Next Article in Journal
An Immunological Glance on Pancreatic Ductal Adenocarcinoma
Previous Article in Journal
A Synergistic Effect of Reactive Oxygen and Reactive Nitrogen Species in Plasma Activated Liquid Media Triggers Astrocyte Wound Healing
Previous Article in Special Issue
Ketone Bodies Promote Amyloid-β1–40 Clearance in a Human in Vitro Blood–Brain Barrier Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanisms of Blood–Brain Barrier Dysfunction in Traumatic Brain Injury

1
The Department of Biomedical Sciences and Pathobiology, Virginia-Maryland College of Veterinary Medicine, Blacksburg, VA 24061, USA
2
The Center for Regenerative Medicine, Virginia-Maryland College of Veterinary Medicine, Blacksburg, VA 24061, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(9), 3344; https://doi.org/10.3390/ijms21093344
Submission received: 30 March 2020 / Revised: 2 May 2020 / Accepted: 4 May 2020 / Published: 8 May 2020

Abstract

:
Traumatic brain injuries (TBIs) account for the majority of injury-related deaths in the United States with roughly two million TBIs occurring annually. Due to the spectrum of severity and heterogeneity in TBIs, investigation into the secondary injury is necessary in order to formulate an effective treatment. A mechanical consequence of trauma involves dysregulation of the blood–brain barrier (BBB) which contributes to secondary injury and exposure of peripheral components to the brain parenchyma. Recent studies have shed light on the mechanisms of BBB breakdown in TBI including novel intracellular signaling and cell–cell interactions within the BBB niche. The current review provides an overview of the BBB, novel detection methods for disruption, and the cellular and molecular mechanisms implicated in regulating its stability following TBI.

1. Traumatic Brain Injury

1.1. Overview of Traumatic Brain Injury

Traumatic brain injury (TBI) is the leading cause of injury-related death and disability in the United States (CDC; TBI Surveillance System). TBI results from external forces that hinder normal brain function. These range from mild, to moderate, to severe. The metrics for classification of injury severity include a variety of factors including level of consciousness, amnesia status, and neuroimaging modalities. The commonly-used Glasgow Coma Scale (GCS) evaluates consciousness on a scale of 1–15 as a measurement of eye opening, verbal, and motor responses [1,2]. As defined by the Veterans Health Administration and the Department of Defense, mild injuries are predominantly sport and concussion-related injuries characterized by a brief alteration of consciousness. Whereas, moderate to severe diagnoses are harder to classify since they involve criteria that are more diverse. These are commonly based on loss of consciousness for greater than 30 min, at least a day of amnesia, and scores below 13 on the GCS [3]. Additionally, severe injuries are associated with serious physical and mental impairment and a GCS score of ≥ 8 [2]. While GCS is a valuable tool used to index TBI severity, it relies on observable outcomes and, therefore, a growing need exists to improve the physiological tools to complement the traditional classification system. Other, less common metrics, include time to follow command, loss of consciousness, post-traumatic amnesia [4], and the Abbreviated Injury Score [4,5,6]. While these approaches focus on clinical neurological scoring, additional measures such as blood–brain barrier (BBB) assessment may aide in diagnosis and long-term management of TBI. However, such strategies will require the use of new imaging modalities and detection of circulating brain-derived molecules. Mechanical insult to the brain causes immediate tissue deformation, shear stress, and damage to surrounding blood vessels [7,8]. Following the primary impact, BBB disruption contributes to tissue damage, subsequent edema, inflammation, and neural dysfunction [1,9]. BBB breakdown has long-lasting effects and is associated with neurodegeneration or other comorbidities such as Alzheimer’s disease and epilepsy [10,11,12,13,14], making it a crucial target for treatment.

1.2. Blood–Brain Barrier Disruption in TBI

BBB disruption is a known consequence of TBI and is associated with poorer outcomes [9]. BBB disruption is an early event, occurring within hours following injury [15,16,17,18,19] but can persist for years [19,20,21], and is considered a major risk factor for high mortality and morbidity (Table 1). Extravasation of the serum protein, fibrinogen (FBG), and immunoglobulin G (IgG), both markers for BBB disruption, were observed in the brains of human patients that died in the acute phase following TBI, as well as in those that survived at least a year [19]. Additional findings showed increased fibrinogen in the human brain 6–72 h following severe TBI [22]. Early restoration of BBB integrity may also aid in preventing the sequelae of other co-morbidities associated with TBI such as post-traumatic epilepsy (PTE) and neurodegenerative disease [9]. It has been shown that focal motor seizures occur immediately after osmotically-induced BBB disruption in human patients, contralateral to the site of disruption which was confirmed in porcine models [23]. Persistent BBB breakdown in PTE patients may contribute to its pathogenesis [24]. Moreover, fluorescein isothiocyanate (FITC)-labelled plasma component Abeta42, implicated in Alzheimer’s disease (AD), crossed the BBB and was found to be accumulated in the brains of AD patients [25,26]. Emphasis on the multifaceted role of the BBB in acute and chronic outcomes following TBI will help advance new detection methods and prognostic and treatment strategies.
Disruption of two key biological processes contributes to BBB dysfunction. The first involves an increase in paracellular transport indicated by a loss of tight junction (TJ) proteins, allowing passage of molecules that would usually be restricted [9,39,40]. This includes an influx of immune cells such as neutrophils that can exacerbate the inflammatory response. The second process occurs due to an increase in transcytosis across the endothelial cell (EC), leading to transport of larger molecules and serum proteins, such as albumin, which are normally restricted from entering the brain [9,39,40]. In addition, a major consequence of BBB disruption is cerebral edema or swelling due to excess water accumulation in the brain. Generation of cerebral edema following injury is heterogeneous, based on severity and the brain region most affected. Two types of edema can occur following brain injury, cytotoxic and vasogenic. Following BBB breakdown, vasogenic edema results following fluid accumulation in the peri-vascular space leading to changes in cerebral blood flow (CBF), and increased intracranial pressure (ICP) [41,42]. On the other hand, cytotoxic edema is caused by activation of ion channels that drive influx of water into the intracellular space of various cell types resulting in further disruption to the BBB [42,43,44,45]. Elevated ICP and changes in CBF that are not immediately controlled can result in irreversible tissue damage and cell death, which contributes to the high mortality in severe cases of TBI [42,45].

1.3. Animal Models in the Study of BBB Disruption Following TBI

Numerous animal models have demonstrated enhanced BBB permeability within hours following TBI [46,47,48] (Table 1). The BBB is disrupted as early as 1-6 h following fluid-percussion injury in rabbits, 1 h following murine controlled cortical impact (CCI), 1 h following severe TBI in swine, and immediately following closed head injury in rats [17,18,22,28,46]. A timeline of BBB permeability at the site of injury demonstrates extravasation of IgG which peaked 24-72 h, remained increased at 7 days post-injury, then was restored at 60–180 days following TBI in rats [30]. Employment of large-animal models, which more closely resemble human TBI, have also contributed to our understanding of BBB breakdown (Table 1). Immunohistochemical (IHC) analysis of serum FBG in the swine showed staining in the cortex indicating significant BBB disruption at 6-72 h following diffuse TBI that overlapped with axonal injury and paralleled similar FBG extravasation in post-mortem human tissue [33]. Dysregulation of the BBB correlates with reduced expression of TJ proteins such as occludin, claudin-5, and zona occludens [22,30,33,47,49]. It is important to note that animal models come with inherent limitations when translating to human TBI, such as differences in anatomy and brain size. While mice and rat studies far exceed other models due to efficiency, they have obvious dissimilarities to humans. Not only are they quadrupeds with vastly different brain sizes, but mice and rats are lissencephalic, whereas, humans are gyrencephalic [50]. Therefore, complementary studies using large animal models such as swine and non-human primates with gyrencephalic brains is warranted. However, these models impose their own constraints such as increased cost, and space and ethical issues. BBB disruption occurs across numerous species following TBI; however, the converging cellular and molecular mechanism(s) regulating this response require further investigation.

1.4. Clinical Assessment of BBB Disruption in TBI

Human studies of BBB function require the ability to visualize and detect the dynamic changes that may occur in response to perturbation. Non-invasive imaging techniques represent a highly attractive method to assess and monitor BBB breakdown in patients following TBI. Computed tomography (CT) scans are a non-invasive test that can distinguish acute intracranial pathology such as sites of hemorrhage, tissue swelling, and foreign bodies [51,52]. Hemorrhage, or fluid (blood) accumulation, is observed as regions of abnormal hyperdensity [52]. CT is superior to other imaging in that it is accessible and highly sensitive in the acute stages following injury. It is better able to distinguish clot from brain parenchyma as compared to the magnetic resonance imaging (MRI) [53,54]. However, once the composition of hemorrhage changes in the days following TBI, subjective analysis becomes difficult [55]. Moreover, certain contrast CT scans, such as dynamic contrast-enhanced CT (DCE-CT) use ionizing radiation and an iodine contrast that may cause adverse reactions [56]. The most widely used brain imaging modality is MRI, in particular T1-weighted dynamic contrast-enhanced MRI [17,57,58,59,60,61,62]. This is a sensitive, minimally-invasive method used to quantify the functional integrity of the BBB [57] by evaluating high-intensity acute reperfusion marker (HARM), and is found to be the most influential factor in predicting early BBB injury in stroke [59,60,61] and TBI [17,62]. Importantly, mapping water exchange across the BBB can be achieved using 3D diffusion-prepared or multiple echo time arterial-spin-labeled (ASL) perfusion MRI. This non-invasive approach bypasses the use of exogenous contrast agents and utilizes water exchange rate across the BBB, which may potentially serve as a measure of BBB function [63,64,65].
More advanced imaging modalities such as diffusion tensor imaging (DTI) and susceptibility-weighted imaging (SWI) may be valuable as prognostic tools in evaluating BBB integrity as they provide information regarding water diffusion in multiple spatial directions [66,67]. DTI provides fractional anisotropy (FA) values which indicate the overall directionality of water diffusion [68,69,70,71,72]. DTI can also differentiate between increased diffusivity resulting from vasogenic edema compared to decreased diffusivity shown by cytotoxic edema. This is due to water molecules moving more freely within the extracellular vs. intracellular space, respectively. Increased FA values are seen in human cases 72 h following mild TBI [73], and persistent depressed FA values in human patients weeks to years following mild and severe brain injuries [74,75,76,77]. These changes are associated with poorer outcomes and, therefore, may be a valuable prognostic tool in TBI [73,74,75,76,77]. There is mounting evidence that even in cases of mild concussion, there are areas of focal cerebral microbleeds, which may increase the risk of intracranial hemorrhage and neurological dysfunction [78,79]. These are detectable on magnetic resonance SWI [78,79,80,81]. SWI is hypersensitive compared to other weighted MRI modalities and microbleeds appear as circular hypointense lesions [80,82]. However, disadvantages of MRI are that it is expensive, time-consuming, requires a contrast agent, and may have contraindications such as pacemakers or other metal implants [83,84]. Evaluation of BBB permeability in TBI may also be possible using perfusion CT, as well as, single-photon emission computerized tomography-diethylenetriaminepentaacetic acid (SPECT-DCTA) as seen in stroke [85,86,87,88,89]. Furthermore, DCE-MRI is capable of picking up subtle changes in BBB disruption following stroke [90] and in Alzheimer’s disease [57,91] suggesting its use in TBI may be valuable. Therefore, combinations of MRI or CT may provide the best picture of the BBB pathophysiology in order to assess and monitor its function following TBI or in response to treatment paradigms [87].
Importantly, these imaging techniques are being complimented by evidence-based evaluation of serum biomarkers. Protein biomarkers represent a highly attractive diagnostic tool with the potential to detect substantial changes in BBB disruption that may be predictive of injury severity and outcome (Table 2). Biomarkers of BBB damage may include vascular-related structural proteins such as fragmented TJ proteins (occludin, claudin-5, and zona occludens), increased cerebral spinal fluid (CSF) [92], and increased occludin levels in the blood [93] as seen following ischemic stroke. Plasma albumin and brain-specific proteins such as glial fibrillary acid protein (GFAP), ubiquitin carboxyl-terminal hydrolase isozyme L1 (UCH-L1), and S100 calcium-binding protein B (S100B) are elevated in the serum following TBI which correlates with BBB dysfunction [36,37]. Recent studies combined the use of DTI and serum assessment of S100B and S100B autoantibodies to evaluate low-force head impact and BBB disruption [35]. Furthermore, in a mouse model of TBI, shear stress resulted in the release of extracellular microvesicles (eMVs) containing TJ proteins, specifically occludin, within 24 h following injury from the brain endothelium [94]. Similarly, microRNA (miRNA) alterations in serum and plasma, as a measure of BBB disruption, have recently gained attention in TBI [38,95]. miRNAs are small, endogenous, post-transcriptional gene regulators that play critical roles in various biological processes such as translational suppression or degradation of mRNA, and have also been suggested to contribute to many pathological conditions [95,96]. In human patients, a microarray analysis of 108 microRNAs showed that 52 miRNAs were altered 24 h following severe TBI as compared to healthy individuals [38]. These studies suggest miRNA detection may represent a new diagnostic and prognostic tool for TBI [97,98,99,100]. Additional studies are needed to determine whether these markers meet specificity, sensitivity, and reliability requirements. Improving our understanding of the mechanisms of BBB breakdown will also reveal additional cellular or biochemical targets that are highly sensitive and specific for BBB damage following TBI.

2. Blood–Brain Barrier (BBB)

Overview of BBB

BBB function is critical in maintaining brain homeostasis by suppressing entry of peripheral immune cells and providing nutrient delivery and toxic substance removal, as well as acting as a solute exchange barrier between blood and brain. The BBB provides a semi-permeable barrier separating the circulating blood from the brain environment to regulate molecules that undergo influx and efflux. The BBB is composed of specialized endothelial cells that are fenestrated by transmembrane proteins known as tight junction proteins (occludin and claudins) or junctional adhesion molecules (JAMs) that restrict paracellular permeability [156,157,158,159]. These transmembrane proteins are anchored to the cytoplasmic surface via scaffolding proteins, zonula occludens (ZO) [160]. Select compounds such as small lipid soluble molecules, water, and oxygen passively diffuse across the BBB, whereas, other larger molecules and nutrients such as amino acids, insulin, and plasma proteins require transporters or endocytosis to traverse the BBB. In order for these transporters to mediate homeostasis, assembly within the plasma membrane pore of the endothelial cell (EC) is paramount. The ECs have both a luminal (blood-facing) and abluminal (brain-facing) assembly of transporters that are required for proper BBB function [161].
Endothelial cells that make up the BBB are regulated by surrounding cells such as pericytes, astrocytes, neurons, and microglia in order to maintain ionic balance and homeostasis [162]. They influence ECs directly through secretion of soluble factors or interactions of cell-to-cell contact proteins [163]. Each cell type plays an integral role in the development and maintenance of the BBB. Formation of the murine BBB begins during embryonic day 12 (E12) with sprouting angiogenesis to form a vascular network within the neuroectoderm that is mediated, in part, by a vascular endothelial growth factor (VEGF) gradient [164,165]. At this stage, vessels contain immature barrier properties such as tight junctions, nutrient transporters, and leukocyte adhesion molecules [165]. Pericytes follow vessel sprouting around E19 and ensheath blood vessels to mediate capillary diameter, stability, and blood flow [165,166,167]. Pericytes provide the initial support to ECs, influencing transcellular transport; however, they do not affect TJ integrity [168,169]. Mice born with pericyte deficiencies exhibited increased BBB permeability through higher rates of endothelial transcytosis [170]. Further studies utilizing null mice for platelet-derived growth factor receptor (Pdgfrβ), show reduced pericyte numbers and BBB dysfunction. These findings demonstrate that pericyte coverage of the vasculature is necessary for tight junction formation, vesicle trafficking, and suppression of genes that promote BBB permeability, such as angiopoietin-2 (Angpt2), plasmalemma vesicle-associated protein (Plvap), and leukocyte adhesion molecules [166,171]. Pericytes are also integral as the intermediary between astrocytes and ECs by guiding astrocytic end-feet to the endothelium, creating the necessary polarity and subsequent maturation of the BBB [165].
Following birth, astrocytes proliferate and mature to ensheath ECs with distal projections known as end-feet. Maturation and stability of the BBB occurs through astrocyte release of soluble factors, such as transforming growth factor beta (TGFb), glial-derived neurotrophic factor (GDNF), basic fibroblast growth factor (bFGF), and angiopoietin-1 (Angpt1) which promote BBB integrity [172,173,174,175]. These astrocyte-derived factors are crucial to BBB maturation as confirmed using endothelial cell cultures. ECs co-cultured with astrocytes exhibited a BBB phenotype and improved barrier function, including expression of junctional proteins such as zona occludens (ZO-1), as compared to cultures of ECs alone [176,177,178]. Signaling pathways essential during early development of the BBB include Angpt1/Tie2, Eph/ephrin, and Hedgehog [179,180,181,182] (Table 2). Astrocytes directly regulate endothelial cell expression of tight junction proteins through release of Src-suppressed C kinase substrate (SSeCKS), Angpt1, which in turn, acts on endothelial-derived Tie2 receptor to promote barrier integrity [183,184]. Angpt1 expression has been implicated in barrier maintenance through tyrosine dephosphorylation of occludin, promoting interaction between occludin and ZO-1 [185]. The Sonic-Hedgehog (Shh) pathway is also an important developmental mediator of BBB integrity [182]. Mice devoid of Shh show embryonic lethality, which is associated with decreased expression of claudin-5 and occludin proteins [182]. Furthermore, confirmation of astrocyte-derived Shh on up-regulation of TJ proteins in human endothelial cells was confirmed through in vitro studies [182]. Astrocytes continue to maintain the BBB through control of water and ion gradients via aquaporin 4 (AQP4) and Kir4.1, respectively [165,186]. Various neurological conditions can alter homeostasis within the BBB niche, which may contribute to neural dysfunction.

3. Cellular and Molecular Mechanisms Regulating BBB Disruption Following TBI

3.1. Neuroinflammation

The secondary injury response including neuroinflammation and glial activation can contribute to disruption of the BBB (Table 2). TBI elicits activation of the endothelium and a neuroinflammatory response within minutes to hours post-injury, which is demonstrated by recruitment and up-regulation of cytokines, chemokines, neutrophils, and other pro-inflammatory mediators [110,187,188,189]. Interestingly, pericytes have recently gained attention as inflammatory mediators of the BBB following injury [190]. However, our understanding of this response is sparse due to difficulties with how quickly pericytes change phenotypes following injury and complications in identifying their structure and function. Innate immune cells also drive neuroinflammation and BBB dysregulation. Microglia, resident innate immune cells, undergo activation by albumin that has entered the brain via transcytosis leading to release of IL-1β, TNF, TGFβ, and MCP-1 [110,111,112,113] which influence BBB permeability and TJ distribution [191,192,193]. Albumin extravasation also causes astrocytes to release matrix metalloproteinases (MMPs) which degrade the basement membrane leading to BBB permeability [114], and increased vasogenic edema following TBI [103]. This is further highlighted by studies showing that suppression of CypA-MMP-9 signaling by apolipoprotein-E (ApoE) regulates BBB integrity following TBI in an isoform-dependent manner [27]. ApoE polymorphisms have gained attention in TBI due to their association with BBB breakdown in cases of deficiency or deletion [194,195], and correlation to late-onset Alzheimer’s disease (ApoE4) [194,196]. This suggests differential ApoE isoform expression may result in divergent patient-specific BBB outcomes following TBI.
Cytokine release by glial cells activates downstream pathways such as Rho/ROCK, PKC, and MAPK, which affect phosphorylation of TJ proteins and contribute to increased paracellular permeability of the BBB [115,116,117,118]. MAPK continues to gain attention for its role in temporal lobe epilepsy and intracerebral hemorrhage (ICH) [119,120]. Transcriptomic analysis of human temporal lobe epilepsy patients showed increased AQP4 expression that may be regulated by the MAPK signaling pathway. When human astrocytes from patients with temporal lobe epilepsy were treated in culture with p-38 MAPK inhibitors, expression of AQP4 was downregulated [119]. Additionally, mice treated with propagermanium, a chemokine CC ligand 2 (CCL2) inhibitor, decreased the expression of MAPK and AQP4 which correlated with a reduction in edema and behavioral deficits [120]. Moreover, studies show monocytes are regulators of BBB permeability through release of oncostatin M and VEGF, which activate glial cells and endothelial cells [121,122,123,124]. Human studies have also shown an increase in malondialdehyde, a byproduct of lipid peroxidation and oxidative stress, immediately following trauma. Oxidative stress is a known disrupter of the BBB and may be the initial catalyst for increased permeability [125,126]. Following mitochondrial damage, reactive oxygen species (ROS) released from astrocytes, microglia, and neurons further activate glial release of cytokines and chemokines [126,127,128,129]. ROS affect downstream pathways that decrease TJ expression, and increase MMPs to enhance paracellular permeability through lipid peroxidation [197,198]. Microglia are also responsible for releasing IL-1 and IL-6, which enhance intracellular adhesion molecule-1 (ICAM-1), P-selectin, and E-selectin expression, thus allowing increased leukocyte adhesion and migration across the brain endothelium to elicit additional peripheral-derived neuroinflammatory responses [118,130,131].

3.2. Vascular-Astrocyte Coupling

Communication between endothelial cells and astrocytes has been termed vascular–astrocyte coupling. Recent studies have suggested that the location and distribution of astrocytic end-feet coverage on the microvasculature may influence BBB permeability [104,175,199]. Furthermore, displacement of end-feet from the endothelium results in disruption of the BBB in cases of invading gliomas, as well as, multiple sclerosis [175]. However, recent findings demonstrate that selective loss of astrocytes using laser ablation methods under naïve conditions was not sufficient to increase vascular leakage [200].
Surprisingly, our understanding of vascular-astrocyte coupling in BBB disruption following TBI is limited. Aquaporin 4 (AQP4), a water-channel protein predominantly expressed at the junction of ECs and astrocytic end-feet, maintains ion concentrations and fluid homeostasis, while mediating edema and brain swelling in TBI [101,105,132]. In response to changes in tonicity associated with brain swelling, rat primary cortical astrocytes exhibited re-localization of AQP4 [133]. This perivascular channel is beneficial in water clearance during vasogenic edema but can exacerbate cytotoxic edema [102], therefore, further examination is needed into surface and protein-level expression in pathological conditions. Increased expression of AQP4 is present in rats at 1, 4, and 24 h following TBI [134]. Recent studies have evaluated expression of AQP4 following mild and severe human TBI and report increased expression in tissues and cerebral spinal fluid (CSF) [135,136,201]. Studies show that deletion of AQP4 attenuated BBB disruption, edema, and loss of TJ expression in ischemic conditions [137,202,203]. This supports the notion that up-regulation and redistribution of AQP4 is correlated to BBB disruption in cases of glioblastoma and cerebral ischemia [204,205]. However, the effects of AQP4 deletion on the cerebral vascular may be context dependent. For example, deletion of AQP4 results in a chronic increase in cerebral vascularization in adult mice due to maladaptive water exchange across the BBB in order to maintain CBF [137]. Additional studies using astroglial-conditional Aqp4 knockout mice further highlights the importance of astrocytic AQP4 in brain water uptake without disruption of barrier function to macromolecules in response to hypoosmotic stress. Moreover, global Aqp4 knockout mice show reduced expression of perivascular glial scaffolding proteins while BBB function remained intact under normal conditions [138,206].
It remains unclear how these developmentally-driven changes influence the outcome in models of brain injury. Nonetheless, AQP4 remains a potential therapeutic target in the acute and chronic management of BBB disruption, which could influence the onset of other comorbidities. However, additional studies are needed to improve our understanding of how changes in the overall expression and subcellular localization controls BBB function following TBI. Changes in AQP4 subcellular localization, either at the end-foot or mis-localized to other membranes, has also been shown to contribute to BBB dysfunction [106,207,208]. Under certain conditions, modulation of AQP4 expression and its redistribution may be mutually exclusive events [207,209]. For example, when exposed to hypothermic conditions, human primary cortical astrocytes in culture showed increased surface localization without accompanying increases in protein expression level [209]. On the other hand, increased expression and redistribution of AQP4 from the perivascular end-foot to the neuropil was demonstrated in mice that developed PTE following TBI [207].
A primary role of astrocytes is uptake of glutamate through transporters, EAAT1 and EAAT2 [147]. Decreased expression of these transporters is seen in human TBI and may contribute to neurotoxicity [125,148]. Excessive glutamate leads to disruption of the BBB through its activation of NMDA receptors, which enhances vascular permeability and seizures in rats, while NMDA antagonists reduced BBB permeability [149]. Overall, these studies suggest glial-derived factors play an important functional role in BBB homeostasis and TBI-induced disruption. Astrocytes also influence endothelial activity through release of soluble molecules. In particular, MMPs, VEGF, endothelin-1 (ET-1), and glutamate [114,139,140,142,200] released by astrocytes have been linked to BBB disruption. Increased release of MMP-9, an enzyme that degrades the extracellular matrix (ECM), following brain injury has been associated with increased BBB permeability through degradation of TJ proteins, occludin, and claudin-5 [114,210]. Astrocytes also influence the brain endothelium through VEGF signaling. Release of VEGF-A increased BBB disruption through down-regulation of claudin-5 and occludin in a mouse model of cerebral inflammation [141]. VEGF-A interacts with thymidine phosphorylase (TYMP), another astrocyte-derived pro-permeability factor, to promote breakdown through repression of TJ proteins in human microvascular ECs [211]. Interestingly, blocking VEGF resulted in decreased edema formation and injury following ischemia [212]. Finally, ET-1 is a potent vasoconstrictor that is implicated in poorer outcomes following brain insults, and it binds to endothelial-cell-specific ETB receptors. Enhanced expression occurs as early as 4 h following TBI [143]. Over-expression of ET-1 in astrocytes increases vasogenic edema, vasospasms, and reactive gliosis [142,144]. Intriguingly, administration of an ETB antagonist improved BBB permeability and edema following traumatic brain injury in correlation with decreased expression of MMP-9 and VEGF-A, indicating a potential upstream mechanism of BBB breakdown by these molecules [145]. These findings highlight a greater need to evaluate the mechanisms driving vascular–astrocyte crosstalk and its influence over BBB function following TBI.

3.3. Endothelial-Derived Influences on the BBB Niche

Endothelial cells interact with perivascular cells in numerous ways to regulate the BBB. Endothelial intracellular signaling is modulated through direct mechanical injury and through activation of receptors or transmembrane proteins such as ETB, Ephs, ICAM, and Mfsd2a [146,150,151,213]. Endothelial-specific ETB activation via its ligand, ET-1, causes increased transendothelial transport of monocytes [146]. Early activation of the endothelium following TBI also causes up-regulation of ICAM-1, a cell adhesion molecule on endothelial cells important for leukocyte trafficking and BBB regulation [150,151]. Similarly, major facilitator superfamily domain containing 2a (Mfsd2a), a transmembrane protein that is integral to the development and maintenance of an intact BBB [214], is decreased following brain injury in conjunction with an up-regulation in vesicle trafficking proteins such as caveolin-1, and BBB disruption [152,153,154]. Over-expression of Mfsd2a following injury attenuated these effects, therefore, Mfsd2a provides protection to the BBB by reducing vesicular transcytosis [152,153,154]. These findings indicate that the endothelial cell response is a key driver of neuroinflammation and subsequent BBB disruption. Moreover, Eph receptor signaling, the largest family of receptor tyrosine kinases, has been shown to mediate secondary injury and may also influence BBB function [155,213]. Deletion of the class B receptor, EphB3, increased BBB integrity, endothelial cell survival, and enhanced astrocyte-EC interactions in mice following CCI injury [213]. Pharmacological inhibition of certain endothelial cell-specific transporters has also proven beneficial for outcome following brain injury. Suppression of the Na(+)-K(+)-2Cl(-) cotransporter, NKCC1, and the trauma-/ischemia-induced SUR1-regulated NC(Ca-ATP) (SUR1/TRPM4) channel by bumetanide and glibenclamide, respectively, show reduction of capillary failure in rats following TBI and ischemia [107]. Additionally, blocking SUR1 using glibenclamide reduces progressive secondary hemorrhage, necrotic lesion, and neurobehavioral deficits following TBI [108], as well as improved rapid learning and long-term protection in the hippocampus in rats after TBI [109]. Further exploration of the cellular and molecular mechanisms involved in BBB dysfunction will lead to improved targets for BBB therapy.

3.4. Age-Dependent Responses

Divergent age-at-injury effects on the BBB response has also been implicated in functional outcome following TBI. Secondary injury elicited across the age spectrum may substantially change the course of BBB disruption and prognosis for patients. A whole-blood transcriptomic profile of juvenile mice at 4 days following TBI suggests suppression of neuroinflammation through attenuation of the innate immune activation and pattern recognition receptor (PRR) signaling such as Dectin-1 compared to P60-80 adult mice [215]. Moreover, aged mice show increased MMP-9 activation concomitant with decreased BBB repair [216], as well as decreased motor function, increased edema, and prolonged opening of the BBB [217]. This disparity may be attributable to changes in AQP4 or TJ protein expression between ages, as juveniles have been reported to have delayed increases in AQP4 and preserved expression of collagen-IV, laminin, claudin-5, occludin, and ZO-1 following brain injury [218,219]. Furthermore, juvenile mice exhibited decreased BBB permeability 4 days post-TBI in correlation with decreased lesion volume, improved behavioral function, and restored cerebral blood flow [220]. Pharmacological inhibition of Tie2 receptor signaling in juvenile mice subjected to injury reversed the BBB-protective phenotype, therefore, providing a potential age-related mechanism through Tie2/angiopoietin signaling [220]. Interestingly, vascular integrity was assessed in plasma protein from human TBI patients which showed decreased Ang1 expression, as well as, Ang1/Ang2 ratio as compared to controls [221]. This pathway may represent an additional key predictive marker of vascular impairment in TBI. Human studies that further evaluate age-related changes in vascular-specific biomarkers may help expand prognostic and therapeutic approaches across the age spectrum.

3.5. Therapeutic Targeting of BBB Disruption

Animal models have been instrumental in understanding and testing novel treatments aimed at preventing BBB breakdown after TBI (Table 3). In a model of murine mild TBI, administration of calpain III attenuated BBB breakdown as shown using Evans blue (EB) analysis, a method that quantifies albumin extravasation in the brain. This was confirmed using intravital microscopy to analyze vascular leakage in the brain [28,222]. Similarly, diffuse injury using weight drop in mice, showed that administration of basic fibroblast growth factor (bFGF) decreased BBB permeability 24 h following injury when assessed using FITC-dextran and was associated with increased expression and colocalization of ZO-1, claudin-5, and occludin with the vessel marker platelet/endothelial cell adhesion molecule-1 (PECAM-1 or CD31) [29,222]. Catechin treatment has been another promising therapy as studies have reported decreased water accumulation, decreased inflammatory markers, and increased BBB integrity through expression of TJ proteins in rat models of TBI [31,222]. Intravenous adrenomedullin, an endogenous peptide that plays a role in BBB integrity, is also a promising therapeutic target because it exhibits anti-inflammatory and anti-apoptotic properties following fluid percussion injury (FPI) in rats by decreasing TNF, IL-1β, and IL-6 levels and brain edema, and by increasing BBB stability [32]. Additionally, combination treatment with valproic acid and fresh-frozen plasma following CCI and hemorrhagic shock in swine resulted in increased TJ proteins, ZO-1, and claudin-5, along with increased laminin protein in the extracellular matrix [34]. Further targeting of tight-junction protein expression is seen in novel studies addressing miRNA-mediated therapy following brain injury [223,224]. In rats, infusion with miR-21 agomir following TBI improved neurological outcome, activation of the Ang1/Tie2 axis, and subsequent promotion of tight junction expression [223]. Likewise, administration of miR-501-3p following chronic cerebral hypoperfusion was shown to prevent the loss of tight junction, ZO-1 [224]. These studies suggest miRNA modulation may represent a novel treatment strategy for preventing disruption of the BBB following brain injury.
Targeting BBB-related astrocytic responses using therapeutics is also an encouraging avenue. Treatment with acetazolamide, an FDA-approved drug already administered for disorders such as glaucoma, epilepsy, and heart failure, was shown to prevent redistribution of AQP4 and concurrently diminish cytotoxic edema in a murine model of TBI [208]. AQP4 modulators represent a favorable approach to reducing BBB disruption and secondary injury following TBI. However, aquaporins have many isoforms in humans, which could lead to off-target effects. Their structure also leads to poor druggability and discrepancies in screening assays. Unfortunately, highly-specific modulators of AQP4 have yet to be developed [227,228]. Lastly, targeting the Ang/Tie2 axis may be an attractive target for BBB therapy. Although not assessed in TBI, over-expression of Ang1 following ischemia resulted in decreased infarct volume and increased expression of TJ proteins [225,226]. Suppression of Tie2 function also was shown to attenuate BBB protection in a juvenile model of TBI [220] suggesting that promoting Tie2 activation may represent a strategy for limiting BBB breakdown. These and other mechanistic targets are described in Table 3.

4. Discussion

Despite continued advancements in the characterization and understanding of secondary injury responses following TBI, there is a growing need for advanced assessments of BBB function. A novel tool for assessing immediate changes may include the development of a “BBB on a chip”. This technique would mimic the physiology and function of the BBB under different pathological states such as neuroinflammation or hypoxia. Additionally, enhanced BBB chip models have been developed that recapitulate receptor-mediated transcytosis across the barrier, which could provide insights into disruption caused by increased trafficking, as well as, an innovative analysis into drug delivery [229,230]. Another unresolved issue is rapid assessment of BBB breakdown following insult. Considerations should be made into advancing neuroimaging, as well as potential intravenous tracers to define more precise measurements and locations of BBB disruption. The current accepted tracer used in MRI detection of injury severity is gadolinium; however, incorporating other tracers similar to those used in animal models may provide greater assessments of BBB breakdown in the brain parenchyma. Furthermore, exploration into the chronic effects of BBB dysfunction may aid our understanding of comorbidities associated with TBI such as Alzheimer’s disease, post-traumatic epilepsy, and chronic traumatic encephalopathy. Pre-clinical and clinical models of TBI will be paramount in future studies to address cellular and molecular changes within the BBB niche in order to advance discovery of biomarkers for early, non-invasive detection in human patients.

Author Contributions

Conceptualization, A.C. and M.H.T.; writing-original draft preparation, A.C.; writing-review and editing, A.C. and M.H.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Institute of Neurological Disorders and Stroke of the National Institutes of Health, R01NS096281 (MHT).

Acknowledgments

We recognize the Regenerative Medicine-Interdisciplinary Graduate Education Program at Virginia Tech, The Institute of Critical Technology and Applied Sciences Programs, the Virginia-Maryland Regional College of Veterinary Medicine, and the support from Virginia Tech’s Open Access Subvention Fund (OASF).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Alzheimer’s DiseaseAD
AngptAngiopoeitin
ApoEApolipoprotein E
AQP4Aquaporin 4
ASLArterial spin labeling
BBBBlood–brain barrier
bFGFBasic fibroblast growth factor
CBFCerebral blood flow
CCIControlled cortical impact
CSFCerebral spinal fluid
CTComputed tomography
DCE-CTDynamic contrast-enhanced computed tomography
DPIDays post injury
DTIDiffuse tensor imaging
EBEvans blue
ECEndothelial cells
ECMExtracellular matrix
eMVsExtracellular microvesicles
ET-1Endothelin-1
FAFractional anisotropy
FBGFibrinogen
FPIFluid percussion injury
GCSGlasgow Coma Scale
GFAPGlial fibrillary acid protein
ICAMIntracellular adhesion molecule
ICPIntracranial pressure
IgGImmunoglobulin G
IHCImmunohistochemistry
ILInterleukin
MAPKMitogen-activated protein kinase
MCPMonocyte chemoattractant protein
Mfsd2aMajor facilitator superfamily domain containing 2a
miRNAMicroRNA
MMPMatrix metalloproteinases
MRIMagnetic resonance imaging
PdgfrPlatelet-derived growth factor receptor
PKCProtein kinase C
PlvapPlasmalemma vesicle-associated protein
PTEPost-traumatic epilepsy
ROSReactive oxygen species
ShhSonic Hedgehog
SWI Susceptibility-weighted imaging
TBITraumatic brain injury
TGFβTransforming growth factor beta
TNFTumor necrosis factor
TJTight junction
UCH-L1Ubiquitin carboxyl-terminal hydrolase isozyme L1
VEGFVascular endothelial growth factor
ZOZonula occludens

References

  1. Saatman, K.E.; Duhaime, A.C.; Bullock, R.; Maas, A.I.; Valadka, A.; Manley, G.T. Classification of traumatic brain injury for targeted therapies. J. Neurotrauma 2008, 25, 719–738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hawryluk, G.W.; Manley, G.T. Classification of traumatic brain injury: Past, present, and future. Handb Clin. Neurol. 2015, 127, 15–21. [Google Scholar] [PubMed]
  3. O’Neil, M.E.; Carlson, K.; Storzbach, D.; Brenner, L.; Freeman, M.; Quinones, A.; Motu’apuaka, M.; Ensley, M.; Kansagara, D. VA evidence-based synthesis program reports. In Complications of Mild Traumatic Brain Injury in Veterans and Military Personnel: A Systematic Review; Department of Veterans Affairs (US): Washington, DC, USA, 2013. [Google Scholar]
  4. Corrigan, J.D.; Kreider, S.; Cuthbert, J.; Whyte, J.; Dams-O’Connor, K.; Faul, M.; Harrison-Felix, C.; Whiteneck, G.; Pretz, C.R. Components of traumatic brain injury severity indices. J. Neurotrauma 2014, 31, 1000–1007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Rating the severity of tissue damage. I. The abbreviated scale. Jama 1971, 215, 277–280. [Google Scholar] [CrossRef]
  6. Association for the Advancement of Automotive Medicine. The abbreviated injury scale. 1990. Available online: https://www.aaam.org/abbreviated-injury-scale-ais/.
  7. Werner, C.; Engelhard, K. Pathophysiology of traumatic brain injury. Br. J. Anaesth. 2007, 99, 4–9. [Google Scholar] [CrossRef] [Green Version]
  8. Kaur, P.; Sharma, S. Recent advances in pathophysiology of traumatic brain injury. Curr. Neuropharmacol. 2018, 16, 1224–1238. [Google Scholar] [CrossRef]
  9. Price, L.; Wilson, C.; Grant, G. Frontiers in neuroscience blood-brain barrier pathophysiology following traumatic brain injury. In Translational Research in Traumatic Brain Injury; Laskowitz, D., Grant, G., Eds.; CRC Press/Taylor and Francis Group©: Boca Raton, FL, USA, 2016. [Google Scholar]
  10. Annegers, J.F.; Hauser, W.A.; Coan, S.P.; Rocca, W.A. A population-based study of seizures after traumatic brain injuries. N. Engl. J. Med. 1998, 338, 20–24. [Google Scholar] [CrossRef]
  11. Ruttan, L.; Martin, K.; Liu, A.; Colella, B.; Green, R.E. Long-term cognitive outcome in moderate to severe traumatic brain injury: A meta-analysis examining timed and untimed tests at 1 and 4.5 or more years after injury. Arch. Phys. Med. Rehabil. 2008, 89, S69–S76. [Google Scholar] [CrossRef]
  12. Herman, S.T. Epilepsy after brain insult: Targeting epileptogenesis. Neurology 2002, 59, S21–S26. [Google Scholar] [CrossRef] [PubMed]
  13. Zlokovic, B.V. The blood-brain barrier in health and chronic neurodegenerative disorders. Neuron 2008, 57, 178–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Shlosberg, D.; Benifla, M.; Kaufer, D.; Friedman, A. Blood-brain barrier breakdown as a therapeutic target in traumatic brain injury. Nat. Rev. Neurol 2010, 6, 393–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Rinder, L.; Olsson, Y. Studies on vascular permeability changes in experimental brain concussion. I. Distribution of circulating fluorescent indicators in brain and cervical cord after sudden mechanical loading of the brain. Acta Neuropathol. 1968, 11, 183–200. [Google Scholar] [PubMed]
  16. Hekmatpanah, J.; Hekmatpanah, C.R. Microvascular alterations following cerebral contusion in rats. Light, scanning, and electron microscope study. J. Neurosurg. 1985, 62, 888–897. [Google Scholar] [CrossRef]
  17. Barzo, P.; Marmarou, A.; Fatouros, P.; Corwin, F.; Dunbar, J. Magnetic resonance imaging-monitored acute blood-brain barrier changes in experimental traumatic brain injury. J. Neurosurg. 1996, 85, 1113–1121. [Google Scholar] [CrossRef]
  18. Habgood, M.D.; Bye, N.; Dziegielewska, K.M.; Ek, C.J.; Lane, M.A.; Potter, A.; Morganti-Kossmann, C.; Saunders, N.R. Changes in blood-brain barrier permeability to large and small molecules following traumatic brain injury in mice. Eur J. Neurosci. 2007, 25, 231–238. [Google Scholar] [CrossRef] [PubMed]
  19. Hay, J.R.; Johnson, V.E.; Young, A.M.; Smith, D.H.; Stewart, W. Blood-brain barrier disruption is an early event that may persist for many years after traumatic brain injury in humans. J. Neuropathol. Exp. Neurol. 2015, 74, 1147–1157. [Google Scholar]
  20. Baskaya, M.K.; Rao, A.M.; Dogan, A.; Donaldson, D.; Dempsey, R.J. The biphasic opening of the blood-brain barrier in the cortex and hippocampus after traumatic brain injury in rats. Neurosci. Lett. 1997, 226, 33–36. [Google Scholar] [CrossRef]
  21. Glushakova, O.Y.; Johnson, D.; Hayes, R.L. Delayed increases in microvascular pathology after experimental traumatic brain injury are associated with prolonged inflammation, blood-brain barrier disruption, and progressive white matter damage. J. Neurotrauma 2014, 31, 1180–1193. [Google Scholar] [CrossRef]
  22. Johnson, V.E.; Weber, M.T.; Xiao, R.; Cullen, D.K.; Meaney, D.F.; Stewart, W.; Smith, D.H. Mechanical disruption of the blood-brain barrier following experimental concussion. Acta Neuropathol. 2018, 135, 711–726. [Google Scholar] [CrossRef] [Green Version]
  23. Marchi, N.; Angelov, L.; Masaryk, T.; Fazio, V.; Granata, T.; Hernandez, N.; Hallene, K.; Diglaw, T.; Franic, L.; Najm, I.; et al. Seizure-promoting effect of blood-brain barrier disruption. Epilepsia 2007, 48, 732–742. [Google Scholar] [CrossRef] [Green Version]
  24. Tomkins, O.F.A.; Benifla, M.; Cohen, A.; Friedman, A.; Shelef, I. Blood -brain barrier breakdown following traumatic brain injury: A possible role in posttraumatic epilepsy. Cardiovasc. Psychiatry Neurol. 2011, 2011, 765923. [Google Scholar]
  25. Clifford, P.M.; Zarrabi, S.; Siu, G.; Kinsler, K.J.; Kosciuk, M.C.; Venkataraman, V.; D’Andrea, M.R.; Dinsmore, S.; Nagele, R.G. Abeta peptides can enter the brain through a defective blood-brain barrier and bind selectively to neurons. Brain Res. 2007, 1142, 223–236. [Google Scholar] [CrossRef] [PubMed]
  26. Toropova, A.P.; Toropov, A.A.; Begum, S.; Achary, P.G.R. Blood brain barrier and Alzheimer’s disease: Similarity and dissimilarity of molecular alerts. Curr. Neuropharmacol. 2018, 16, 769–785. [Google Scholar] [CrossRef] [PubMed]
  27. Main, B.S.; Villapol, S.; Sloley, S.S.; Barton, D.J.; Parsadanian, M.; Agbaegbu, C.; Stefos, K.; McCann, M.S.; Washington, P.M.; Rodriguez, O.C.; et al. Apolipoprotein E4 impairs spontaneous blood brain barrier repair following traumatic brain injury. Mol. Neurodegener. 2018, 13, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Alluri, H.; Grimsley, M.; Anasooya Shaji, C.; Varghese, K.P.; Zhang, S.L.; Peddaboina, C.; Robinson, B.; Beeram, M.R.; Huang, J.H.; Tharakan, B. Attenuation of blood-brain barrier breakdown and hyperpermeability by Calpain inhibition. J. Biol. Chem. 2016, 291, 26958–26969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Wang, Z.G.; Cheng, Y.; Yu, X.C.; Ye, L.B.; Xia, Q.H.; Johnson, N.R.; Wei, X.; Chen, D.Q.; Cao, G.; Fu, X.B.; et al. bFGF protects against blood-brain barrier damage through junction protein regulation via PI3K-Akt-Rac1 pathway following traumatic brain injury. Mol. Neurobiol. 2016, 53, 7298–7311. [Google Scholar] [CrossRef] [Green Version]
  30. Pop, V.; Badaut, J. A neurovascular perspective for long-term changes after brain trauma. Transl. Stroke Res. 2011, 2, 533–545. [Google Scholar] [CrossRef] [Green Version]
  31. Jiang, Z.; Zhang, J.; Cai, Y.; Huang, J.; You, L. Catechin attenuates traumatic brain injury-induced blood-brain barrier damage and improves longer-term neurological outcomes in rats. Exp. Physiol. 2017, 102, 1269–1277. [Google Scholar] [CrossRef]
  32. Gao, W.; Ju, Y.N.; Chen, J.F.; Zhou, Q.; Song, C.Y.; Wang, Y.Z.; Cao, H.L.; Yang, W.C. Adrenomedullin reduces secondary injury and improves outcome in rats with fluid percussion brain injury. World Neurosurg 2018, 119, e765–e773. [Google Scholar] [CrossRef]
  33. Williams, A.M.; Bhatti, U.F.; Brown, J.F.; Biesterveld, B.E.; Kathawate, R.G.; Graham, N.J.; Chtraklin, K.; Siddiqui, A.Z.; Dekker, S.E.; Andjelkovic, A.; et al. Early single-dose treatment with exosomes provides neuroprotection and improves blood-brain barrier integrity in swine model of traumatic brain injury and hemorrhagic shock. J. Trauma Acute Care Surg. 2020, 88, 207–218. [Google Scholar] [CrossRef]
  34. Nikolian, V.C.; Dekker, S.E.; Bambakidis, T.; Higgins, G.A.; Dennahy, I.S.; Georgoff, P.E.; Williams, A.M.; Andjelkovic, A.V.; Alam, H.B. Improvement of blood-brain barrier integrity in traumatic brain injury and hemorrhagic shock following treatment with valproic acid and fresh frozen plasma. Crit. Care Med. 2018, 46, e59–e66. [Google Scholar] [CrossRef] [PubMed]
  35. Marchi, N.; Bazarian, J.J.; Puvenna, V.; Janigro, M.; Ghosh, C.; Zhong, J.; Zhu, T.; Blackman, E.; Stewart, D.; Ellis, J.; et al. Consequences of repeated blood-brain barrier disruption in football players. PLoS ONE 2013, 8, e56805. [Google Scholar] [CrossRef] [PubMed]
  36. Blyth, B.J.; Farhavar, A.; Gee, C.; Hawthorn, B.; He, H.; Nayak, A.; Stocklein, V.; Bazarian, J.J. Validation of serum markers for blood-brain barrier disruption in traumatic brain injury. J. Neurotrauma 2009, 26, 1497–1507. [Google Scholar] [CrossRef] [PubMed]
  37. Blyth, B.J.; Farahvar, A.; He, H.; Nayak, A.; Yang, C.; Shaw, G.; Bazarian, J.J. Elevated serum ubiquitin carboxy-terminal hydrolase L1 is associated with abnormal blood-brain barrier function after traumatic brain injury. J. Neurotrauma 2011, 28, 2453–2462. [Google Scholar] [CrossRef] [PubMed]
  38. Redell, J.B.; Moore, A.N.; Ward, N.H.; Hergenroeder, G.W.; Dash, P.K. Human traumatic brain injury alters plasma microRNA levels. J. Neurotrauma 2010, 27, 2147–2156. [Google Scholar] [CrossRef]
  39. Hawkins, B.T.; Davis, T.P. The blood-brain barrier/neurovascular unit in health and disease. Pharmacol. Rev. 2005, 57, 173–185. [Google Scholar] [CrossRef]
  40. Keaney, J.; Campbell, M. The dynamic blood-brain barrier. Febs J. 2015, 282, 4067–4079. [Google Scholar] [CrossRef]
  41. Badaut, J.; Ashwal, S.; Obenaus, A. Aquaporins in cerebrovascular disease: A target for treatment of brain edema? Cerebrovasc. Dis. 2011, 31, 521–531. [Google Scholar] [CrossRef] [Green Version]
  42. Jha, R.M.; Kochanek, P.M.; Simard, J.M. Pathophysiology and treatment of cerebral edema in traumatic brain injury. Neuropharmacology 2019, 145, 230–246. [Google Scholar] [CrossRef]
  43. Stokum, J.A.; Gerzanich, V.; Simard, J.M. Molecular pathophysiology of cerebral edema. J. Cereb. Blood Flow Metab. 2016, 36, 513–538. [Google Scholar] [CrossRef] [Green Version]
  44. Hudak, A.M.; Peng, L.; Marquez de la Plata, C.; Thottakara, J.; Moore, C.; Harper, C.; McColl, R.; Babcock, E.; Diaz-Arrastia, R. Cytotoxic and vasogenic cerebral oedema in traumatic brain injury: Assessment with FLAIR and DWI imaging. Brain Inj. 2014, 28, 1602–1609. [Google Scholar] [CrossRef] [PubMed]
  45. Winkler, E.A.; Minter, D.; Yue, J.K.; Manley, G.T. Cerebral edema in traumatic brain injury: Pathophysiology and prospective therapeutic targets. Neurosurg. Clin. N. Am. 2016, 27, 473–488. [Google Scholar] [CrossRef] [PubMed]
  46. Hartl, R.; Medary, M.; Ruge, M.; Arfors, K.E.; Ghajar, J. Blood-brain barrier breakdown occurs early after traumatic brain injury and is not related to white blood cell adherence. Acta Neurochir. Suppl. 1997, 70, 240–242. [Google Scholar] [PubMed]
  47. Nag, S.; Venugopalan, R.; Stewart, D.J. Increased caveolin-1 expression precedes decreased expression of occludin and claudin-5 during blood-brain barrier breakdown. Acta Neuropathol. 2007, 114, 459–469. [Google Scholar] [CrossRef]
  48. Li, W.; Watts, L.; Long, J.; Zhou, W.; Shen, Q.; Jiang, Z.; Li, Y.; Duong, T.Q. Spatiotemporal changes in blood-brain barrier permeability, cerebral blood flow, T2 and diffusion following mild traumatic brain injury. Brain Res. 2016, 1646, 53–61. [Google Scholar] [CrossRef]
  49. Nag, S.; Manias, J.L.; Stewart, D.J. Expression of endothelial phosphorylated caveolin-1 is increased in brain injury. Neuropathol. Appl. Neurobiol. 2009, 35, 417–426. [Google Scholar] [CrossRef]
  50. Sun, T.; Hevner, R.F. Growth and folding of the mammalian cerebral cortex: From molecules to malformations. Nature Rev. 2014, 15, 217–233. [Google Scholar] [CrossRef] [Green Version]
  51. Manley, G.T.; Diaz-Arrastia, R.; Brophy, M.; Engel, D.; Goodman, C.; Gwinn, K.; Veenstra, T.D.; Ling, G.; Ottens, A.K.; Tortella, F.; et al. Common data elements for traumatic brain injury: Recommendations from the biospecimens and biomarkers working group. Arch. Phys. Med. Rehabil. 2010, 91, 1667–1672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Heit, J.J.; Iv, M.; Wintermark, M. Imaging of intracranial hemorrhage. J. Stroke 2017, 19, 11–27. [Google Scholar] [CrossRef] [Green Version]
  53. Bradley, W.G., Jr. MR appearance of hemorrhage in the brain. Radiology 1993, 189, 15–26. [Google Scholar] [CrossRef]
  54. Lee, B.; Newberg, A. Neuroimaging in traumatic brain imaging. NeuroRx 2005, 2, 372–383. [Google Scholar] [CrossRef] [PubMed]
  55. Zimmerman, R.D. Stroke wars: Episode IV CT strikes back. AJNR Am. J. Neuroradiol. 2004, 25, 1304–1309. [Google Scholar] [PubMed]
  56. Veksler, R.; Shelef, I.; Friedman, A. Blood-brain barrier imaging in human neuropathologies. Arch. Med. Res. 2014, 45, 646–652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Heye, A.K.; Culling, R.D.; Valdes Hernandez Mdel, C.; Thrippleton, M.J.; Wardlaw, J.M. Assessment of blood-brain barrier disruption using dynamic contrast-enhanced MRI. A systematic review. Neuroimage Clin. 2014, 6, 262–274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Chassidim, Y.; Veksler, R.; Lublinsky, S.; Pell, G.S.; Friedman, A.; Shelef, I. Quantitative imaging assessment of blood-brain barrier permeability in humans. Fluids Barriers CNS 2013, 10, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Villringer, K.; Sanz Cuesta, B.E.; Ostwaldt, A.C.; Grittner, U.; Brunecker, P.; Khalil, A.A.; Schindler, K.; Eisenblatter, O.; Audebert, H.; Fiebach, J.B. DCE-MRI blood-brain barrier assessment in acute ischemic stroke. Neurology 2017, 88, 433–440. [Google Scholar] [CrossRef]
  60. Merali, Z.; Huang, K.; Mikulis, D.; Silver, F.; Kassner, A. Evolution of blood-brain-barrier permeability after acute ischemic stroke. PLoS ONE 2017, 12, e0171558. [Google Scholar] [CrossRef] [Green Version]
  61. Yan, G.; Xuan, Y.; Dai, Z.; Zhang, G.; Xu, H.; Mikulis, D.; Wu, R. Evolution of blood-brain barrier damage associated with changes in brain metabolites following acute ischemia. Neuroreport 2015, 26, 945–951. [Google Scholar] [CrossRef]
  62. Li, W.; Long, J.A.; Watts, L.T.; Jiang, Z.; Shen, Q.; Li, Y.; Duong, T.Q. A quantitative MRI method for imaging blood-brain barrier leakage in experimental traumatic brain injury. PLoS ONE 2014, 9, e114173. [Google Scholar] [CrossRef] [Green Version]
  63. Lin, Z.; Li, Y.; Su, P.; Mao, D.; Wei, Z.; Pillai, J.J.; Moghekar, A.; van Osch, M.; Ge, Y.; Lu, H. Non-contrast MR imaging of blood-brain barrier permeability to water. Magn. Reson. Med. 2018, 80, 1507–1520. [Google Scholar] [CrossRef]
  64. Shao, X.; Ma, S.J.; Casey, M.; D’Orazio, L.; Ringman, J.M.; Wang, D.J.J. Mapping water exchange across the blood-brain barrier using 3D diffusion-prepared arterial spin labeled perfusion MRI. Magn. Reson. Med. 2019, 81, 3065–3079. [Google Scholar] [CrossRef] [PubMed]
  65. Ohene, Y.; Harrison, I.F.; Nahavandi, P.; Ismail, O.; Bird, E.V.; Ottersen, O.P.; Nagelhus, E.A.; Thomas, D.L.; Lythgoe, M.F.; Wells, J.A. Non-invasive MRI of brain clearance pathways using multiple echo time arterial spin labelling: An aquaporin-4 study. Neuroimage 2019, 188, 515–523. [Google Scholar] [CrossRef] [PubMed]
  66. Mukherjee, P.; Berman, J.I.; Chung, S.W.; Hess, C.P.; Henry, R.G. Diffusion tensor MR imaging and fiber tractography: Theoretic underpinnings. AJNR Am. J. Neuroradiol. 2008, 29, 632–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Mukherjee, P.; Chung, S.W.; Berman, J.I.; Hess, C.P.; Henry, R.G. Diffusion tensor MR imaging and fiber tractography: Technical considerations. AJNR Am. J. Neuroradiol. 2008, 29, 843–852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Chenevert, T.L.; Brunberg, J.A.; Pipe, J.G. Anisotropic diffusion in human white matter: Demonstration with MR techniques in vivo. Radiology 1990, 177, 401–405. [Google Scholar] [CrossRef] [PubMed]
  69. Basser, P.J.; Mattiello, J.; LeBihan, D. Estimation of the effective self-diffusion tensor from the NMR spin echo. J. Magn. Reson. B. 1994, 103, 247–254. [Google Scholar] [CrossRef]
  70. Basser, P.J.; Mattiello, J.; LeBihan, D. MR diffusion tensor spectroscopy and imaging. Biophys. J. 1994, 66, 259–267. [Google Scholar] [CrossRef] [Green Version]
  71. Basser, P.J.M.J.; Turner, R.; Le Bihan, D. Diffusion tensor echo-planar imaging of human brain. In Proceedings of the SMRM, New York, NY, USA, 14–20 August 1993; p. 584. [Google Scholar]
  72. Le Bihan, D.; Mangin, J.F.; Poupon, C.; Clark, C.A.; Pappata, S.; Molko, N.; Chabriat, H. Diffusion tensor imaging: Concepts and applications. J. Magn. Reson. Imaging 2001, 13, 534–546. [Google Scholar] [CrossRef]
  73. Bazarian, J.J.; Zhong, J.; Blyth, B.; Zhu, T.; Kavcic, V.; Peterson, D. Diffusion tensor imaging detects clinically important axonal damage after mild traumatic brain injury: A pilot study. J. Neurotrauma 2007, 24, 1447–1459. [Google Scholar] [CrossRef]
  74. Wilde, E.A.; McCauley, S.R.; Hunter, J.V.; Bigler, E.D.; Chu, Z.; Wang, Z.J.; Hanten, G.R.; Troyanskaya, M.; Yallampalli, R.; Li, X.; et al. Diffusion tensor imaging of acute mild traumatic brain injury in adolescents. Neurology 2008, 70, 948–955. [Google Scholar] [CrossRef]
  75. Wilde, E.A.; Ayoub, K.W.; Bigler, E.D.; Chu, Z.D.; Hunter, J.V.; Wu, T.C.; McCauley, S.R.; Levin, H.S. Diffusion tensor imaging in moderate-to-severe pediatric traumatic brain injury: Changes within an 18 month post-injury interval. Brain Imaging Behav. 2012, 6, 404–416. [Google Scholar] [CrossRef] [PubMed]
  76. Sidaros, A.; Engberg, A.W.; Sidaros, K.; Liptrot, M.G.; Herning, M.; Petersen, P.; Paulson, O.B.; Jernigan, T.L.; Rostrup, E. Diffusion tensor imaging during recovery from severe traumatic brain injury and relation to clinical outcome: A longitudinal study. Brain 2008, 131, 559–572. [Google Scholar] [CrossRef] [PubMed]
  77. Hart, J., Jr.; Kraut, M.A.; Womack, K.B.; Strain, J.; Didehbani, N.; Bartz, E.; Conover, H.; Mansinghani, S.; Lu, H.; Cullum, C.M. Neuroimaging of cognitive dysfunction and depression in aging retired National Football League players: A cross-sectional study. JAMA Neurol. 2013, 70, 326–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Liu, J.; Kou, Z.; Tian, Y. Diffuse axonal injury after traumatic cerebral microbleeds: An evaluation of imaging techniques. Neural Regen. Res. 2014, 9, 1222–1230. [Google Scholar] [PubMed]
  79. Lawrence, T.P.; Pretorius, P.M.; Ezra, M.; Cadoux-Hudson, T.; Voets, N.L. Early detection of cerebral microbleeds following traumatic brain injury using MRI in the hyper-acute phase. Neurosci. Lett. 2017, 655, 143–150. [Google Scholar] [CrossRef] [PubMed]
  80. Van den Heuvel, T.L.; van der Eerden, A.W.; Manniesing, R.; Ghafoorian, M.; Tan, T.; Andriessen, T.M.; Vande Vyvere, T.; van den Hauwe, L.; Ter Haar Romeny, B.M.; Goraj, B.M.; et al. Automated detection of cerebral microbleeds in patients with traumatic brain injury. Neuroimage Clin. 2016, 12, 241–251. [Google Scholar] [CrossRef] [Green Version]
  81. Lotan, E.; Morley, C.; Newman, J.; Qian, M.; Abu-Amara, D.; Marmar, C.; Lui, Y.W. Prevalence of cerebral microhemorrhage following chronic blast-related mild traumatic brain injury in military service members using susceptibility-weighted MRI. AJNR Am. J. Neuroradiol. 2018, 39, 1222–1225. [Google Scholar] [CrossRef]
  82. Cheng, A.L.; Batool, S.; McCreary, C.R.; Lauzon, M.L.; Frayne, R.; Goyal, M.; Smith, E.E. Susceptibility-weighted imaging is more reliable than T2*-weighted gradient-recalled echo MRI for detecting microbleeds. Stroke 2013, 44, 2782–2786. [Google Scholar] [CrossRef]
  83. Shetty, V.S.; Reis, M.N.; Aulino, J.M.; Berger, K.L.; Broder, J.; Choudhri, A.F.; Kendi, A.T.; Kessler, M.M.; Kirsch, C.F.; Luttrull, M.D.; et al. ACR appropriateness criteria head trauma. J. Am. Coll. Radiol. 2016, 13, 668–679. [Google Scholar] [CrossRef]
  84. Wintermark, M.; Sanelli, P.C.; Anzai, Y.; Tsiouris, A.J.; Whitlow, C.T. Imaging evidence and recommendations for traumatic brain injury: Conventional neuroimaging techniques. J. Am. Coll. Radiol. 2015, 12, e1–e14. [Google Scholar] [CrossRef]
  85. Ozkul-Wermester, O.; Guegan-Massardier, E.; Triquenot, A.; Borden, A.; Perot, G.; Gerardin, E. Increased blood-brain barrier permeability on perfusion computed tomography predicts hemorrhagic transformation in acute ischemic stroke. Eur. Neurol. 2014, 72, 45–53. [Google Scholar] [CrossRef] [PubMed]
  86. Hom, J.; Dankbaar, J.W.; Soares, B.P.; Schneider, T.; Cheng, S.C.; Bredno, J.; Lau, B.C.; Smith, W.; Dillon, W.P.; Wintermark, M. Blood-brain barrier permeability assessed by perfusion CT predicts symptomatic hemorrhagic transformation and malignant edema in acute ischemic stroke. AJNR Am. J. Neuroradiol. 2011, 32, 41–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Edgell, R.C.; Vora, N.A. Neuroimaging markers of hemorrhagic risk with stroke reperfusion therapy. Neurology 2012, 79, S100–S104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Gilad, R.; Lampl, Y.; Eilam, A.; Boaz, M.; Loyberboim, M. SPECT-DTPA as a tool for evaluating the blood-brain barrier in post-stroke seizures. J. Neurol. 2012, 259, 2041–2044. [Google Scholar] [CrossRef] [PubMed]
  89. Lorberboym, M.; Lampl, Y.; Sadeh, M. Correlation of 99mTc-DTPA SPECT of the blood-brain barrier with neurologic outcome after acute stroke. J. Nucl. Med. 2003, 44, 1898–1904. [Google Scholar] [PubMed]
  90. Israeli, D.; Tanne, D.; Daniels, D.; Last, D.; Shneor, R.; Guez, D.; Landau, E.; Roth, Y.; Ocherashvilli, A.; Bakon, M.; et al. The application of MRI for depiction of subtle blood brain barrier disruption in stroke. Int J. Biol. Sci. 2010, 7, 1–8. [Google Scholar] [CrossRef] [Green Version]
  91. Starr, J.M.; Farrall, A.J.; Armitage, P.; McGurn, B.; Wardlaw, J. Blood-brain barrier permeability in Alzheimer’s disease: A case-control MRI study. Psychiatry Res. 2009, 171, 232–241. [Google Scholar] [CrossRef]
  92. Jiao, X.; He, P.; Li, Y.; Fan, Z.; Si, M.; Xie, Q.; Chang, X.; Huang, D. The role of circulating tight junction proteins in evaluating blood brain barrier disruption following intracranial hemorrhage. Dis. Markers 2015, 2015, 860120. [Google Scholar] [CrossRef] [Green Version]
  93. Pan, R.; Yu, K.; Weatherwax, T.; Zheng, H.; Liu, W.; Liu, K.J. Blood occludin level as a potential biomarker for early blood brain barrier damage following ischemic stroke. Sci Rep. 2017, 7, 40331. [Google Scholar] [CrossRef]
  94. Andrews, A.M.; Lutton, E.M.; Merkel, S.F.; Razmpour, R.; Ramirez, S.H. Mechanical injury induces brain endothelial-derived microvesicle release: Implications for cerebral vascular injury during traumatic brain injury. Front. Cell Neurosci. 2016, 10, 43. [Google Scholar] [CrossRef] [Green Version]
  95. Di Pietro, V.; Yakoub, K.M.; Scarpa, U.; Di Pietro, C.; Belli, A. MicroRNA signature of traumatic brain injury: From the biomarker discovery to the point-of-care. Front. Neurol. 2018, 9, 429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Felekkis, K.; Touvana, E.; Stefanou, C.; Deltas, C. microRNAs: A newly described class of encoded molecules that play a role in health and disease. Hippokratia 2010, 14, 236–240. [Google Scholar] [PubMed]
  97. Di Pietro, V.; Ragusa, M.; Davies, D.; Su, Z.; Hazeldine, J.; Lazzarino, G.; Hill, L.J.; Crombie, N.; Foster, M.; Purrello, M.; et al. MicroRNAs as novel biomarkers for the diagnosis and prognosis of mild and severe traumatic brain injury. J. Neurotrauma 2017, 34, 1948–1956. [Google Scholar] [CrossRef] [PubMed]
  98. Bhomia, M.; Balakathiresan, N.S.; Wang, K.K.; Papa, L.; Maheshwari, R.K. A panel of serum miRNA biomarkers for the diagnosis of severe to mild traumatic brain injury in humans. Sci. Rep. 2016, 6, 28148. [Google Scholar] [CrossRef] [PubMed]
  99. Yang, T.; Song, J.; Bu, X.; Wang, C.; Wu, J.; Cai, J.; Wan, S.; Fan, C.; Zhang, C.; Wang, J. Elevated serum miR-93, miR-191, and miR-499 are noninvasive biomarkers for the presence and progression of traumatic brain injury. J. Neurochem. 2016, 137, 122–129. [Google Scholar] [CrossRef] [PubMed]
  100. Mitra, B.; Rau, T.F.; Surendran, N.; Brennan, J.H.; Thaveenthiran, P.; Sorich, E.; Fitzgerald, M.C.; Rosenfeld, J.V.; Patel, S.A. Plasma micro-RNA biomarkers for diagnosis and prognosis after traumatic brain injury: A pilot study. J. Clin. Neurosci. 2017, 38, 37–42. [Google Scholar] [CrossRef] [PubMed]
  101. Kitchen, P.; Day, R.E.; Salman, M.M.; Conner, M.T.; Bill, R.M.; Conner, A.C. Beyond water homeostasis: Diverse functional roles of mammalian aquaporins. Biochim. Biophys. Acta 2015, 1850, 2410–2421. [Google Scholar] [CrossRef] [Green Version]
  102. Papadopoulos, M.C.; Manley, G.T.; Krishna, S.; Verkman, A.S. Aquaporin-4 facilitates reabsorption of excess fluid in vasogenic brain edema. Faseb J. 2004, 18, 1291–1293. [Google Scholar] [CrossRef]
  103. Guilfoyle, M.R.; Carpenter, K.L.; Helmy, A.; Pickard, J.D.; Menon, D.K.; Hutchinson, P.J. Matrix metalloproteinase expression in contusional traumatic brain injury: A paired microdialysis study. J. Neurotrauma 2015, 32, 1553–1559. [Google Scholar] [CrossRef] [Green Version]
  104. Eilam, R.; Segal, M.; Malach, R.; Sela, M.; Arnon, R.; Aharoni, R. Astrocyte disruption of neurovascular communication is linked to cortical damage in an animal model of multiple sclerosis. Glia 2018, 66, 1098–1117. [Google Scholar] [CrossRef]
  105. Tang, G.; Yang, G.Y. Aquaporin-4: A potential therapeutic target for cerebral edema. Int. J. Mol. Sci. 2016, 17, 1413. [Google Scholar] [CrossRef] [PubMed]
  106. Amiry-Moghaddam, M.; Williamson, A.; Palomba, M.; Eid, T.; de Lanerolle, N.C.; Nagelhus, E.A.; Adams, M.E.; Froehner, S.C.; Agre, P.; Ottersen, O.P. Delayed K+ clearance associated with aquaporin-4 mislocalization: Phenotypic defects in brains of alpha-syntrophin-null mice. Proc. Natl. Acad. Sci. USA 2003, 100, 13615–13620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Simard, J.M.; Kahle, K.T.; Gerzanich, V. Molecular mechanisms of microvascular failure in central nervous system injury-synergistic roles of NKCC1 and SUR1/TRPM4. J. Neurosurg. 2010, 113, 622–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Simard, J.M.; Kilbourne, M.; Tsymbalyuk, O.; Tosun, C.; Caridi, J.; Ivanova, S.; Keledjian, K.; Bochicchio, G.; Gerzanich, V. Key role of sulfonylurea receptor 1 in progressive secondary hemorrhage after brain contusion. J. Neurotrauma 2009, 26, 2257–2267. [Google Scholar] [CrossRef] [PubMed]
  109. Patel, A.D.; Gerzanich, V.; Geng, Z.; Simard, J.M. Glibenclamide reduces hippocampal injury and preserves rapid spatial learning in a model of traumatic brain injury. J. Neuropathol. Exp. Neurol. 2010, 69, 1177–1190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Corrigan, F.; Mander, K.A.; Leonard, A.V.; Vink, R. Neurogenic inflammation after traumatic brain injury and its potentiation of classical inflammation. J. Neuroinflammation 2016, 13, 264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Calvo, C.F.; Amigou, E.; Tence, M.; Yoshimura, T.; Glowinski, J. Albumin stimulates monocyte chemotactic protein-1 expression in rat embryonic mixed brain cells. J. Neurosci. Res. 2005, 80, 707–714. [Google Scholar] [CrossRef]
  112. Hooper, C.; Taylor, D.L.; Pocock, J.M. Pure albumin is a potent trigger of calcium signalling and proliferation in microglia but not macrophages or astrocytes. J. Neurochem. 2005, 92, 1363–1376. [Google Scholar] [CrossRef]
  113. Hooper, C.; Pinteaux-Jones, F.; Fry, V.A.; Sevastou, I.G.; Baker, D.; Heales, S.J.; Pocock, J.M. Differential effects of albumin on microglia and macrophages; implications for neurodegeneration following blood-brain barrier damage. J. Neurochem. 2009, 109, 694–705. [Google Scholar] [CrossRef]
  114. Ralay Ranaivo, H.; Hodge, J.N.; Choi, N.; Wainwright, M.S. Albumin induces upregulation of matrix metalloproteinase-9 in astrocytes via MAPK and reactive oxygen species-dependent pathways. J. Neuroinflammation 2012, 9, 68. [Google Scholar] [CrossRef] [Green Version]
  115. Krizbai, I.A.; Deli, M.A. Signalling pathways regulating the tight junction permeability in the blood-brain barrier. Cell Mol. Biol. 2003, 49, 23–31. [Google Scholar] [PubMed]
  116. Fujibe, M.; Chiba, H.; Kojima, T.; Soma, T.; Wada, T.; Yamashita, T.; Sawada, N. Thr203 of claudin-1, a putative phosphorylation site for MAP kinase, is required to promote the barrier function of tight junctions. Exp. Cell Res. 2004, 295, 36–47. [Google Scholar] [CrossRef] [PubMed]
  117. Beckers, C.M.; van Hinsbergh, V.W.; van Nieuw Amerongen, G.P. Driving Rho GTPase activity in endothelial cells regulates barrier integrity. Thromb. Haemost. 2010, 103, 40–55. [Google Scholar] [CrossRef] [PubMed]
  118. JIang, X.; Andjelkovic, A.V.; Zhu, L.; Yang, T.; Bennett, M.; Chen, J.; Keep, R.F.; Shi, Y. Blood-brain barrier dysfunction and recovery after ischemic stroke. Prog. Neurobiol. 2018, 163–164, 144–171. [Google Scholar] [CrossRef]
  119. Salman, M.M.; Sheilabi, M.A.; Bhattacharyya, D.; Kitchen, P.; Conner, A.C.; Bill, R.M.; Woodroofe, M.N.; Conner, M.T.; Princivalle, A.P. Transcriptome analysis suggests a role for the differential expression of cerebral aquaporins and the MAPK signalling pathway in human temporal lobe epilepsy. Eur J. Neurosci. 2017, 46, 2121–2132. [Google Scholar] [CrossRef]
  120. Guo, F.; Xu, D.; Lin, Y.; Wang, G.; Wang, F.; Gao, Q.; Wei, Q.; Lei, S. Chemokine CCL2 contributes to BBB disruption via the p38 MAPK signaling pathway following acute intracerebral hemorrhage. Faseb J. 2020, 34, 1872–1884. [Google Scholar] [CrossRef] [Green Version]
  121. Van Wagoner, N.J.; Choi, C.; Repovic, P.; Benveniste, E.N. Oncostatin M regulation of interleukin-6 expression in astrocytes: Biphasic regulation involving the mitogen-activated protein kinases ERK1/2 and p38. J. Neurochem. 2000, 75, 563–575. [Google Scholar] [CrossRef]
  122. Ruprecht, K.; Kuhlmann, T.; Seif, F.; Hummel, V.; Kruse, N.; Bruck, W.; Rieckmann, P. Effects of oncostatin M on human cerebral endothelial cells and expression in inflammatory brain lesions. J. Neuropathol. Exp. Neurol. 2001, 60, 1087–1098. [Google Scholar] [CrossRef] [Green Version]
  123. Repovic, P.; Mi, K.; Benveniste, E.N. Oncostatin M enhances the expression of prostaglandin E2 and cyclooxygenase-2 in astrocytes: Synergy with interleukin-1beta, tumor necrosis factor-alpha, and bacterial lipopolysaccharide. Glia 2003, 42, 433–446. [Google Scholar] [CrossRef]
  124. Rodrigues, S.F.; Granger, D.N. Blood cells and endothelial barrier function. Tissue Barriers 2015, 3, e978720. [Google Scholar] [CrossRef] [Green Version]
  125. Van Landeghem, F.K.; Weiss, T.; Oehmichen, M.; von Deimling, A. Decreased expression of glutamate transporters in astrocytes after human traumatic brain injury. J. Neurotrauma 2006, 23, 1518–1528. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, Q.G.; Laird, M.D.; Han, D.; Nguyen, K.; Scott, E.; Dong, Y.; Dhandapani, K.M.; Brann, D.W. Critical role of NADPH oxidase in neuronal oxidative damage and microglia activation following traumatic brain injury. PLoS ONE 2012, 7, e34504. [Google Scholar] [CrossRef] [PubMed]
  127. Abdul-Muneer, P.M.; Schuetz, H.; Wang, F.; Skotak, M.; Jones, J.; Gorantla, S.; Zimmerman, M.C.; Chandra, N.; Haorah, J. Induction of oxidative and nitrosative damage leads to cerebrovascular inflammation in an animal model of mild traumatic brain injury induced by primary blast. Free Radic. Biol. Med. 2013, 60, 282–291. [Google Scholar] [CrossRef] [PubMed]
  128. Abdul-Muneer, P.M.; Chandra, N.; Haorah, J. Interactions of oxidative stress and neurovascular inflammation in the pathogenesis of traumatic brain injury. Mol. Neurobiol. 2015, 51, 966–979. [Google Scholar] [CrossRef]
  129. Logsdon, A.F.; Lucke-Wold, B.P.; Turner, R.C.; Huber, J.D.; Rosen, C.L.; Simpkins, J.W. Role of microvascular disruption in brain damage from traumatic brain injury. Compr. Physiol. 2015, 5, 1147–1160. [Google Scholar]
  130. Wang, Q.; Doerschuk, C.M. The signaling pathways induced by neutrophil-endothelial cell adhesion. Antioxid. Redox Signal. 2002, 4, 39–47. [Google Scholar] [CrossRef]
  131. McColl, B.W.; Rothwell, N.J.; Allan, S.M. Systemic inflammation alters the kinetics of cerebrovascular tight junction disruption after experimental stroke in mice. J. Neurosci. 2008, 28, 9451–9462. [Google Scholar] [CrossRef] [Green Version]
  132. Nielsen, S.; Nagelhus, E.A.; Amiry-Moghaddam, M.; Bourque, C.; Agre, P.; Ottersen, O.P. Specialized membrane domains for water transport in glial cells: High-resolution immunogold cytochemistry of aquaporin-4 in rat brain. J. Neurosci. 1997, 17, 171–180. [Google Scholar] [CrossRef] [Green Version]
  133. Kitchen, P.; Day, R.E.; Taylor, L.H.; Salman, M.M.; Bill, R.M.; Conner, M.T.; Conner, A.C. Identification and molecular mechanisms of the rapid tonicity-induced relocalization of the aquaporin 4 channel. J. Biol. Chem. 2015, 290, 16873–16881. [Google Scholar] [CrossRef] [Green Version]
  134. Sun, M.C.; Honey, C.R.; Berk, C.; Wong, N.L.; Tsui, J.K. Regulation of aquaporin-4 in a traumatic brain injury model in rats. J. Neurosurg. 2003, 98, 565–569. [Google Scholar] [CrossRef] [Green Version]
  135. Hu, H.; Yao, H.T.; Zhang, W.P.; Zhang, L.; Ding, W.; Zhang, S.H.; Chen, Z.; Wei, E.Q. Increased expression of aquaporin-4 in human traumatic brain injury and brain tumors. J. Zhejiang Univ. Sci. B. 2005, 6, 33–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Lo Pizzo, M.; Schiera, G.; Di Liegro, I.; Di Liegro, C.M.; Pal, J.; Czeiter, E.; Sulyok, E.; Doczi, T. Aquaporin-4 distribution in control and stressed astrocytes in culture and in the cerebrospinal fluid of patients with traumatic brain injuries. Neurol. Sci. 2013, 34, 1309–1314. [Google Scholar] [CrossRef] [PubMed]
  137. Zhang, Y.; Xu, K.; Liu, Y.; Erokwu, B.O.; Zhao, P.; Flask, C.A.; Ramos-Estebanez, C.; Farr, G.W.; LaManna, J.C.; Boron, W.F.; et al. Increased cerebral vascularization and decreased water exchange across the blood-brain barrier in aquaporin-4 knockout mice. PLoS ONE 2019, 14, e0218415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Haj-Yasein, N.N.; Vindedal, G.F.; Eilert-Olsen, M.; Gundersen, G.A.; Skare, O.; Laake, P.; Klungland, A.; Thoren, A.E.; Burkhardt, J.M.; Ottersen, O.P.; et al. Glial-conditional deletion of aquaporin-4 (Aqp4) reduces blood-brain water uptake and confers barrier function on perivascular astrocyte endfeet. Proc. Natl. Acad. Sci. USA 2011, 108, 17815–17820. [Google Scholar] [CrossRef] [Green Version]
  139. Karve, I.P.; Taylor, J.M.; Crack, P.J. The contribution of astrocytes and microglia to traumatic brain injury. Br. J. Pharmacol. 2016, 173, 692–702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Michinaga, S.; Koyama, Y. Dual roles of astrocyte-derived factors in regulation of blood-brain barrier function after brain damage. Int. J. Mol. Sci. 2019, 20, 571. [Google Scholar] [CrossRef] [Green Version]
  141. Argaw, A.T.; Gurfein, B.T.; Zhang, Y.; Zameer, A.; John, G.R. VEGF-mediated disruption of endothelial CLN-5 promotes blood-brain barrier breakdown. Proc. Natl. Acad. Sci. USA 2009, 106, 1977–1982. [Google Scholar] [CrossRef] [Green Version]
  142. Yeung, P.K.; Shen, J.; Chung, S.S.; Chung, S.K. Targeted over-expression of endothelin-1 in astrocytes leads to more severe brain damage and vasospasm after subarachnoid hemorrhage. BMC Neurosci. 2013, 14, 131. [Google Scholar] [CrossRef] [Green Version]
  143. Armstead, W.M.; Kreipke, C.W. Endothelin-1 is upregulated after traumatic brain injury: A cross-species, cross-model analysis. Neurol. Res. 2011, 33, 133–136. [Google Scholar] [CrossRef]
  144. Gadea, A.; Schinelli, S.; Gallo, V. Endothelin-1 regulates astrocyte proliferation and reactive gliosis via a JNK/c-Jun signaling pathway. J. Neurosci. 2008, 28, 2394–2408. [Google Scholar] [CrossRef]
  145. Michinaga, S.; Kimura, A.; Hatanaka, S.; Minami, S.; Asano, A.; Ikushima, Y.; Matsui, S.; Toriyama, Y.; Fujii, M.; Koyama, Y. Delayed administration of BQ788, an ETB antagonist, after experimental traumatic brain injury promotes recovery of blood-brain barrier function and a reduction of cerebral edema in mice. J. Neurotrauma 2018, 35, 1481–1494. [Google Scholar] [CrossRef] [PubMed]
  146. Reijerkerk, A.; Lakeman, K.A.; Drexhage, J.A.; van Het Hof, B.; van Wijck, Y.; van der Pol, S.M.; Kooij, G.; Geerts, D.; de Vries, H.E. Brain endothelial barrier passage by monocytes is controlled by the endothelin system. J. Neurochem. 2012, 121, 730–737. [Google Scholar] [CrossRef] [PubMed]
  147. Rothstein, J.D.; Dykes-Hoberg, M.; Pardo, C.A.; Bristol, L.A.; Jin, L.; Kuncl, R.W.; Kanai, Y.; Hediger, M.A.; Wang, Y.; Schielke, J.P.; et al. Knockout of glutamate transporters reveals a major role for astroglial transport in excitotoxicity and clearance of glutamate. Neuron 1996, 16, 675–686. [Google Scholar] [CrossRef] [Green Version]
  148. Beschorner, R.; Dietz, K.; Schauer, N.; Mittelbronn, M.; Schluesener, H.J.; Trautmann, K.; Meyermann, R.; Simon, P. Expression of EAAT1 reflects a possible neuroprotective function of reactive astrocytes and activated microglia following human traumatic brain injury. Histol. Histopathol. 2007, 22, 515–526. [Google Scholar]
  149. Vazana, U.; Veksler, R.; Pell, G.S.; Prager, O.; Fassler, M.; Chassidim, Y.; Roth, Y.; Shahar, H.; Zangen, A.; Raccah, R.; et al. Glutamate-mediated blood-brain barrier opening: Implications for neuroprotection and drug delivery. J. Neurosci. 2016, 36, 7727–7739. [Google Scholar] [CrossRef]
  150. Dietrich, J.B. The adhesion molecule ICAM-1 and its regulation in relation with the blood-brain barrier. J. Neuroimmunol. 2002, 128, 58–68. [Google Scholar] [CrossRef]
  151. Lutton, E.M.; Farney, S.K.; Andrews, A.M.; Shuvaev, V.V.; Chuang, G.Y.; Muzykantov, V.R.; Ramirez, S.H. Endothelial targeted strategies to combat oxidative stress: Improving outcomes in traumatic brain injury. Front. Neurol. 2019, 10, 582. [Google Scholar] [CrossRef]
  152. Yang, Y.R.; Xiong, X.Y.; Liu, J.; Wu, L.R.; Zhong, Q.; Zhou, K.; Meng, Z.Y.; Liu, L.; Wang, F.X.; Gong, Q.; et al. Mfsd2a (Major Facilitator Superfamily Domain Containing 2a) attenuates intracerebral hemorrhage-induced blood-brain barrier disruption by inhibiting vesicular transcytosis. J. Am. Heart Assoc. 2017, 6, e005811. [Google Scholar] [CrossRef]
  153. Zhao, C.; Ma, J.; Wang, Z.; Li, H.; Shen, H.; Li, X.; Chen, G. Mfsd2a attenuates blood-brain barrier disruption after sub-arachnoid hemorrhage by inhibiting caveolae-mediated transcellular transport in rats. Transl. Stroke Res. 2020, 1–16. [Google Scholar] [CrossRef]
  154. Eser Ocak, P.; Ocak, U.; Sherchan, P.; Gamdzyk, M.; Tang, J.; Zhang, J.H. Overexpression of Mfsd2a attenuates blood brain barrier dysfunction via Cav-1/Keap-1/Nrf-2/HO-1 pathway in a rat model of surgical brain injury. Exp. Neurol. 2020, 326, 113203. [Google Scholar] [CrossRef]
  155. Malik, V.A.; Di Benedetto, B. The blood-brain barrier and the EphR/Ephrin system: Perspectives on a link between neurovascular and neuropsychiatric disorders. Front. Mol. Neurosci. 2018, 11, 127. [Google Scholar] [CrossRef] [PubMed]
  156. Citi, S.; Cordenonsi, M. Tight junction proteins. Biochim. Biophys. Acta 1998, 1448, 1–11. [Google Scholar] [CrossRef] [Green Version]
  157. Martin-Padura, I.; Lostaglio, S.; Schneemann, M.; Williams, L.; Romano, M.; Fruscella, P.; Panzeri, C.; Stoppacciaro, A.; Ruco, L.; Villa, A.; et al. Junctional adhesion molecule, a novel member of the immunoglobulin superfamily that distributes at intercellular junctions and modulates monocyte transmigration. J. Cell Biol. 1998, 142, 117–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Gonzalez-Mariscal, L.; Betanzos, A.; Nava, P.; Jaramillo, B.E. Tight junction proteins. Prog. Biophys. Mol. Biol. 2003, 81, 1–44. [Google Scholar] [CrossRef]
  159. Stamatovic, S.M.; Keep, R.F.; Andjelkovic, A.V. Brain endothelial cell-cell junctions: How to "open" the blood brain barrier. Curr. Neuropharmacol. 2008, 6, 179–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. McNeil, E.; Capaldo, C.T.; Macara, I.G. Zonula occludens-1 function in the assembly of tight junctions in Madin-Darby canine kidney epithelial cells. Mol. Biol. Cell 2006, 17, 1922–1932. [Google Scholar] [CrossRef] [Green Version]
  161. Serlin, Y.; Shelef, I.; Knyazer, B.; Friedman, A. Anatomy and physiology of the blood-brain barrier. Semin. Cell Dev. Biol. 2015, 38, 2–6. [Google Scholar] [CrossRef] [Green Version]
  162. Yamamizu, K.; Iwasaki, M.; Takakubo, H.; Sakamoto, T.; Ikuno, T.; Miyoshi, M.; Kondo, T.; Nakao, Y.; Nakagawa, M.; Inoue, H.; et al. In vitro modeling of blood-brain barrier with human iPSC-derived endothelial cells, pericytes, neurons, and astrocytes via Notch signaling. Stem Cell Reports 2017, 8, 634–647. [Google Scholar] [CrossRef]
  163. Mizee, M.R.; de Vries, H.E. Blood-brain barrier regulation: Environmental cues controlling the onset of barrier properties. Tissue Barriers 2013, 1, e26882. [Google Scholar] [CrossRef] [Green Version]
  164. Liebner, S.; Czupalla, C.J.; Wolburg, H. Current concepts of blood-brain barrier development. Int. J. Dev. Biol. 2011, 55, 467–476. [Google Scholar] [CrossRef]
  165. Obermeier, B.; Daneman, R.; Ransohoff, R.M. Development, maintenance and disruption of the blood-brain barrier. Nat. Med. 2013, 19, 1584–1596. [Google Scholar] [CrossRef] [PubMed]
  166. Daneman, R.; Zhou, L.; Kebede, A.A.; Barres, B.A. Pericytes are required for blood-brain barrier integrity during embryogenesis. Nature 2010, 468, 562–566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Winkler, E.A.; Bell, R.D.; Zlokovic, B.V. Central nervous system pericytes in health and disease. Nat. Neurosci. 2011, 14, 1398–1405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Sweeney, M.D.; Ayyadurai, S.; Zlokovic, B.V. Pericytes in the neurovascular unit: Key functions and signaling pathways. Nat. Neurosci. 2017, 19, 771–783. [Google Scholar] [CrossRef]
  169. Leibner, S.; Dijkhuizen, R.M.; Reiss, Y.; Plate, K.H.; Agalliu, D.; Constantin, G. Functional morphology of the blood brain barrier in health and disease. Acta Neuropathol. 2018, 135, 311–336. [Google Scholar] [CrossRef] [Green Version]
  170. Armulik, A.; Genove, G.; Mae, M.; Nisancioglu, M.H.; Wallgard, E.; Niaudet, C.; He, L.; Norlin, J.; Lindblom, P.; Strittmatter, K.; et al. Pericytes regulate the blood-brain barrier. Nature 2010, 468, 557–561. [Google Scholar] [CrossRef] [Green Version]
  171. Daneman, R.; Prat, A. The blood-brain barrier. Cold Spring Harb. Perspect. Biol. 2015, 7, a020412. [Google Scholar] [CrossRef] [Green Version]
  172. Abbott, N.J.; Ronnback, L.; Hansson, E. Astrocyte-endothelial interactions at the blood-brain barrier. Nat. Rev. Neurosci. 2006, 7, 41–53. [Google Scholar] [CrossRef]
  173. Abbott, N.J. Astrocyte-endothelial interactions and blood-brain barrier permeability. J. Anat. 2002, 200, 629–638. [Google Scholar] [CrossRef]
  174. Igarashi, Y.; Utsumi, H.; Chiba, H.; Yamada-Sasamori, Y.; Tobioka, H.; Kamimura, Y.; Furuuchi, K.; Kokai, Y.; Nakagawa, T.; Mori, M.; et al. Glial cell line-derived neurotrophic factor induces barrier function of endothelial cells forming the blood-brain barrier. Biochem. Biophys. Res. Commun. 1999, 261, 108–112. [Google Scholar] [CrossRef]
  175. Watkins, S.; Robel, S.; Kimbrough, I.F.; Robert, S.M.; Ellis-Davies, G.; Sontheimer, H. Disruption of astrocyte-vascular coupling and the blood-brain barrier by invading glioma cells. Nat. Commun. 2014, 5, 4196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Hayashi, Y.; Nomura, M.; Yamagishi, S.; Harada, S.; Yamashita, J.; Yamamoto, H. Induction of various blood-brain barrier properties in non-neural endothelial cells by close apposition to co-cultured astrocytes. Glia 1997, 19, 13–26. [Google Scholar] [CrossRef]
  177. Dehouck, M.P.; Meresse, S.; Delorme, P.; Fruchart, J.C.; Cecchelli, R. An easier, reproducible, and mass-production method to study the blood-brain barrier in vitro. J. Neurochem. 1990, 54, 1798–1801. [Google Scholar] [CrossRef] [PubMed]
  178. Rubin, L.L.; Hall, D.E.; Porter, S.; Barbu, K.; Cannon, C.; Horner, H.C.; Janatpour, M.; Liaw, C.W.; Manning, K.; Morales, J.; et al. A cell culture model of the blood-brain barrier. J. Cell Biol. 1991, 115, 1725–1735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Suri, C.; Jones, P.F.; Patan, S.; Bartunkova, S.; Maisonpierre, P.C.; Davis, S.; Sato, T.N.; Yancopoulos, G.D. Requisite role of angiopoietin-1, a ligand for the TIE2 receptor, during embryonic angiogenesis. Cell 1996, 87, 1171–1180. [Google Scholar] [CrossRef] [Green Version]
  180. Pfaff, D.; Fiedler, U.; Augustin, H.G. Emerging roles of the Angiopoietin-Tie and the ephrin-Eph systems as regulators of cell trafficking. J. Leukoc. Biol. 2006, 80, 719–726. [Google Scholar] [CrossRef]
  181. Nagase, T.; Nagase, M.; Machida, M.; Fujita, T. Hedgehog signalling in vascular development. Angiogenesis 2008, 11, 71–77. [Google Scholar] [CrossRef]
  182. Alvarez, J.I.; Dodelet-Devillers, A.; Kebir, H.; Ifergan, I.; Fabre, P.J.; Terouz, S.; Sabbagh, M.; Wosik, K.; Bourbonniere, L.; Bernard, M.; et al. The Hedgehog pathway promotes blood-brain barrier integrity and CNS immune quiescence. Science 2011, 334, 1727–1731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Lee, S.W.; Kim, W.J.; Choi, Y.K.; Song, H.S.; Son, M.J.; Gelman, I.H.; Kim, Y.J.; Kim, K.W. SSeCKS regulates angiogenesis and tight junction formation in blood-brain barrier. Nat. Med. 2003, 9, 900–906. [Google Scholar] [CrossRef] [PubMed]
  184. Alvarez, J.I.; Katayama, T.; Prat, A. Glial influence on the blood brain barrier. Glia 2013, 61, 1939–1958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Siddiqui, M.R.; Mayanil, C.S.; Kim, K.S.; Tomita, T. Angiopoietin-1 regulates brain endothelial permeability through PTPN-2 mediated tyrosine dephosphorylation of occludin. PLoS ONE 2015, 10, e0130857. [Google Scholar] [CrossRef] [PubMed]
  186. Neuhaus, J. Orthogonal arrays of particles in astroglial cells: Quantitative analysis of their density, size, and correlation with intramembranous particles. Glia 1990, 3, 241–251. [Google Scholar] [CrossRef] [PubMed]
  187. Carlos, T.M.; Clark, R.S.; Franicola-Higgins, D.; Schiding, J.K.; Kochanek, P.M. Expression of endothelial adhesion molecules and recruitment of neutrophils after traumatic brain injury in rats. J. Leukoc. Biol. 1997, 61, 279–285. [Google Scholar] [CrossRef]
  188. Balabanov, R.; Goldman, H.; Murphy, S.; Pellizon, G.; Owen, C.; Rafols, J.; Dore-Duffy, P. Endothelial cell activation following moderate traumatic brain injury. Neurol. Res. 2001, 23, 175–182. [Google Scholar] [CrossRef] [PubMed]
  189. Ziebell, J.M.; Morganti-Kossmann, M.C. Involvement of pro- and anti-inflammatory cytokines and chemokines in the pathophysiology of traumatic brain injury. Neurotherapeutics 2010, 7, 22–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Rustenhoven, J.; Jansson, D.; Smyth, L.C.; Dragunow, M. Brain pericytes as mediators of neuroinflammation. Trends Pharmacol. Sci. 2017, 38, 291–304. [Google Scholar] [CrossRef]
  191. Shen, W.; Li, S.; Chung, S.H.; Zhu, L.; Stayt, J.; Su, T.; Couraud, P.O.; Romero, I.A.; Weksler, B.; Gillies, M.C. Tyrosine phosphorylation of VE-cadherin and claudin-5 is associated with TGF-beta1-induced permeability of centrally derived vascular endothelium. Eur. J. Cell Biol. 2011, 90, 323–332. [Google Scholar] [CrossRef]
  192. Mankertz, J.; Tavalali, S.; Schmitz, H.; Mankertz, A.; Riecken, E.O.; Fromm, M.; Schulzke, J.D. Expression from the human occludin promoter is affected by tumor necrosis factor alpha and interferon gamma. J. Cell Sci. 2000, 113, 2085–2090. [Google Scholar]
  193. Bolton, S.J.; Anthony, D.C.; Perry, V.H. Loss of the tight junction proteins occludin and zonula occludens-1 from cerebral vascular endothelium during neutrophil-induced blood-brain barrier breakdown in vivo. Neuroscience 1998, 86, 1245–1257. [Google Scholar] [CrossRef]
  194. Methia, N.; Andre, P.; Hafezi-Moghadam, A.; Economopoulos, M.; Thomas, K.L.; Wagner, D.D. ApoE deficiency compromises the blood brain barrier especially after injury. Mol. Med. 2001, 7, 810–815. [Google Scholar] [CrossRef]
  195. Hafezi-Moghadam, A.; Thomas, K.L.; Wagner, D.D. ApoE deficiency leads to a progressive age-dependent blood-brain barrier leakage. Am. J. Physiol. Cell Physiol. 2007, 292, C1256–C1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Mayeux, R.; Ottman, R.; Maestre, G.; Ngai, C.; Tang, M.X.; Ginsberg, H.; Chun, M.; Tycko, B.; Shelanski, M. Synergistic effects of traumatic head injury and apolipoprotein-epsilon 4 in patients with Alzheimer’s disease. Neurology 1995, 45, 555–557. [Google Scholar] [CrossRef] [PubMed]
  197. Pun, P.B.; Lu, J.; Moochhala, S. Involvement of ROS in BBB dysfunction. Free Radic. Res. 2009, 43, 348–364. [Google Scholar] [CrossRef] [PubMed]
  198. Mertsch, K.; Blasig, I.; Grune, T. 4-Hydroxynonenal impairs the permeability of an in vitro rat blood-brain barrier. Neurosci. Lett. 2001, 314, 135–138. [Google Scholar] [CrossRef]
  199. Anzabi, M.; Ardalan, M.; Iversen, N.K.; Rafati, A.H.; Hansen, B.; Ostergaard, L. Hippocampal atrophy following subarachnoid hemorrhage correlates with disruption of astrocyte morphology and capillary coverage by AQP4. Front. Cell Neurosci. 2018, 12, 19. [Google Scholar] [CrossRef] [Green Version]
  200. Kubotera, H.; Ikeshima-Kataoka, H.; Hatashita, Y.; Allegra Mascaro, A.L.; Pavone, F.S.; Inoue, T. Astrocytic endfeet re-cover blood vessels after removal by laser ablation. Sci. Rep. 2019, 9, 1263. [Google Scholar] [CrossRef] [Green Version]
  201. Neri, M.; Frati, A.; Turillazzi, E.; Cantatore, S.; Cipolloni, L.; Di Paolo, M.; Frati, P.; La Russa, R.; Maiese, A.; Scopetti, M.; et al. Immunohistochemical evaluation of Aquaporin-4 and its correlation with CD68, IBA-1, HIF-1alpha, GFAP, and CD15 Expressions in Fatal Traumatic Brain Injury. Int. J. Mol. Sci. 2018, 19, 3544. [Google Scholar] [CrossRef] [Green Version]
  202. Manley, G.T.; Fujimura, M.; Ma, T.; Noshita, N.; Filiz, F.; Bollen, A.W.; Chan, P.; Verkman, A.S. Aquaporin-4 deletion in mice reduces brain edema after acute water intoxication and ischemic stroke. Nat. Med. 2000, 6, 159–163. [Google Scholar] [CrossRef]
  203. Yao, X.; Derugin, N.; Manley, G.T.; Verkman, A.S. Reduced brain edema and infarct volume in aquaporin-4 deficient mice after transient focal cerebral ischemia. Neurosci. Lett. 2015, 584, 368–372. [Google Scholar] [CrossRef] [Green Version]
  204. Wolburg, H.; Noell, S.; Fallier-Becker, P.; Mack, A.F.; Wolburg-Buchholz, K. The disturbed blood-brain barrier in human glioblastoma. Mol. Aspects Med. 2012, 33, 579–589. [Google Scholar] [CrossRef]
  205. Tang, G.; Liu, Y.; Zhang, Z.; Lu, Y.; Wang, Y.; Huang, J.; Li, Y.; Chen, X.; Gu, X.; Wang, Y.; et al. Mesenchymal stem cells maintain blood-brain barrier integrity by inhibiting aquaporin-4 upregulation after cerebral ischemia. Stem Cells 2014, 32, 3150–3162. [Google Scholar] [CrossRef] [PubMed]
  206. Eilert-Olsen, M.; Haj-Yasein, N.N.; Vindedal, G.F.; Enger, R.; Gundersen, G.A.; Hoddevik, E.H.; Petersen, P.H.; Haug, F.M.; Skare, O.; Adams, M.E.; et al. Deletion of aquaporin-4 changes the perivascular glial protein scaffold without disrupting the brain endothelial barrier. Glia 2012, 60, 432–440. [Google Scholar] [CrossRef] [PubMed]
  207. Szu, J.I.; Chaturvedi, S.; Patel, D.D.; Binder, D.K. Aquaporin-4 Dysregulation in a Controlled Cortical Impact Injury Model of Posttraumatic Epilepsy. Neuroscience 2020, 428, 140–153. [Google Scholar] [CrossRef] [PubMed]
  208. Glober, N.K.; Sprague, S.; Ahmad, S.; Mayfield, K.G.; Fletcher, L.M.; Digicaylioglu, M.H.; Sayre, N.L. Acetazolamide treatment prevents redistribution of astrocyte Aquaporin 4 after murine traumatic brain injury. Neurosci. J. 2019, 2019, 2831501. [Google Scholar] [CrossRef] [Green Version]
  209. Salman, M.M.; Kitchen, P.; Woodroofe, M.N.; Brown, J.E.; Bill, R.M.; Conner, A.C.; Conner, M.T. Hypothermia increases aquaporin 4 (AQP4) plasma membrane abundance in human primary cortical astrocytes via a calcium/transient receptor potential vanilloid 4 (TRPV4)- and calmodulin-mediated mechanism. Eur. J. Neurosci. 2017, 46, 2542–2547. [Google Scholar] [CrossRef] [Green Version]
  210. Yang, Y.; Rosenberg, G.A. MMP-mediated disruption of claudin-5 in the blood-brain barrier of rat brain after cerebral ischemia. Methods Mol. Biol. 2011, 762, 333–345. [Google Scholar]
  211. Chapouly, C.; Tadesse Argaw, A.; Horng, S.; Castro, K.; Zhang, J.; Asp, L.; Loo, H.; Laitman, B.M.; Mariani, J.N.; Straus Farber, R.; et al. Astrocytic TYMP and VEGFA drive blood-brain barrier opening in inflammatory central nervous system lesions. Brain 2015, 138, 1548–1567. [Google Scholar] [CrossRef] [Green Version]
  212. Van Bruggen, N.; Thibodeaux, H.; Palmer, J.T.; Lee, W.P.; Fu, L.; Cairns, B.; Tumas, D.; Gerlai, R.; Williams, S.P.; van Lookeren Campagne, M.; et al. VEGF antagonism reduces edema formation and tissue damage after ischemia/reperfusion injury in the mouse brain. J. Clin. Invest. 1999, 104, 1613–1620. [Google Scholar] [CrossRef] [Green Version]
  213. Assis-Nascimento, P.; Tsenkina, Y.; Liebl, D.J. EphB3 signaling induces cortical endothelial cell death and disrupts the blood-brain barrier after traumatic brain injury. Cell Death Dis. 2018, 9, 7. [Google Scholar] [CrossRef] [Green Version]
  214. Ben-Zvi, A.; Lacoste, B.; Kur, E.; Andreone, B.J.; Mayshar, Y.; Yan, H.; Gu, C. Mfsd2a is critical for the formation and function of the blood-brain barrier. Nature 2014, 509, 507–511. [Google Scholar] [CrossRef] [Green Version]
  215. Hazy, A.; Bochicchio, L.; Oliver, A.; Xie, E.; Geng, S.; Brickler, T.; Xie, H.; Li, L.; Allen, I.C.; Theus, M.H. Divergent age-dependent peripheral immune transcriptomic profile following traumatic brain injury. Sci. Rep. 2019, 9, 8564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Lee, P.; Kim, J.; Williams, R.; Sandhir, R.; Gregory, E.; Brooks, W.M.; Berman, N.E. Effects of aging on blood brain barrier and matrix metalloproteases following controlled cortical impact in mice. Exp. Neurol. 2012, 234, 50–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Onyszchuk, G.; He, Y.Y.; Berman, N.E.; Brooks, W.M. Detrimental effects of aging on outcome from traumatic brain injury: A behavioral, magnetic resonance imaging, and histological study in mice. J. Neurotrauma 2008, 25, 153–171. [Google Scholar] [CrossRef] [PubMed]
  218. Fernandez-Lopez, D.; Faustino, J.; Daneman, R.; Zhou, L.; Lee, S.Y.; Derugin, N.; Wendland, M.F.; Vexler, Z.S. Blood-brain barrier permeability is increased after acute adult stroke but not neonatal stroke in the rat. J. Neurosci. 2012, 32, 9588–9600. [Google Scholar] [CrossRef]
  219. Fukuda, A.M.; Pop, V.; Spagnoli, D.; Ashwal, S.; Obenaus, A.; Badaut, J. Delayed increase of astrocytic aquaporin 4 after juvenile traumatic brain injury: Possible role in edema resolution? Neuroscience 2012, 222, 366–378. [Google Scholar] [CrossRef] [Green Version]
  220. Brickler, T.R.; Hazy, A.; Guilhaume Correa, F.; Dai, R.; Kowalski, E.J.A.; Dickerson, R.; Chen, J.; Wang, X.; Morton, P.D.; Whittington, A.; et al. Angiopoietin/Tie2 Axis Regulates the Age-at-Injury Cerebrovascular Response to Traumatic Brain Injury. J. Neurosci. 2018, 38, 9618–9634. [Google Scholar] [CrossRef]
  221. Gong, Y.; Campbell, C.; Qu, B.-X.; Silverman, E.; Moore, C.; Kenney, K.; Diaz-Arrastia, R. Biomarkers of vascular integrity after traumatic brain injury (TBI), and correlation with cerebrovascular reactivity and phosphodiesterase 5 inhibition. Neurology 2016, 86, I11.009. [Google Scholar]
  222. Sulhan, S.; Lyon, K.A.; Shapiro, L.A.; Huang, J.H. Neuroinflammation and blood-brain barrier disruption following traumatic brain injury: Pathophysiology and potential therapeutic targets. J. Neurosci. Res. 2020, 98, 19–28. [Google Scholar] [CrossRef] [Green Version]
  223. Ge, X.; Han, Z.; Chen, F.; Wang, H.; Zhang, B.; Jiang, R.; Lei, P.; Zhang, J. MiR-21 alleviates secondary blood-brain barrier damage after traumatic brain injury in rats. Brain Res. 2015, 1603, 150–157. [Google Scholar] [CrossRef]
  224. Toyama, K.; Spin, J.M.; Deng, A.C.; Huang, T.T.; Wei, K.; Wagenhauser, M.U.; Yoshino, T.; Nguyen, H.; Mulorz, J.; Kundu, S.; et al. MicroRNA-mediated therapy modulating blood-brain barrier disruption improves vascular cognitive impairment. Arterioscler Thromb. Vasc. Biol. 2018, 38, 1392–1406. [Google Scholar] [CrossRef] [Green Version]
  225. Shen, F.; Walker, E.J.; Jiang, L.; Degos, V.; Li, J.; Sun, B.; Heriyanto, F.; Young, W.L.; Su, H. Coexpression of angiopoietin-1 with VEGF increases the structural integrity of the blood-brain barrier and reduces atrophy volume. J. Cereb. Blood Flow Metab. 2011, 31, 2343–2351. [Google Scholar] [CrossRef] [PubMed]
  226. Meng, Z.; Li, M.; He, Q.; Jiang, S.; Zhang, X.; Xiao, J.; Bai, Y. Ectopic expression of human angiopoietin-1 promotes functional recovery and neurogenesis after focal cerebral ischemia. Neuroscience 2014, 267, 135–146. [Google Scholar] [CrossRef] [PubMed]
  227. Verkman, A.S.; Anderson, M.O.; Papadopoulos, M.C. Aquaporins: Important but elusive drug targets. Nature Reviews 2014, 13, 259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Abir-Awan, M.; Kitchen, P.; Salman, M.M.; Conner, M.T.; Conner, A.C.; Bill, R.M. Inhibitors of Mammalian Aquaporin Water Channels. Int J. Mol. Sci. 2019, 20, 1589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Park, T.E.; Mustafaoglu, N.; Herland, A.; Hasselkus, R.; Mannix, R.; FitzGerald, E.A.; Prantil-Baun, R.; Watters, A.; Henry, O.; Benz, M.; et al. Hypoxia-enhanced blood-brain barrier chip recapitulates human barrier function and shuttling of drugs and antibodies. Nat. Commun. 2019, 10, 2621. [Google Scholar] [CrossRef] [PubMed]
  230. Wevers, N.R.; Kasi, D.G.; Gray, T.; Wilschut, K.J.; Smith, B.; van Vught, R.; Shimizu, F.; Sano, Y.; Kanda, T.; Marsh, G.; et al. A perfused human blood-brain barrier on-a-chip for high-throughput assessment of barrier function and antibody transport. Fluids Barriers CNS 2018, 15, 23. [Google Scholar] [CrossRef] [Green Version]
Table 1. Models and Mechanisms of Blood–Brain Barrier (BBB) Disruption following Traumatic Brain Injury (TBI).
Table 1. Models and Mechanisms of Blood–Brain Barrier (BBB) Disruption following Traumatic Brain Injury (TBI).
SpeciesMethod of Evaluation and ModelTimepointMajor FindingsReference
Pre-clinicalMouse
  • Evans blue (EB) extravasation-controlled cortical impact (CCI)
  • EB extravasation and intravital microscopy-mild TBI (mTBI) EB
  • Extravasation and FITC-dextran-weight drop TBI
  • 1–21 days post injury (dpi) 1 and 3 dpi *
  • 60 min *
  • 1 dpi *
  • Apolipoprotein E4 (Apoe4) impairs BBB repair through decreased pericyte and tight junction (TJ) expression with increased matrix metalloproteinase-9 (MMP-9)
  • Calpain III administration before or after decreases BBB permeability
  • Basic fibroblast growth factor (bFGF) given intranasally prior to TBI decreased BBB permeability and increased expression and colocalization of TJ proteins
[27,28,29]
Rat
  • Anti-IgG stain for extravasation-CCI
  • EB extravasation-CCI
  • BBB permeability-FPI
  • 1 *, 3 *, 7, 60, 180 dpi
  • 1 dpi *
  • 2 *, 3 *, and 7 * dpi
  • Decreased expression of TJ proteins
  • Catechin administered via oral gavage decreased BBB leakiness, swelling, and inflammation
  • Adrenomedullin treatment following TBI decreased BBB permeability and increased aquaporin 4 (AQP4) expression
[30,31,32]
Pig
  • Evans blue albumin extravasation-severe TBI
  • Serum protein fibrinogen (FBG) IHC-concussion
  • Immunofluoroscopic evaluation-CCI
  • 6 h *
  • 6–72 h *
  • 6 h *
  • Exosome treatment 1 h following CCI decreased BBB permeability and increased TJ protein levels
  • BBB disruption overlap with axonal injury pathology, similarities to human
  • Valproic acid and fresh-frozen plasma combination treatment following TBI increased TJ expression
[22,33,34]
ClinicalAcute phase
  • Serum protein fibrinogen (FBG) IHC—severe TBI (post-mortem)
  • FBG immunoreactivity and IgG IHC
  • Serum S100b and MRI-DTI-sub-concussive events
  • Serum S100b—severe TBI
  • Serum UCHL1-moderate to severe TBI
  • Plasma miRNA-Severe TBI
  • 6–72 h *
  • Survival 10 h – 13 dpi *
  • Immediately following game, 6 months for DTI
  • 12 h *
  • 12 h *
  • 24 h
  • FBG extravasation but no axonal injury parallels
  • Multifocal FBG extravasation and IgG immunoreactivity
  • Increased serum S100b indicating BBB disruption and persistent abnormalities on magnetic resonance imaging-diffusion tensor imaging (MRI-DTI)
  • Increased S100b levels indicating BBB disruption
  • Increased serum Ubiquitin C-Terminal Hydrolase L1 (UCHL1)
  • Alterations in expression
[19,33,35,36,37,38]
Long-term
post-mortem
FBG immunoreactivity and IgG IHC Survival 1–47 years post-injuryMultifocal FBG extravasation and IgG immunoreactivity[19]
(* indicates significant difference from control).
Table 2. Cellular and Molecular Mechanisms of BBB Breakdown following TBI.
Table 2. Cellular and Molecular Mechanisms of BBB Breakdown following TBI.
ExpressionOrigin/Cell Type(s)FindingsReference
Vasogenic edemaN/AEndothelial cells
  • Increased transendothelial extravasation of serum proteins
  • Increased vascular endothelial growth factor (VEGF)
  • Increased intracranial pressure (ICP) and water accumulation in extracellular space
  • Increased BBB breakdown
[41,42,45,101,102,103]
Cytotoxic edemaN/AAstrocytes, endothelial cells, neurons
  • Astrocytic AQP4 redistribution
  • Sur1/Trpm4 upregulated after injury in endothelial cells (ECs), neurons, and astrocytes
  • Increased NKCC1 expression in neurons and glia
  • Increased ICP and BBB breakdown
  • Increased water accumulation in cells
[42,45,104,105,106,107,108,109]
CaveolaeIncreased Endothelium
  • Albumin extravasation
  • Increased neuroinflammation
  • Plasma Protein Influx
  • Breakdown TJ proteins
  • Increased transcellular permeability
[49,110]
AlbuminIncreased presence in brainPeripheral blood
  • Glial activation: Microglial and astrocyte release of chemokine, cytokines and MMPs
  • Further BBB breakdown
[110,111,112,113,114]
Rho/Rock, PKC, MAPK pathwaysIncreased downstream activationGlial cells
  • Breakdown of TJ proteins
  • Increased Cytokine signaling
  • Increased BBB breakdown
[115,116,117,118,119,120]
Oncostatin MIncreased following TBIMonocytes
  • Increased IL-6 and ERK 1/2 expression in astrocytes
  • Increased prostaglandin E2 and cyclooxygenase-2 in astrocytes
  • Activation of glial and endothelial cells leading to pro-inflammation
  • BBB breakdown
[121,122,123,124]
Reactive oxygen species (ROS)Increased following mitochondrial damageAstrocytes, microglia, ECs, and neurons
  • Glial activation
  • Glial release of MMPs, IL-6, IL-1
  • Increase in ICAM-1, leukocyte adhesion and migration
  • BBB breakdown
[125,126,127,128,129,130,131]
Aquaporin-4 (AQP4)Varied and redistributed Astrocytic end-feet
  • Vasogenic and cytotoxic edema
  • Decreased expression leads to BBB breakdown
  • Redistribution from end-feet to other membranes
  • Disrupted transport and clearance of water/ion leading to accumulation
[101,102,105,132,133,134,135,136,137,138]
MMPsIncreased Astrocytes
  • Breakdown of TJ proteins, occludin and claudin-5
  • Breakdown BBB
[103,139]
Vascular endothelial growth factor (VEGF)IncreasedMonocytes
Astrocytes (VEGF-A)
  • Reduced claudin-5 and occludin expression
  • Interaction with TYMP, represses TJ expression
  • Induces pro-inflammation
  • BBB breakdown
[140,141]
Endothelin signaling Potent vasoconstrictor-upregulated Astrocytes
Endothelial cells (ETB receptor)
  • Increased transendothelial transport of monocytes (ETB)
  • Increased vasogenic edema and vasospasms
  • Glial activation
  • BBB breakdown
[142,143,144,145,146]
Glutamate Increased accumulationAstrocytes
  • Decreased expression of glutamate transporters EAAT1 and EAAT2
  • Activation of N-methyl-D-aspartate (NMDA) receptors and excitotoxicity in neurons and endothelial cells
  • Increased vascular permeability
[125,147,148,149]
Immune cell adhesion molecule-1 (ICAM-1)increasedEndothelial cells
  • Increase leukocyte trafficking
  • BBB breakdown
[150,151]
Major Facilitator Superfamily Domain Containing 2A (Mfsd2a)DecreasedEndothelial cells
  • Increased vesicle trafficking and transcytosis
  • BBB breakdown
[152,153,154]
Eph/Ephrin signaling Increased Endothelial cells, Astrocytes pericytes, microglia, immune cells
  • Decreased TJ protein, zona occludens
  • Increased neuroinflammation
  • BBB breakdown
[155]
Table 3. Genetic and Pharmacological Approaches to BBB Modification.
Table 3. Genetic and Pharmacological Approaches to BBB Modification.
Pharmacological/Genetic ModificationOrigin/
Cell Type(s)
FindingsReference
Ang1Over-expression (adeno-associated and lentivirus vector)Glial
Immune cells
  • Up-regulation of TJ protein, claudin-5, occludin, and ZO-1
  • Reduced infarct volume
[225,226]
Tie2Pharmacological inhibition (soluble Tie2 inhibitor)Endothelial cell
Subset immune cells
  • Reduced occludin expression in endothelial cells
  • Increased VEGF in endothelial cells
  • Attenuated BBB breakdown and neuroprotection in juvenile mice
[220]
Calpain IIIPharmacological inhibitionUbiquitously
  • Improved BBB integrity
  • Displaced ZO-1
[28]
Basic fibroblast growth factor (bFGF)Pharmacological over-expressionNeural stem cells
Capillary endothelial cells
Vascularized tissue (tumors)
  • Increased colocalization of ZO-1, claudin-5, and occludin
  • Improved BBB integrity
[222,29]
CatechinPharmacological administration of tea flavonoid, antioxidantHigh affinity binding to laminin receptors
Potentiate
brain derived neurotrophic factor (BDNF)
  • Decreased water accumulation
  • Decreased inflammation
  • Increased expression of TJ proteins
  • Increased BBB integrity
[31]
EphB3Genetic knockdownAstrocytes
  • Increased pericyte-EC interactions
  • Increased astrocyte-EC interactions
  • Increased endothelial cell survival
  • Increased BBB integrity
[213]
ETB Antagonist administrationEndothelial cells
Astrocytes
  • Increased anti-inflammatory response
  • Decreased transendothelial passage of monocytes
[145]
miRNAsAgomir administration
(i.e., miR-21, miR-501-3p)
Antagomir administration
(i.e., miR-21 antagomir)
N/A
  • Activation of Ang1/Tie2 axis, expression of TJ proteins
  • Suppressed TNF, increased expression of ZO-1
  • Neuroprotection
  • Increased BBB stability
[223,224]
AQP4AQP4 inhibitor: acetazolamide
Inhibition by CCL2/p-38 MAPK inhibitor
Astrocytes
  • Eliminated cytotoxic edema
  • Reduced edema
  • Prevented AQP4 redistribution
[208,120]

Share and Cite

MDPI and ACS Style

Cash, A.; Theus, M.H. Mechanisms of Blood–Brain Barrier Dysfunction in Traumatic Brain Injury. Int. J. Mol. Sci. 2020, 21, 3344. https://doi.org/10.3390/ijms21093344

AMA Style

Cash A, Theus MH. Mechanisms of Blood–Brain Barrier Dysfunction in Traumatic Brain Injury. International Journal of Molecular Sciences. 2020; 21(9):3344. https://doi.org/10.3390/ijms21093344

Chicago/Turabian Style

Cash, Alison, and Michelle H. Theus. 2020. "Mechanisms of Blood–Brain Barrier Dysfunction in Traumatic Brain Injury" International Journal of Molecular Sciences 21, no. 9: 3344. https://doi.org/10.3390/ijms21093344

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