated with the severity and the neurological outcome [297]. In addition, reduced levels of vitamin C are associated with an increase in lesion volume [298]. Despite the apparent disruption of vitamin C function in TBI, relatively few studies in humans have attempted direct vitamin C supplementation. Two studies have shown that treatment with vitamin C preserves ascorbic acid at appropriate levels in rat models of TBI, and maintained the production of superoxide dismutase with improved motor function in the course of the trauma [299]. Combined with vitamin E, ascorbic acid could improve cellular defense mechanisms and protect against oxidative effects of TBI [300]. In addition to its free radical scavenger activity, vitamin C may promote the integrity of the BBB by regulating the balance between metalloproteinase 9 (MMP-9) and free radical scavengers like Nrf2 [301,302]. These results warrant further study. Prophylactic ascorbic acid 2-glucoside significantly limited TBI-induced oxidative stress and mitigated motor dysfunction while the effects of therapeutic treatment were limited [303]. Clinically, the effect of vitamin C on critically ill patients is not particularly beneficial [304]. Literature on ascorbic acid treatment in TBI patients is generally consistent with this finding but inherently limited. The only randomized controlled trial on high-dose vitamin C administration for TBI was conducted in 2011 with 23 patients, but the authors did not observe any improvement in outcomes [305]. With its high safety, low cost profile, however, as well as the antioxidant properties above with potential to target TBI mechanisms, ascorbic acid remains a promising candidate for the acute stage of TBI management, e.g., in prehospital administration [306].

5.3. Vitamin D

A fat-soluble vitamin found in food, vitamin D is a secosteroid associated with peripherical calcium homeostasis. It is synthesized by the human body from a derivative of cholesterol and converted to its active form via ultraviolet radiation from the sun. Much of what is known about the neuroprotective effects of vitamin D comes from data on vitamin D deficiency [307] which suggest that it modulates apoptosis and reduces oxidative stress, inflammation, and excitotoxicity.
In the context of TBI, vitamin D was initially explored with progesterone [308] for its potential to act synergistically, as well as to study the relationship between age-related vitamin D decline and brain damage [309]. Other work not only observes an improvement in cognitive function [310] but also a reduction in inflammation and neuronal loss [311]. Although effective in adult rats, it appears that this combination may be most beneficial in middle-aged animals, potentially due to greater existing vitamin D deficiencies in that subpopulation. In older animals, this combination significantly reduces astrocyte proliferation and reduces neuronal loss [312]. The reason for the synergy of vitamin D and progesterone has not yet been fully elucidated, but a study suggests that it is the combination of a decrease in astrocyte activation and phosphorylation of NFκB [311].
Yang et al. demonstrated in 2021 that vitamin D supplementation reduces brain edema and inflammation while improving BBB integrity and behavioral function post-TBI [313].
Clinically, vitamin D supplementations in mild to moderate TBI patients at the acute phase of injury may improve long-term performance and cognitive outcomes evaluated with the Mini-Mental Status examination and the GOSE [314].
In 2020, thirty-five patients with moderate TBI were randomly allocated to a one-time oral dose of vitamin D; an improvement in level of consciousness was observed after 7 days in the vitamin D treatment group [315], confirming results of a previous study. Although further studies are necessary to validate the synergistic effects of vitamin D with progesterone, there is increasing evidence that this combination is an interesting therapeutic strategy and unlikely to exhibit toxicity. However, further exploration of the effects in younger animals, a better understanding of the therapeutic window, and more robust characterizations of functional recovery must be established before proceeding with further clinical trials.

5.4. Vitamin E

Vitamin E is a fat-soluble vitamin comprising a set of eight molecules, the most active biological form of which is alpha-tocopherol. It acts as an antioxidant against oxygen derivatives and in particular those resulting from the oxidation of fatty acids [316]. In combination with polyethylene glycol, it reduced mortality by 50% in TBI models and improved motor function [317]. Similar results were seen on neurocognitive function with alpha tocopherol alone given up to 90 days after trauma [318,319]. Alpha tocopherol reduced microscopic brain damage and also promoted nerve regeneration probably by reducing Nogo-A and NgR expression [320]. Although these behavioral effects are important, a study showed limited efficacy of vitamin E on lipid peroxidation in the acute post-injury phase [321]; others have demonstrated a later improvement in markers of oxidative stress. Vitamin E could also present neuroprotective effects for TBI associated dementia [322]. However, despite its high fat solubility and low toxicity, it takes a considerable time to reach effective concentrations in the central nervous system [323] and at these doses can be responsible for bleeding.

6. Ions

6.1. Magnésium

Over the past decades, many studies have suggested the interest of magnesium (Mg2+) in post-TBI recovery through improved cognitive functions [324]. The efficacy of Mg2+ therapy in promoting functional recovery across a variety of animal models of TBI is well demonstrated [325,326,327]. Administration of Mg2+ in animals with a normal diet not only improves sensorimotor function but also reduces certain histological damage such as rupture of the BBB, cerebral edema, neuronal death [328], apoptosis [329] glial proliferation [330,331,332,333] and brain damage [334]. Magnesium pre-treatment prevented injury impairments in working and reference memory via hippocampal ERK activation, and neuronal loss. At 1 week post-injury, magnesium treatment improved posttraumatic anxiety and depression [335]; in another study, four weeks after magnesium treatment improved also sensorimotor performance and recovery [336]. We have to note that some authors tried to compare magnesium sulfate and magnesium gluconate and they conclude that both have neuroprotective effects post-TBI [337], although there have been recent clinical trial failures for both TBI and stroke [338,339]. In 2017, Natale et al. showed that magnesium sulfate administration did not modify mean arterial pressure or alter cardiac conduction [340] in children population.

6.2. Zinc

Zinc plays a controversial role in the pathophysiology of TBI [341]. Many studies have identified increased and toxic levels of zinc after an experimental TBI while others have highlighted zinc deficiency to be deleterious and have shown zinc supplementation to be an effective therapy. Moreover recent studies showed that TBI did not link to drastic change of brain zinc level [342]. Several studies nevertheless link the accumulation of post-traumatic zinc to neuronal cell death by excitotoxicity [343,344] that could be due to rapid shifts in zinc localization [343,344]. The zinc mechanisms of toxicity is not well elucidated but it seems that like microglial dysfunction, it could induce reactive oxygen species [345] leading to mitochondrial disruption and neuroinflammation [346]. Other hypotheses are raised in the literature like mitochondrial dysfunction [347] or poly (ADP-ribose) polymerase (PARP) activation [348]. Mice without vesicular zinc presented more damages after TBI compared to wild type controls [349]. Zinc supplementation could also promote neurogenesis [350].Thus, the elimination of excess zinc, by chelation for example, is evaluated through several studies with a mixture of beneficial results [351], no effect [350] or harm (Need ref for harm).
Because the observation of TBI patients shows zinc deficiency [352], zinc supplementation was evaluated in patients and rats and indeed shows an improvement in cognitive and motor functions [353,354]. In a rat fluid percussion model, intraventricular injection of Ca EDTA provided neuroprotection in the CA1 region of the hippocampus, and dentate [343] and upregulated neuroprotective genes coding for heat shock proteins 27 and 70, and anti-apoptotic protein p21 [351]. Nonetheless one study demonstrated that zinc chelation participated to neuronal damages when hippocampus neurons were overexcited [355]. In addition, zinc chelation did not improve neurological outcome after TBI and increased the pro-apoptotic proteins BAX and caspase 3 two weeks after the injury leading to a second wave of injury weeks after the initial trauma [356]. On the other hand zinc deficiency was associated with ateration in matrix metalloproteinases, BBB disruption, inflammation and angiogenesis [357] but did not worsen behavioral outcome.
In a model of controlled cortical impact to the medial frontal cortex, rats were fed with a zinc adequate or a zinc supplemented diet: none of these treatment reduced anxiety behaviors but the dietary zinc supplementation improved learning and memory and the combination of intraperitoneal zinc and zinc dietary supplementation was necessary to reduce depression-like behaviors [358]. In another study, chronic zinc supplementation provided behavioral resiliency. Clinically zinc supplementation treatment for three months improved the Glasgow Coma Scale at 2 weeks and 3 months [354]. More recently a double-blind controlled study confirmed this results: zinc supplemented patients had a higher Sequential Organ Failure Assessment and Glasgow Coma Scale [359].

7. Omega-3

Omega-3 fatty acids are polyunsaturated fatty acids found in fatty fish and in certain plants; they have received a lot of attention, particularly in the prevention of cancer and cardiovascular risk factors. Regarding the CNS, they play multiple protective roles: both antioxidants, anti-inflammatories, and modulators of neurotransmission [360]. They have been the subject of numerous works in a traumatic context [361,362]. They have multiple modes of action: first of all, they reduce neuronal death by increasing the level of BDNF [363,364], then, they are powerful anti-inflammatory agents capable of lowering the level of cytokines such as TNF-α, IL-6 or reactive C protein [365], and finally they have other more controversial effects such as a decrease in excitotoxicity via the modulation of AMPA receptors [366,367]. While the biochemical evidence is quite promising, there are relatively few studies that examine the functional results associated with omega-3 acids and TBI. Some studies show that the depletion of omega-3 acid reserves leads to worsening motor and memory deficits [368], conversely, others show that supplementation before an injury leads to an improvement in motor and learning capacity [369,370], preserved the integrity of myelinand maintained post-TBI conductivity [370]. Dietary supplementation with docosahexaenoic acid reduced brain injury (measure with marker of cellular injury and apoptosis like APP and caspase-3), inflammation (CD68 positive cells) and improved memory assessment [371]. Omega-3 polyunsaturated fatty acids treatment promoted a shift from microglial pro-inflammatory phenotype to anti-inflammatory on, reducing inflammatory factors level probably mediated by decreasing HMGB1-medication activation of the NFkB pathway [372]. Recently some authors showed that pretreatment with omega-3 in TBI mice improved glymphatic clearance and suppressed expression of aquaporin 4 [373]. The accumulation of evidence in favor of the neuroprotective effect of omega-3 fatty [374] acids is quite promising in the setting of TBI but further study is still needed.

8. Discussion

Traumatic brain injury remains a significant health public problem and the leading cause of disability in patients under 40 years old. Post-traumatic encephalopathy can present with a broad range of symptoms, including somatic, neuropsychiatric, motor, and cognitive disorders, with lifelong consequences. These long-term consequences have been the target of clinical research for the last ten decades. In spite of a recent better understanding of the underlying neuroinflammatory mechanisms, there is currently no treatment proven to reduce post-traumatic handicap. In this review we have aimed to establish a list of current therapeutic drugs tested in both animal models of TBI and in humans, including anti-inflammatory medications such as glucocorticoids and NSAIDs, more specialized anti-inflammatory treatments, anesthetic agents, hormonal treatments, vitamins and ions (magnesium, zinc), and stem cell therapies. The problem remains that currently none of these drugs have been demonstrated to convincingly improve TBI patient outcomes, and therefore none is routinely used for TBI patients.
Even if the mechanisms responsible for the progression of secondary lesions to tertiary lesions are not well elucidated, the best post-traumatic time period for intervention seems to be the acute phase; later, injuries become established. Treatment during intensive care unit hospitalization is therefore a key element of managing these lesions; for maximum efficacy, it should be administered too late: the earlier the better. As the goal of treatment is to block the immune response before it gets out of control and creates irreversible damage, the best window of opportunity is during hospitalization in intensive care or just before. Once the lesions are already constituted, treatment seems without benefit. However, the ICU resuscitation period is also a risky period for use of anti-inflammatory treatment, since the patient is extremely fragile and at increased risk of opportunistic infections. Hence the importance of choosing a treatment that controls brain inflammation without weakening the patient’s immune defenses against infections.
The difficulty lies mainly in the heterogeneity of the brain damage; it is indeed a multifactorial pathology involving several cell types, mainly neurons, but also the brain’s own immune system (microglia, astrocytes) and the peripheral immune system, which communicate with each other owing to the hyperpermeability of the blood-brain barrier. As it seems unlikely that a single drug could effectively act on all the actors involved in the inflammatory response, one solution could be to use a combination of several treatments with different targets and modes of action. One of the most studied cellular targets is microglia. Indeed, some studies on other models of brain inflammation show that blocking microglial activation is sufficient to prevent brain damage, particularly white matter damage [375]. We can therefore imagine a therapeutic strategy that targets only microglia and which could improve TBI patients’ prognosis.
Aside from consideration of medical intervention, rehabilitation also seems to be crucial. Early intervention produces better outcomes: Andelic et al. showed that patients with early rehabilitation training had higher GOSE and Disability Rating Scale (DRS) at 12 months. Several technologies have been developed recently [376], including transcranial magnetic stimulation (TMS) and transcranial direct current stimulation (tDCS). These procedures are painless, noninvasive, and without major adverse effects. For example, in a case study, throughout five courses of transcranial magnetic stimulation, tinnitus symptoms improved [377]. In a double-blind randomized controlled trial published in 2015, repetitive TMS showed significant effect on depression symptoms and cognitive function [378]. These results were confirmed by two reviews [379,380], which found TMS to improve depression and anhedonia symptoms, auditory hallucinations, tinnitus, and autonomy in activities of daily living.
Among other novel rehabilitation techniques, some studies have highlighted the potential of virtual reality training. Computer-aided training can stimulate auditory, visual, and cognitive functions such as memory, attention, concentration, and proprioception. The recreational and game-like features of this technique promote participation and patient adherence to treatment. Such computer-based cognitive rehabilitation can improve the use of compensatory strategies, enhance memory, and alleviate post-traumatic mood disorders [381,382]. Another kinesthetic, multisensory rehabilitation method described recently is music therapy [383], shown to stabilize emotions [384] and promote enhancement of executive functions [385]. And finally, functional electrical stimulation, a low-frequency pulse current, can promote functional reconstruction, with task-oriented functional electrical stimulation demonstrating particular benefit in patients with hemiplegia. Social support—the belief that one is cared for and loved, that one belongs to a network of communication and mutual obligations—plays an essential role in recovery as well. Sinnakaruppan et al. showed in 2005 that caregiver and family member education programs help to relieve stress and anxiety and to promote recovery [386].
The main limitation of this review is that it is not exhaustive or systematic. It allows discussion and citation of the major publications in the field and for the main drugs tested in head trauma. Despite the importance of the literature on the subject, none of the treatments cited in this review has been proven to be neuroprotective and therefore cannot be recommended. The literature is still burgeoning and many studies are currently underway in humans to evaluate some of these strategies.
Several studies suspect the role of neuroinflammation and in particular microglial activation in the generation of tertiary post-traumatic lesions, and therefore in longer-term post-traumatic disability.
Table 1. Principal therapeutic drugs for traumatic brain injury and their mode of action.
Table 1. Principal therapeutic drugs for traumatic brain injury and their mode of action.
DrugsMode of ActionPediatric Studies
InflammationMicroglial/
Astrocytic Activation
ExcitotoxicityAnti-OxydativeApoptoseŒdemaMitochondriaNeuronal Death/NeurogenesisCerebral MetabolismBBB
Antiinflammatory drugsGC [387][44,50]
NSAI[56]
Statins[70,71,72][64][388] [59,60][389] [61,62,66] [389]
Melatonine[390][390] [78] [86,391,392,393] [394][91]
Minocycline[97,395][97,110] [396] [395,397,398]
Ciclosporin [113][399][112,115,116,117]
Oxytocin
Anti-TNF-α[136,137,139][136] [400] [135]
Anti-IL1[140][401]
Anti-IL-1β[146][146] [147]
Anti IL-6[151]
Anti-HMGB1[156,402][155] [156,157] [402,403,404,405,406]
Anesthesic agentsHallogenous
Argon[194]
Xenon[186,407][183] [186]
Propofol[199,203][202] [408,409] [198]
HormonsŒstrogenes [231,232] [228,229][227] [227]
Progesterone[262] [242,262] [244,245,246][403]
Vitaminic supplementationVitamin B2 [271] [269] [271]
Vitamin B3 [281,410][281,410][281][410] [281][411]
Vitamin B6 [284,286][284,286]
Vitamin B9 [291]
Vitamin C [294][293]
Vitamin D[311] [311][312][412]
Vitamin E [318] [413]
IonsMagnesium [332][331,333] [331]
Zinc [414]
Omega-3[365] [366,367]
BBB: blood brain barrier; IL: interleukine; NSAI: non-steroid anti-inflammatory.

9. Conclusions

Although there is currently no specific immunologically-targeted treatment for adult or pediatric head trauma, this review shows the substantial number of preclinical and clinical trials conducted to date. The only recommendations are those involving the prevention of secondary injury and treatment of intracranial hypertension. This multimodal and multidisciplinary management has effectively improved TBI mortality while increasing the rate of late disability. It therefore seems essential to implement and enrich research in this field.

Funding

This research was supported by grants the ERANET-NEURONS (TRAINS & ICON-TBI) and the Fondation des Gueules Cassées.

Acknowledgments

Figure 1 was generated using the Biorender application.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Menon, D.K.; Schwab, K.; Wright, D.W.; Maas, A.I. Demographics and Clinical Assessment Working Group of the International and Interagency Initiative toward Common Data Elements for Research on Traumatic Brain Injury and Psychological Health Position statement: Definition of traumatic brain injury. Arch. Phys. Med. Rehabil. 2010, 91, 1637–1640. [Google Scholar] [CrossRef]
  2. Asken, B.M.; Sullan, M.J.; DeKosky, S.T.; Jaffee, M.S.; Bauer, R.M. Research Gaps and Controversies in Chronic Traumatic Encephalopathy: A Review. JAMA Neurol. 2017, 74, 1255–1262. [Google Scholar] [CrossRef]
  3. Roozenbeek, B.; Maas, A.I.R.; Menon, D.K. Changing patterns in the epidemiology of traumatic brain injury. Nat. Rev. Neurol. 2013, 9, 231–236. [Google Scholar] [CrossRef]
  4. Peeters, W.; van den Brande, R.; Polinder, S.; Brazinova, A.; Steyerberg, E.W.; Lingsma, H.F.; Maas, A.I.R. Epidemiology of traumatic brain injury in Europe. Acta Neurochir. 2015, 157, 1683–1696. [Google Scholar] [CrossRef]
  5. Dewan, M.C.; Rattani, A.; Gupta, S.; Baticulon, R.E.; Hung, Y.-C.; Punchak, M.; Agrawal, A.; Adeleye, A.O.; Shrime, M.G.; Rubiano, A.M.; et al. Estimating the global incidence of traumatic brain injury. J. Neurosurg. 2018, 1–18. [Google Scholar] [CrossRef]
  6. Bayen, É.; Jourdan, C.; Azouvi, P.; Weiss, J.-J.; Pradat-Diehl, P. Prise en charge après lésion cérébrale acquise de type traumatisme crânien. Inf. Psychiatr. 2012, 88, 331–337. [Google Scholar] [CrossRef]
  7. Cambier, J.; Masson, M.; Masson, C.; Henri, D. (Eds.) Traumatismes crâniens. In Neurologie, 13th ed.; Elsevier Masson: Paris, France, 2012; pp. 408–417. ISBN 978-2-294-71451-1. [Google Scholar]
  8. Tazarourte, K.; Bensalah, N.; Rebillard, L.; Vigué, B. Epidémiologie des Traumatismes Crâniens; MAPAR: Paris, France, 2008. [Google Scholar]
  9. Masson, F.; Thicoipe, M.; Mokni, T.; Aye, P.; Erny, P.; Dabadie, P. Aquaitaine Group for Severe Brain Injury Study Epidemiology of traumatic comas: A prospective population-based study. Brain Inj. 2003, 17, 279–293. [Google Scholar] [CrossRef]
  10. Collaborators, M.C.T. Predicting outcome after traumatic brain injury: Practical prognostic models based on large cohort of international patients. BMJ 2008, 336, 425–429. [Google Scholar] [CrossRef]
  11. Davis, D.P.; Serrano, J.A.; Vilke, G.M.; Sise, M.J.; Kennedy, F.; Eastman, A.B.; Velky, T.; Hoyt, D.B. The predictive value of field versus arrival Glasgow Coma Scale score and TRISS calculations in moderate-to-severe traumatic brain injury. J. Trauma 2006, 60, 985–990. [Google Scholar] [CrossRef]
  12. Regel, G.; Lobenhoffer, P.; Grotz, M.; Pape, H.C.; Lehmann, U.; Tscherne, H. Treatment results of patients with multiple trauma: An analysis of 3406 cases treated between 1972 and 1991 at a German Level I Trauma Center. J. Trauma 1995, 38, 70–78. [Google Scholar] [CrossRef]
  13. Hills, M.W.; Deane, S.A. Head injury and facial injury: Is there an increased risk of cervical spine injury? J. Trauma 1993, 34, 549–553, discussion 553–554. [Google Scholar] [CrossRef]
  14. Sandiego, C.M.; Gallezot, J.-D.; Pittman, B.; Nabulsi, N.; Lim, K.; Lin, S.-F.; Matuskey, D.; Lee, J.-Y.; O’Connor, K.C.; Huang, Y.; et al. Imaging robust microglial activation after lipopolysaccharide administration in humans with PET. Proc. Natl. Acad. Sci. USA 2015, 112, 12468–12473. [Google Scholar] [CrossRef]
  15. Tagliaferri, F.; Compagnone, C.; Korsic, M.; Servadei, F.; Kraus, J. A systematic review of brain injury epidemiology in Europe. Acta Neurochir. 2006, 148, 255–268, discussion 268. [Google Scholar] [CrossRef]
  16. Adams, J.H.; Doyle, D.; Ford, I.; Gennarelli, T.A.; Graham, D.I.; McLellan, D.R. Diffuse axonal injury in head injury: Definition, diagnosis and grading. Histopathology 1989, 15, 49–59. [Google Scholar] [CrossRef]
  17. Li, X.-Y.; Feng, D.-F. Diffuse axonal injury: Novel insights into detection and treatment. J. Clin. Neurosci. 2009, 16, 614–619. [Google Scholar] [CrossRef]
  18. Carpentier, A.; Galanaud, D.; Puybasset, L.; Muller, J.-C.; Lescot, T.; Boch, A.-L.; Riedl, V.; Riedl, V.; Cornu, P.; Coriat, P.; et al. Early morphologic and spectroscopic magnetic resonance in severe traumatic brain injuries can detect “invisible brain stem damage” and predict “vegetative states”. J. Neurotrauma 2006, 23, 674–685. [Google Scholar] [CrossRef]
  19. Hammoud, D.A.; Wasserman, B.A. Diffuse axonal injuries: Pathophysiology and imaging. Neuroimaging Clin. N. Am. 2002, 12, 205–216. [Google Scholar] [CrossRef]
  20. Scheid, R.; Walther, K.; Guthke, T.; Preul, C.; von Cramon, D.Y. Cognitive sequelae of diffuse axonal injury. Arch. Neurol. 2006, 63, 418–424. [Google Scholar] [CrossRef]
  21. Chesnut, R.M.; Marshall, L.F.; Klauber, M.R.; Blunt, B.A.; Baldwin, N.; Eisenberg, H.M.; Jane, J.A.; Marmarou, A.; Foulkes, M.A. The role of secondary brain injury in determining outcome from severe head injury. J. Trauma 1993, 34, 216–222. [Google Scholar] [CrossRef]
  22. Johannigman, J.A.; Zonies, D.; Dubose, J.; Blakeman, T.C.; Hanseman, D.; Branson, R.D. Reducing Secondary Insults in Traumatic Brain Injury. Mil. Med. 2015, 180, 50–55. [Google Scholar] [CrossRef] [Green Version]
  23. Groom, R.; Oakley, P.A. Secondary brain injury: Mechanisms and prevention. Curr. Anaesth. Crit. Care 1997, 8, 248–253. [Google Scholar] [CrossRef]
  24. Corps, K.N.; Roth, T.L.; McGavern, D.B. Inflammation and neuroprotection in traumatic brain injury. JAMA Neurol. 2015, 72, 355–362. [Google Scholar] [CrossRef]
  25. Kumar, A.; Loane, D.J. Neuroinflammation after traumatic brain injury: Opportunities for therapeutic intervention. Brain Behav. Immun. 2012, 26, 1191–1201. [Google Scholar] [CrossRef]
  26. Simon, D.W.; McGeachy, M.J.; Bayır, H.; Clark, R.S.B.; Loane, D.J.; Kochanek, P.M. The far-reaching scope of neuroinflammation after traumatic brain injury. Nat. Rev. Neurol. 2017, 13, 572. [Google Scholar] [CrossRef]
  27. Smith, C.; Gentleman, S.M.; Leclercq, P.D.; Murray, L.S.; Griffin, W.S.T.; Graham, D.I.; Nicoll, J.A.R. The neuroinflammatory response in humans after traumatic brain injury: Neuroinflammation after brain injury. Neuropathol. Appl. Neurobiol. 2013, 39, 654–666. [Google Scholar] [CrossRef]
  28. Johnson, V.E.; Stewart, J.E.; Begbie, F.D.; Trojanowski, J.Q.; Smith, D.H.; Stewart, W. Inflammation and white matter degeneration persist for years after a single traumatic brain injury. Brain 2013, 136, 28–42. [Google Scholar] [CrossRef]
  29. Martland, H.S. Punch Drunk. JAMA 1928, 91, 1103–1107. [Google Scholar] [CrossRef]
  30. Smith, D.H.; Johnson, V.E.; Trojanowski, J.Q.; Stewart, W. Chronic traumatic encephalopathy—Confusion and controversies. Nat. Rev. Neurol. 2019, 15, 179–183. [Google Scholar] [CrossRef]
  31. Omalu, B.I.; DeKosky, S.T.; Minster, R.L.; Kamboh, M.I.; Hamilton, R.L.; Wecht, C.H. Chronic traumatic encephalopathy in a National Football League player. Neurosurgery 2005, 57, 128–134, discussion 128–134. [Google Scholar] [CrossRef]
  32. Omalu, B.I.; DeKosky, S.T.; Hamilton, R.L.; Minster, R.L.; Kamboh, M.I.; Shakir, A.M.; Wecht, C.H. Chronic traumatic encephalopathy in a national football league player: Part II. Neurosurgery 2006, 59, 1086–1092, discussion 1092–1093. [Google Scholar] [CrossRef]
  33. Omalu, B.I.; Fitzsimmons, R.P.; Hammers, J.; Bailes, J. Chronic traumatic encephalopathy in a professional American wrestler. J. Forensic. Nurs. 2010, 6, 130–136. [Google Scholar] [CrossRef] [PubMed]
  34. McKee, A.C.; Cantu, R.C.; Nowinski, C.J.; Hedley-Whyte, E.T.; Gavett, B.E.; Budson, A.E.; Santini, V.E.; Lee, H.-S.; Kubilus, C.A.; Stern, R.A. Chronic traumatic encephalopathy in athletes: Progressive tauopathy after repetitive head injury. J. Neuropathol. Exp. Neurol. 2009, 68, 709–735. [Google Scholar] [CrossRef] [PubMed]
  35. Maroon, J.C.; Winkelman, R.; Bost, J.; Amos, A.; Mathyssek, C.; Miele, V. Chronic traumatic encephalopathy in contact sports: A systematic review of all reported pathological cases. PLoS ONE 2015, 10, e0117338. [Google Scholar] [CrossRef]
  36. Gardner, A.; Iverson, G.L.; McCrory, P. Chronic traumatic encephalopathy in sport: A systematic review. Br. J. Sports Med. 2014, 48, 84–90. [Google Scholar] [CrossRef]
  37. Broshek, D.K.; De Marco, A.P.; Freeman, J.R. A review of post-concussion syndrome and psychological factors associated with concussion. Brain Inj. 2015, 29, 228–237. [Google Scholar] [CrossRef]
  38. Azouvi, P.; Jourdan, C.; Bayen, E.; Darnoux, E.; Ghout, I.; Azerad, S.; Vallat-Azouvi, C.; Ruet, A.; Weiss, J.J.; Aegerter, P.; et al. L’étude PariS-TBI: Suivi longitudinal d’une cohorte de blessés après un traumatisme crânien (TC) sévère. Ann. Phys. Rehabil. Med. 2014, 57, e75. [Google Scholar] [CrossRef]
  39. Whitnall, L.; McMillan, T.M.; Murray, G.D.; Teasdale, G.M. Disability in young people and adults after head injury: 5–7 year follow up of a prospective cohort study. J. Neurol. Neurosurg. Psychiatry 2006, 77, 640–645. [Google Scholar] [CrossRef]
  40. Christensen, B.K.; Colella, B.; Inness, E.; Hebert, D.; Monette, G.; Bayley, M.; Green, R.E. Recovery of cognitive function after traumatic brain injury: A multilevel modeling analysis of Canadian outcomes. Arch. Phys. Med. Rehabil. 2008, 89, S3–S15. [Google Scholar] [CrossRef]
  41. Holbrook, T.L.; Anderson, J.P.; Sieber, W.J.; Browner, D.; Hoyt, D.B. Outcome after major trauma: Discharge and 6-month follow-up results from the Trauma Recovery Project. J. Trauma 1998, 45, 315–323, discussion 323–324. [Google Scholar] [CrossRef]
  42. Geeraerts, T.; Velly, L.; Abdennour, L.; Asehnoune, K.; Audibert, G.; Bouzat, P.; Bruder, N.; Carrillon, R.; Cottenceau, V.; Cotton, F.; et al. Management of severe traumatic brain injury (first 24hours). Anaesth. Crit. Care Pain Med. 2018, 37, 171–186. [Google Scholar] [CrossRef] [PubMed]
  43. Hawryluk, G.W.J.; Rubiano, A.M.; Totten, A.M.; O’Reilly, C.; Ullman, J.S.; Bratton, S.L.; Chesnut, R.; Harris, O.A.; Kissoon, N.; Shutter, L.; et al. Guidelines for the Management of Severe Traumatic Brain Injury: 2020 Update of the Decompressive Craniectomy Recommendations. Neurosurgery 2020, 87, 427–434. [Google Scholar] [CrossRef]
  44. Michinaga, S.; Koyama, Y. Pathogenesis of brain edema and investigation into anti-edema drugs. Int. J. Mol. Sci. 2015, 16, 9949–9975. [Google Scholar] [CrossRef]
  45. Visser, J.; van Boxel-Dezaire, A.; Methorst, D.; Brunt, T.; de Kloet, E.R.; Nagelkerken, L. Differential regulation of interleukin-10 (IL-10) and IL-12 by glucocorticoids in vitro. Blood 1998, 91, 4255–4264. [Google Scholar] [CrossRef]
  46. Salmon, J.E.; Kapur, S.; Meryhew, N.L.; Runquist, O.A.; Kimberly, R.P. High-dose, pulse intravenous methylprednisolone enhances Fc gamma receptor-mediated mononuclear phagocyte function in systemic lupus erythematosus. Arthritis Rheum. 1989, 32, 717–725. [Google Scholar] [CrossRef] [PubMed]
  47. Pitzalis, C.; Pipitone, N.; Bajocchi, G.; Hall, M.; Goulding, N.; Lee, A.; Kingsley, G.; Lanchbury, J.; Panayi, G. Corticosteroids inhibit lymphocyte binding to endothelium and intercellular adhesion: An additional mechanism for their anti-inflammatory and immunosuppressive effect. J. Immunol. 1997, 158, 5007–5016. [Google Scholar]
  48. Huizenga, N.A.T.M.; Koper, J.W.; de Lange, P.; Pols, H.A.P.; Stolk, R.P.; Burger, H.; Grobbee, D.E.; Brinkmann, A.O.; de Jong, F.H.; Lamberts, S.W.J. A Polymorphism in the Glucocorticoid Receptor Gene May Be Associated with an Increased Sensitivity to Glucocorticoids in vivo. J. Clin. Endocrinol. Metab. 1998, 83, 144–151. [Google Scholar] [CrossRef] [PubMed]
  49. Hall, E.D. High-dose glucocorticoid treatment improves neurological recovery in head-injured mice. J. Neurosurg. 1985, 62, 882–887. [Google Scholar] [CrossRef]
  50. Galicich, J.H.; French, L.A. Use of dexamethasone in the treatment of cerebral edema resulting from brain tumors and brain surgery. Am. Pract. Dig. Treat. 1961, 12, 169–174. [Google Scholar]
  51. Jeevaratnam, D.R.; Menon, D.K. Survey of intensive care of severely head injured patients in the United Kingdom. BMJ 1996, 312, 944–947. [Google Scholar] [CrossRef]
  52. Effect of intravenous corticosteroids on death within 14 days in 10 008 adults with clinically significant head injury (MRC CRASH trial): Randomised placebo-controlled trial. Lancet 2004, 364, 1321–1328. [CrossRef]
  53. Alderson, P.; Roberts, I. Corticosteroids for acute traumatic brain injury. Cochrane Database Syst. Rev. 2005. [Google Scholar] [CrossRef] [PubMed]
  54. Saul, T.G.; Ducker, T.B.; Salcman, M.; Carro, E. Steroids in severe head injury: A prospective randomized clinical trial. J. Neurosurg. 1981, 54, 596–600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Grumme, T.; Baethmann, A.; Kolodziejczyk, D.; Krimmer, J.; Fischer, M.; Rothe, B.v.E.; Pelka, R.; Bennefeld, H.; Pöllauer, E.; Kostron, H.; et al. Treatment of patients with severe head injury by triamcinolone: A prospective, controlled multicenter clinical trial of 396 cases. Res. Exp. Med. 1995, 195, 217–229. [Google Scholar] [CrossRef] [PubMed]
  56. Bergold, P.J. Treatment of traumatic brain injury with anti-inflammatory drugs. Exp. Neurol. 2016, 275, 367–380. [Google Scholar] [CrossRef] [PubMed]
  57. Cheng, G.; Wei, L.; Sun, Z.-d.; Zhao, S.-g.; Liu, X.-z. Atorvastatin ameliorates cerebral vasospasm and early brain injury after subarachnoid hemorrhage and inhibits caspase-dependent apoptosis pathway. BMC Neurosci. 2009, 10, 7. [Google Scholar] [CrossRef]
  58. Amin-Hanjani, S.; Stagliano, N.E.; Yamada, M.; Huang, P.L.; Liao, J.K.; Moskowitz, M.A. Mevastatin, an HMG-CoA Reductase Inhibitor, Reduces Stroke Damage and Upregulates Endothelial Nitric Oxide Synthase in Mice. Stroke 2001, 32, 980–986. [Google Scholar] [CrossRef]
  59. Wu, H.; Lu, D.; Jiang, H.; Xiong, Y.; Qu, C.; Li, B.; Mahmood, A.; Zhou, D.; Chopp, M. Increase in phosphorylation of Akt and its downstream signaling targets and suppression of apoptosis by simvastatin after traumatic brain injury: Laboratory investigation. J. Neurosurg. 2008, 109, 691–698. [Google Scholar] [CrossRef] [PubMed]
  60. Lu, D.; Qu, C.; Goussev, A.; Jiang, H.; Lu, C.; Schallert, T.; Mahmood, A.; Chen, J.; Li, Y.; Chopp, M. Statins Increase Neurogenesis in the Dentate Gyrus, Reduce Delayed Neuronal Death in the Hippocampal CA3 Region, and Improve Spatial Learning in Rat after Traumatic Brain Injury. J. Neurotrauma 2007, 24, 1132–1146. [Google Scholar] [CrossRef]
  61. Lu, D.; Goussev, A.; Chen, J.; Pannu, P.; Li, Y.; Mahmood, A.; Chopp, M. Atorvastatin reduces neurological deficit and increases synaptogenesis, angiogenesis, and neuronal survival in rats subjected to traumatic brain injury. J. Neurotrauma 2004, 21, 21–32. [Google Scholar] [CrossRef]
  62. Wang, H.; Lynch, J.R.; Song, P.; Yang, H.-J.; Yates, R.B.; Mace, B.; Warner, D.S.; Guyton, J.R.; Laskowitz, D.T. Simvastatin and atorvastatin improve behavioral outcome, reduce hippocampal degeneration, and improve cerebral blood flow after experimental traumatic brain injury. Exp. Neurol. 2007, 206, 59–69. [Google Scholar] [CrossRef]
  63. Qu, C.; Lu, D.; Goussev, A.; Schallert, T.; Mahmood, A.; Chopp, M. Effect of atorvastatin on spatial memory, neuronal survival, and vascular density in female rats after traumatic brain injury. J. Neurosurg. 2005, 103, 695–701. [Google Scholar] [CrossRef] [PubMed]
  64. Li, B.; Mahmood, A.; Lu, D.; Wu, H.; Xiong, Y.; Qu, C.; Chopp, M. Simvastatin attenuates microglial cells and astrocyte activation and decreases interleukin-1beta level after traumatic brain injury. Neurosurgery 2009, 65, 179–185, discussion 185–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Xu, X.; Gao, W.; Cheng, S.; Yin, D.; Li, F.; Wu, Y.; Sun, D.; Zhou, S.; Wang, D.; Zhang, Y.; et al. Anti-inflammatory and immunomodulatory mechanisms of atorvastatin in a murine model of traumatic brain injury. J. Neuroinflamm. 2017, 14, 167. [Google Scholar] [CrossRef] [PubMed]
  66. Wu, H.; Lu, D.; Jiang, H.; Xiong, Y.; Qu, C.; Li, B.; Mahmood, A.; Zhou, D.; Chopp, M. Simvastatin-mediated upregulation of VEGF and BDNF, activation of the PI3K/Akt pathway, and increase of neurogenesis are associated with therapeutic improvement after traumatic brain injury. J. Neurotrauma 2008, 25, 130–139. [Google Scholar] [CrossRef]
  67. Chen, J.; Zhang, C.; Jiang, H.; Li, Y.; Zhang, L.; Robin, A.; Katakowski, M.; Lu, M.; Chopp, M. Atorvastatin induction of VEGF and BDNF promotes brain plasticity after stroke in mice. J. Cereb. Blood Flow Metab. 2005, 25, 281–290. [Google Scholar] [CrossRef]
  68. Lu, D.; Mahmood, A.; Goussev, A.; Schallert, T.; Qu, C.; Zhang, Z.G.; Li, Y.; Lu, M.; Chopp, M. Atorvastatin reduction of intravascular thrombosis, increase in cerebral microvascular patency and integrity, and enhancement of spatial learning in rats subjected to traumatic brain injury. J. Neurosurg. 2004, 101, 813–821. [Google Scholar] [CrossRef]
  69. Lynch, J.R.; Wang, H.; McGirt, M.J.; Floyd, J.; Friedman, A.H.; Coon, A.L.; Blessing, R.; Alexander, M.J.; Graffagnino, C.; Warner, D.S.; et al. Simvastatin reduces vasospasm after aneurysmal subarachnoid hemorrhage: Results of a pilot randomized clinical trial. Stroke 2005, 36, 2024–2026. [Google Scholar] [CrossRef]
  70. Chen, S.-F.; Hung, T.-H.; Chen, C.-C.; Lin, K.-H.; Huang, Y.-N.; Tsai, H.-C.; Wang, J.-Y. Lovastatin improves histological and functional outcomes and reduces inflammation after experimental traumatic brain injury. Life Sci. 2007, 81, 288–298. [Google Scholar] [CrossRef]
  71. Wang, K.-W.; Chen, H.-J.; Lu, K.; Liliang, P.-C.; Liang, C.-L.; Tsai, Y.-D.; Cho, C.-L. Simvastatin Attenuates the Cerebral Vascular Endothelial Inflammatory Response in a Rat Traumatic Brain Injury. Ann. Clin. Lab. Sci. 2014, 44, 145–150. [Google Scholar] [CrossRef]
  72. Chen, G.; Zhang, S.; Shi, J.; Ai, J.; Qi, M.; Hang, C. Simvastatin reduces secondary brain injury caused by cortical contusion in rats: Possible involvement of TLR4/NF-κB pathway. Exp. Neurol. 2009, 216, 398–406. [Google Scholar] [CrossRef]
  73. Erdös, B.; Snipes, J.A.; Tulbert, C.D.; Katakam, P.; Miller, A.W.; Busija, D.W. Rosuvastatin improves cerebrovascular function in Zucker obese rats by inhibiting NAD(P)H oxidase-dependent superoxide production. Am. J. Physiol. Heart Circ. Physiol. 2006, 290, H1264–H1270. [Google Scholar] [CrossRef] [PubMed]
  74. Turkoglu, O.F.; Eroglu, H.; Okutan, O.; Gurcan, O.; Bodur, E.; Sargon, M.F.; Öner, L.; Beskonaklı, E. Atorvastatin efficiency after traumatic brain injury in rats. Surg. Neurol. 2009, 72, 146–152, discussion 152. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, X.R.; Besson, V.C.; Beziaud, T.; Plotkine, M.; Marchand-Leroux, C. Combination therapy with fenofibrate, a peroxisome proliferator-activated receptor alpha agonist, and simvastatin, a 3-hydroxy-3-methylglutaryl-coenzyme A reductase inhibitor, on experimental traumatic brain injury. J Pharmacol. Exp. Ther. 2008, 326, 966–974. [Google Scholar] [CrossRef] [PubMed]
  76. Sultan, W.; Sapkota, A.; Khurshid, H.; Qureshi, I.A.; Jahan, N.; Went, T.R.; Dominic, J.L.; Win, M.; Kannan, A.; Tara, A.; et al. Statins’ Effect on Cognitive Outcome After Traumatic Brain Injury: A Systematic Review. Cureus 2021, 13, e16953. [Google Scholar] [CrossRef] [PubMed]
  77. Osier, N.; McGreevy, E.; Pham, L.; Puccio, A.; Ren, D.; Conley, Y.P.; Alexander, S.; Dixon, C.E. Melatonin as a Therapy for Traumatic Brain Injury: A Review of Published Evidence. Int. J. Mol. Sci. 2018, 19, 1539. [Google Scholar] [CrossRef]
  78. Reiter, R.J.; Tan, D.-X.; Mayo, J.C.; Sainz, R.M.; Leon, J.; Czarnocki, Z. Melatonin as an antioxidant: Biochemical mechanisms and pathophysiological implications in humans. Acta Biochim. Pol. 2003, 50, 1129–1146. [Google Scholar] [CrossRef]
  79. Lin, H.-W.; Lee, E.-J. Effects of melatonin in experimental stroke models in acute, sub-acute, and chronic stages. Neuropsychiatr. Dis. Treat. 2009, 5, 157–162. [Google Scholar] [CrossRef]
  80. Pei, Z.; Pang, S.F.; Cheung, R.T.F. Administration of Melatonin After Onset of Ischemia Reduces the Volume of Cerebral Infarction in a Rat Middle Cerebral Artery Occlusion Stroke Model. Stroke 2003, 34, 770–775. [Google Scholar] [CrossRef]
  81. Borlongan, C.V.; Yamamoto, M.; Takei, N.; Kumazaki, M.; Ungsuparkorn, C.; Hida, H.; Sanberg, P.R.; Nishino, H. Glial cell survival is enhanced during melatonin-induced neuroprotection against cerebral ischemia. FASEB J. 2000, 14, 1307–1317. [Google Scholar]
  82. Shinozuka, K.; Staples, M.; Borlongan, C.V. Melatonin-based therapeutics for neuroprotection in stroke. Int. J. Mol. Sci. 2013, 14, 8924–8947. [Google Scholar] [CrossRef]
  83. Ding, K.; Wang, H.; Xu, J.; Li, T.; Zhang, L.; Ding, Y.; Zhu, L.; He, J.; Zhou, M. Melatonin stimulates antioxidant enzymes and reduces oxidative stress in experimental traumatic brain injury: The Nrf2–ARE signaling pathway as a potential mechanism. Free Radic. Biol. Med. 2014, 73, 1–11. [Google Scholar] [CrossRef] [PubMed]
  84. Senol, N.; Nazıroğlu, M. Melatonin reduces traumatic brain injury-induced oxidative stress in the cerebral cortex and blood of rats. Neural Regen. Res. 2014, 9, 1112–1116. [Google Scholar] [CrossRef] [PubMed]
  85. Campolo, M.; Ahmad, A.; Crupi, R.; Impellizzeri, D.; Morabito, R.; Esposito, E.; Cuzzocrea, S. Combination therapy with melatonin and dexamethasone in a mouse model of traumatic brain injury. J. Endocrinol. 2013, 217, 291–301. [Google Scholar] [CrossRef] [PubMed]
  86. Tan, D.-X.; Manchester, L.C.; Qin, L.; Reiter, R.J. Melatonin: A Mitochondrial Targeting Molecule Involving Mitochondrial Protection and Dynamics. Int. J. Mol. Sci. 2016, 17, 2124. [Google Scholar] [CrossRef] [PubMed]
  87. Salman, M.; Kaushik, P.; Tabassum, H.; Parvez, S. Melatonin Provides Neuroprotection Following Traumatic Brain Injury-Promoted Mitochondrial Perturbation in Wistar Rat. Cell. Mol. Neurobiol. 2021, 41, 765–781. [Google Scholar] [CrossRef]
  88. Rehman, S.U.; Ikram, M.; Ullah, N.; Alam, S.I.; Park, H.Y.; Badshah, H.; Choe, K.; Kim, M.O. Neurological Enhancement Effects of Melatonin against Brain Injury-Induced Oxidative Stress, Neuroinflammation, and Neurodegeneration via AMPK/CREB Signaling. Cells 2019, 8, 760. [Google Scholar] [CrossRef] [PubMed]
  89. Naseem, M.; Parvez, S. Role of melatonin in traumatic brain injury and spinal cord injury. Sci. World J. 2014, 2014, 586270. [Google Scholar] [CrossRef]
  90. Wang, J.; Jiang, C.; Zhang, K.; Lan, X.; Chen, X.; Zang, W.; Wang, Z.; Guan, F.; Zhu, C.; Yang, X.; et al. Melatonin receptor activation provides cerebral protection after traumatic brain injury by mitigating oxidative stress and inflammation via the Nrf2 signaling pathway. Free Radic. Biol. Med. 2019, 131, 345–355. [Google Scholar] [CrossRef]
  91. Ozdemir, D.; Uysal, N.; Gonenc, S.; Acikgoz, O.; Sonmez, A.; Topcu, A.; Ozdemir, N.; Duman, M.; Semin, I.; Ozkan, H. Effect of melatonin on brain oxidative damage induced by traumatic brain injury in immature rats. Physiol. Res. 2005, 54, 631–637. [Google Scholar] [CrossRef]
  92. Barlow, K.M.; Brooks, B.L.; MacMaster, F.P.; Kirton, A.; Seeger, T.; Esser, M.; Crawford, S.; Nettel-Aguirre, A.; Zemek, R.; Angelo, M.; et al. A double-blind, placebo-controlled intervention trial of 3 and 10 mg sublingual melatonin for post-concussion syndrome in youths (PLAYGAME): Study protocol for a randomized controlled trial. Trials 2014, 15, 271. [Google Scholar] [CrossRef]
  93. Barlow, K.M.; Esser, M.J.; Veidt, M.; Boyd, R. Melatonin as a Treatment after Traumatic Brain Injury: A Systematic Review and Meta-Analysis of the Pre-Clinical and Clinical Literature. J. Neurotrauma 2019, 36, 523–537. [Google Scholar] [CrossRef] [PubMed]
  94. Grima, N.A.; Rajaratnam, S.M.W.; Mansfield, D.; Sletten, T.L.; Spitz, G.; Ponsford, J.L. Efficacy of melatonin for sleep disturbance following traumatic brain injury: A randomised controlled trial. BMC Med. 2018, 16, 8. [Google Scholar] [CrossRef] [PubMed]
  95. Osier, N.D.; Pham, L.; Pugh, B.J.; Puccio, A.; Ren, D.; Conley, Y.P.; Alexander, S.; Dixon, C.E. Brain injury results in lower levels of melatonin receptors subtypes MT1 and MT2. Neurosci. Lett. 2017, 650, 18–24. [Google Scholar] [CrossRef] [Green Version]
  96. Arvin, K.L.; Han, B.H.; Du, Y.; Lin, S.; Paul, S.M.; Holtzman, D.M. Minocycline markedly protects the neonatal brain against hypoxic-ischemic injury. Ann. Neurol. 2002, 52, 54–61. [Google Scholar] [CrossRef] [PubMed]
  97. Kobayashi, K.; Imagama, S.; Ohgomori, T.; Hirano, K.; Uchimura, K.; Sakamoto, K.; Hirakawa, A.; Takeuchi, H.; Suzumura, A.; Ishiguro, N.; et al. Minocycline selectively inhibits M1 polarization of microglia. Cell Death Dis. 2013, 4, e525. [Google Scholar] [CrossRef]
  98. Yrjanheikki, J.; Keinanen, R.; Pellikka, M.; Hokfelt, T.; Koistinaho, J. Tetracyclines inhibit microglial activation and are neuroprotective in global brain ischemia. Proc. Natl. Acad. Sci. USA 1998, 95, 15769–15774. [Google Scholar] [CrossRef]
  99. Crack, P.J.; Gould, J.; Bye, N.; Ross, S.; Ali, U.; Habgood, M.D.; Morganti-Kossman, C.; Saunders, N.R.; Hertzog, P.J. Victorian Neurotrauma Research Group The genomic profile of the cerebral cortex after closed head injury in mice: Effects of minocycline. J. Neural. Transm. 2009, 116, 1–12. [Google Scholar] [CrossRef]
  100. Lin, S.; Zhang, Y.; Dodel, R.; Farlow, M.R.; Paul, S.M.; Du, Y. Minocycline blocks nitric oxide-induced neurotoxicity by inhibition p38 MAP kinase in rat cerebellar granule neurons. Neurosci. Lett. 2001, 315, 61–64. [Google Scholar] [CrossRef]
  101. Tikka, T.M.; Koistinaho, J.E. Minocycline provides neuroprotection against N-methyl-D-aspartate neurotoxicity by inhibiting microglia. J. Immunol. 2001, 166, 7527–7533. [Google Scholar] [CrossRef]
  102. Kovesdi, E.; Kamnaksh, A.; Wingo, D.; Ahmed, F.; Grunberg, N.E.; Long, J.B.; Kasper, C.E.; Agoston, D.V. Acute Minocycline Treatment Mitigates the Symptoms of Mild Blast-Induced Traumatic Brain Injury. Front. Neurol. 2012, 3. [Google Scholar] [CrossRef]
  103. Hanlon, L.A.; Huh, J.W.; Raghupathi, R. Minocycline Transiently Reduces Microglia/Macrophage Activation but Exacerbates Cognitive Deficits Following Repetitive Traumatic Brain Injury in the Neonatal Rat. J. Neuropathol. Exp. Neurol. 2016, 75, 214–226. [Google Scholar] [CrossRef] [PubMed]
  104. Pechacek, K.M.; Reck, A.M.; Frankot, M.A.; Vonder Haar, C. Minocycline fails to treat chronic traumatic brain injury-induced impulsivity and attention deficits. Exp. Neurol. 2021, 348, 113924. [Google Scholar] [CrossRef] [PubMed]
  105. Sangobowale, M.; Nikulina, E.; Bergold, P.J. Minocycline plus N-acetylcysteine protect oligodendrocytes when first dosed 12 h after closed head injury in mice. Neurosci. Lett. 2018, 682, 16–20. [Google Scholar] [CrossRef] [PubMed]
  106. Sangobowale, M.A.; Grin’kina, N.M.; Whitney, K.; Nikulina, E.; St. Laurent-Ariot, K.; Ho, J.S.; Bayzan, N.; Bergold, P.J. Minocycline plus N-Acetylcysteine Reduce Behavioral Deficits and Improve Histology with a Clinically Useful Time Window. J. Neurotrauma 2018, 35, 907–917. [Google Scholar] [CrossRef] [PubMed]
  107. Abdel Baki, S.G.; Schwab, B.; Haber, M.; Fenton, A.A.; Bergold, P.J. Minocycline Synergizes with N-Acetylcysteine and Improves Cognition and Memory Following Traumatic Brain Injury in Rats. PLoS ONE 2010, 5, e12490. [Google Scholar] [CrossRef] [PubMed]
  108. Haber, M.; James, J.; Kim, J.; Sangobowale, M.; Irizarry, R.; Ho, J.; Nikulina, E.; Grin’kina, N.M.; Ramadani, A.; Hartman, I.; et al. Minocycline plus N-acteylcysteine induces remyelination, synergistically protects oligodendrocytes and modifies neuroinflammation in a rat model of mild traumatic brain injury. J. Cereb. Blood Flow Metab. 2018, 38, 1312–1326. [Google Scholar] [CrossRef] [PubMed]
  109. Koulaeinejad, N.; Haddadi, K.; Ehteshami, S.; Shafizad, M.; Salehifar, E.; Emadian, O.; Ali Mohammadpour, R.; Ala, S. Effects of Minocycline on Neurological Outcomes In Patients with Acute Traumatic Brain Injury: A Pilot Study. Iran. J. Pharm. Res. 2019, 18, 1086–1096. [Google Scholar] [CrossRef]
  110. Scott, G.; Zetterberg, H.; Jolly, A.; Cole, J.H.; De Simoni, S.; Jenkins, P.O.; Feeney, C.; Owen, D.R.; Lingford-Hughes, A.; Howes, O.; et al. Minocycline reduces chronic microglial activation after brain trauma but increases neurodegeneration. Brain 2018, 141, 459–471. [Google Scholar] [CrossRef]
  111. Russell, G.; Graveley, R.; Seid, J.; Al-Humidan, A.K.; Skjodt, H. Mechanisms of action of cyclosporine and effects on connective tissues. Semin. Arthritis Rheum. 1992, 21, 16–22. [Google Scholar] [CrossRef]
  112. Li, P.A.; Kristián, T.; He, Q.P.; Siesjö, B.K. Cyclosporin A Enhances Survival, Ameliorates Brain Damage, and Prevents Secondary Mitochondrial Dysfunction after a 30-Minute Period of Transient Cerebral Ischemia. Exp. Neurol. 2000, 165, 153–163. [Google Scholar] [CrossRef]
  113. Sullivan, P.G.; Thompson, M.; Scheff, S.W. Continuous Infusion of Cyclosporin A Postinjury Significantly Ameliorates Cortical Damage Following Traumatic Brain Injury. Exp. Neurol. 2000, 161, 631–637. [Google Scholar] [CrossRef]
  114. Mbye, L.H.; Singh, I.N.; Sullivan, P.G.; Springer, J.E.; Hall, E.D. Attenuation of acute mitochondrial dysfunction after traumatic brain injury in mice by NIM811, a non-immunosuppressive cyclosporin A analog. Exp. Neurol. 2008, 209, 243–253. [Google Scholar] [CrossRef] [PubMed]
  115. Sullivan, P.G.; Thompson, M.B.; Scheff, S.W. Cyclosporin A attenuates acute mitochondrial dysfunction following traumatic brain injury. Exp. Neurol. 1999, 160, 226–234. [Google Scholar] [CrossRef] [PubMed]
  116. Kilbaugh, T.J.; Bhandare, S.; Lorom, D.H.; Saraswati, M.; Robertson, C.L.; Margulies, S.S. Cyclosporin A preserves mitochondrial function after traumatic brain injury in the immature rat and piglet. J. Neurotrauma 2011, 28, 763–774. [Google Scholar] [CrossRef] [PubMed]
  117. Signoretti, S.; Marmarou, A.; Tavazzi, B.; Dunbar, J.; Amorini, A.M.; Lazzarino, G.; Vagnozzi, R. The Protective Effect of Cyclosporin A upon N-Acetylaspartate and Mitochondrial Dysfunction following Experimental Diffuse Traumatic Brain Injury. J. Neurotrauma 2004, 21, 1154–1167. [Google Scholar] [CrossRef] [PubMed]
  118. Dölen, G.; Darvishzadeh, A.; Huang, K.W.; Malenka, R.C. Social reward requires coordinated activity of nucleus accumbens oxytocin and serotonin. Nature 2013, 501, 179–184. [Google Scholar] [CrossRef] [PubMed]
  119. Yuan, L.; Liu, S.; Bai, X.; Gao, Y.; Liu, G.; Wang, X.; Liu, D.; Li, T.; Hao, A.; Wang, Z. Oxytocin inhibits lipopolysaccharide-induced inflammation in microglial cells and attenuates microglial activation in lipopolysaccharide-treated mice. J. Neuroinflamm. 2016, 13, 77. [Google Scholar] [CrossRef]
  120. Wang, S.-C.; Lin, C.-C.; Chen, C.-C.; Tzeng, N.-S.; Liu, Y.-P. Effects of Oxytocin on Fear Memory and Neuroinflammation in a Rodent Model of Posttraumatic Stress Disorder. Int. J. Mol. Sci. 2018, 19, 3848. [Google Scholar] [CrossRef]
  121. Mairesse, J.; Zinni, M.; Pansiot, J.; Hassan-Abdi, R.; Demene, C.; Colella, M.; Charriaut-Marlangue, C.; Rideau Batista Novais, A.; Tanter, M.; Maccari, S.; et al. Oxytocin receptor agonist reduces perinatal brain damage by targeting microglia. Glia 2019, 67, 345–359. [Google Scholar] [CrossRef]
  122. Karelina, K.; Stuller, K.A.; Jarrett, B.; Zhang, N.; Wells, J.; Norman, G.J.; DeVries, A.C. Oxytocin Mediates Social Neuroprotection After Cerebral Ischemia. Stroke 2011, 42, 3606–3611. [Google Scholar] [CrossRef]
  123. Zinni, M.; Colella, M.; Batista Novais, A.R.; Baud, O.; Mairesse, J. Modulating the Oxytocin System During the Perinatal Period: A New Strategy for Neuroprotection of the Immature Brain? Front. Neurol. 2018, 9, 229. [Google Scholar] [CrossRef] [PubMed]
  124. 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]
  125. Digicaylioglu, M.; Lipton, S.A. Erythropoietin-mediated neuroprotection involves cross-talk between Jak2 and NF-kappaB signalling cascades. Nature 2001, 412, 641–647. [Google Scholar] [CrossRef] [PubMed]
  126. Byts, N.; Samoylenko, A.; Fasshauer, T.; Ivanisevic, M.; Hennighausen, L.; Ehrenreich, H.; Sirén, A.-L. Essential role for Stat5 in the neurotrophic but not in the neuroprotective effect of erythropoietin. Cell Death Differ. 2008, 15, 783–792. [Google Scholar] [CrossRef]
  127. Lu, D.; Mahmood, A.; Qu, C.; Goussev, A.; Schallert, T.; Chopp, M. Erythropoietin enhances neurogenesis and restores spatial memory in rats after traumatic brain injury. J. Neurotrauma 2005, 22, 1011–1017. [Google Scholar] [CrossRef]
  128. Peng, W.; Xing, Z.; Yang, J.; Wang, Y.; Wang, W.; Huang, W. The efficacy of erythropoietin in treating experimental traumatic brain injury: A systematic review of controlled trials in animal models. J. Neurosurg. 2014, 121, 653–664. [Google Scholar] [CrossRef]
  129. Xiong, Y.; Mahmood, A.; Meng, Y.; Zhang, Y.; Qu, C.; Schallert, T.; Chopp, M. Delayed administration of erythropoietin reducing hippocampal cell loss, enhancing angiogenesis and neurogenesis, and improving functional outcome following traumatic brain injury in rats: Comparison of treatment with single and triple dose. J. Neurosurg. 2010, 113, 598–608. [Google Scholar] [CrossRef]
  130. Nichol, A.; French, C.; Little, L.; Haddad, S.; Presneill, J.; Arabi, Y.; Bailey, M.; Cooper, D.J.; Duranteau, J.; Huet, O.; et al. Erythropoietin in traumatic brain injury (EPO-TBI): A double-blind randomised controlled trial. Lancet 2015, 386, 2499–2506. [Google Scholar] [CrossRef]
  131. Katiyar, V.; Chaturvedi, A.; Sharma, R.; Gurjar, H.K.; Goda, R.; Singla, R.; Ganeshkumar, A. Meta-Analysis with Trial Sequential Analysis on the Efficacy and Safety of Erythropoietin in Traumatic Brain Injury: A New Paradigm. World Neurosurg. 2020, 142, 465–475. [Google Scholar] [CrossRef]
  132. Jantzie, L.; El Demerdash, N.; Newville, J.C.; Robinson, S. Time to reconsider extended erythropoietin treatment for infantile traumatic brain injury? Exp. Neurol. 2019, 318, 205–215. [Google Scholar] [CrossRef]
  133. Tuttolomondo, A.; Pecoraro, R.; Pinto, A. Studies of selective TNF inhibitors in the treatment of brain injury from stroke and trauma: A review of the evidence to date. Drug Des. Dev. Ther. 2014, 8, 2221–2238. [Google Scholar] [CrossRef] [PubMed]
  134. Chio, C.-C.; Lin, J.-W.; Chang, M.-W.; Wang, C.-C.; Kuo, J.-R.; Yang, C.-Z.; Chang, C.-P. Therapeutic evaluation of etanercept in a model of traumatic brain injury. J. Neurochem. 2010, 115, 921–929. [Google Scholar] [CrossRef] [PubMed]
  135. Cheong, C.-U.; Chang, C.-P.; Chao, C.-M.; Cheng, B.-C.; Yang, C.-Z.; Chio, C.-C. Etanercept attenuates traumatic brain injury in rats by reducing brain TNF- α contents and by stimulating newly formed neurogenesis. Mediat. Inflamm. 2013, 2013, 620837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Chio, C.-C.; Chang, C.-H.; Wang, C.-C.; Cheong, C.-U.; Chao, C.-M.; Cheng, B.-C.; Yang, C.-Z.; Chang, C.-P. Etanercept attenuates traumatic brain injury in rats by reducing early microglial expression of tumor necrosis factor-α. BMC Neurosci. 2013, 14, 33. [Google Scholar] [CrossRef]
  137. Tobinick, E.; Rodriguez-Romanacce, H.; Levine, A.; Ignatowski, T.A.; Spengler, R.N. Immediate neurological recovery following perispinal etanercept years after brain injury. Clin. Drug Investig. 2014, 34, 361–366. [Google Scholar] [CrossRef]
  138. Ignatowski, T.A.; Spengler, R.N.; Dhandapani, K.M.; Folkersma, H.; Butterworth, R.F.; Tobinick, E. Perispinal etanercept for post-stroke neurological and cognitive dysfunction: Scientific rationale and current evidence. CNS Drugs 2014, 28, 679–697. [Google Scholar] [CrossRef]
  139. Tobinick, E.; Kim, N.M.; Reyzin, G.; Rodriguez-Romanacce, H.; DePuy, V. Selective TNF inhibition for chronic stroke and traumatic brain injury: An observational study involving 629 consecutive patients treated with perispinal etanercept. CNS Drugs 2012, 26, 1051–1070. [Google Scholar] [CrossRef]
  140. Tehranian, R.; Andell-Jonsson, S.; Beni, S.M.; Yatsiv, I.; Shohami, E.; Bartfai, T.; Lundkvist, J.; Iverfeldt, K. Improved recovery and delayed cytokine induction after closed head injury in mice with central overexpression of the secreted isoform of the interleukin-1 receptor antagonist. J. Neurotrauma 2002, 19, 939–951. [Google Scholar] [CrossRef]
  141. Jones, N.C.; Prior, M.J.W.; Burden-Teh, E.; Marsden, C.A.; Morris, P.G.; Murphy, S. Antagonism of the interleukin-1 receptor following traumatic brain injury in the mouse reduces the number of nitric oxide synthase-2-positive cells and improves anatomical and functional outcomes. Eur. J. Neurosci. 2005, 22, 72–78. [Google Scholar] [CrossRef]
  142. Flygt, J.; Ruscher, K.; Norberg, A.; Mir, A.; Gram, H.; Clausen, F.; Marklund, N. Neutralization of Interleukin-1β following Diffuse Traumatic Brain Injury in the Mouse Attenuates the Loss of Mature Oligodendrocytes. J. Neurotrauma 2018, 35, 2837–2849. [Google Scholar] [CrossRef]
  143. Newell, E.A.; Todd, B.P.; Mahoney, J.; Pieper, A.A.; Ferguson, P.J.; Bassuk, A.G. Combined Blockade of Interleukin-1α and -1β Signaling Protects Mice from Cognitive Dysfunction after Traumatic Brain Injury. eNeuro 2018, 5, 1–15. [Google Scholar] [CrossRef] [PubMed]
  144. Knoblach, S.M.; Faden, A.I. Cortical interleukin-1 beta elevation after traumatic brain injury in the rat: No effect of two selective antagonists on motor recovery. Neurosci. Lett. 2000, 289, 5–8. [Google Scholar] [CrossRef]
  145. Evans, L.P.; Woll, A.W.; Wu, S.; Todd, B.P.; Hehr, N.; Hedberg-Buenz, A.; Anderson, M.G.; Newell, E.A.; Ferguson, P.J.; Mahajan, V.B.; et al. Modulation of Post-Traumatic Immune Response Using the IL-1 Receptor Antagonist Anakinra for Improved Visual Outcomes. J. Neurotrauma 2020, 37, 1463–1480. [Google Scholar] [CrossRef]
  146. Clausen, F.; Hånell, A.; Björk, M.; Hillered, L.; Mir, A.K.; Gram, H.; Marklund, N. Neutralization of interleukin-1beta modifies the inflammatory response and improves histological and cognitive outcome following traumatic brain injury in mice. Eur. J. Neurosci. 2009, 30, 385–396. [Google Scholar] [CrossRef] [PubMed]
  147. Clausen, F.; Hånell, A.; Israelsson, C.; Hedin, J.; Ebendal, T.; Mir, A.K.; Gram, H.; Marklund, N. Neutralization of interleukin-1β reduces cerebral edema and tissue loss and improves late cognitive outcome following traumatic brain injury in mice. Eur. J. Neurosci. 2011, 34, 110–123. [Google Scholar] [CrossRef]
  148. Helmy, A.; Guilfoyle, M.R.; Carpenter, K.L.H.; Pickard, J.D.; Menon, D.K.; Hutchinson, P.J. Recombinant human interleukin-1 receptor antagonist in severe traumatic brain injury: A phase II randomized control trial. J. Cereb. Blood Flow Metab. 2014, 34, 845–851. [Google Scholar] [CrossRef]
  149. Yang, S.H.; Gustafson, J.; Gangidine, M.; Stepien, D.; Schuster, R.; Pritts, T.A.; Goodman, M.D.; Remick, D.G.; Lentsch, A.B. A murine model of mild traumatic brain injury exhibiting cognitive and motor deficits. J. Surg. Res. 2013, 184, 981–988. [Google Scholar] [CrossRef]
  150. Hergenroeder, G.W.; Moore, A.N.; McCoy, J.P.; Samsel, L.; Ward, N.H.; Clifton, G.L.; Dash, P.K. Serum IL-6: A candidate biomarker for intracranial pressure elevation following isolated traumatic brain injury. J. Neuroinflamm. 2010, 7, 19. [Google Scholar] [CrossRef]
  151. Yang, S.H.; Gangidine, M.; Pritts, T.A.; Goodman, M.D.; Lentsch, A.B. Interleukin 6 mediates neuroinflammation and motor coordination deficits after mild traumatic brain injury and brief hypoxia in mice. Shock 2013, 40, 471–475. [Google Scholar] [CrossRef]
  152. Müller, S.; Bianchi, M.E.; Knapp, S. Thermodynamics of HMGB1 interaction with duplex DNA. Biochemistry 2001, 40, 10254–10261. [Google Scholar] [CrossRef]
  153. Wang, H.; Vishnubhakat, J.M.; Bloom, O.; Zhang, M.; Ombrellino, M.; Sama, A.; Tracey, K.J. Proinflammatory cytokines (tumor necrosis factor and interleukin 1) stimulate release of high mobility group protein-1 by pituicytes. Surgery 1999, 126, 389–392. [Google Scholar] [CrossRef]
  154. Gao, H.-M.; Zhou, H.; Zhang, F.; Wilson, B.C.; Kam, W.; Hong, J.-S. HMGB1 acts on microglia Mac1 to mediate chronic neuroinflammation that drives progressive neurodegeneration. J. Neurosci. 2011, 31, 1081–1092. [Google Scholar] [CrossRef] [PubMed]
  155. Gao, T.; Chen, Z.; Chen, H.; Yuan, H.; Wang, Y.; Peng, X.; Wei, C.; Yang, J.; Xu, C. Inhibition of HMGB1 mediates neuroprotection of traumatic brain injury by modulating the microglia/macrophage polarization. Biochem. Biophys. Res. Commun. 2018, 497, 430–436. [Google Scholar] [CrossRef] [PubMed]
  156. Laird, M.D.; Shields, J.S.; Sukumari-Ramesh, S.; Kimbler, D.E.; Fessler, R.D.; Shakir, B.; Youssef, P.; Yanasak, N.; Vender, J.R.; Dhandapani, K.M. High mobility group box protein-1 promotes cerebral edema after traumatic brain injury via activation of toll-like receptor 4: HMGB1-TLR4 Signaling Promotes Brain Edema after TBI. Glia 2014, 62, 26–38. [Google Scholar] [CrossRef]
  157. Yang, L.; Wang, F.; Yang, L.; Yuan, Y.; Chen, Y.; Zhang, G.; Fan, Z. HMGB1 a-Box Reverses Brain Edema and Deterioration of Neurological Function in a Traumatic Brain Injury Mouse Model. Cell Physiol. Biochem. 2018, 46, 2532–2542. [Google Scholar] [CrossRef]
  158. Okuma, Y.; Wake, H.; Teshigawara, K.; Takahashi, Y.; Hishikawa, T.; Yasuhara, T.; Mori, S.; Takahashi, H.K.; Date, I.; Nishibori, M. Anti-High Mobility Group Box 1 Antibody Therapy May Prevent Cognitive Dysfunction After Traumatic Brain Injury. World Neurosurg. 2019, 122, e864–e871. [Google Scholar] [CrossRef]
  159. Cousin, M. Le cent cinquantenaire du chloroforme. Un agent anesthésique plus merveilleux et terrible encore que l’éther. Ann. Françaises Anesth. Reanim. 1997, 16, 1037–1044. [Google Scholar] [CrossRef]
  160. Alkire, M.T.; Haier, R.J.; Fallon, J.H. Toward a Unified Theory of Narcosis: Brain Imaging Evidence for a Thalamocortical Switch as the Neurophysiologic Basis of Anesthetic-Induced Unconsciousness. Conscious. Cogn. 2000, 9, 370–386. [Google Scholar] [CrossRef]
  161. Eilers, H.; Kindler, C.H.; Bickler, P.E. Different Effects of Volatile Anesthetics and Polyhalogenated Alkanes on Depolarization-Evoked Glutamate Release in Rat Cortical Brain Slices. Anesth. Analg. 1999, 88, 1168–1174. [Google Scholar] [CrossRef]
  162. Larsen, M.; Grøndahl, T.Ø.; Haugstad, T.S.; Langmoen, I.A. The effect of the volatile anesthetic isoflurane on Ca2+-dependent glutamate release from rat cerebral cortex. Brain Res. 1994, 663, 335–337. [Google Scholar] [CrossRef]
  163. Jones, M.V.; Harrison, N.L. Effects of volatile anesthetics on the kinetics of inhibitory postsynaptic currents in cultured rat hippocampal neurons. J. Neurophysiol. 1993, 70, 1339–1349. [Google Scholar] [CrossRef] [PubMed]
  164. Lenz, C.; Frietsch, T.; Fütterer, C.; Rebel, A.; van Ackern, K.; Kuschinsky, W.; Waschke, K.F. Local Coupling of Cerebral Blood Flow to Cerebral Glucose Metabolism during Inhalational Anesthesia in Rats: Desflurane versus Isoflurane. Anesthesiology 1999, 91, 1720. [Google Scholar] [CrossRef] [PubMed]
  165. Mielck, F.; Stephan, H.; Buhre, W.; Weyland, A.; Sonntag, H. Effects of 1 MAC desflurane on cerebral metabolism, blood flow and carbon dioxide reactivity in humans. Br. J. Anaesth. 1998, 81, 155–160. [Google Scholar] [CrossRef] [PubMed]
  166. Yin, J.; Li, H.; Feng, C.; Zuo, Z. Inhibition of brain ischemia-caused notch activation in microglia may contribute to isoflurane postconditioning-induced neuroprotection in male rats. CNS Neurol. Disord. Drug Targets 2014, 13, 718–732. [Google Scholar] [CrossRef] [PubMed]
  167. Yuan, M.; Ge, M.; Yin, J.; Dai, Z.; Xie, L.; Li, Y.; Liu, X.; Peng, L.; Zhang, G.; Si, J.; et al. Isoflurane post-conditioning down-regulates expression of aquaporin 4 in rats with cerebral ischemia/reperfusion injury and is possibly related to bone morphogenetic protein 4/Smad1/5/8 signaling pathway. Biomed. Pharmacother. 2018, 97, 429–438. [Google Scholar] [CrossRef]
  168. Wang, S.; Yin, J.; Ge, M.; Dai, Z.; Li, Y.; Si, J.; Ma, K.; Li, L.; Yao, S. Transforming growth-beta 1 contributes to isoflurane postconditioning against cerebral ischemia-reperfusion injury by regulating the c-Jun N-terminal kinase signaling pathway. Biomed. Pharmacother. 2016, 78, 280–290. [Google Scholar] [CrossRef]
  169. Statler, K.D.; Alexander, H.; Vagni, V.; Holubkov, R.; Dixon, C.E.; Clark, R.S.B.; Jenkins, L.; Kochanek, P.M. Isoflurane exerts neuroprotective actions at or near the time of severe traumatic brain injury. Brain Res. 2006, 1076, 216–224. [Google Scholar] [CrossRef]
  170. Statler, K.D.; Kochanek, P.M.; Dixon, C.E.; Alexander, H.L.; Warner, D.S.; Clark, R.S.; Wisniewski, S.R.; Graham, S.H.; Jenkins, L.W.; Marion, D.W.; et al. Isoflurane improves long-term neurologic outcome versus fentanyl after traumatic brain injury in rats. J. Neurotrauma 2000, 17, 1179–1189. [Google Scholar] [CrossRef]
  171. Hertle, D.; Beynon, C.; Zweckberger, K.; Vienenkötter, B.; Jung, C.S.; Kiening, K.; Unterberg, A.; Sakowitz, O.W. Influence of isoflurane on neuronal death and outcome in a rat model of traumatic brain injury. Acta Neurochir. Suppl. 2012, 114, 383–386. [Google Scholar] [CrossRef]
  172. Beck-Schimmer, B.; Baumann, L.; Restin, T.; Eugster, P.; Hasler, M.; Booy, C.; Schläpfer, M. Sevoflurane attenuates systemic inflammation compared with propofol, but does not modulate neuro-inflammation: A laboratory rat study. Eur. J. Anaesthesiol. 2017, 34, 764–775. [Google Scholar] [CrossRef]
  173. Hwang, J.-W.; Jeon, Y.-T.; Lim, Y.-J.; Park, H.-P. Sevoflurane Postconditioning-Induced Anti-Inflammation via Inhibition of the Toll-Like Receptor-4/Nuclear Factor Kappa B Pathway Contributes to Neuroprotection against Transient Global Cerebral Ischemia in Rats. Int. J. Mol. Sci. 2017, 18, 2347. [Google Scholar] [CrossRef] [PubMed]
  174. Dang, D.-D.; Saiyin, H.; Yu, Q.; Liang, W.-M. Effects of sevoflurane preconditioning on microglia/macrophage dynamics and phagocytosis profile against cerebral ischemia in rats. CNS Neurosci. Ther. 2018, 24, 564–571. [Google Scholar] [CrossRef]
  175. Chen, X.; Zhou, X.; Yang, L.; Miao, X.; Lu, D.-H.; Yang, X.-Y.; Zhou, Z.-B.; Kang, W.-B.; Chen, K.-Y.; Zhou, L.-H.; et al. Neonatal Exposure to Low-Dose (1.2%) Sevoflurane Increases Rats’ Hippocampal Neurogenesis and Synaptic Plasticity in Later Life. Neurotox. Res. 2018, 34, 188–197. [Google Scholar] [CrossRef] [PubMed]
  176. Lai, Z.; Zhang, L.; Su, J.; Cai, D.; Xu, Q. Sevoflurane postconditioning improves long-term learning and memory of neonatal hypoxia-ischemia brain damage rats via the PI3K/Akt-mPTP pathway. Brain Res. 2016, 1630, 25–37. [Google Scholar] [CrossRef] [PubMed]
  177. Zhang, H.; Xiong, X.; Liu, J.; Gu, L.; Li, F.; Wan, Y.; Xu, S. Emulsified Isoflurane Protects Against Transient Focal Cerebral Ischemia Injury in Rats via the PI3K/Akt Signaling Pathway. Anesth. Analg. 2016, 122, 1377–1384. [Google Scholar] [CrossRef]
  178. Pi, Z.; Lin, H.; Yang, J. Isoflurane reduces pain and inhibits apoptosis of myocardial cells through the phosphoinositide 3-kinase/protein kinase B signaling pathway in mice during cardiac surgery. Mol. Med. Rep. 2018, 17, 6497–6505. [Google Scholar] [CrossRef]
  179. Thal, S.C.; Luh, C.; Schaible, E.-V.; Timaru-Kast, R.; Hedrich, J.; Luhmann, H.J.; Engelhard, K.; Zehendner, C.M. Volatile anesthetics influence blood-brain barrier integrity by modulation of tight junction protein expression in traumatic brain injury. PLoS ONE 2012, 7, e50752. [Google Scholar] [CrossRef]
  180. Statler, K.D.; Alexander, H.; Vagni, V.; Dixon, C.E.; Clark, R.S.B.; Jenkins, L.; Kochanek, P.M. Comparison of Seven Anesthetic Agents on Outcome after Experimental Traumatic Brain Injury in Adult, Male Rats. J. Neurotrauma 2006, 23, 97–108. [Google Scholar] [CrossRef]
  181. Neag, M.-A.; Mitre, A.-O.; Catinean, A.; Mitre, C.-I. An Overview on the Mechanisms of Neuroprotection and Neurotoxicity of Isoflurane and Sevoflurane in Experimental Studies. Brain Res. Bull. 2020, 165, 281–289. [Google Scholar] [CrossRef]
  182. Maze, M.; Laitio, T. Neuroprotective Properties of Xenon. Mol. Neurobiol. 2020, 57, 118–124. [Google Scholar] [CrossRef]
  183. Ma, D.; Wilhelm, S.; Maze, M.; Franks, N.P. Neuroprotective and neurotoxic properties of the “inert” gas, xenon. Br. J. Anaesth. 2002, 89, 739–746. [Google Scholar] [CrossRef] [PubMed]
  184. Banks, P.; Franks, N.P.; Dickinson, R. Competitive inhibition at the glycine site of the N-methyl-D-aspartate receptor mediates xenon neuroprotection against hypoxia-ischemia. Anesthesiology 2010, 112, 614–622. [Google Scholar] [CrossRef] [PubMed]
  185. Campos-Pires, R.; Armstrong, S.P.; Sebastiani, A.; Luh, C.; Gruss, M.; Radyushkin, K.; Hirnet, T.; Werner, C.; Engelhard, K.; Franks, N.P.; et al. Xenon improves neurologic outcome and reduces secondary injury following trauma in an in vivo model of traumatic brain injury. Crit. Care Med. 2015, 43, 149–158. [Google Scholar] [CrossRef] [Green Version]
  186. Campos-Pires, R.; Hirnet, T.; Valeo, F.; Ong, B.E.; Radyushkin, K.; Aldhoun, J.; Saville, J.; Edge, C.J.; Franks, N.P.; Thal, S.C.; et al. Xenon improves long-term cognitive function, reduces neuronal loss and chronic neuroinflammation, and improves survival after traumatic brain injury in mice. Br. J. Anaesth. 2019, 123, 60–73. [Google Scholar] [CrossRef] [PubMed]
  187. Campos-Pires, R.; Onggradito, H.; Ujvari, E.; Karimi, S.; Valeo, F.; Aldhoun, J.; Edge, C.J.; Franks, N.P.; Dickinson, R. Xenon treatment after severe traumatic brain injury improves locomotor outcome, reduces acute neuronal loss and enhances early beneficial neuroinflammation: A randomized, blinded, controlled animal study. Crit. Care 2020, 24, 667. [Google Scholar] [CrossRef] [PubMed]
  188. Filev, A.D.; Silachev, D.N.; Ryzhkov, I.A.; Lapin, K.N.; Babkina, A.S.; Grebenchikov, O.A.; Pisarev, V.M. Effect of Xenon Treatment on Gene Expression in Brain Tissue after Traumatic Brain Injury in Rats. Brain Sci. 2021, 11, 889. [Google Scholar] [CrossRef] [PubMed]
  189. Abraini, J.H.; Kriem, B.; Balon, N.; Rostain, J.-C.; Risso, J.-J. Gamma-aminobutyric acid neuropharmacological investigations on narcosis produced by nitrogen, argon, or nitrous oxide. Anesth. Analg. 2003, 96, 746–749. [Google Scholar] [CrossRef]
  190. Sanders, R.D.; Ma, D.; Maze, M. Argon neuroprotection. Crit. Care 2010, 14, 117. [Google Scholar] [CrossRef]
  191. David, H.N.; Haelewyn, B.; Degoulet, M.; Colomb, D.G.; Risso, J.-J.; Abraini, J.H. Ex Vivo and In Vivo Neuroprotection Induced by Argon When Given after an Excitotoxic or Ischemic Insult. PLoS ONE 2012, 7, e30934. [Google Scholar] [CrossRef]
  192. Zhuang, L.; Yang, T.; Zhao, H.; Fidalgo, A.R.; Vizcaychipi, M.P.; Sanders, R.D.; Yu, B.; Takata, M.; Johnson, M.R.; Ma, D. The protective profile of argon, helium, and xenon in a model of neonatal asphyxia in rats. Crit. Care Med. 2012, 40, 1724–1730. [Google Scholar] [CrossRef]
  193. Creed, J.; Cantillana-Riquelme, V.; Yan, B.H.; Ma, S.; Chu, D.; Wang, H.; Turner, D.A.; Laskowitz, D.T.; Hoffmann, U. Argon Inhalation for 24 h After Closed-Head Injury Does not Improve Recovery, Neuroinflammation, or Neurologic Outcome in Mice. Neurocrit. Care 2021, 34, 833–843. [Google Scholar] [CrossRef] [PubMed]
  194. Moro, F.; Fossi, F.; Magliocca, A.; Pascente, R.; Sammali, E.; Baldini, F.; Tolomeo, D.; Micotti, E.; Citerio, G.; Stocchetti, N.; et al. Efficacy of acute administration of inhaled argon on traumatic brain injury in mice. Br. J. Anaesth. 2021, 126, 256–264. [Google Scholar] [CrossRef] [PubMed]
  195. Kawaguchi, M.; Furuya, H.; Patel, P.M. Neuroprotective effects of anesthetic agents. J. Anesth. 2005, 19, 150–156. [Google Scholar] [CrossRef] [PubMed]
  196. Kaur, J.; Flores Gutiérrez, J.; Nistri, A. Neuroprotective effect of propofol against excitotoxic injury to locomotor networks of the rat spinal cord in vitro. Eur. J. Neurosci. 2016, 44, 2418–2430. [Google Scholar] [CrossRef]
  197. Koerner, I.P.; Brambrink, A.M. Brain protection by anesthetic agents. Curr. Opin. Anaesthesiol. 2006, 19, 481–486. [Google Scholar] [CrossRef] [PubMed]
  198. Eberspcher, E.; Heimann, K.; Hollweck, R.; Werner, C.; Schneider, G.; Engelhard, K. The Effect of Electroencephalogram-Targeted High- and Low-Dose Propofol Infusion on Histopathological Damage After Traumatic Brain Injury in the Rat. Anesth. Analg. 2006, 103, 1527–1533. [Google Scholar] [CrossRef] [PubMed]
  199. Ma, J.; Xiao, W.; Wang, J.; Wu, J.; Ren, J.; Hou, J.; Gu, J.; Fan, K.; Yu, B. Propofol Inhibits NLRP3 Inflammasome and Attenuates Blast-Induced Traumatic Brain Injury in Rats. Inflammation 2016, 39, 2094–2103. [Google Scholar] [CrossRef]
  200. Peters, C.E.; Korcok, J.; Gelb, A.W.; Wilson, J.X. Anesthetic Concentrations of Propofol Protect against Oxidative Stress in Primary Astrocyte Cultures. Anesthesiology 2001, 94, 313–321. [Google Scholar] [CrossRef] [PubMed]
  201. Ding, Z.; Zhang, J.; Xu, J.; Sheng, G.; Huang, G. Propofol administration modulates AQP-4 expression and brain edema after traumatic brain injury. Cell Biochem. Biophys. 2013, 67, 615–622. [Google Scholar] [CrossRef] [PubMed]
  202. Luo, T.; Wu, J.; Kabadi, S.V.; Sabirzhanov, B.; Guanciale, K.; Hanscom, M.; Faden, J.; Cardiff, K.; Bengson, C.J.; Faden, A.I. Propofol limits microglial activation after experimental brain trauma through inhibition of nicotinamide adenine dinucleotide phosphate oxidase. Anesthesiology 2013, 119, 1370–1388. [Google Scholar] [CrossRef]
  203. Liu, F.; Chen, M.-R.; Liu, J.; Zou, Y.; Wang, T.-Y.; Zuo, Y.-X.; Wang, T.-H. Propofol administration improves neurological function associated with inhibition of pro-inflammatory cytokines in adult rats after traumatic brain injury. Neuropeptides 2016, 58, 1–6. [Google Scholar] [CrossRef] [PubMed]
  204. Thal, S.C.; Timaru-Kast, R.; Wilde, F.; Merk, P.; Johnson, F.; Frauenknecht, K.; Sebastiani, A.; Sommer, C.; Staib-Lasarzik, I.; Werner, C.; et al. Propofol impairs neurogenesis and neurologic recovery and increases mortality rate in adult rats after traumatic brain injury. Crit. Care Med. 2014, 42, 129–141. [Google Scholar] [CrossRef] [PubMed]
  205. Sun, J.; Wang, L.; Shen, J.; Wang, Z.; Qian, Y. Effect of propofol on mucous permeability and inflammatory mediators expression in the intestine following traumatic brain injury in rats. Cytokine 2007, 40, 151–156. [Google Scholar] [CrossRef] [PubMed]
  206. Wang, D.; Xu, X.; Wu, Y.-G.; Lyu, L.; Zhou, Z.-W.; Zhang, J.-N. Dexmedetomidine attenuates traumatic brain injury: Action pathway and mechanisms. Neural. Regen. Res. 2018, 13, 819–826. [Google Scholar] [CrossRef]
  207. Zhang, M.-H.; Zhou, X.-M.; Cui, J.-Z.; Wang, K.-J.; Feng, Y.; Zhang, H.-A. Neuroprotective effects of dexmedetomidine on traumatic brain injury: Involvement of neuronal apoptosis and HSP70 expression. Mol. Med. Rep. 2018, 17, 8079–8086. [Google Scholar] [CrossRef] [PubMed]
  208. Wu, J.; Vogel, T.; Gao, X.; Lin, B.; Kulwin, C.; Chen, J. Neuroprotective effect of dexmedetomidine in a murine model of traumatic brain injury. Sci. Rep. 2018, 8, 4935. [Google Scholar] [CrossRef]
  209. Shen, M.; Wang, S.; Wen, X.; Han, X.-R.; Wang, Y.-J.; Zhou, X.-M.; Zhang, M.-H.; Wu, D.-M.; Lu, J.; Zheng, Y.-L. Dexmedetomidine exerts neuroprotective effect via the activation of the PI3K/Akt/mTOR signaling pathway in rats with traumatic brain injury. Biomed. Pharmacother. 2017, 95, 885–893. [Google Scholar] [CrossRef]
  210. Qiu, Z.; Lu, P.; Wang, K.; Zhao, X.; Li, Q.; Wen, J.; Zhang, H.; Li, R.; Wei, H.; Lv, Y.; et al. Dexmedetomidine Inhibits Neuroinflammation by Altering Microglial M1/M2 Polarization Through MAPK/ERK Pathway. Neurochem. Res. 2020, 45, 345–353. [Google Scholar] [CrossRef]
  211. Peng, J.; Zhang, P.; Zheng, H.; Ren, Y.-Q.; Yan, H. Dexmedetomidine reduces hippocampal microglia inflammatory response induced by surgical injury through inhibiting NLRP3. Chin. J. Traumatol. 2019, 22, 161–165. [Google Scholar] [CrossRef]
  212. Zheng, B.; Zhang, S.; Ying, Y.; Guo, X.; Li, H.; Xu, L.; Ruan, X. Administration of Dexmedetomidine inhibited NLRP3 inflammasome and microglial cell activities in hippocampus of traumatic brain injury rats. Biosci. Rep. 2018, 38, BSR20180892. [Google Scholar] [CrossRef] [PubMed]
  213. Ding, M.; Chen, Y.; Luan, H.; Zhang, X.; Zhao, Z.; Wu, Y. Dexmedetomidine reduces inflammation in traumatic brain injury by regulating the inflammatory responses of macrophages and splenocytes. Exp. Ther. Med. 2019, 18, 2323–2331. [Google Scholar] [CrossRef] [PubMed]
  214. Réus, G.Z.; Matias, B.I.; Maciel, A.L.; Abelaira, H.M.; Ignácio, Z.M.; de Moura, A.B.; Matos, D.; Danielski, L.G.; Petronilho, F.; Carvalho, A.F.; et al. Mechanism of synergistic action on behavior, oxidative stress and inflammation following co-treatment with ketamine and different antidepressant classes. Pharmacol. Rep. 2017, 69, 1094–1102. [Google Scholar] [CrossRef] [PubMed]
  215. Luggya, T.S.; Roche, T.; Ssemogerere, L.; Kintu, A.; Kasumba, J.M.; Kwizera, A.; Tindimwebwa, J.V. Effect of low-dose ketamine on post-operative serum IL-6 production among elective surgical patients: A randomized clinical trial. Afr. Health Sci. 2017, 17, 500–507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Marcoux, F.W.; Goodrich, J.E.; Dominick, M.A. Ketamine prevents ischemic neuronal injury. Brain Res. 1988, 452, 329–335. [Google Scholar] [CrossRef]
  217. Chang, L.C.; Raty, S.R.; Ortiz, J.; Bailard, N.S.; Mathew, S.J. The emerging use of ketamine for anesthesia and sedation in traumatic brain injuries. CNS Neurosci. Ther. 2013, 19, 390–395. [Google Scholar] [CrossRef]
  218. Liang, J.; Wu, S.; Xie, W.; He, H. Ketamine ameliorates oxidative stress-induced apoptosis in experimental traumatic brain injury via the Nrf2 pathway. DDDT 2018, 12, 845–853. [Google Scholar] [CrossRef]
  219. Wagner, A.K.; McCullough, E.H.; Niyonkuru, C.; Ozawa, H.; Loucks, T.L.; Dobos, J.A.; Brett, C.A.; Santarsieri, M.; Dixon, C.E.; Berga, S.L.; et al. Acute serum hormone levels: Characterization and prognosis after severe traumatic brain injury. J. Neurotrauma 2011, 28, 871–888. [Google Scholar] [CrossRef]
  220. Bazarian, J.J.; Blyth, B.; Mookerjee, S.; He, H.; McDermott, M.P. Sex differences in outcome after mild traumatic brain injury. J. Neurotrauma 2010, 27, 527–539. [Google Scholar] [CrossRef]
  221. Groswasser, Z.; Cohen, M.; Keren, O. Female TBI patients recover better than males. Brain Inj. 1998, 12, 805–808. [Google Scholar] [CrossRef]
  222. Garringer, J.A.; Niyonkuru, C.; McCullough, E.H.; Loucks, T.; Dixon, C.E.; Conley, Y.P.; Berga, S.; Wagner, A.K. Impact of Aromatase Genetic Variation on Hormone Levels and Global Outcome after Severe TBI. J. Neurotrauma 2013, 30, 1415–1425. [Google Scholar] [CrossRef]
  223. Roof, R.L.; Duvdevani, R.; Braswell, L.; Stein, D.G. Progesterone Facilitates Cognitive Recovery and Reduces Secondary Neuronal Loss Caused by Cortical Contusion Injury in Male Rats. Exp. Neurol. 1994, 129, 64–69. [Google Scholar] [CrossRef] [PubMed]
  224. Gatson, J.W.; Warren, V.; Abdelfattah, K.; Wolf, S.; Hynan, L.S.; Moore, C.; Diaz-Arrastia, R.; Minei, J.P.; Madden, C.; Wigginton, J.G. Detection of β-amyloid oligomers as a predictor of neurological outcome after brain injury: Laboratory investigation. J. Neurosurg. 2013, 118, 1336–1342. [Google Scholar] [CrossRef] [PubMed]
  225. Pepe, P.; Wigginton, J.; Gatson, J.; Simpkins, J.; Maass, D.; AbdelFattah, K.; Idris, A.; Warren, V.; Minei, J. Single-dose estrogen infusion can amplify brain levels of Sonic hedgehog, a signal protein for neuro stem cells and repair following the indirect brain injury resulting after severe torso burns. Crit. Care 2013, 17. [Google Scholar] [CrossRef] [Green Version]
  226. McClean, J.; Nuñez, J.L. 17α-Estradiol is neuroprotective in male and female rats in a model of early brain injury. Exp. Neurol. 2008, 210, 41–50. [Google Scholar] [CrossRef] [PubMed]
  227. Neese, S.L.; Clough, R.W.; Banz, W.J.; Smith, D.C. Z-Bisdehydrodoisynolic acid (Z-BDDA): An estrogenic seco-steroid that enhances behavioral recovery following moderate fluid percussion brain injury in male rats. Brain Res. 2010, 1362, 93–101. [Google Scholar] [CrossRef]
  228. Behl, C.; Manthey, D. Neuroprotective activites of estrogen: An update. J. Neurocytol. 2000, 29, 351–358. [Google Scholar] [CrossRef]
  229. Green, P.S.; Yang, S.-H.; Simpkins, J.W. Neuroprotective Effects of Phenolic A Ring Oestrogens. In Novartis Foundation Symposia; Chadwick, D.J., Goode, J.A., Eds.; John Wiley & Sons, Ltd: Chichester, UK, 2008; pp. 202–220. ISBN 978-0-471-49203-0. [Google Scholar]
  230. Brann, D.W.; Dhandapani, K.; Wakade, C.; Mahesh, V.B.; Khan, M.M. Neurotrophic and neuroprotective actions of estrogen: Basic mechanisms and clinical implications. Steroids 2007, 72, 381–405. [Google Scholar] [CrossRef] [PubMed]
  231. Arevalo, M.-A.; Santos-Galindo, M.; Bellini, M.-J.; Azcoitia, I.; Garcia-Segura, L.M. Actions of estrogens on glial cells: Implications for neuroprotection. Biochim. Biophys. Acta 2010, 1800, 1106–1112. [Google Scholar] [CrossRef]
  232. Bruce-Keller, A.J.; Keeling, J.L.; Keller, J.N.; Huang, F.F.; Camondola, S.; Mattson, M.P. Antiinflammatory Effects of Estrogen on Microglial Activation. Endocrinology 2000, 141, 3646–3656. [Google Scholar] [CrossRef]
  233. Emerson, C.S.; Headrick, J.P.; Vink, R. Estrogen improves biochemical and neurologic outcome following traumatic brain injury in male rats, but not in females. Brain Res. 1993, 608, 95–100. [Google Scholar] [CrossRef]
  234. Kim, H.; Cam-Etoz, B.; Zhai, G.; Hubbard, W.J.; Zinn, K.R.; Chaudry, I.H. Salutary Effects of Estrogen Sulfate for Traumatic Brain Injury. J. Neurotrauma 2015, 32, 1210–1216. [Google Scholar] [CrossRef] [PubMed]
  235. Wang, J.; Hou, Y.; Zhang, L.; Liu, M.; Zhao, J.; Zhang, Z.; Ma, Y.; Hou, W. Estrogen Attenuates Traumatic Brain Injury by Inhibiting the Activation of Microglia and Astrocyte-Mediated Neuroinflammatory Responses. Mol. Neurobiol. 2021, 58, 1052–1061. [Google Scholar] [CrossRef]
  236. Li, L.-Z.; Bao, Y.-J.; Zhao, M. 17beta-estradiol attenuates programmed cell death in cortical pericontusional zone following traumatic brain injury via upregulation of ERalpha and inhibition of caspase-3 activation. Neurochem. Int. 2011, 58, 126–133. [Google Scholar] [CrossRef]
  237. Kövesdi, E.; Szabó-Meleg, E.; Abrahám, I.M. The Role of Estradiol in Traumatic Brain Injury: Mechanism and Treatment Potential. Int. J. Mol. Sci. 2020, 22, 11. [Google Scholar] [CrossRef] [PubMed]
  238. Lan, Y.-L.; Wang, X.; Zou, Y.-J.; Xing, J.-S.; Lou, J.-C.; Zou, S.; Ma, B.-B.; Ding, Y.; Zhang, B. Bazedoxifene protects cerebral autoregulation after traumatic brain injury and attenuates impairments in blood-brain barrier damage: Involvement of anti-inflammatory pathways by blocking MAPK signaling. Inflamm. Res. 2019, 68, 311–323. [Google Scholar] [CrossRef] [PubMed]
  239. Asl, S.Z.; Khaksari, M.; Khachki, A.S.; Shahrokhi, N.; Nourizade, S. Contribution of estrogen receptors alpha and beta in the brain response to traumatic brain injury. J. Neurosurg. 2013, 119, 353–361. [Google Scholar] [CrossRef]
  240. Khaksari, M.; Abbasloo, E.; Dehghan, F.; Soltani, Z.; Asadikaram, G. The brain cytokine levels are modulated by estrogen following traumatic brain injury: Which estrogen receptor serves as modulator? Int. Immunopharmacol. 2015, 28, 279–287. [Google Scholar] [CrossRef]
  241. Roof, R.L.; Hall, E.D. Gender Differences in Acute CNS Trauma and Stroke: Neuroprotective Effects of Estrogen and Progesterone. J. Neurotrauma 2000, 17, 367–388. [Google Scholar] [CrossRef] [PubMed]
  242. Singh, M.; Su, C. Progesterone and neuroprotection. Horm. Behav. 2013, 63, 284–290. [Google Scholar] [CrossRef]
  243. Stein, D.G. A clinical/translational perspective: Can a developmental hormone play a role in the treatment of traumatic brain injury? Horm. Behav. 2013, 63, 291–300. [Google Scholar] [CrossRef]
  244. Roof, R.L.; Duvdevani, R.; Heyburn, J.W.; Stein, D.G. Progesterone Rapidly Decreases Brain Edema: Treatment Delayed up to 24 Hours Is Still Effective. Exp. Neurol. 1996, 138, 246–251. [Google Scholar] [CrossRef] [PubMed]
  245. Wright, D.W.; Bauer, M.E.; Hoffman, S.W.; Stein, D.G. Serum Progesterone Levels Correlate with Decreased Cerebral Edema after Traumatic Brain Injury in Male Rats. J. Neurotrauma 2001, 18, 901–909. [Google Scholar] [CrossRef]
  246. Guo, Q.; Sayeed, I.; Baronne, L.M.; Hoffman, S.W.; Guennoun, R.; Stein, D.G. Progesterone administration modulates AQP4 expression and edema after traumatic brain injury in male rats. Exp. Neurol. 2006, 198, 469–478. [Google Scholar] [CrossRef]
  247. Vink, R.; Van Den Heuvel, C. Recent advances in the development of multifactorial therapies for the treatment of traumatic brain injury. Expert Opin. Investig. Drugs 2004, 13, 1263–1274. [Google Scholar] [CrossRef]
  248. Chen, G.; Shi, J.; Jin, W.; Wang, L.; Xie, W.; Sun, J.; Hang, C. Progesterone administration modulates TLRs/NF-kappaB signaling pathway in rat brain after cortical contusion. Ann. Clin. Lab. Sci. 2008, 38, 65–74. [Google Scholar] [PubMed]
  249. Pan, D.-S.; Liu, W.-G.; Yang, X.-F.; Cao, F. Inhibitory effect of progesterone on inflammatory factors after experimental traumatic brain injury. Biomed. Environ. Sci. 2007, 20, 432–438. [Google Scholar] [PubMed]
  250. Roof, R.L.; Hoffman, S.W.; Stein, D.G. Progesterone protects against lipid peroxidation following traumatic brain injury in rats. Mol. Chem. Neuropathol. 1997, 31, 1–11. [Google Scholar] [CrossRef]
  251. Ghoumari, A.M.; Ibanez, C.; El-Etr, M.; Leclerc, P.; Eychenne, B.; O’Malley, B.W.; Baulieu, E.E.; Schumacher, M. Progesterone and its metabolites increase myelin basic protein expression in organotypic slice cultures of rat cerebellum. J. Neurochem. 2003, 86, 848–859. [Google Scholar] [CrossRef]
  252. Koenig, H.L.; Schumacher, M.; Ferzaz, B.; Thi, A.N.; Ressouches, A.; Guennoun, R.; Jung-Testas, I.; Robel, P.; Akwa, Y.; Baulieu, E.E. Progesterone synthesis and myelin formation by Schwann cells. Science 1995, 268, 1500–1503. [Google Scholar] [CrossRef]
  253. Ibanez, C.; Shields, S.A.; El-Etr, M.; Baulieu, E.-E.; Schumacher, M.; Franklin, R.J.M. Systemic progesterone administration results in a partial reversal of the age-associated decline in CNS remyelination following toxin-induced demyelination in male rats: Progesterone and CNS remyelination. Neuropathol. Appl. Neurobiol. 2004, 30, 80–89. [Google Scholar] [CrossRef]
  254. Wagner, C.K. Progesterone receptors and neural development: A gap between bench and bedside? Endocrinology 2008, 149, 2743–2749. [Google Scholar] [CrossRef]
  255. Wagner, C.K. The many faces of progesterone: A role in adult and developing male brain. Front. Neuroendocrinol. 2006, 27, 340–359. [Google Scholar] [CrossRef]
  256. Quadros, P.S.; Wagner, C.K. Regulation of Progesterone Receptor Expression by Estradiol Is Dependent on Age, Sex and Region in the Rat Brain. Endocrinology 2008, 149, 3054–3061. [Google Scholar] [CrossRef] [Green Version]
  257. Quadros, P.S.; Pfau, J.L.; Wagner, C.K. Distribution of progesterone receptor immunoreactivity in the fetal and neonatal rat forebrain. J. Comp. Neurol. 2007, 504, 42–56. [Google Scholar] [CrossRef] [PubMed]
  258. Uysal, N.; Baykara, B.; Kiray, M.; Cetin, F.; Aksu, I.; Dayi, A.; Gurpinar, T.; Ozdemir, D.; Arda, M.N. Combined treatment with progesterone and magnesium sulfate positively affects traumatic brain injury in immature rats. Turk. Neurosurg. 2013, 23, 129–137. [Google Scholar] [CrossRef]
  259. Baykara, B.; Aksu, I.; Buyuk, E.; Kiray, M.; Sisman, A.; Baykara, B.; Dayi, A.; Tas, A.; Ozdemir, D.; Arda, M.; et al. Progesterone treatment decreases traumatic brain injury induced anxiety and is correlated with increased serum IGF-1 levels; prefrontal cortex, amygdala, hippocampus neuron density; and reduced serum corticosterone levels in immature rats. Biotech. Histochem. 2013, 88, 250–257. [Google Scholar] [CrossRef]
  260. Trotter, A.; Bokelmann, B.; Sorgo, W.; Bechinger-Kornhuber, D.; Heinemann, H.; Schmücker, G.; Oesterle, M.; Köhntop, B.; Brisch, K.H.; Pohlandt, F. Follow-up examination at the age of 15 months of extremely preterm infants after postnatal estradiol and progesterone replacement. J. Clin. Endocrinol. Metab. 2001, 86, 601–603. [Google Scholar] [CrossRef]
  261. Trotter, A.; Steinmacher, J.; Kron, M.; Pohlandt, F. Neurodevelopmental follow-up at five years corrected age of extremely low birth weight infants after postnatal replacement of 17β-estradiol and progesterone. J. Clin. Endocrinol. Metab. 2012, 97, 1041–1047. [Google Scholar] [CrossRef]
  262. Stein, D.G. Progesterone exerts neuroprotective effects after brain injury. Brain Res. Rev. 2008, 57, 386–397. [Google Scholar] [CrossRef]
  263. Wright, D.W.; Kellermann, A.L.; Hertzberg, V.S.; Clark, P.L.; Frankel, M.; Goldstein, F.C.; Salomone, J.P.; Dent, L.L.; Harris, O.A.; Ander, D.S.; et al. ProTECT: A Randomized Clinical Trial of Progesterone for Acute Traumatic Brain Injury. Ann. Emerg. Med. 2007, 49, 391–402.e2. [Google Scholar] [CrossRef] [PubMed]
  264. Xiao, G.; Wei, J.; Yan, W.; Wang, W.; Lu, Z. Improved outcomes from the administration of progesterone for patients with acute severe traumatic brain injury: A randomized controlled trial. Crit. Care 2008, 12, R61. [Google Scholar] [CrossRef] [PubMed]
  265. Nikbakht, H.; Aminmansour, B.; Ghorbani, A.; Rahmani, P.; Nourian, M.; Rezvani, M.; Torkashvand, M.; Moradi, M. Comparison of the administration of progesterone versus progesterone and vitamin D in improvement of outcomes in patients with traumatic brain injury: A randomized clinical trial with placebo group. Adv. Biomed. Res. 2012, 1, 58. [Google Scholar] [CrossRef] [PubMed]
  266. Shakeri, M.; Boustani, M.R.; Pak, N.; Panahi, F.; Salehpour, F.; Lotfinia, I.; Meshkini, A.; Daghighi, S.; vahedi, P.; Khani, M.; et al. Effect of progesterone administration on prognosis of patients with diffuse axonal injury due to severe head trauma. Clin. Neurol. Neurosurg. 2013, 115, 2019–2022. [Google Scholar] [CrossRef]
  267. Wright, D.W.; Yeatts, S.D.; Silbergleit, R.; Palesch, Y.Y.; Hertzberg, V.S.; Frankel, M.; Goldstein, F.C.; Caveney, A.F.; Howlett-Smith, H.; Bengelink, E.M.; et al. Very Early Administration of Progesterone for Acute Traumatic Brain Injury. N. Engl. J. Med. 2014, 371, 2457–2466. [Google Scholar] [CrossRef]
  268. Skolnick, B.E.; Maas, A.I.; Narayan, R.K.; van der Hoop, R.G.; MacAllister, T.; Ward, J.D.; Nelson, N.R.; Stocchetti, N. A Clinical Trial of Progesterone for Severe Traumatic Brain Injury. N. Engl. J. Med. 2014, 371, 2467–2476. [Google Scholar] [CrossRef]
  269. Hultquist, D.E.; Xu, F.; Quandt, K.S.; Shlafer, M.; Mack, C.P.; Till, G.O.; Seekamp, A.; Betz, A.L.; Ennis, S.R. Evidence that NADPH-dependent methemoglobin reductase and administered riboflavin protect tissues from oxidative injury. Am. J. Hematol. 1993, 42, 13–18. [Google Scholar] [CrossRef]
  270. Lin, Y.; Desbois, A.; Jiang, S.; Hou, S.T. Group B vitamins protect murine cerebellar granule cells from glutamate/NMDA toxicity. Neuroreport 2004, 15, 2241–2244. [Google Scholar] [CrossRef]
  271. Hoane, M.R.; Wolyniak, J.G.; Akstulewicz, S.L. Administration of riboflavin improves behavioral outcome and reduces edema formation and glial fibrillary acidic protein expression after traumatic brain injury. J. Neurotrauma 2005, 22, 1112–1122. [Google Scholar] [CrossRef]
  272. Barbre, A.B.; Hoane, M.R. Magnesium and riboflavin combination therapy following cortical contusion injury in the rat. Brain Res. Bull. 2006, 69, 639–646. [Google Scholar] [CrossRef]
  273. Vonder Haar, C. The Use of Nicotinamide as a Treatment for Experimental Traumatic Brain Injury and Stroke: A Review and Evaluation. Clin. Pharm. Biopharm. 2013, S1. [Google Scholar] [CrossRef]
  274. Maiese, K.; Chong, Z.Z.; Hou, J.; Shang, Y.C. The vitamin nicotinamide: Translating nutrition into clinical care. Molecules 2009, 14, 3446–3485. [Google Scholar] [CrossRef] [PubMed]
  275. Hoane, M.R.; Pierce, J.L.; Holland, M.A.; Anderson, G.D. Nicotinamide treatment induces behavioral recovery when administered up to 4 h following cortical contusion injury in the rat. Neuroscience 2008, 154, 861–868. [Google Scholar] [CrossRef] [PubMed]
  276. Vonder Haar, C.; Maass, W.R.; Jacobs, E.A.; Hoane, M.R. Deficits in discrimination after experimental frontal brain injury are mediated by motivation and can be improved by nicotinamide administration. J. Neurotrauma 2014, 31, 1711–1720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Hoane, M.R.; Akstulewicz, S.L.; Toppen, J. Treatment with vitamin B3 improves functional recovery and reduces GFAP expression following traumatic brain injury in rats. J. Neurotrauma 2003, 20, 1189–1199. [Google Scholar] [CrossRef] [PubMed]
  278. Goffus, A.M.; Anderson, G.D.; Hoane, M. Sustained delivery of nicotinamide limits cortical injury and improves functional recovery following traumatic brain injury. Oxid. Med. Cell Longev. 2010, 3, 145–152. [Google Scholar] [CrossRef]
  279. Hoane, M.R.; Pierce, J.L.; Kaufman, N.A.; Beare, J.E. Variation in chronic nicotinamide treatment after traumatic brain injury can alter components of functional recovery independent of histological damage. Oxid. Med. Cell Longev. 2008, 1, 46–53. [Google Scholar] [CrossRef]
  280. Hoane, M.R.; Tan, A.A.; Pierce, J.L.; Anderson, G.D.; Smith, D.C. Nicotinamide treatment reduces behavioral impairments and provides cortical protection after fluid percussion injury in the rat. J. Neurotrauma 2006, 23, 1535–1548. [Google Scholar] [CrossRef]
  281. Hoane, M.R.; Kaplan, S.A.; Ellis, A.L. The effects of nicotinamide on apoptosis and blood-brain barrier breakdown following traumatic brain injury. Brain Res. 2006, 1125, 185–193. [Google Scholar] [CrossRef]
  282. Vonder Haar, C.; Anderson, G.D.; Hoane, M.R. Continuous nicotinamide administration improves behavioral recovery and reduces lesion size following bilateral frontal controlled cortical impact injury. Behav. Brain Res. 2011, 224, 311–317. [Google Scholar] [CrossRef]
  283. Bender, D.A. Non-nutritional uses of vitamin B6. Br. J. Nutr. 1999, 81, 7–20. [Google Scholar] [CrossRef]
  284. Hwang, I.K.; Yoo, K.-Y.; Kim, D.H.; Lee, B.-H.; Kwon, Y.-G.; Won, M.H. Time course of changes in pyridoxal 5’-phosphate (vitamin B6 active form) and its neuroprotection in experimental ischemic damage. Exp. Neurol. 2007, 206, 114–125. [Google Scholar] [CrossRef] [PubMed]
  285. Oka, T. Modulation of gene expression by vitamin B6. Nutr. Res. Rev. 2001, 14, 257–266. [Google Scholar] [CrossRef] [PubMed]
  286. Cabrini, L.; Bergami, R.; Fiorentini, D.; Marchetti, M.; Landi, L.; Tolomelli, B. Vitamin B6 deficiency affects antioxidant defences in rat liver and heart. Biochem. Mol. Biol. Int. 1998, 46, 689–697. [Google Scholar] [CrossRef]
  287. Kuypers, N.J.; Hoane, M.R. Pyridoxine administration improves behavioral and anatomical outcome after unilateral contusion injury in the rat. J. Neurotrauma 2010, 27, 1275–1282. [Google Scholar] [CrossRef]
  288. Xu, Y.; Sladky, J.T.; Brown, M.J. Dose-dependent expression of neuronopathy after experimental pyridoxine intoxication. Neurology 1989, 39, 1077–1083. [Google Scholar] [CrossRef] [PubMed]
  289. Krinke, G.; Schaumburg, H.H.; Spencer, P.S.; Suter, J.; Thomann, P.; Hess, R. Pyridoxine megavitaminosis produces degeneration of peripheral sensory neurons (sensory neuronopathy) in the dog. Neurotoxicology 1981, 2, 13–24. [Google Scholar]
  290. Fenech, M. The role of folic acid and Vitamin B12 in genomic stability of human cells. Mutat. Res. 2001, 475, 57–67. [Google Scholar] [CrossRef]
  291. Naim, M.Y.; Friess, S.; Smith, C.; Ralston, J.; Ryall, K.; Helfaer, M.A.; Margulies, S.S. Folic acid enhances early functional recovery in a piglet model of pediatric head injury. Dev. Neurosci. 2010, 32, 466–479. [Google Scholar] [CrossRef]
  292. Vonder Haar, C.; Emery, M.A.; Hoane, M.R. Chronic folic acid administration confers no treatment effects in either a high or low dose following unilateral controlled cortical impact injury in the rat. Restor. Neurol. Neurosci. 2012, 30, 291–302. [Google Scholar] [CrossRef]
  293. Grünewald, R.A. Ascorbic acid in the brain. Brain Res. Brain Res. Rev. 1993, 18, 123–133. [Google Scholar] [CrossRef]
  294. Rice, M.E. Ascorbate regulation and its neuroprotective role in the brain. Trends Neurosci. 2000, 23, 209–216. [Google Scholar] [CrossRef]
  295. Awasthi, D.; Church, D.F.; Torbati, D.; Carey, M.E.; Pryor, W.A. Oxidative stress following traumatic brain injury in rats. Surg. Neurol. 1997, 47, 575–581, discussion 581–582. [Google Scholar] [CrossRef]
  296. Tyurin, V.A.; Tyurina, Y.Y.; Borisenko, G.G.; Sokolova, T.V.; Ritov, V.B.; Quinn, P.J.; Rose, M.; Kochanek, P.; Graham, S.H.; Kagan, V.E. Oxidative stress following traumatic brain injury in rats: Quantitation of biomarkers and detection of free radical intermediates. J. Neurochem. 2000, 75, 2178–2189. [Google Scholar] [CrossRef] [PubMed]
  297. Polidori, M.C.; Mecocci, P.; Frei, B. Plasma vitamin C levels are decreased and correlated with brain damage in patients with intracranial hemorrhage or head trauma. Stroke 2001, 32, 898–902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Moor, E.; Shohami, E.; Kanevsky, E.; Grigoriadis, N.; Symeonidou, C.; Kohen, R. Impairment of the ability of the injured aged brain in elevating urate and ascorbate. Exp. Gerontol. 2006, 41, 303–311. [Google Scholar] [CrossRef] [PubMed]
  299. Wang, K.-W.; Wang, H.-K.; Chen, H.-J.; Liliang, P.-C.; Liang, C.-L.; Tsai, Y.-D.; Cho, C.-L.; Lu, K. Simvastatin Combined with Antioxidant Attenuates the Cerebral Vascular Endothelial Inflammatory Response in a Rat Traumatic Brain Injury. BioMed Res. Int. 2014, 2014, 1–6. [Google Scholar] [CrossRef]
  300. Ishaq, G.M.; Saidu, Y.; Bilbis, L.S.; Muhammad, S.A.; Jinjir, N.; Shehu, B.B. Effects of α-tocopherol and ascorbic acid in the severity and management of traumatic brain injury in albino rats. J. Neurosci. Rural Pract. 2013, 4, 292–297. [Google Scholar] [CrossRef]
  301. Lin, J.-L.; Huang, Y.-H.; Shen, Y.-C.; Huang, H.-C.; Liu, P.-H. Ascorbic acid prevents blood-brain barrier disruption and sensory deficit caused by sustained compression of primary somatosensory cortex. J. Cereb. Blood Flow Metab. 2010, 30, 1121–1136. [Google Scholar] [CrossRef]
  302. Zhao, J.; Moore, A.N.; Redell, J.B.; Dash, P.K. Enhancing expression of Nrf2-driven genes protects the blood brain barrier after brain injury. J. Neurosci. 2007, 27, 10240–10248. [Google Scholar] [CrossRef]
  303. Maekawa, T.; Uchida, T.; Nakata-Horiuchi, Y.; Kobayashi, H.; Kawauchi, S.; Kinoshita, M.; Saitoh, D.; Sato, S. Oral ascorbic acid 2-glucoside prevents coordination disorder induced via laser-induced shock waves in rat brain. PLoS ONE 2020, 15, e0230774. [Google Scholar] [CrossRef]
  304. Putzu, A.; Daems, A.-M.; Lopez-Delgado, J.C.; Giordano, V.F.; Landoni, G. The Effect of Vitamin C on Clinical Outcome in Critically Ill Patients: A Systematic Review With Meta-Analysis of Randomized Controlled Trials. Crit. Care Med. 2019, 47, 774–783. [Google Scholar] [CrossRef] [PubMed]
  305. Razmkon, A.; Sadidi, A.; Sherafat-Kazemzadeh, E.; Mehrafshan, A.; Jamali, M.; Malekpour, B.; Saghafinia, M. Administration of vitamin C and vitamin E in severe head injury: A randomized double-blind controlled trial. Clin. Neurosurg. 2011, 58, 133–137. [Google Scholar] [CrossRef] [PubMed]
  306. Leichtle, S.W.; Sarma, A.K.; Strein, M.; Yajnik, V.; Rivet, D.; Sima, A.; Brophy, G.M. High-Dose Intravenous Ascorbic Acid: Ready for Prime Time in Traumatic Brain Injury? Neurocrit. Care 2020, 32, 333–339. [Google Scholar] [CrossRef] [PubMed]
  307. Holick, M.F. Vitamin D Deficiency. N. Engl. J. Med. 2007, 357, 266–281. [Google Scholar] [CrossRef] [PubMed]
  308. Lawrence, D.W.; Sharma, B. A review of the neuroprotective role of vitamin D in traumatic brain injury with implications for supplementation post-concussion. Brain Inj. 2016, 30, 960–968. [Google Scholar] [CrossRef] [PubMed]
  309. Cekic, M.; Stein, D.G. Traumatic brain injury and aging: Is a combination of progesterone and vitamin D hormone a simple solution to a complex problem? Neurotherapeutics 2010, 7, 81–90. [Google Scholar] [CrossRef]
  310. Hua, F.; Reiss, J.I.; Tang, H.; Wang, J.; Fowler, X.; Sayeed, I.; Stein, D.G. Progesterone and low-dose vitamin D hormone treatment enhances sparing of memory following traumatic brain injury. Horm. Behav. 2012, 61, 642–651. [Google Scholar] [CrossRef]
  311. Tang, H.; Hua, F.; Wang, J.; Yousuf, S.; Atif, F.; Sayeed, I.; Stein, D.G. Progesterone and vitamin D combination therapy modulates inflammatory response after traumatic brain injury. Brain Inj. 2015, 29, 1165–1174. [Google Scholar] [CrossRef]
  312. Tang, H.; Hua, F.; Wang, J.; Sayeed, I.; Wang, X.; Chen, Z.; Yousuf, S.; Atif, F.; Stein, D.G. Progesterone and vitamin D: Improvement after traumatic brain injury in middle-aged rats. Horm. Behav. 2013, 64, 527–538. [Google Scholar] [CrossRef] [PubMed]
  313. Yang, J.; Wang, K.; Hu, T.; Wang, G.; Wang, W.; Zhang, J. Vitamin D3 Supplement Attenuates Blood-Brain Barrier Disruption and Cognitive Impairments in a Rat Model of Traumatic Brain Injury. Neuromol. Med. 2021, 23, 491–499. [Google Scholar] [CrossRef] [PubMed]
  314. Lee, J.M.; Jeong, S.W.; Kim, M.Y.; Park, J.B.; Kim, M.S. The Effect of Vitamin D Supplementation in Patients with Acute Traumatic Brain Injury. World Neurosurg. 2019, 126, e1421–e1426. [Google Scholar] [CrossRef] [PubMed]
  315. Sharma, S.; Kumar, A.; Choudhary, A.; Sharma, S.; Khurana, L.; Sharma, N.; Kumar, V.; Bisht, A. Neuroprotective Role of Oral Vitamin D Supplementation on Consciousness and Inflammatory Biomarkers in Determining Severity Outcome in Acute Traumatic Brain Injury Patients: A Double-Blind Randomized Clinical Trial. Clin. Drug Investig. 2020, 40, 327–334. [Google Scholar] [CrossRef] [PubMed]
  316. Brigelius-Flohé, R.; Traber, M.G. Vitamin E: Function and metabolism. FASEB J. 1999, 13, 1145–1155. [Google Scholar] [CrossRef] [PubMed]
  317. Clifton, G.L.; Lyeth, B.G.; Jenkins, L.W.; Taft, W.C.; DeLorenzo, R.J.; Hayes, R.L. Effect of D, alpha-tocopheryl succinate and polyethylene glycol on performance tests after fluid percussion brain injury. J. Neurotrauma 1989, 6, 71–81. [Google Scholar] [CrossRef]
  318. Wu, A.; Ying, Z.; Gomez-Pinilla, F. Vitamin E protects against oxidative damage and learning disability after mild traumatic brain injury in rats. Neurorehabil. Neural Repair 2010, 24, 290–298. [Google Scholar] [CrossRef]
  319. Stein, D.G.; Halks-Miller, M.; Hoffman, S.W. Intracerebral administration of alpha-tocopherol-containing liposomes facilitates behavioral recovery in rats with bilateral lesions of the frontal cortex. J. Neurotrauma 1991, 8, 281–292. [Google Scholar] [CrossRef] [PubMed]
  320. Yang, J.; Han, Y.; Ye, W.; Liu, F.; Zhuang, K.; Wu, G. Alpha tocopherol treatment reduces the expression of Nogo-A and NgR in rat brain after traumatic brain injury. J. Surg. Res. 2013, 182, e69–e77. [Google Scholar] [CrossRef] [PubMed]
  321. Koç, R.K.; Kurtsoy, A.; Paşaoğlu, H.; Karaküçük, E.I.; Oktem, I.S.; Meral, M. Lipid peroxidation and oedema in experimental brain injury: Comparison of treatment with methylprednisolone, tirilazad mesylate and vitamin E. Res. Exp. Med. 1999, 199, 21–28. [Google Scholar] [CrossRef] [PubMed]
  322. Dobrovolny, J.; Smrcka, M.; Bienertova-Vasku, J. Therapeutic potential of vitamin E and its derivatives in traumatic brain injury-associated dementia. Neurol. Sci. 2018, 39, 989–998. [Google Scholar] [CrossRef]
  323. Hall, E.D.; Yonkers, P.A.; Andrus, P.K.; Cox, J.W.; Anderson, D.K. Biochemistry and pharmacology of lipid antioxidants in acute brain and spinal cord injury. J. Neurotrauma 1992, 9, S425–S442. [Google Scholar]
  324. Huskisson, E.; Maggini, S.; Ruf, M. The influence of micronutrients on cognitive function and performance. J. Int. Med. Res. 2007, 35, 1–19. [Google Scholar] [CrossRef]
  325. van den Heuvel, C.; Vink, R. The role of magnesium in traumatic brain injury. Clin. Calcium 2004, 14, 9–14. [Google Scholar] [CrossRef] [PubMed]
  326. Sen, A.P.; Gulati, A. Use of magnesium in traumatic brain injury. Neurotherapeutics 2010, 7, 91–99. [Google Scholar] [CrossRef]
  327. Vink, R.; Cook, N.L.; van den Heuvel, C. Magnesium in acute and chronic brain injury: An update. Magnes. Res. 2009, 22, 158S–162S. [Google Scholar] [CrossRef]
  328. Saatman, K.E.; Bareyre, F.M.; Grady, M.S.; McIntosh, T.K. Acute cytoskeletal alterations and cell death induced by experimental brain injury are attenuated by magnesium treatment and exacerbated by magnesium deficiency. J. Neuropathol. Exp. Neurol. 2001, 60, 183–194. [Google Scholar] [CrossRef] [Green Version]
  329. Muir, J.K.; Raghupathi, R.; Emery, D.L.; Bareyre, F.M.; McIntosh, T.K. Postinjury magnesium treatment attenuates traumatic brain injury-induced cortical induction of p53 mRNA in rats. Exp. Neurol. 1999, 159, 584–593. [Google Scholar] [CrossRef] [PubMed]
  330. Enomoto, T.; Osugi, T.; Satoh, H.; McIntosh, T.K.; Nabeshima, T. Pre-Injury magnesium treatment prevents traumatic brain injury-induced hippocampal ERK activation, neuronal loss, and cognitive dysfunction in the radial-arm maze test. J. Neurotrauma 2005, 22, 783–792. [Google Scholar] [CrossRef] [PubMed]
  331. Esen, F.; Erdem, T.; Aktan, D.; Kalayci, R.; Cakar, N.; Kaya, M.; Telci, L. Effects of magnesium administration on brain edema and blood-brain barrier breakdown after experimental traumatic brain injury in rats. J. Neurosurg. Anesthesiol. 2003, 15, 119–125. [Google Scholar] [CrossRef] [PubMed]
  332. Park, C.O.; Hyun, D.K. Apoptotic change in response to magnesium therapy after moderate diffuse axonal injury in rats. Yonsei Med. J. 2004, 45, 908–916. [Google Scholar] [CrossRef]
  333. Ghabriel, M.N.; Thomas, A.; Vink, R. Magnesium restores altered aquaporin-4 immunoreactivity following traumatic brain injury to a pre-injury state. Acta Neurochir. Suppl. 2006, 96, 402–406. [Google Scholar] [CrossRef]
  334. Bareyre, F.M.; Saatman, K.E.; Raghupathi, R.; McIntosh, T.K. Postinjury treatment with magnesium chloride attenuates cortical damage after traumatic brain injury in rats. J. Neurotrauma 2000, 17, 1029–1039. [Google Scholar] [CrossRef]
  335. Fromm, L.; Heath, D.L.; Vink, R.; Nimmo, A.J. Magnesium attenuates post-traumatic depression/anxiety following diffuse traumatic brain injury in rats. J. Am. Coll. Nutr. 2004, 23, 529S–533S. [Google Scholar] [CrossRef]
  336. Vink, R.; O’Connor, C.A.; Nimmo, A.J.; Heath, D.L. Magnesium attenuates persistent functional deficits following diffuse traumatic brain injury in rats. Neurosci. Lett. 2003, 336, 41–44. [Google Scholar] [CrossRef]
  337. Turner, R.J.; Dasilva, K.W.; O’Connor, C.; van den Heuvel, C.; Vink, R. Magnesium gluconate offers no more protection than magnesium sulphate following diffuse traumatic brain injury in rats. J. Am. Coll. Nutr. 2004, 23, 541S–544S. [Google Scholar] [CrossRef]
  338. Temkin, N.R.; Anderson, G.D.; Winn, H.R.; Ellenbogen, R.G.; Britz, G.W.; Schuster, J.; Lucas, T.; Newell, D.W.; Mansfield, P.N.; Machamer, J.E.; et al. Magnesium sulfate for neuroprotection after traumatic brain injury: A randomised controlled trial. Lancet Neurol. 2007, 6, 29–38. [Google Scholar] [CrossRef]
  339. Saver, J.L.; Starkman, S.; Eckstein, M.; Stratton, S.J.; Pratt, F.D.; Hamilton, S.; Conwit, R.; Liebeskind, D.S.; Sung, G.; Kramer, I.; et al. Prehospital use of magnesium sulfate as neuroprotection in acute stroke. N. Engl. J. Med. 2015, 372, 528–536. [Google Scholar] [CrossRef]
  340. Natale, J.E.; Guerguerian, A.-M.; Joseph, J.G.; McCarter, R.; Shao, C.; Slomine, B.; Christensen, J.; Johnston, M.V.; Shaffner, D.H. Pilot study to determine the hemodynamic safety and feasibility of magnesium sulfate infusion in children with severe traumatic brain injury. Pediatr. Crit. Care Med. 2007, 8, 1–9. [Google Scholar] [CrossRef]
  341. Levenson, C.W. Zinc and Traumatic Brain Injury: From Chelation to Supplementation. Med. Sci. 2020, 8, 36. [Google Scholar] [CrossRef]
  342. Portbury, S.D.; Hare, D.J.; Sgambelloni, C.; Finkelstein, D.I.; Adlard, P.A. A time-course analysis of changes in cerebral metal levels following a controlled cortical impact. Metallomics 2016, 8, 193–200. [Google Scholar] [CrossRef]
  343. Suh, S.W.; Chen, J.W.; Motamedi, M.; Bell, B.; Listiak, K.; Pons, N.F.; Danscher, G.; Frederickson, C.J. Evidence that synaptically-released zinc contributes to neuronal injury after traumatic brain injury. Brain Res. 2000, 852, 268–273. [Google Scholar] [CrossRef]
  344. Hellmich, H.L.; Eidson, K.A.; Capra, B.A.; Garcia, J.M.; Boone, D.R.; Hawkins, B.E.; Uchida, T.; Dewitt, D.S.; Prough, D.S. Injured Fluoro-Jade-positive hippocampal neurons contain high levels of zinc after traumatic brain injury. Brain Res. 2007, 1127, 119–126. [Google Scholar] [CrossRef]
  345. Stork, C.J.; Li, Y.V. Elevated Cytoplasmic Free Zinc and Increased Reactive Oxygen Species Generation in the Context of Brain Injury. Acta Neurochir. Suppl. 2016, 121, 347–353. [Google Scholar] [CrossRef]
  346. Isaev, N.K.; Stelmashook, E.V.; Genrikhs, E.E. Role of zinc and copper ions in the pathogenetic mechanisms of traumatic brain injury and Alzheimer’s disease. Rev. Neurosci. 2020, 31, 233–243. [Google Scholar] [CrossRef]
  347. Sensi, S.L.; Yin, H.Z.; Weiss, J.H. AMPA/kainate receptor-triggered Zn2+ entry into cortical neurons induces mitochondrial Zn2+ uptake and persistent mitochondrial dysfunction. Eur. J. Neurosci. 2000, 12, 3813–3818. [Google Scholar] [CrossRef]
  348. Kim, Y.-H.; Koh, J.-Y. The role of NADPH oxidase and neuronal nitric oxide synthase in zinc-induced poly(ADP-ribose) polymerase activation and cell death in cortical culture. Exp. Neurol. 2002, 177, 407–418. [Google Scholar] [CrossRef]
  349. Doering, P.; Stoltenberg, M.; Penkowa, M.; Rungby, J.; Larsen, A.; Danscher, G. Chemical blocking of zinc ions in CNS increases neuronal damage following traumatic brain injury (TBI) in mice. PLoS ONE 2010, 5, e10131. [Google Scholar] [CrossRef]
  350. Choi, B.Y.; Kim, J.H.; Kim, H.J.; Lee, B.E.; Kim, I.Y.; Sohn, M.; Suh, S.W. Zinc chelation reduces traumatic brain injury-induced neurogenesis in the subgranular zone of the hippocampal dentate gyrus. J. Trace Elem. Med. Biol. 2014, 28, 474–481. [Google Scholar] [CrossRef]
  351. Hellmich, H.L.; Frederickson, C.J.; DeWitt, D.S.; Saban, R.; Parsley, M.O.; Stephenson, R.; Velasco, M.; Uchida, T.; Shimamura, M.; Prough, D.S. Protective effects of zinc chelation in traumatic brain injury correlate with upregulation of neuroprotective genes in rat brain. Neurosci. Lett. 2004, 355, 221–225. [Google Scholar] [CrossRef]
  352. McClain, C.J.; Twyman, D.L.; Ott, L.G.; Rapp, R.P.; Tibbs, P.A.; Norton, J.A.; Kasarskis, E.J.; Dempsey, R.J.; Young, B. Serum and urine zinc response in head-injured patients. J. Neurosurg. 1986, 64, 224–230. [Google Scholar] [CrossRef]
  353. Cope, E.C.; Morris, D.R.; Scrimgeour, A.G.; VanLandingham, J.W.; Levenson, C.W. Zinc supplementation provides behavioral resiliency in a rat model of traumatic brain injury. Physiol. Behav. 2011, 104, 942–947. [Google Scholar] [CrossRef]
  354. Young, B.; Ott, L.; Kasarskis, E.; Rapp, R.; Moles, K.; Dempsey, R.J.; Tibbs, P.A.; Kryscio, R.; McClain, C. Zinc supplementation is associated with improved neurologic recovery rate and visceral protein levels of patients with severe closed head injury. J. Neurotrauma 1996, 13, 25–34. [Google Scholar] [CrossRef] [PubMed]
  355. Domínguez, M.I.; Blasco-Ibáñez, J.M.; Crespo, C.; Marqués-Marí, A.I.; Martínez-Guijarro, F.J. Zinc chelation during non-lesioning overexcitation results in neuronal death in the mouse hippocampus. Neuroscience 2003, 116, 791–806. [Google Scholar] [CrossRef]
  356. Hellmich, H.L.; Eidson, K.; Cowart, J.; Crookshanks, J.; Boone, D.K.; Shah, S.; Uchida, T.; DeWitt, D.S.; Prough, D.S. Chelation of neurotoxic zinc levels does not improve neurobehavioral outcome after traumatic brain injury. Neurosci. Lett. 2008, 440, 155–159. [Google Scholar] [CrossRef]
  357. Scrimgeour, A.G.; Carrigan, C.T.; Condlin, M.L.; Urso, M.L.; van den Berg, R.M.; van Helden, H.P.M.; Montain, S.J.; Joosen, M.J.A. Dietary Zinc Modulates Matrix Metalloproteinases in Traumatic Brain Injury. J. Neurotrauma 2018, 35, 2495–2506. [Google Scholar] [CrossRef]
  358. Cope, E.C.; Morris, D.R.; Scrimgeour, A.G.; Levenson, C.W. Use of zinc as a treatment for traumatic brain injury in the rat: Effects on cognitive and behavioral outcomes. Neurorehabil. Neural Repair 2012, 26, 907–913. [Google Scholar] [CrossRef]
  359. Khazdouz, M.; Mazidi, M.; Ehsaei, M.-R.; Ferns, G.; Kengne, A.P.; Norouzy, A.-R. Impact of Zinc Supplementation on the Clinical Outcomes of Patients with Severe Head Trauma: A Double-Blind Randomized Clinical Trial. J. Diet Suppl. 2018, 15, 1–10. [Google Scholar] [CrossRef]
  360. Niemoller, T.D.; Stark, D.T.; Bazan, N.G. Omega-3 fatty acid docosahexaenoic acid is the precursor of neuroprotectin D1 in the nervous system. World Rev. Nutr. Diet 2009, 99, 46–54. [Google Scholar] [CrossRef]
  361. Hasadsri, L.; Wang, B.H.; Lee, J.V.; Erdman, J.W.; Llano, D.A.; Barbey, A.K.; Wszalek, T.; Sharrock, M.F.; Wang, H.J. Omega-3 fatty acids as a putative treatment for traumatic brain injury. J. Neurotrauma 2013, 30, 897–906. [Google Scholar] [CrossRef]
  362. Michael-Titus, A.T.; Priestley, J.V. Omega-3 fatty acids and traumatic neurological injury: From neuroprotection to neuroplasticity? Trends Neurosci. 2014, 37, 30–38. [Google Scholar] [CrossRef]
  363. Wu, A.; Ying, Z.; Gomez-Pinilla, F. Dietary omega-3 fatty acids normalize BDNF levels, reduce oxidative damage, and counteract learning disability after traumatic brain injury in rats. J. Neurotrauma 2004, 21, 1457–1467. [Google Scholar] [CrossRef]
  364. Kumar, P.R.; Essa, M.M.; Al-Adawi, S.; Dradekh, G.; Memon, M.A.; Akbar, M.; Manivasagam, T. Omega-3 Fatty acids could alleviate the risks of traumatic brain injury—A mini review. J. Tradit. Complement Med. 2014, 4, 89–92. [Google Scholar] [CrossRef]
  365. Ferrucci, L.; Cherubini, A.; Bandinelli, S.; Bartali, B.; Corsi, A.; Lauretani, F.; Martin, A.; Andres-Lacueva, C.; Senin, U.; Guralnik, J.M. Relationship of plasma polyunsaturated fatty acids to circulating inflammatory markers. J. Clin. Endocrinol. Metab. 2006, 91, 439–446. [Google Scholar] [CrossRef]
  366. Vreugdenhil, M.; Bruehl, C.; Voskuyl, R.A.; Kang, J.X.; Leaf, A.; Wadman, W.J. Polyunsaturated fatty acids modulate sodium and calcium currents in CA1 neurons. Proc. Natl. Acad. Sci. USA 1996, 93, 12559–12563. [Google Scholar] [CrossRef]
  367. Ménard, C.; Patenaude, C.; Gagné, A.-M.; Massicotte, G. AMPA receptor-mediated cell death is reduced by docosahexaenoic acid but not by eicosapentaenoic acid in area CA1 of hippocampal slice cultures. J. Neurosci. Res. 2009, 87, 876–886. [Google Scholar] [CrossRef]
  368. Desai, A.; Kevala, K.; Kim, H.-Y. Depletion of brain docosahexaenoic acid impairs recovery from traumatic brain injury. PLoS ONE 2014, 9, e86472. [Google Scholar] [CrossRef] [Green Version]
  369. Wang, T.; Van, K.C.; Gavitt, B.J.; Grayson, J.K.; Lu, Y.-C.; Lyeth, B.G.; Pichakron, K.O. Effect of fish oil supplementation in a rat model of multiple mild traumatic brain injuries. Restor. Neurol. Neurosci. 2013, 31, 647–659. [Google Scholar] [CrossRef]
  370. Pu, H.; Guo, Y.; Zhang, W.; Huang, L.; Wang, G.; Liou, A.K.; Zhang, J.; Zhang, P.; Leak, R.K.; Wang, Y.; et al. Omega-3 polyunsaturated fatty acid supplementation improves neurologic recovery and attenuates white matter injury after experimental traumatic brain injury. J. Cereb. Blood Flow Metab. 2013, 33, 1474–1484. [Google Scholar] [CrossRef]
  371. Mills, J.D.; Hadley, K.; Bailes, J.E. Dietary supplementation with the omega-3 fatty acid docosahexaenoic acid in traumatic brain injury. Neurosurgery 2011, 68, 474–481, discussion 481. [Google Scholar] [CrossRef]
  372. Chen, X.; Chen, C.; Fan, S.; Wu, S.; Yang, F.; Fang, Z.; Fu, H.; Li, Y. Omega-3 polyunsaturated fatty acid attenuates the inflammatory response by modulating microglia polarization through SIRT1-mediated deacetylation of the HMGB1/NF-κB pathway following experimental traumatic brain injury. J. Neuroinflamm. 2018, 15, 116. [Google Scholar] [CrossRef]
  373. Zhang, E.; Wan, X.; Yang, L.; Wang, D.; Chen, Z.; Chen, Y.; Liu, M.; Zhang, G.; Wu, J.; Han, H.; et al. Omega-3 Polyunsaturated Fatty Acids Alleviate Traumatic Brain Injury by Regulating the Glymphatic Pathway in Mice. Front. Neurol. 2020, 11, 707. [Google Scholar] [CrossRef]
  374. Barrett, E.C.; McBurney, M.I.; Ciappio, E.D. ω-3 fatty acid supplementation as a potential therapeutic aid for the recovery from mild traumatic brain injury/concussion. Adv. Nutr. 2014, 5, 268–277. [Google Scholar] [CrossRef] [PubMed]
  375. Van Steenwinckel, J.; Schang, A.-L.; Krishnan, M.L.; Degos, V.; Delahaye-Duriez, A.; Bokobza, C.; Csaba, Z.; Verdonk, F.; Montané, A.; Sigaut, S.; et al. Decreased microglial Wnt/β-catenin signalling drives microglial pro-inflammatory activation in the developing brain. Brain 2019, 142, 3806–3833. [Google Scholar] [CrossRef] [PubMed]
  376. Dang, B.; Chen, W.; He, W.; Chen, G. Rehabilitation Treatment and Progress of Traumatic Brain Injury Dysfunction. Neural Plast. 2017, 2017, 1582182. [Google Scholar] [CrossRef] [PubMed]
  377. Kreuzer, P.M.; Landgrebe, M.; Frank, E.; Langguth, B. Repetitive transcranial magnetic stimulation for the treatment of chronic tinnitus after traumatic brain injury: A case study. J. Head Trauma Rehabil. 2013, 28, 386–389. [Google Scholar] [CrossRef] [PubMed]
  378. Neville, I.S.; Hayashi, C.Y.; El Hajj, S.A.; Zaninotto, A.L.C.; Sabino, J.P.; Sousa, L.M.; Nagumo, M.M.; Brunoni, A.R.; Shieh, B.D.F.S.; Amorim, R.L.O.; et al. Repetitive Transcranial Magnetic Stimulation (rTMS) for the cognitive rehabilitation of traumatic brain injury (TBI) victims: Study protocol for a randomized controlled trial. Trials 2015, 16, 440. [Google Scholar] [CrossRef] [Green Version]
  379. Fregni, F.; Li, S.; Zaninotto, A.; Santana Neville, I.; Paiva, W.; Nunn, D. Clinical utility of brain stimulation modalities following traumatic brain injury: Current evidence. NDT 2015, 1573. [Google Scholar] [CrossRef]
  380. Dhaliwal, S.K.; Meek, B.P.; Modirrousta, M.M. Non-Invasive Brain Stimulation for the Treatment of Symptoms Following Traumatic Brain Injury. Front. Psychiatry 2015, 6, 119. [Google Scholar] [CrossRef]
  381. Tam, S.-F.; Man, W.-K. Evaluating computer-assisted memory retraining programmes for people with post-head injury amnesia. Brain Inj. 2004, 18, 461–470. [Google Scholar] [CrossRef]
  382. Bergquist, T.; Gehl, C.; Mandrekar, J.; Lepore, S.; Hanna, S.; Osten, A.; Beaulieu, W. The effect of internet-based cognitive rehabilitation in persons with memory impairments after severe traumatic brain injury. Brain Inj. 2009, 23, 790–799. [Google Scholar] [CrossRef] [PubMed]
  383. Mishra, R.; Florez-Perdomo, W.A.; Shrivatava, A.; Chouksey, P.; Raj, S.; Moscote-Salazar, L.R.; Rahman, M.M.; Sutar, R.; Agrawal, A. Role of Music Therapy in Traumatic Brain Injury: A Systematic Review and Meta-analysis. World Neurosurg. 2021, 146, 197–204. [Google Scholar] [CrossRef]
  384. Baker, F.; Wigram, T.; Gold, C. The effects of a song-singing programme on the affective speaking intonation of people with traumatic brain injury. Brain Inj. 2005, 19, 519–528. [Google Scholar] [CrossRef] [PubMed]
  385. Siponkoski, S.-T.; Martínez-Molina, N.; Kuusela, L.; Laitinen, S.; Holma, M.; Ahlfors, M.; Jordan-Kilkki, P.; Ala-Kauhaluoma, K.; Melkas, S.; Pekkola, J.; et al. Music Therapy Enhances Executive Functions and Prefrontal Structural Neuroplasticity after Traumatic Brain Injury: Evidence from a Randomized Controlled Trial. J. Neurotrauma 2020, 37, 618–634. [Google Scholar] [CrossRef] [PubMed]
  386. Sinnakaruppan, I.; Downey, B.; Morrison, S. Head injury and family carers: A pilot study to investigate an innovative community-based educational programme for family carers and patients. Brain Inj. 2005, 19, 283–308. [Google Scholar] [CrossRef]
  387. Brunetti, M.; Martelli, N.; Colasante, A.; Piantelli, M.; Musiani, P.; Aiello, F.B. Spontaneous and glucocorticoid-induced apoptosis in human mature T lymphocytes. Blood 1995, 86, 4199–4205. [Google Scholar] [CrossRef]
  388. Zacco, A.; Togo, J.; Spence, K.; Ellis, A.; Lloyd, D.; Furlong, S.; Piser, T. 3-Hydroxy-3-Methylglutaryl Coenzyme A Reductase Inhibitors Protect Cortical Neurons from Excitotoxicity. J. Neurosci. 2003, 23, 11104–11111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  389. Béziaud, T.; Ru Chen, X.; El Shafey, N.; Fréchou, M.; Teng, F.; Palmier, B.; Beray-Berthat, V.; Soustrat, M.; Margaill, I.; Plotkine, M.; et al. Simvastatin in traumatic brain injury: Effect on brain edema mechanisms. Crit. Care Med. 2011, 39, 2300–2307. [Google Scholar] [CrossRef]
  390. Ding, K.; Wang, H.; Xu, J.; Lu, X.; Zhang, L.; Zhu, L. Melatonin reduced microglial activation and alleviated neuroinflammation induced neuron degeneration in experimental traumatic brain injury: Possible involvement of mTOR pathway. Neurochem. Int. 2014, 76, 23–31. [Google Scholar] [CrossRef]
  391. Feng, Y.-M.; Jia, Y.-F.; Su, L.-Y.; Wang, D.; Lv, L.; Xu, L.; Yao, Y.-G. Decreased mitochondrial DNA copy number in the hippocampus and peripheral blood during opiate addiction is mediated by autophagy and can be salvaged by melatonin. Autophagy 2013, 9, 1395–1406. [Google Scholar] [CrossRef] [PubMed]
  392. Chang, C.-F.; Huang, H.-J.; Lee, H.-C.; Hung, K.-C.; Wu, R.-T.; Lin, A.M.-Y. Melatonin attenuates kainic acid-induced neurotoxicity in mouse hippocampus via inhibition of autophagy and α-synuclein aggregation: Kainic acid and α-synuclein aggregation. J. Pineal Res. 2012, 52, 312–321. [Google Scholar] [CrossRef]
  393. Reiter, R.; Paredes, S.; Korkmaz, A.; Jou, M.-J.; Tan, D.-X. Melatonin combats molecular terrorism at the mitochondrial level. Interdiscip. Toxicol. 2008, 1, 137–149. [Google Scholar] [CrossRef] [PubMed]
  394. Alluri, H.; Wilson, R.L.; Anasooya Shaji, C.; Wiggins-Dohlvik, K.; Patel, S.; Liu, Y.; Peng, X.; Beeram, M.R.; Davis, M.L.; Huang, J.H.; et al. Melatonin Preserves Blood-Brain Barrier Integrity and Permeability via Matrix Metalloproteinase-9 Inhibition. PLoS ONE 2016, 11, e0154427. [Google Scholar] [CrossRef] [PubMed]
  395. Chhor, V.; Moretti, R.; Le Charpentier, T.; Sigaut, S.; Lebon, S.; Schwendimann, L.; Oré, M.-V.; Zuiani, C.; Milan, V.; Josserand, J.; et al. Role of microglia in a mouse model of paediatric traumatic brain injury. Brain Behav. Immun. 2017, 63, 197–209. [Google Scholar] [CrossRef]
  396. Homsi, S.; Federico, F.; Croci, N.; Palmier, B.; Plotkine, M.; Marchand-Leroux, C.; Jafarian-Tehrani, M. Minocycline effects on cerebral edema: Relations with inflammatory and oxidative stress markers following traumatic brain injury in mice. Brain Res. 2009, 1291, 122–132. [Google Scholar] [CrossRef] [PubMed]
  397. Hanlon, L.A.; Raghupathi, R.; Huh, J.W. Differential effects of minocycline on microglial activation and neurodegeneration following closed head injury in the neonate rat. Exp. Neurol. 2017, 290, 1–14. [Google Scholar] [CrossRef]
  398. Simon, D.W.; Aneja, R.K.; Alexander, H.; Bell, M.J.; Bayır, H.; Kochanek, P.M.; Clark, R.S.B. Minocycline Attenuates High Mobility Group Box 1 Translocation, Microglial Activation, and Thalamic Neurodegeneration after Traumatic Brain Injury in Post-Natal Day 17 Rats. J. Neurotrauma 2018, 35, 130–138. [Google Scholar] [CrossRef] [PubMed]
  399. Fukui, S.; Signoretti, S.; Dunbar, J.G.; Marmarou, A. The effect of Cyclosporin A on brain edema formation following experimental cortical contusion. In Brain Edema XII; Kuroiwa, T., Baethmann, A., Czernicki, Z., Hoff, J.T., Ito, U., Katayama, Y., Marmarou, A., Mendelow, B.A.D., Reulen, H.-J., Eds.; Springer: Vienna, Austria, 2003; pp. 301–303. ISBN 978-3-7091-7220-9. [Google Scholar]
  400. Aykanat, Ö.; Karakoyun, D.O.; Türkoğlu, M.E.; Dinç, C. Anti-edematous, anti-inflammatory and neuroprotective effect of etanercept in acute stage in experimental head injury. Ulus. Travma. Acil. Cerrahi. Derg. 2017, 23, 173–180. [Google Scholar] [CrossRef]
  401. Ozen, I.; Ruscher, K.; Nilsson, R.; Flygt, J.; Clausen, F.; Marklund, N. Interleukin-1 Beta Neutralization Attenuates Traumatic Brain Injury-Induced Microglia Activation and Neuronal Changes in the Globus Pallidus. Int. J. Mol. Sci. 2020, 21, 387. [Google Scholar] [CrossRef] [PubMed]
  402. Webster, K.M.; Shultz, S.R.; Ozturk, E.; Dill, L.K.; Sun, M.; Casillas-Espinosa, P.; Jones, N.C.; Crack, P.J.; O’Brien, T.J.; Semple, B.D. Targeting high-mobility group box protein 1 (HMGB1) in pediatric traumatic brain injury: Chronic neuroinflammatory, behavioral, and epileptogenic consequences. Exp. Neurol. 2019, 320, 112979. [Google Scholar] [CrossRef] [PubMed]
  403. Robertson, C.L.; Saraswati, M. Progesterone protects mitochondrial function in a rat model of pediatric traumatic brain injury. J Bioenerg Biomembr 2015, 47, 43–51. [Google Scholar] [CrossRef] [PubMed]
  404. Lengel, D.; Huh, J.W.; Barson, J.R.; Raghupathi, R. Progesterone treatment following traumatic brain injury in the 11-day-old rat attenuates cognitive deficits and neuronal hyperexcitability in adolescence. Exp. Neurol. 2020, 330, 113329. [Google Scholar] [CrossRef] [PubMed]
  405. Geddes, R.I.; Sribnick, E.A.; Sayeed, I.; Stein, D.G. Progesterone treatment shows benefit in a pediatric model of moderate to severe bilateral brain injury. PLoS ONE 2014, 9, e87252. [Google Scholar] [CrossRef]
  406. Robertson, C.L.; Fidan, E.; Stanley, R.M.; Noje, C.; Bayir, H. Progesterone for neuroprotection in pediatric traumatic brain injury. Pediatr. Crit. Care Med. 2015, 16, 236–244. [Google Scholar] [CrossRef]
  407. Harris, K.; Armstrong, S.P.; Campos-Pires, R.; Kiru, L.; Franks, N.P.; Dickinson, R. Neuroprotection against traumatic brain injury by xenon, but not argon, is mediated by inhibition at the N-methyl-D-aspartate receptor glycine site. Anesthesiology 2013, 119, 1137–1148. [Google Scholar] [CrossRef] [PubMed]
  408. Ozturk, E.; Demirbilek, S.; Kadir But, A.; Saricicek, V.; Gulec, M.; Akyol, O.; Ozcan Ersoy, M. Antioxidant properties of propofol and erythropoietin after closed head injury in rats. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2005, 29, 922–927. [Google Scholar] [CrossRef]
  409. Kaptanoglu, E.; Sen, S.; Beskonakli, E.; Surucu, H.S.; Tuncel, M.; Kilinc, K.; Taskin, Y. Antioxidant Actions and Early Ultrastructural Findings of Thiopental and Propofol in Experimental Spinal Cord Injury. J. Neurosurg. Anesthesiol. 2002, 14, 114–122. [Google Scholar] [CrossRef] [PubMed]
  410. Hoane, M.R.; Gilbert, D.R.; Holland, M.A.; Pierce, J.L. Nicotinamide reduces acute cortical neuronal death and edema in the traumatically injured brain. Neurosci. Lett. 2006, 408, 35–39. [Google Scholar] [CrossRef]
  411. Smith, A.C.; Holden, R.C.; Rasmussen, S.M.; Hoane, M.R.; Hylin, M.J. Effects of nicotinamide on spatial memory and inflammation after juvenile traumatic brain injury. Behav. Brain Res. 2019, 364, 123–132. [Google Scholar] [CrossRef]
  412. Cui, C.; Song, S.; Cui, J.; Feng, Y.; Gao, J.; Jiang, P. Vitamin D Receptor Activation Influences NADPH Oxidase (NOX2) Activity and Protects against Neurological Deficits and Apoptosis in a Rat Model of Traumatic Brain Injury. Oxid. Med. Cell Longev. 2017, 2017, 9245702. [Google Scholar] [CrossRef]
  413. Ikeda, Y.; Mochizuki, Y.; Nakamura, Y.; Dohi, K.; Matsumoto, H.; Jimbo, H.; Hayashi, M.; Matsumoto, K.; Yoshikawa, T.; Murase, H.; et al. Protective effect of a novel vitamin E derivative on experimental traumatic brain edema in rats—Preliminary study. Acta Neurochir. Suppl. 2000, 76, 343–345. [Google Scholar] [CrossRef]
  414. Li, Y.; Hawkins, B.E.; DeWitt, D.S.; Prough, D.S.; Maret, W. The relationship between transient zinc ion fluctuations and redox signaling in the pathways of secondary cellular injury: Relevance to traumatic brain injury. Brain Res. 2010, 1330, 131–141. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation of post-traumatic neuroinflammation. The rupture of the blood-brain barrier, the release of alarmines (or DAMPS: Damage Associated Molecular Patterns) by injured cells and the production of cytokines are at the origin of an endothelial, astrocytic, and microglial activation with a change in conformation of the microglial cells which take an amoeboid conformation and migrate towards the injured area. This response to trauma is both localized and generalized with secondary recruitment of the peripheral immune system. BBB: blood brain barrier; CXCL: chemokine ligand; DAMPS: damaged associated molecular patterns; IL: interleukin; HMGB1: high mobility group bow 1; iNOS: inducible nitric oxide synthase; PRR: pattern recognition receptor; PS100: protein S 100; TBI: traumatic brain injury; TGF: transforming growth factor; TNF: tumor necrosis factor.
Figure 1. Schematic representation of post-traumatic neuroinflammation. The rupture of the blood-brain barrier, the release of alarmines (or DAMPS: Damage Associated Molecular Patterns) by injured cells and the production of cytokines are at the origin of an endothelial, astrocytic, and microglial activation with a change in conformation of the microglial cells which take an amoeboid conformation and migrate towards the injured area. This response to trauma is both localized and generalized with secondary recruitment of the peripheral immune system. BBB: blood brain barrier; CXCL: chemokine ligand; DAMPS: damaged associated molecular patterns; IL: interleukin; HMGB1: high mobility group bow 1; iNOS: inducible nitric oxide synthase; PRR: pattern recognition receptor; PS100: protein S 100; TBI: traumatic brain injury; TGF: transforming growth factor; TNF: tumor necrosis factor.
Ijms 23 11193 g001
Figure 2. Chronology of post-traumatic lesions. The TBI is responsible for a primary mechanical lesion that can be transformed into a tertiary lesion via neuro-inflammatory mechanisms and thus generate a chronic post-traumatic encephalopathy. TBI: traumatic brain injury.
Figure 2. Chronology of post-traumatic lesions. The TBI is responsible for a primary mechanical lesion that can be transformed into a tertiary lesion via neuro-inflammatory mechanisms and thus generate a chronic post-traumatic encephalopathy. TBI: traumatic brain injury.
Ijms 23 11193 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jacquens, A.; Needham, E.J.; Zanier, E.R.; Degos, V.; Gressens, P.; Menon, D. Neuro-Inflammation Modulation and Post-Traumatic Brain Injury Lesions: From Bench to Bed-Side. Int. J. Mol. Sci. 2022, 23, 11193. https://doi.org/10.3390/ijms231911193

AMA Style

Jacquens A, Needham EJ, Zanier ER, Degos V, Gressens P, Menon D. Neuro-Inflammation Modulation and Post-Traumatic Brain Injury Lesions: From Bench to Bed-Side. International Journal of Molecular Sciences. 2022; 23(19):11193. https://doi.org/10.3390/ijms231911193

Chicago/Turabian Style

Jacquens, Alice, Edward J. Needham, Elisa R. Zanier, Vincent Degos, Pierre Gressens, and David Menon. 2022. "Neuro-Inflammation Modulation and Post-Traumatic Brain Injury Lesions: From Bench to Bed-Side" International Journal of Molecular Sciences 23, no. 19: 11193. https://doi.org/10.3390/ijms231911193

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