Next Article in Journal
NR3C1 rs6198 Variant May Be Involved in the Relationship of Graves’ Disease with Stressful Events
Next Article in Special Issue
Air Pollution: A Silent Key Driver of Dementia
Previous Article in Journal
Physiological Associations between Vitamin B Deficiency and Diabetic Kidney Disease
Previous Article in Special Issue
Gene–Environment Interactions in Repeat Expansion Diseases: Mechanisms of Environmentally Induced Repeat Instability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Increased Risk of Aging-Related Neurodegenerative Disease after Traumatic Brain Injury

by
Sarah Barker
1,2,3,4,5,
Bindu D. Paul
6,7,8,9 and
Andrew A. Pieper
1,2,3,4,5,10,11,*
1
Center for Brain Health Medicines, Harrington Discovery Institute, University Hospitals Cleveland Medical Center, Cleveland, OH 44106, USA
2
Department of Psychiatry, Case Western Reserve University, Cleveland, OH 44106, USA
3
Geriatric Psychiatry, GRECC, Louis Stokes Cleveland VA Medical Center, Cleveland, OH 44106, USA
4
Institute for Transformative Molecular Medicine, School of Medicine, Case Western Reserve University, Cleveland, OH 44106, USA
5
Department of Pathology, School of Medicine, Case Western Reserve University, Cleveland, OH 44106, USA
6
Department of Pharmacology and Molecular Sciences, Johns Hopkins University School of Medicine, Baltimore, MD 21211, USA
7
Department of Psychiatry and Behavioral Sciences, Johns Hopkins University School of Medicine, Baltimore, MD 21211, USA
8
The Solomon H. Snyder Department of Neuroscience, Johns Hopkins University School of Medicine, Baltimore, MD 21211, USA
9
Lieber Institute for Brain Development, Baltimore, MD 21205, USA
10
Department of Neuroscience, School of Medicine, Case Western Reserve University, Cleveland, OH 44106, USA
11
Translational Therapeutics Core, Cleveland Alzheimer’s Disease Research Center, Cleveland, OH 44106, USA
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(4), 1154; https://doi.org/10.3390/biomedicines11041154
Submission received: 28 February 2023 / Revised: 30 March 2023 / Accepted: 5 April 2023 / Published: 11 April 2023

Abstract

:
Traumatic brain injury (TBI) survivors frequently suffer from chronically progressive complications, including significantly increased risk of developing aging-related neurodegenerative disease. As advances in neurocritical care increase the number of TBI survivors, the impact and awareness of this problem are growing. The mechanisms by which TBI increases the risk of developing aging-related neurodegenerative disease, however, are not completely understood. As a result, there are no protective treatments for patients. Here, we review the current literature surrounding the epidemiology and potential mechanistic relationships between brain injury and aging-related neurodegenerative disease. In addition to increasing the risk for developing all forms of dementia, the most prominent aging-related neurodegenerative conditions that are accelerated by TBI are amyotrophic lateral sclerosis (ALS), frontotemporal dementia (FTD), Parkinson’s disease (PD), and Alzheimer’s disease (AD), with ALS and FTD being the least well-established. Mechanistic links between TBI and all forms of dementia that are reviewed include oxidative stress, dysregulated proteostasis, and neuroinflammation. Disease-specific mechanistic links with TBI that are reviewed include TAR DNA binding protein 43 and motor cortex lesions in ALS and FTD; alpha-synuclein, dopaminergic cell death, and synergistic toxin exposure in PD; and brain insulin resistance, amyloid beta pathology, and tau pathology in AD. While compelling mechanistic links have been identified, significantly expanded investigation in the field is needed to develop therapies to protect TBI survivors from the increased risk of aging-related neurodegenerative disease.

1. Introduction

1.1. Scope of the Problem

Traumatic brain injury (TBI) has an annual incidence of almost 70 million people worldwide and is a major chronic public health problem [1]. Notably, the problem has grown in prevalence as advances in neurocritical care have reduced TBI mortality rates [2]. In the United States alone, 22% of people have sustained at least one TBI with loss of consciousness in their lifetime [3] and there are now approximately 5 million people living with TBI-related disabilities at an estimated annual cost of 80 billion USD [4,5].
TBI is often of multimodal etiology, including aspects of direct contact injury, acceleration/deceleration injury, and blast-wave injury. Regardless of the cause, however, all forms of TBI initiate a complex and unique disease process of primary and secondary injury. Primary injury results from mechanical tissue deformation, including contusion and rapid axonal and vessel shearing [6,7]. Secondary injury is caused by a multitude of interrelated chronic pathologic processes, including ischemia, excitotoxicity, metabolic dysfunction, blood–brain barrier (BBB) deterioration, oxidative stress, and neuroinflammation [7,8]. Over the long term, TBI results in chronic and progressive axonal demyelination and degeneration, leading to neuronal loss and varying degrees of neuropsychiatric impairment [7,9].
Importantly, TBI is also a significant environmental risk factor for developing aging-related neurodegenerative disease [10]. Epidemiologic evidence shows that TBI is associated with a significantly increased risk of developing all forms of dementia. Various studies and meta-analyses found odds ratios ranging from 1.25–1.63, indicating that individuals with a history of TBI have a significantly increased risk of developing dementia [11,12,13]. Indeed, it is estimated that TBI accounts for 5–15% of the attributable risk for dementia [14]. Specifically, TBI is associated with amyotrophic lateral sclerosis (ALS), frontotemporal dementia (FTD), Parkinson’s disease (PD), and Alzheimer’s disease (AD). Among these, ALS and FTD are the least well established as being linked to TBI, while the evidence linking TBI to PD and AD is quite robust. Here, we review the epidemiology and current understanding of the mechanistic links between these complex disease processes in both animal models and humans.

1.2. Challenges in Studying the Role of TBI in the Increased Risk of Aging-Related Neurodegenerative Disease

Numerous epidemiologic studies have assessed TBI as a risk factor for aging-related neurodegenerative disease. Some of these studies have reported conflicting results, which have been attributed to variability in study design, definition of the TBI exposure variable, and ethnicity and age of study subjects [15]. Another complicating factor is that human studies frequently rely on self-reported TBI, which is susceptible to recall bias, in cases that lack administrative data for TBI-related health care. In addition, earlier studies often used small sample sizes due to the low prevalence of clinically diagnosed TBI in the 19th century [16]. Furthermore, the number and/or severity of injuries has not always been monitored in clinical studies. Outcome measures have also not been kept uniform across studies, with utilization of a wide variety of approaches, such as International Classification of Diseases (ICD) codes, National Institute of Neurological and Communicative Disorders and Stroke/Alzheimer’s Disease and Related Disorders Association (NINCDS/ADRDA) AD criteria, Diagnostic and Statistical Manual of Mental Disorders (DSM) classification, and stratification by postmortem neuropathology. Furthermore, reverse causality has not been carefully assessed historically, which is problematic because many neurodegenerative diseases predispose individuals to falls that can result in TBI. Additionally, the duration of the neurodegenerative disease process before diagnosis is often unclear [15]. Importantly, a recent review found that studies with clearly defined TBI exposure variables and longer follow-up periods were more likely to find an association between TBI and dementia [17]. Despite these challenges, however, multiple meta-analyses of the published literature have consistently identified a prominent association between TBI and increased risk of developing aging-related neurodegenerative disease.
To investigate the mechanisms by which TBI increases the risk of aging-related neurodegenerative disease, researchers predominately rely on animal models. Both TBI and neurodegenerative disease are difficult to model in the laboratory, however, due to the heterogeneous nature of these conditions. With respect to TBI, there are four laboratory models of isolated TBI that have historically been most widely used: weight-drop impact acceleration injury [18], controlled cortical impact (CCI) injury [19], fluid percussion injury (FPI) [20,21], and blast injury [22].
In weight drop and CCI injuries, a direct insult is applied to the brain with a weight or rigid impactor. There are many versions of weight drop and CCI models, some of which require a craniotomy and cause cortical tissue loss, and others without craniotomy that cause skull fracture and are associated with high mortality [19]. The FPI model also involves craniotomy, with subsequent infliction of a fluid pressure pulse on the dura. Lastly, blast injuries typically use a compression driven shock tube to simulate exposure to explosive blast, which models some forms of battlefield injury [7,23].
These 4 distinct rodent models of TBI mimic different aspects of clinical TBI in people, with the first three modeling mechanical head trauma and the pure blast model entailing isolated exposure to a blast wave. To model the multimodal nature of most forms of human TBI, we recently characterized an additional rodent model of TBI that uses an overpressure chamber to deliver precise and reproducible aspects of global concussion, acceleration/deceleration injury, and early blast wave exposure without the need for surgery [24]. This model of multimodal TBI initiates acute axonal degeneration and BBB damage that persists chronically and progresses to neuronal cell death and behavioral alterations related to human neuropsychiatric impairment, with metabolomic changes in the blood that mimic changes seen in human TBI patients [25,26,27,28,29,30,31]. Thus, while there is no single laboratory model of TBI that recapitulates all of the disease processes in human TBI, multiple models are valued for their ability to reproduce various aspects of the human condition.
As with TBI, modeling aging-related human neurodegenerative disease in the laboratory is problematic. Here, researchers typically utilize genetic rodent models. However, most cases of aging-related human neurodegenerative disease are sporadic (90–95% for ALS and FTD, 85% for PD, and 92–96% for AD) and not related to any known genetic cause [32,33,34]. Despite the challenges in the field, however, there is much effort to study this important problem with the tools available.

2. Potential Mechanisms of TBI-Mediated Increased Risk of All Forms of Aging-Related Neurodegenerative Disease

While TBI accelerates the age of cognitive decline by approximately three years in people across all forms of dementia, no single reason has been determined for this effect [35]. Indeed, it has become increasingly appreciated that TBI and other neurodegenerative diseases share many pathologic mechanisms, and that several forms of chronic neurodegeneration display mixed and overlapping pathologies. For example, approximately half of patients with probable AD show degenerative pathologies common to other forms of neurodegenerative disease [36]. Likewise, many of the sequelae of TBI overlap with underlying pathology of multiple neurodegenerative diseases. Here, we review current knowledge on the contribution of three different TBI-derived pathological mechanisms that generally increase the risk of all forms of aging-related neurodegenerative disease: oxidative stress, dysregulated proteostasis, and neuroinflammation.

2.1. TBI-Induced Oxidative Stress and Increased Risk of All Forms of Aging-Related Neurodegenerative Disease

Although the human brain accounts for only 2% of total body weight, it voraciously consumes more than 20% of the body’s oxygen [37]. Due to this extraordinarily high rate of metabolism, the brain generates high levels of reactive oxygen species (ROS). Paradoxically, the brain is also among the most vulnerable organs to oxidative damage. This is predominantly due to its high polyunsaturated fatty acid content that is readily altered by ROS and its relatively poor oxidative stress response mechanisms [38]. These factors synergistically generate a multitude of toxic loss-of-function and gain-of-function events in key cellular processes [39,40,41]. The regulation and dysregulation of the antioxidant defense network in the brain, however, is less well characterized.
Particularly damaging ROS in the brain that lead to axonal degeneration and neuronal cell death include superoxide, hydrogen peroxide, hydroxyl radical, nitric oxide, and peroxynitrite. Not surprisingly, cells have evolved multiple levels of antioxidant defense. For example, natural antioxidants in the body that deactivate ROS into stable non-toxic molecules include superoxide dismutase (SOD), catalase, peroxidase, heme oxygenase, and biliverdin reductase, as well as lower molecular weight water- or lipid-soluble free-radical scavengers, such as glutathione (GSH), ascorbic acid, bilirubin, ergothioneine, and cysteine-derived thiols [42,43,44,45]. GSH, one of the most abundant antioxidants in the cell, plays central roles in antioxidant defense and is also a cofactor for several enzymes and important in synaptic transmission [46,47,48]. Similarly, ascorbic acid, also known as vitamin C, participates in a wide variety of neuroprotective signaling pathways [49]. In addition to these molecules, gaseous signaling molecules such as hydrogen sulfide modulate redox balance in the brain through multiple mechanisms [50,51,52,53,54,55]. All of the brain’s antioxidant systems are tightly regulated to provide optimal response for maintaining cellular integrity.
A substantial body of evidence implicates oxidative stress across all forms of aging-related neurodegenerative disease, with lipids, proteins, and DNA all undergoing extensive oxidative damage (Figure 1) [40,42]. Furthermore, proteins that pathologically aggregate in neurodegeneration interact with redox-active metal ions and generate ROS. It is also well documented that TBI robustly generates ROS in both animal models and human patients. In animal models, there is evidence of protein nitration (3-nitrotyrosine: 3-NT), protein oxidation (carbonylation), lipid peroxidation (4-hydoxynonenal: 4-HNE), and DNA oxidation (8-hydroxy-2′-deoxyguanosine: 8-OHdG) within hours after TBI [56,57,58,59]. In human patients, 3-NT accumulates in the cerebrospinal fluid (CSF) and brain after TBI [60]. Importantly, the acute increase in ROS precedes chronic neuronal damage, suggesting a causative role [57,58].
In the brain, oxidative stress has also been linked to dysregulated iron homeostasis (Figure 1). For example, both TBI and aging-related neurodegenerative diseases induce changes in iron metabolism that increase iron deposition and elevate oxidative damage in the brain [61,62,63]. Ferroptosis, an iron-mediated form of cell death, has also been observed in TBI [64,65], and counteracting ferroptosis mitigates TBI symptoms [66]. Importantly, ferroptosis can be countered by several antioxidant proteins and metabolites, such as nuclear factor erythroid 2-related factor 2 (Nrf2) activation, GSH, and cysteine, the metabolism of which is frequently disrupted in neurodegenerative diseases, viral diseases, and aging [67].
Interestingly, increases in ROS are also associated with reduced antioxidant defense mechanisms (GSH, SOD, ascorbate, and protein sulfhydryls) after TBI in rodent models, consistent with the notion that antioxidant molecules and proteins counteract or prevent oxidative damage (Figure 1) [57,68]. In humans, TBI also exhausts the antioxidant reserve over the course of 7 days, as evidenced by reduced ascorbate and GSH [69].
One important regulatory mechanism of antioxidant defense in the brain is the Nrf2 transcription factor that plays a pivotal role in maintaining cellular redox balance in response to oxidative stress. Under basal conditions, Nrf2 is sequestered in the cytosol by Kelch-like ECH-associated-protein 1 (Keap1) and targeted for routine degradation [70]. In response to oxidative stress, however, Nrf2 dissociates from Keap1 and translocates to the nucleus, where it activates transcription of oxidative-stress-response genes (Figure 1) [71]. This phenomenon has been reported after TBI [72], and pharmacologic activation of Nrf2 rescues oxidative damage and neurologic dysfunction after TBI [73]. By contrast, Nrf2 knockout mice show more severe oxidative stress and neurologic dysfunction after TBI [73]. Nrf2 also links redox homeostasis to pain perception mechanisms and tolerance [74].
Overall, it is well-accepted that TBI increases ROS production and depletes antioxidant defense systems in the brain, leading to protein, lipid, and DNA damage that collectively impair neuronal function and eventually lead to neurodegeneration (Figure 1). Thus, TBI-induced oxidative stress represents a link to increased risk of developing all forms of aging-related neurodegenerative disease.

2.2. TBI-Induced Dysregulated Proteostasis and Increased Risk for All Forms of Aging-Related Neurodegenerative Disease

All neurodegenerative diseases share the common pathologic hallmark of aberrant protein aggregation, and there is evidence that TBI dysregulates key proteostatic processes that influence this system. These include the heat shock response, the unfolded protein response, the ubiquitin proteasome system, and autophagy.
Heat shock response is the process by which cellular stress, including elevated ROS, results in transcription-factor-driven expression of heat shock proteins (HSPs). HSPs are chaperones that refold proteins and prevent their aggregation by binding hydrophobic residues [75]. The heat shock response is induced after TBI in both rodent and human TBI [76,77,78,79]. Interestingly, genetically impairing the heat shock response by eliminating Hsp70 confers worse outcomes after TBI [80]. Conversely, genetic and pharmacologic activation of the heat shock response via Hsp70 is protective in rodent models of TBI [80,81].
The unfolded protein response upregulates protein-folding chaperones, reduces protein synthesis, degrades permanently misfolded proteins, and in cases of severe stress, initiates apoptosis. The unfolded protein response is coordinated by three arms: protein kinase RNA-like ER kinase (PERK), inositol-requiring enzyme 1 (IRE-1), and activating transcription factor 6 (ATF-6) in the endoplasmic reticulum (ER). Importantly, this system is induced in rodents after TBI [82,83], and its inhibition is protective [84,85,86].
The ubiquitin proteasome system, which selectively degrades misfolded and damaged proteins, is also dysregulated by TBI. For example, rodent TBI is associated with accumulation of ubiquitinated proteins in the brain, in conjunction with reduced protease activity [87,88]. The ubiquitin proteasome system is also dysregulated after TBI in human patients [89].
Lastly, accumulated misfolded proteins are also normally degraded by autophagy, through sequestration into autophagosomes that fuse with lysosomes for cargo degradation. Notably, autophagic flux is reduced after TBI in both animal models and human patients, and thus contributes to dysregulated proteostasis [89,90,91].
Overall, there is substantial evidence for widespread dysregulation of proteostasis after TBI. Given the commonality of impaired proteostasis across all forms of neurodegeneration, this is considered a general mechanistic link between TBI and aging-related neurodegenerative disease.

2.3. TBI-Induced Neuroinflammation and Increased Risk of All Forms of Aging-Related Neurodegenerative Disease

Neuroinflammation is another prominent component of aging-related neurodegenerative disease [92,93,94,95]. Typically, this process is indicated by reactive morphology of astrocytes and microglia, which secrete inflammatory mediators into the brain parenchyma [96]. Acutely after TBI, the immune response is crucial for clearing debris. However, chronic inflammation, which is seen years after injury in human patients, can perpetuate injury and contribute to aging-related neurodegenerative disease [97]. Recently, TBI has also been linked to development of senescence-associated phenotype (SASP)-derived inflammaging in the brain, which is a chronic low-grade sterile inflammation previously established to contribute to brain aging [98].
During the acute primary injury phase of TBI, there is a rapid release of damage-associated molecular signaling events that stimulate local and infiltrating immune cells to secrete toxic interleukin 6, interleukin 1 beta, and tumor necrosis factor [97]. Notably, blocking aspects of this process has been shown to be neuroprotective in an animal model of TBI [99]. TBI also induces astrogliosis, in which resident immune cells proliferate, change their morphology, and secrete cytokines and chemokines that potentiate neuroinflammation [100]. In addition, necrotic tissue releases double-stranded DNA, which activates the cyclic GMP-AMP synthase and stimulator of interferon genes (cGAS-STING) pathway. Here, cyclic GMP-AMP synthase produces cGAMP, which stimulates expression of pro-inflammatory interferon genes [101]. After TBI, excessive activation of the cGAS-STING pathway mediates neuroinflammation with elevated type-I interferons [102,103]. STING signaling is increased in both rodent and human TBI, and genetically ablating STING is neuroprotective in TBI [104,105].
The proinflammatory milieu induced by TBI also polarizes microglia into an M1-like state with reduced phagocytic activity and enhanced secretion of pro-inflammatory cytokines and free radicals. This toxic transition contributes to neurodegeneration. Conversely, microglia can also be neuroprotective when they are polarized to an M2-like state, in which they aggressively phagocytose debris, secrete anti-inflammatory cytokines and trophic factors, and augment postnatal neurogenesis. In rodent models, TBI induces a mixed M1- and M2-like phenotype in microglia, and much remains to be discovered about how this system could be manipulated to protect patients from TBI-induced acceleration of aging-related neurodegenerative disease [106,107,108].

3. Association of TBI with ALS and FTD

3.1. Epidemiology of TBI-Induced Increased Risk of ALS and FTD

Multiple studies have demonstrated an increased risk of ALS among retired professional athletes of contact sports such as football, soccer, and rugby [109,110,111]. While TBIs are certainly enriched in these individuals compared to the general population, these studies did not specifically account for history of TBI. An analysis of World War II veterans in 1980, however, found that, “men dying of ALS more often had a history of injury [TBI] 15 or more years before death than did the controls during the same period” [112]. Other studies have identified an increased risk for ALS in individuals who had a TBI within one [113] or ten [114] years of disease onset, but did not see an association at all time points. It is difficult to parse the directionality of this relationship because ALS is also associated with motor symptoms that predispose patients to falls. However, one study did not find any association between other physical injuries that could result from falls (trunk, arm, or leg) and ALS, suggesting specificity of the association with TBI [114]. Overall, a meta-analysis of studies published between 1980 and 2011 found insufficient evidence to support an association between a single TBI and ALS [115]. In addition, another study found that a history of TBI does not worsen disease progression or pathology in patients with ALS [116]. Thus, the association between TBI and ALS, while often discussed, remains unclear.
This relationship is perhaps further complicated by the evolution of understanding that ALS and FTD in some cases have overlapping pathology. FTD is a group of disorders associated with progressive damage to the temporal and frontal lobes of the brain associated with personality, language, and behavior, leading to impulsive behavior and communication difficulties. Although FTD is not usually associated with motor symptoms, upwards of 15% of people with FTD do experience motor neuron degeneration, termed FTD with ALS. In these cases, symptoms of FTD traditionally precede ALS. Approximately half of ALS patients also experience some degree of change in their thinking and behavior. Notably, three retrospective case control studies found a significantly increased risk for FTD in people who suffered from TBI [117,118,119]. More recently, it has been shown that a mutation in the noncoding region of the chromosome 9 open reading frame 72 (C9ORF72) locus, where an expansion of the hexanucleotide GGGGCC occurs, is not only a leading cause of ALS but also found in ~40% of cases of FTD [120,121]. Strikingly, prior TBI is specifically associated with advanced onset of disease in patients carrying this hexanucleotide repeat expansion [122]. Future well-controlled prospective studies will be required to elucidate the complex relationship between TBI and risk of developing ALS and FTD.

3.2. Potential Mechanisms of TBI-Induced Increased Risk of ALS and FTD

There is conflicting evidence of whether TBI accelerates or worsens disease in the most common animal model of ALS, which overexpresses a mutant form of human SOD1 (G93A) associated with a small number of human ALS patients. One group found that a single controlled cortical impact TBI did not affect disease onset or survival [123], but that repeated mild TBI caused earlier onset of motor symptoms, chronic motor deficits, and decreased cortical thickness in mutant SOD1 rodents [124,125]. However, another group found that a mild stab-wound injury to the motor cortex did not affect disease onset or progression in three different genetic rodent ALS models (SOD1, TAR DNA binding protein 43 (TDP-43), and fused in sarcoma (FUS)) [126]. It should be noted, however, that stab wounds to the head are one of the least frequent forms of human TBI and have not been specifically epidemiologically linked to aging-associated neurodegenerative disease. Lastly, another study showed that closed head cortical impact did not alter age of onset or lifespan in SOD1 mutant rodents but did worsen disease by more greatly impairing grip strength and increasing electromyography changes [127]. Thus, mechanistic studies show mixed evidence for an association between TBI and increased risk of ALS. However, as described below, substantial evidence indicates that TBI initiates related forms of TDP-43 and motor cortex pathology.

3.2.1. TBI-Induced TDP-43 Pathology and ALS and FTD

A major proposed link by which TBI may increase the risk of ALS and FTD relates to TDP-43. TDP-43, a member of the TAR DNA-binding protein family, binds single-stranded DNA, RNA, and proteins. TDP-43 is implicated in mRNA stability, apoptosis, and cell division. However, its full array of physiologic functions in the brain is not yet fully understood. In both ALS and FTD, TDP-43 aberrantly localizes to the cytoplasm, where it aggregates and forms ubiquitin-positive inclusions [128]. In mouse models of TBI, axonal damage also increases TDP-43 expression, which then mislocalizes to the cytoplasm [129,130,131,132]. Interestingly, TDP-43 pathology after TBI may be transient, as TBI in three different ALS models (SOD1, TDP-43, FUS) acutely increases phosphorylated TDP-43 granules [126]. In human patients as well, a single TBI is sufficient to increase intraneuronal TDP-43, in the absence of the post-translational phosphorylation of TDP-43 that is typically associated with TDP-43 accumulation in ALS and FTD [133].

3.2.2. TBI-Induced Motor Cortex Lesions and ALS and FTD

ALS involves death of motor neurons. In animal models, TBI causes progressive atrophy in the motor cortex and degeneration of the corticospinal tracts, resulting in muscle atrophy and motor deficits reminiscent of ALS [132]. ALS typically presents with focal motor symptoms, with onset in one or more myotomes spreading to adjacent muscles. A small study of 18 cases of sporadic ALS following frontotemporal cortical lesions (latency 8–42 years) found that the lesion site was frequently contralateral to the site of symptom onset, suggesting that TBI may be one “hit” that could lead to disease in conjunction with other genetic and non-genetic factors [134].
In summary, substantial challenges in disease definition and reliable animal models have precluded the firm establishment of whether and how TBI contributes to ALS and FTD. Substantial evidence, however, indicates that this topic is worthy of ongoing investigation.

4. TBI and PD

4.1. Epidemiology of TBI-Induced Increased Risk of PD

Many studies have assessed the relationship between TBI and PD, with a majority finding a significant association with odds ratios ranging from 0.6–6.2 [135]. Conversely, some studies did not find any association between TBI and PD [136,137]. Importantly, two meta-analyses of the published literature found a significant association between TBI and PD (Table 1) [138,139]. Interestingly, higher-severity TBI has been associated with higher risk of developing PD, further supporting a causative association [140,141,142].
It is important to recognize, however, that many epidemiologic studies investigating the association between TBI and PD are confounded by reverse causality, because PD is characterized by motor impairment, which can lead to TBI from falls. Overall, 60% of PD patients report at least one fall and 40% report recurrent falls, compared to 15% of the general older population [143]. While analysis of a large dataset from the Danish national hospital register found an odds ratio of 1.5 (95% confidence interval 1.4–1.7), this was almost completely driven by head injuries that occurred in the three months prior to PD diagnosis (odds ratio 8.0, 95% confidence interval 5.6–11.6) when study subjects were likely to have already experienced motor impairment. When only TBIs more than ten years prior to PD were considered, this study found no association [135]. In a study that controlled for trauma by incorporating non-TBI trauma controls (fractures) in patients over the age of 55, however, TBI patients were significantly more likely to be diagnosed with PD (hazard ratio 1.44) [141]. In summary, further prospective cohort studies are needed to elucidate the temporal relationship between TBI and PD.

4.2. Potential Mechanisms of TBI-Induced Increased Risk of PD

Evidence suggests that TBI initiates pathologic alpha (α)-synuclein loss of function and aggregation and directly harms dopaminergic neurons. Interaction between TBI and toxins known to cause PD provides further evidence for a mechanistic relationship. Further research in animal models that can recapitulate the full spectrum of PD pathology is needed to understand the mechanistic relationship between TBI and PD.

4.2.1. TBI-Induced α-Synuclein Pathology in PD

The α-synuclein protein composes ~1% of all cytoplasmic proteins in neurons and is found predominantly in presynaptic axon terminals where it interacts with phospholipids and proteins [144,145]. Interestingly, α-synuclein was originally identified as the principal non-amyloid beta component of amyloid deposits in AD plaques and blood vessels [146,147,148,149]. Subsequent work showed that α-synuclein also pathologically aggregates into insoluble fibrils that form Lewy body deposits in PD, as well as the related conditions of dementia with Lewy bodies and multiple system atrophy that are variably associated with TBI [150,151,152].
Levels of α-synuclein are elevated five- to ten-fold in cerebral spinal fluid after TBI in human patients [153,154]. In rodent models, TBI also causes α-synuclein accumulation with an age-dependent effect, most prominently in axons [155,156,157]. It has been hypothesized that TBI-induced oxidative and nitrosative stress causes conformationally modified forms of α-synuclein in both TBI and aging, which could contribute to PD [39]. Indeed, axonal α-synuclein pathology is one of the initial changes seen in PD [158]. Neuropathologic studies also show that severe TBI (loss of consciousness greater than 1 hour) is associated with increased accumulation of Lewy bodies in the substantia nigra and frontotemporal cortex. Importantly, this association remained even when only cases with a TBI before the age of 25 were included, controlling for reverse causality [159]. Furthermore, it is hypothesized that α-synuclein loses its function of regulating presynaptic neurotransmitter release in PD pathogenesis [160], and it has been shown that levels of monomeric alpha-synuclein are reduced regionally after TBI in a manner that correlates with synaptic loss [161]. Unfortunately, this topic is challenging to study in the laboratory because there are no TBI rodent models that on their own faithfully recapitulate the aggregated forms of α-synuclein observed in PD.

4.2.2. TBI-Induced Dopaminergic Cell Loss in PD

Loss of dopaminergic neurons in the substantia nigra, which underlies akinesia, rigidity, and postural tremor in PD, is also thought to be a potential mechanism by which TBI increases the risk of PD. For example, dopamine transporter imaging revealed that human patients with TBI show a similar reduction in dopamine transporter levels in the caudate compared with PD [162]. Complementary to this, TBI in rodents leads to reduced expression of tyrosine hydroxylase and dopamine transporter in the substantia nigra [157], as well as reduced dopamine transporter in the cortex ipsilateral to the injury site [163]. Conversely, other studies found an increase in dopamine [164] and tyrosine hydroxylase [165] levels after TBI. Further research is needed to clarify the effect of TBI on dopaminergic cells.

4.2.3. Synergistic Effect of Toxins with TBI in PD

Environmental exposure to toxins such as pesticides, solvents, metals, and pollutants increase the risk of PD [146] and other forms of neurodegenerative disease [147,148]. Remarkably, TBI increases the vulnerability of rodents to toxins known to be associated with PD [149,150]. For example, while neither a low-dose toxin or mild TBI alone were damaging, the combination of both together caused accumulation of α-synuclein and loss of tyrosine hydroxylase positive cells in rats [150].

5. TBI and AD

5.1. Epidemiology of TBI-Induced Increased Risk of AD

Of all forms of aging-related neurodegenerative disease, the association between TBI and AD is the most well-established. This relationship was described as early as 1939 in various case studies that described AD-like clinical and pathologic presentations in patients after a single TBI [166,167]. Since then, many epidemiological studies have assessed the relationship between TBI and AD, often with varying results due to the methodologic challenges discussed above. Meta-analyses of the published literature revealed an approximately 50% increased risk of AD in TBI patients (Table 2). Although one published meta-analysis did not find an association between TBI and AD, the authors noted that it was not properly powered to assess this association [168]. Interestingly, two large meta-analyses found a male-specific effect [13,169].
In addition to increasing the risk of developing AD, TBI also accelerates its age of onset. For example, a population-based study found that time to develop AD was 8 years faster than expected in individuals with a history of TBI [171], and a recent study found that a single remote TBI accelerates the onset of AD by 3–4 years [172]. In addition, a history of TBI is also associated with greater amyloid beta deposition and more severe cortical thinning in AD patients [172].

5.2. Potential Mechanisms of TBI-Induced Increased Risk of AD

Although epidemiological and clinical studies have uncovered a clear link between TBI and an elevated risk of AD, the mechanisms underpinning this phenomenon are not completely understood. Studies assessing the effect of TBI in AD mouse models found that TBI worsens AD pathology. Most of these studies used 3xTg and APP/PS1 mouse models, which display amyloid beta and tau pathology, and the controlled cortical impact or weight-drop model of TBI. These studies found that TBI increased AD-like pathologies, including amyloid beta [173,174,175,176,177], phosphorylated tau [173,178], neuronal loss [174,179], and neuroinflammation [178,179,180]. Few studies performed cognitive behavioral tests, and those that did found cognitive deficits after TBI in AD models in contextual fear conditioning [181], radial-arm water maze [174], and Morris water maze [177,182]. Unfortunately, most studies have applied TBI at older ages (6–12 months) after AD pathology has developed in these models, and only analyzed at short timepoints (1–14 days). By contrast, there are only two published studies that applied TBI before the onset of severe pathology in AD models [129,181]. In this section, we review the molecular changes accompanying TBI that may result in developing AD, including TBI-induced brain insulin resistance, amyloid beta pathology, and tau pathology.

5.2.1. TBI-Induced Brain Insulin Resistance in AD

Insulin, a peptide hormone secreted by the pancreas, crosses the BBB and engages insulin receptors in the brain. Indeed, emerging evidence shows that insulin plays an important role in brain metabolism, synaptic transmission, neuroinflammation, and vascular function. In addition, type 2 diabetes and central insulin resistance are known risk factors for AD [183]. Remarkably, there is growing evidence for pathologic insulin resistance in the brain in patients with AD, with increasing insulin resistance correlated with increased AD Braak stage [184]. Intranasal insulin administration also improves cognitive function in humans with mild cognitive impairment and AD [185,186].
Notably, TBI also induces brain insulin resistance in animal models [187] and human patients [188,189], which has been proposed to interfere with the ability of insulin to protect synapses against amyloid beta and tau oligomers [187,190]. Furthermore, brain insulin resistance in mice causes greater neurologic impairment after TBI [191]. In summary, TBI-induced insulin resistance may increase neuronal vulnerability to toxic amyloid beta and tau oligomers, increasing the risk of developing AD.

5.2.2. TBI-Induced Amyloid Beta Pathology in AD

It is well documented, in both human patients and mouse models, that TBI induces amyloid plaques similar to AD [173,176,192,193]. In fact, amyloid plaques are found in approximately one-third of acute brain injuries [194]. Furthermore, long-term survivors of a single TBI have significantly more amyloid plaques than age-matched control subjects [193]. Surprisingly, decreased amyloid pathology has been reported after controlled cortical impact TBI in mice that carry an amyloid precursor protein mutation associated with early onset AD [195,196]. Additionally, a recent study found no significant difference in amyloid beta levels in Vietnam war veterans with or without a TBI [197]. Nonetheless, most of the literature reports an association of TBI with amyloid plaque accumulation in the brain.
A potential mechanism for this increase in amyloid beta is axonal injury caused by TBI. TBI, especially rotational injuries, can injure neurons through stretching tissue. Axonal injury disrupts the cytoskeleton and axonal transport, creating disconnected axon terminals. Here, amyloid precursor protein accumulates in the axons and can be cleaved into amyloid beta by presenilin-1 and beta secretase, which are also found in pathological axonal swellings [198,199]. It has been proposed that amyloid beta is released when damaged axons fully deteriorate, allowing formation of amyloid plaques [200,201].
There are also many studies showing that TBI acutely increases expression of amyloid precursor protein in injured axons in mice [202] and humans [203]. Additionally, caspases are activated after TBI [204]. For example, caspase 3 increases APP processing via disruption of beta secretase trafficking and degradation. Interestingly, TBI elevates beta secretase levels and activity [205]. Therefore, elevated beta secretase activity results in increased levels of pathogenic amyloid beta. Furthermore, caspase 3 also cleaves APP after TBI, generating more amyloid beta [206].

5.2.3. TBI-Induced Tau Pathology in AD

Tau is a microtubule-binding protein essential for neuronal health, and abnormal accumulation of tau has been noted in axonal swellings in humans after TBI [31,32]. In pathologic states, including AD and TBI, tau undergoes post translational modifications (PTMs) that interfere with microtubule binding and increase its aberrant aggregation. These tau aggregates, termed neurofibrillary tangles (NFTs), are hyperphosphorylated and are a pathologic hallmark of AD. NFTs are found in approximately one-third of patients following moderate to severe TBI [193], and levels of brain phosphorylated tau correlate to TBI severity measured by the Glasgow Coma Scale [207].
In mice, TBI acutely increases levels of oligomeric and phosphorylated tau [208]. In the PS19 tauopathy mouse model, controlled cortical impact TBI also acutely increases phosphorylated tau pathology (AT8), which persists at least six months after injury [209]. Furthermore, in wild-type mice, TBI induces tau pathology that propagates and leads to synaptic loss and memory deficits [210]. While many studies have investigated tau PTMs after TBI in animal models (Table 3), these studies primarily rely on antibody-based PTM detection and focus primarily on phosphorylation. There is a need for a non-biased approach, such as mass spectrometry, to fully characterize the tau PTM landscape after TBI.
In addition to phosphorylation, tau is also acetylated in neurodegenerative disease. For example, AD brains show significant increases in tau acetylation at K274 and K281 [211,212]. In human brains, acetylated tau (ac-tau) is localized to thioflavin-s positive intracellular inclusions [213]. Ac-tau is also an early pathologic event in AD, beginning at the earliest Braak stages (one and two). Conversely, tau phosphorylation does not appear until Braak stages five and six [213,214,215]. Additionally, levels of ac-tau positively correlate with stage of disease and levels of tau accumulation, suggesting that tau acetylation is a driver of pathology [213]. The Gan group generated mice with acetylation mimics at two key pathologic residues, lysines 274 and 281. These mice, termed KQ mice, show significant tau pathology, impaired hippocampal synaptic plasticity, and memory deficits [216]. Notably, tau acetylation has also been linked to another risk factor for AD, female sex [217,218].
Table 3. Known tau PTMs induced by TBI in animal models.
Table 3. Known tau PTMs induced by TBI in animal models.
PTMResidueSpeciesModelTimepointReference
PhosphorylationS198RatFPIAcute[219]
Chronic[219]
S199RatCCIAcute[220,221]
Subacute[220]
S202MouseCCIAcute[222,223,224,225]
Chronic[210,224]
WDChronic[226,227]
BlastAcute[228]
Subacute[227,229,230]
RatCCIAcute[220]
Subacute[220]
Chronic[231]
FPIAcute[208,221]
Subacute[208,221]
Chronic[221,232]
T205MouseCCIAcute[222,223]
Chronic[210]
WDChronic[226]
BlastAcute[228,233]
Subacute[229,230]
RatCCIChronic[231]
FPIAcute[208,221]
Subacute[208,221]
Chronic[232]
S212MouseBlastAcute[229]
S214MouseWDChronic[226,227]
BlastAcute[229]
T231MouseCCIAcute[222,223]
Chronic[210]
WDAcute[226,227]
Subacute[226,227]
Chronic[226,227]
BlastAcute[227]
Subacute[8,9]
Chronic[227]
RatWDAcute[234,235]
Chronic[235]
FPIAcute[208,221]
Subacute[221]
Chronic[221]
S262MouseBlastAcute[228]
S396MouseCCIAcute[236]
Chronic[210]
WDChronic[227]
BlastAcute[228,229]
Subacute[229]
RatWDAcute[234,237]
BlastAcute[237]
S404MouseCCIAcute[238]
Chronic[210,238]
WDChronic[227]
BlastAcute[228]
S416MouseCCIAcute[222,223]
S422MouseCCIAcute[236]
AcetylationK274MouseCCIAcute[25]
Subacute[25]
BlastAcute[25]
Subacute[25]
Chronic[25]
K281MouseCCIAcute[25]
Subacute[25]
BlastAcute[25]
Subacute[25]
Chronic[25]
CCI = controlled cortical impact; WD = weight drop, FPI = fluid percussion injury, Blast = blast neurotrauma model, shock tube TBI, overpressure wave TBI, jet-flow overpressure multimodal TBI; Acute: <1 week; Subacute: 1 week–1 month; Chronic: >1 month.
We recently reported that tau is rapidly acetylated at these same sites after TBI as well, and that this acetylation occurs rapidly, persists chronically, and drives axonal degeneration [25]. Mechanistically, TBI induces GAPDH S-nitrosylation, which is the triggering pathological event that activates p300/CBP acetyltransferase to acetylate tau. S-nitrosylated GAPDH also inactivates SIRT1, one of the primary enzymes that removes acetyl groups from tau. Together, increased production and decreased clearance of ac-tau causes pathologic mislocalization of tau in the soma, which occurs following axonal initial segment degeneration, as established by the Rasband [239] and Gan laboratories [212]. This axon degeneration triggered by ac-tau also leads to neurodegeneration and cognitive impairment after TBI, and pharmacologically reducing ac-tau after TBI via multiple mechanisms rescued neurodegeneration and cognitive impairment [25]. Notably, ac-tau in the brain is more elevated in people with AD and a history of TBI, compared to people with AD and no history of TBI and healthy controls [25]. Therefore, ac-tau is an early and persisting pathologic event after TBI that may contribute to acceleration of AD. Importantly, blood levels of ac-tau directly correlate with brain levels, suggesting that ac-tau could be the first blood-based biomarker of neurodegeneration that directly reflects the abundance of a therapeutic target in the brain after TBI [25]. Furthermore, the rapid and persistent accumulation of ac-tau after TBI mimics the early disease state of AD, such that AD could be accelerated by TBI through this mechanism.
In summary, there is substantial evidence that TBI increases the risk and accelerates the onset of AD. Mechanistic studies suggest that multiple pathologic processes, including insulin resistance and pathology related to amyloid beta and tau, mediate this relationship. Further research should test whether therapeutically targeting these mechanisms will prevent the accelerated onset of AD after TBI in rodent models.

6. Conclusions

Despite methodologic challenges, the current body of research overwhelmingly shows that TBI is associated with a significantly increased risk of age-related neurodegenerative disease. The National Institutes of Health 2022 triennial Alzheimer’s Disease and Related Dementias (ADRD) Summit to inform the national research agenda underscored this problem by including TBI [240]. Specifically, four priority areas were determined that together conceptualize TBI as a major contributor to dementia. These recommendations included (1) promoting interdisciplinary approaches to accelerate clinically meaningful research in this area; (2) characterizing clinical and biological phenotypes of neurodegenerative disease after TBI across diverse populations and histories of TBI, including validation of multimodal biomarkers; (3) establishing and strengthening infrastructure to support standardized methods with common data elements for antemortem and postmortem clinical and neuropathological characterization; and (4) extending basic and translational research to determine the mechanistic pathways and clinical manifestations of post-TBI neurodegenerative disease.
It is well established that TBI initiates multiple pathologic processes that drive increased risk of neurodegenerative disease, including oxidative stress, impaired proteostasis, and both acute and chronic neuroinflammation, mechanisms that are common to varying degrees across all forms of neurodegenerative disease (Figure 2). Additional research testing whether therapeutically targeting these common pathologic changes after TBI will block the increased risk of aging-related neurodegenerative disease would greatly advance our knowledge of this problem and point towards potential neuroprotective therapies. Regarding specific forms of aging-related neurodegenerative disease, there is mixed evidence supporting the relationship between TBI and the increased risk of ALS and FTD. Further research should focus on large prospective epidemiology studies to assess this potential relationship. There is highly compelling evidence, however, that TBI increases the risk of developing PD and AD.
With respect to PD, mechanistic studies provide evidence that TBI dysregulates α-synuclein, harms dopaminergic neurons, and synergizes with environmental toxins to accelerate the disease (Figure 2). Further research is needed to more rigorously establish potential mechanisms by which TBI increases the risk of PD, from which therapeutics can be designed. The largest body of research supports the association between TBI and AD. Preliminary mechanistic studies suggest that central insulin resistance may mediate this relationship, and there is also strong evidence that TBI initiates amyloid and tau pathology related to AD. Further research should investigate whether targeting these pathologic mechanisms initiated by TBI can prevent the increased risk and accelerated onset of AD in animal models.
Overall, to understand the complex interplay between TBI and neurodegenerative disease, it is imperative for the field to generate and validate new animal models that combine TBI and aging-related neurodegenerative disease, in order to ultimately develop new neuroprotective therapies for patients.

Author Contributions

S.B., B.D.P. and A.A.P. conceptualized, wrote, and edited this review. All authors have read and agreed to the published version of the manuscript.

Funding

AAP was supported as the Case Western Reserve University Rebecca E. Barchas, MD, Professor in Translational Psychiatry and as the University Hospitals Morley-Mather Chair in Neuropsychiatry. AAP also acknowledges support from the Brockman Foundation, Department of Veterans Affairs Merit Award I01BX005976, NIH/NIA RO1AG065240 and NIH/NIA R01AG074346, NIH/NIGMS RM1 GM142002, NIH/NIA RO1AG066707, NIH/NIA 1 U01 AG073323, the Elizabeth Ring Mather & William Gwinn Mather Fund, S. Livingston Samuel Mather Trust, G.R. Lincoln Family Foundation, Wick Foundation, Gordon & Evie Safran, the Leonard Krieger Fund of the Cleveland Foundation, the Maxine and Lester Stoller Parkinson’s Research Fund, the Louis Stokes VA Medical Center resources and facilities, and the Translational Therapeutics Core of the Cleveland Alzheimer’s Disease Research Center (NIH/NIA 1 P30 AGO62428-01). AAP and BDP were supported by the American Heart Association and Paul Allen Foundation Initiative in Brain Health and Cognitive Impairment (19PABH134580006) and by NIH/NIA 1R01AG071512. BDP was supported by NIH NIDA grant P50 DA044123, NIH 1R21AG073684-01, and the Catalyst Award from Johns Hopkins University. SB was supported by The Alzheimer’s Disease Translational Data Science Training Program NIH T32 AG071474 and by Case Western Medical Scientist Training program NIH T32 GM007250.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge BioRender for the use of icons in figures.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dewan, M.C.; Rattani, A.; Gupta, S.; Baticulon, R.E.; Hung, Y.C.; Punchak, M.; Agrawal, A.; Adeleye, A.O.; Shrime, M.G.; Rubiano, A.M.; et al. Estimating the Global Incidence of Traumatic Brain Injury. J. Neurosurg. 2019, 130, 1080–1097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Stein, S.C.; Georgoff, P.; Meghan, S.; Mizra, K.; Sonnad, S.S. 150 Years of Treating Severe Traumatic Brain Injury: A Systematic Review of Progress in Mortality. J. Neurotrauma 2010, 27, 1343–1353. [Google Scholar] [CrossRef]
  3. Corrigan, J.D.; Yang, J.; Singichetti, B.; Manchester, K.; Bogner, J. Lifetime Prevalence of Traumatic Brain Injury with Loss of Consciousness. Inj. Prev. 2018, 24, 396–404. [Google Scholar] [CrossRef]
  4. Ma, V.Y.; Chan, L.; Carruthers, K.J. The Incidence, Prevalence, Costs and Impact on Disability of Common Conditions Requiring Rehabilitation in the US: Stroke, Spinal Cord Injury, Traumatic Brain Injury, Multiple Sclerosis, Osteoarthritis, Rheumatoid Arthritis, Limb Loss, and Back Pain. Arch. Phys. Med. Rehabil. 2014, 95, 986–995. [Google Scholar] [CrossRef] [Green Version]
  5. Corrigan, J.D.; Hammond, F.M. Traumatic Brain Injury as a Chronic Health Condition. Arch. Phys. Med. Rehabil. 2013, 94, 1199–1201. [Google Scholar] [CrossRef]
  6. Gaetz, M. The Neurophysiology of Brain Injury. Clin. Neurophysiol. 2004, 115, 4–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Johnson, V.E.; Meaney, D.F.; Cullen, D.K.; Smith, D.H. Animal Models of Traumatic Brain Injury. Handb. Clin. Neurol. 2015, 127, 115–128. [Google Scholar] [CrossRef] [Green Version]
  8. Werner, C.; Engelhard, K. Pathophysiology of Traumatic Brain Injury. BJA Br. J. Anaesth. 2007, 99, 4–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Bramlett, H.M.; Dietrich, W.D. Long-Term Consequences of Traumatic Brain Injury: Current Status of Potential Mechanisms of Injury and Neurological Outcomes. J. Neurotrauma 2015, 32, 1834–1848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Livingston, G.; Huntley, J.; Sommerlad, A.; Ames, D.; Ballard, C.; Banerjee, S.; Brayne, C.; Burns, A.; Cohen-Mansfield, J.; Cooper, C.; et al. Dementia Prevention, Intervention, and Care: 2020 Report of the Lancet Commission. Lancet 2020, 396, 413. [Google Scholar] [CrossRef] [PubMed]
  11. Fann, J.R.; Ribe, A.R.; Pedersen, H.S.; Fenger-Grøn, M.; Christensen, J.; Benros, M.E.; Vestergaard, M. Long-Term Risk of Dementia among People with Traumatic Brain Injury in Denmark: A Population-Based Observational Cohort Study. Lancet Psychiatry 2018, 5, 424–431. [Google Scholar] [CrossRef] [PubMed]
  12. Barnes, D.E.; Kaup, A.; Kirby, K.A.; Byers, A.L.; Diaz-Arrastia, R.; Yaffe, K. Traumatic Brain Injury and Risk of Dementia in Older Veterans. Neurology 2014, 83, 312–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Li, Y.; Li, Y.; Li, X.; Zhang, S.; Zhao, J.; Zhu, X.; Tian, G. Head Injury as a Risk Factor for Dementia and Alzheimer’s Disease: A Systematic Review and Meta-Analysis of 32 Observational Studies. PLoS ONE 2017, 12, e0169650. [Google Scholar] [CrossRef] [Green Version]
  14. Shively, S.; Scher, A.I.; Perl, D.P.; Diaz-Arrastia, R. Dementia Resulting from Traumatic Brain Injury: What Is the Pathology? Arch. Neurol. 2012, 69, 1245–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Hicks, A.J.; James, A.C.; Spitz, G.; Ponsford, J.L. Traumatic Brain Injury as a Risk Factor for Dementia and Alzheimer Disease: Critical Review of Study Methodologies. J. Neurotrauma 2019, 36, 3191–3219. [Google Scholar] [CrossRef]
  16. Lye, T.C.; Shores, E.A. Traumatic Brain Injury as a Risk Factor for Alzheimer’s Disease: A Review. Neuropsychol. Rev. 2000, 10, 1376–1381. [Google Scholar] [CrossRef]
  17. Hanrahan, J.G.; Burford, C.; Nagappan, P.; Adegboyega, G.; Rajkumar, S.; Kolias, A.; Helmy, A.; Hutchinson, P.J. Is Dementia More Likely Following Traumatic Brain Injury? A Systematic Review. J. Neurol. 2023; ahead-of-print. [Google Scholar] [CrossRef]
  18. Flierl, M.A.; Stahel, P.F.; Beauchamp, K.M.; Morgan, S.J.; Smith, W.R.; Shohami, E. Mouse Closed Head Injury Model Induced by a Weight-Drop Device. Nat. Protoc. 2009, 4, 1328–1337. [Google Scholar] [CrossRef]
  19. Osier, N.; Dixon, C.E. The Controlled Cortical Impact Model of Experimental Brain Trauma: Overview, Research Applications, and Protocol. Methods Mol. Biol. 2016, 1462, 177–192. [Google Scholar] [CrossRef] [Green Version]
  20. Kabadi, S.v.; Hilton, G.D.; Stoica, B.A.; Zapple, D.N.; Faden, A.I. Fluid-Percussion-Induced Traumatic Brain Injury Model in Rats. Nat. Protoc. 2010, 5, 1552–1563. [Google Scholar] [CrossRef] [Green Version]
  21. Alder, J.; Fujioka, W.; Lifshitz, J.; Crockett, D.P.; Thakker-Varia, S. Lateral Fluid Percussion: Model of Traumatic Brain Injury in Mice. J. Vis. Exp. 2011, 54, e3063. [Google Scholar] [CrossRef] [Green Version]
  22. Shah, A.; Stemper, B.D.; Pintar, F.A. Development and Characterization of an Open-Ended Shock Tube for the Study of Blast Mtbi. Biomed. Sci. Instrum. 2012, 48, 393–400. [Google Scholar]
  23. Ma, X.; Aravind, A.; Pfister, B.J.; Chandra, N.; Haorah, J. Animal Models of Traumatic Brain Injury and Assessment of Injury Severity. Mol. Neurobiol. 2019, 56, 5332–5345. [Google Scholar] [CrossRef]
  24. Shin, M.-K.; Vázquez-Rosa, E.; Cintrón-Pérez, C.J.; Riegel, W.A.; Harper, M.M.; Ritzel, D.; Pieper, A.A. Characterization of the Jet-Flow Overpressure Model of Traumatic Brain Injury in Mice. Neurotrauma Rep. 2021, 2, 1–13. [Google Scholar] [CrossRef]
  25. Shin, M.K.; Vázquez-Rosa, E.; Koh, Y.; Dhar, M.; Chaubey, K.; Cintrón-Pérez, C.J.; Barker, S.; Miller, E.; Franke, K.; Noterman, M.F.; et al. Reducing Acetylated Tau Is Neuroprotective in Brain Injury. Cell 2021, 184, 2715–2732.e23. [Google Scholar] [CrossRef] [PubMed]
  26. Vázquez-Rosa, E.; Shin, M.K.; Dhar, M.; Chaubey, K.; Cintrón-Pérez, C.J.; Tang, X.; Liao, X.; Miller, E.; Koh, Y.; Barker, S.; et al. P7C3-A20 Treatment One Year after TBI in Mice Repairs the Blood-Brain Barrier, Arrests Chronic Neurodegeneration, and Restores Cognition. Proc. Natl. Acad. Sci. USA 2020, 117, 27667–27675. [Google Scholar] [CrossRef] [PubMed]
  27. Harper, M.M.; Rudd, D.; Meyer, K.J.; Kanthasamy, A.G.; Anantharam, V.; Pieper, A.A.; Vázquez-Rosa, E.; Shin, M.K.; Chaubey, K.; Koh, Y.; et al. Identification of Chronic Brain Protein Changes and Protein Targets of Serum Auto-Antibodies after Blast-Mediated Traumatic Brain Injury. Heliyon 2020, 6, e03374. [Google Scholar] [CrossRef] [Green Version]
  28. Yin, T.C.; Britt, J.K.; de Jesús-Cortés, H.; Lu, Y.; Genova, R.M.; Khan, M.Z.; Voorhees, J.R.; Shao, J.; Katzman, A.C.; Huntington, P.J.; et al. P7C3 Neuroprotective Chemicals Block Axonal Degeneration and Preserve Function after Traumatic Brain Injury. Cell Rep. 2014, 8, 1731–1740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Dutca, L.M.; Stasheff, S.F.; Hedberg-Buenz, A.; Rudd, D.S.; Batra, N.; Blodi, F.R.; Yorek, M.S.; Yin, T.; Shankar, M.; Herlein, J.A.; et al. Early Detection of Subclinical Visual Damage after Blast-Mediated TBI Enables Prevention of Chronic Visual Deficit by Treatment with P7C3-S243. Investig. Ophthalmol. Vis. Sci. 2014, 55, 8330–8341. [Google Scholar] [CrossRef] [PubMed]
  30. Yin, T.C.; Voorhees, J.R.; Genova, R.M.; Davis, K.C.; Madison, A.M.; Britt, J.K.; Cintrón-Pérez, C.J.; McDaniel, L.; Harper, M.M.; Pieper, A.A. Acute Axonal Degeneration Drives Development of Cognitive, Motor, and Visual Deficits after Blast-Mediated Traumatic Brain Injury in Mice. eNeuro 2016, 3, ENEURO.0220-16.2016. [Google Scholar] [CrossRef] [Green Version]
  31. Vázquez-Rosa, E.; Watson, M.R.; Sahn, J.J.; Hodges, T.R.; Schroeder, R.E.; Cintrón-Pérez, C.J.; Shin, M.K.; Yin, T.C.; Emery, J.L.; Martin, S.F.; et al. Neuroprotective Efficacy of a Sigma 2 Receptor/TMEM97 Modulator (DKR-1677) after Traumatic Brain Injury. ACS Chem. Neurosci. 2019, 10, 1595–1602. [Google Scholar] [CrossRef] [PubMed]
  32. Chapman, P.F.; Falinska, A.M.; Knevett, S.G.; Ramsay, M.F. Genes, Models and Alzheimer’s Disease. Trends Genet. 2001, 17, 254–261. [Google Scholar] [CrossRef] [PubMed]
  33. Kim, G.; Gautier, O.; Tassoni-Tsuchida, E.; Ma, X.R.; Gitler, A.D. ALS Genetics: Gains, Losses, and Implications for Future Therapies. Neuron 2020, 108, 822–842. [Google Scholar] [CrossRef]
  34. Tran, J.; Anastacio, H.; Bardy, C. Genetic Predispositions of Parkinson’s Disease Revealed in Patient-Derived Brain Cells. Park. Dis. 2020, 6, 8. [Google Scholar] [CrossRef] [Green Version]
  35. Iacono, D.; Raiciulescu, S.; Olsen, C.; Perl, D.P. Traumatic Brain Injury Exposure Lowers Age of Cognitive Decline in AD and Non-AD Conditions. Front. Neurol. 2021, 12, 573401. [Google Scholar] [CrossRef]
  36. Kapasi, A.; DeCarli, C.; Schneider, J.A. Impact of Multiple Pathologies on the Threshold for Clinically Overt Dementia. Acta Neuropathol. 2017, 134, 171–186. [Google Scholar] [CrossRef]
  37. Bélanger, M.; Allaman, I.; Magistretti, P.J. Brain Energy Metabolism: Focus on Astrocyte-Neuron Metabolic Cooperation. Cell Metab. 2011, 14, 724–738. [Google Scholar] [CrossRef] [Green Version]
  38. Popa-Wagner, A.; Mitran, S.; Sivanesan, S.; Chang, E.; Buga, A.M. ROS and Brain Diseases: The Good, the Bad, and the Ugly. Oxid. Med. Cell. Longev. 2013, 2013, 963520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Barnham, K.J.; Masters, C.L.; Bush, A.I. Neurodegenerative Diseases and Oxidative Stress. Nat. Rev. Drug 2004, 3, 205–214. [Google Scholar] [CrossRef]
  40. Gandhi, S.; Abramov, A.Y. Mechanism of Oxidative Stress in Neurodegeneration. Oxid. Med. Cell. Longev. 2012, 2012, 428010. [Google Scholar] [CrossRef] [Green Version]
  41. Gadoth, N.; Göbel, H.H. Oxidative Stress and Free Radical Damage in Neurology. J. Pharmacol. Exp. Ther. 2017, 360, 201–205. [Google Scholar]
  42. Sbodio, J.I.; Snyder, S.H.; Paul, B.D. Redox Mechanisms in Neurodegeneration: From Disease Outcomes to Therapeutic Opportunities. Antioxid. Redox Signal. 2019, 30, 1450–1499. [Google Scholar] [CrossRef]
  43. Paul, B.D.; Snyder, S.H. The Unusual Amino Acid L-Ergothioneine Is a Physiologic Cytoprotectant. Cell Death Differ. 2010, 17, 1134–1140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Paul, B.D. Ergothioneine: A Stress Vitamin with Antiaging, Vascular, and Neuroprotective Roles? Antioxid. Redox Signal. 2022, 36, 1306–1317. [Google Scholar] [CrossRef] [PubMed]
  45. Vasavda, C.; Kothari, R.; Malla, A.P.; Tokhunts, R.; Lin, A.; Ji, M.; Ricco, C.; Xu, R.; Saavedra, H.G.; Sbodio, J.I.; et al. Bilirubin Links Heme Metabolism to Neuroprotection by Scavenging Superoxide. Cell Chem. Biol. 2019, 26, 1450–1460.e7. [Google Scholar] [CrossRef] [PubMed]
  46. Sedlak, T.W.; Paul, B.D.; Parker, G.M.; Hester, L.D.; Snowman, A.M.; Taniguchi, Y.; Kamiya, A.; Snyder, S.H.; Sawa, A. The Glutathione Cycle Shapes Synaptic Glutamate Activity. Proc. Natl. Acad. Sci. USA 2019, 116, 2701–2706. [Google Scholar] [CrossRef] [Green Version]
  47. Sedlak, T.W.; Saleh, M.; Higginson, D.S.; Paul, B.D.; Juluri, K.R.; Snyder, S.H. Bilirubin and Glutathione Have Complementary Antioxidant and Cytoprotective Roles. Proc. Natl. Acad. Sci. USA 2009, 106, 5171–5176. [Google Scholar] [CrossRef] [Green Version]
  48. Janáky, R.; Cruz-Aguado, R.; Oja, S.S.; Shaw, C.A. Glutathione in the Nervous System: Roles in Neural Function and Health and Implications for Neurological Disease. In Handbook of Neurochemistry and Molecular Neurobiology: Amino Acids and Peptides in the Nervous System; Springer: Cham, Switzerland, 2007; pp. 347–399. [Google Scholar] [CrossRef]
  49. Ballaz, S.J.; Rebec, G.v. Neurobiology of Vitamin C: Expanding the Focus from Antioxidant to Endogenous Neuromodulator. Pharmacol. Res. 2019, 146, 104321. [Google Scholar] [CrossRef]
  50. Paul, B.D. Neuroprotective Roles of the Reverse Transsulfuration Pathway in Alzheimer’s Disease. Front. Aging Neurosci. 2021, 13, 659402. [Google Scholar] [CrossRef] [PubMed]
  51. Paul, B.D.; Snyder, S.H.; Kashfi, K. Effects of Hydrogen Sulfide on Mitochondrial Function and Cellular Bioenergetics. Redox Biol. 2021, 38, 101772. [Google Scholar] [CrossRef]
  52. Paul, B.D.; Snyder, S.H. H2S: A Novel Gasotransmitter That Signals by Sulfhydration. Trends Biochem. Sci. 2015, 40, 687–700. [Google Scholar] [CrossRef] [Green Version]
  53. Giovinazzo, D.; Bursac, B.; Sbodio, J.I.; Nalluru, S.; Vignane, T.; Snowman, A.M.; Albacarys, L.M.; Sedlak, T.W.; Torregrossa, R.; Whiteman, M.; et al. Hydrogen Sulfide Is Neuroprotective in Alzheimer’s Disease by Sulfhydrating GSK3β and Inhibiting Tau Hyperphosphorylation. Proc. Natl. Acad. Sci. USA 2021, 118, e2017225118. [Google Scholar] [CrossRef]
  54. Paul, B.D.; Snyder, S.H. Modes of Physiologic H2S Signaling in the Brain and Peripheral Tissues. Antioxid. Redox Signal. 2015, 22, 411–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Paul, B.D.; Snyder, S.H. H₂S Signalling through Protein Sulfhydration and Beyond. Nat. Rev. Mol. Cell Biol. 2012, 13, 499–507. [Google Scholar] [CrossRef]
  56. Solaroglu, I.; Okutan, O.; Kaptanoglu, E.; Beskonakli, E.; Kilinc, K. Increased Xanthine Oxidase Activity after Traumatic Brain Injury in Rats. J. Clin. Neurosci. 2005, 12, 273–275. [Google Scholar] [CrossRef] [PubMed]
  57. Ansari, M.A.; Roberts, K.N.; Scheff, S.W. Oxidative Stress and Modification of Synaptic Proteins in Hippocampus after Traumatic Brain Injury. Free Radic. Biol. Med. 2008, 45, 443–452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Hall, E.D.; Detloff, M.R.; Johnson, K.; Kupina, N.C. Peroxynitrite-Mediated Protein Nitration and Lipid Peroxidation in a Mouse Model of Traumatic Brain Injury. J. Neurotrauma 2004, 21, 9–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Wu, A.; Ying, Z.; Gomez-Pinilla, F. Dietary Curcumin Counteracts the Outcome of Traumatic Brain Injury on Oxidative Stress, Synaptic Plasticity, and Cognition. Exp. Neurol. 2006, 197, 309–317. [Google Scholar] [CrossRef]
  60. Darwish, R.S.; Amiridze, N.; Aarabi, B. Nitrotyrosine as an Oxidative Stress Marker: Evidence for Involvement in Neurologic Outcome in Human Traumatic Brain Injury. J. Trauma-Inj. Infect. Crit. Care 2007, 63, 439–442. [Google Scholar] [CrossRef]
  61. Cheng, H.; Wang, N.; Ma, X.; Wang, P.; Dong, W.; Chen, Z.; Wu, M.; Wang, Z.; Wang, L.; Guan, D.; et al. Spatial-Temporal Changes of Iron Deposition and Iron Metabolism after Traumatic Brain Injury in Mice. Front. Mol. Neurosci. 2022, 15, 949573. [Google Scholar] [CrossRef] [PubMed]
  62. Robicsek, S.A.; Bhattacharya, A.; Rabai, F.; Shukla, K.; Doré, S. Blood-Related Toxicity after Traumatic Brain Injury: Potential Targets for Neuroprotection. Mol. Neurobiol. 2020, 57, 159–178. [Google Scholar] [CrossRef] [PubMed]
  63. Nisenbaum, E.J.; Novikov, D.S.; Lui, Y.W. The Presence and Role of Iron in Mild Traumatic Brain Injury: An Imaging Perspective. J. Neurotrauma 2014, 31, 301–307. [Google Scholar] [CrossRef] [PubMed]
  64. Stockwell, B.R.; Friedmann Angeli, J.P.; Bayir, H.; Bush, A.I.; Conrad, M.; Dixon, S.J.; Fulda, S.; Gascón, S.; Hatzios, S.K.; Kagan, V.E.; et al. Ferroptosis: A Regulated Cell Death Nexus Linking Metabolism, Redox Biology, and Disease. Cell 2017, 171, 273–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Fang, J.; Yuan, Q.; Du, Z.; Fei, M.; Zhang, Q.; Yang, L.; Wang, M.; Yang, W.; Yu, J.; Wu, G.; et al. Ferroptosis in Brain Microvascular Endothelial Cells Mediates Blood-Brain Barrier Disruption after Traumatic Brain Injury. Biochem. Biophys. Res. Commun. 2022, 619, 34–41. [Google Scholar] [CrossRef]
  66. Hu, X.; Xu, Y.; Xu, H.; Jin, C.; Zhang, H.; Su, H.; Li, Y.; Zhou, K.; Ni, W. Progress in Understanding Ferroptosis and Its Targeting for Therapeutic Benefits in Traumatic Brain and Spinal Cord Injuries. Front. Cell Dev. Biol. 2021, 9, 705786. [Google Scholar] [CrossRef]
  67. Fujii, J.; Homma, T.; Kobayashi, S. Ferroptosis Caused by Cysteine Insufficiency and Oxidative Insult. Free Radic. Res. 2019, 54, 969–980. [Google Scholar] [CrossRef]
  68. Tyurin, V.A.; Tyurina, Y.Y.; Borisenko, G.G.; Sokolova, T.v.; Ritov, V.B.; Quinn, P.J.; Rose, M.; Kochanek, P.; Graham, S.H.; Kagan, V.E. Oxidative Stress Following Traumatic Brain Injury in Rats. J. Neurochem. 2000, 75, 2178–2189. [Google Scholar] [CrossRef] [Green Version]
  69. Bayir, H.; Kagan, V.E.; Tyurina, Y.Y.; Tyurin, V.; Ruppel, R.A.; Adelson, D.D.; Graham, S.H.; Janesko, K.; Clark, R.S.B.; Kochanek, P.M. Assessment of Antioxidant Reserves and Oxidative Stress in Cerebrospinal Fluid after Severe Traumatic Brain Injury in Infants and Children. Pediatr. Res. 2002, 51, 571–578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Yamamoto, M.; Kensler, T.W.; Motohashi, H. The KEAP1-NRF2 System: A Thiol-Based Sensor-Effector Apparatus for Maintaining Redox Homeostasis. Physiol. Rev. 2018, 98, 1169–1203. [Google Scholar] [CrossRef] [Green Version]
  71. Ishii, T.; Itoh, K.; Takahashi, S.; Sato, H.; Yanagawa, T.; Katoh, Y.; Bannai, S.; Yamamoto, M. Transcription Factor Nrf2 Coordinately Regulates a Group of Oxidative Stress-Inducible Genes in Macrophages. J. Biol. Chem. 2000, 275, 16023–16029. [Google Scholar] [CrossRef] [Green Version]
  72. Yan, W.; Wang, H.D.; Hu, Z.G.; Wang, Q.F.; Yin, H.X. Activation of Nrf2–ARE Pathway in Brain after Traumatic Brain Injury. Neurosci. Lett. 2008, 431, 150–154. [Google Scholar] [CrossRef]
  73. Zhang, J.M.; Hong, Y.; Yan, W.; Chen, S.; Sun, C.R. The Role of Nrf2 Signaling in the Regulation of Antioxidants and Detoxifying Enzymes after Traumatic Brain Injury in Rats and Mice. Acta Pharmacol. Sin. 2010, 31, 1421–1430. [Google Scholar] [CrossRef] [Green Version]
  74. Vasavda, C.; Xu, R.; Liew, J.; Kothari, R.; Dhindsa, R.S.; Semenza, E.R.; Paul, B.D.; Green, D.P.; Sabbagh, M.F.; Shin, J.Y.; et al. Identification of the NRF2 Transcriptional Network as a Therapeutic Target for Trigeminal Neuropathic Pain. Sci. Adv. 2022, 8, eabo5633. [Google Scholar] [CrossRef] [PubMed]
  75. Hartl, F.U.; Bracher, A.; Hayer-Hartl, M. Molecular Chaperones in Protein Folding and Proteostasis. Nature 2011, 475, 324–332. [Google Scholar] [CrossRef]
  76. Lai, Y.; Kochanek, P.M.; Adelson, P.D.; Janesko, K.; Ruppel, R.A.; Clark, R.S.B. Induction of the Stress Response after Inflicted and Non-Inflicted Traumatic Brain Injury in Infants and Children. J. Neurotrauma 2004, 21, 229–237. [Google Scholar] [CrossRef]
  77. Michael, D.B.; Byers, D.M.; Irwin, L.N. Gene Expression Following Traumatic Brain Injury in Humans: Analysis by Microarray. J. Clin. Neurosci. 2005, 12, 284–290. [Google Scholar] [CrossRef]
  78. Truettner, J.S.; Hu, B.; Alonso, O.F.; Bramlett, H.M.; Kokame, K.; Dietrich, W.D. Subcellular Stress Response after Traumatic Brain Injury. J. Neurotrauma 2007, 24, 599–612. [Google Scholar] [CrossRef]
  79. Seidberg, N.A.; Clark, R.S.B.; Zhang, X.; Lai, Y.; Chen, M.; Graham, S.H.; Kochanek, P.M.; Watkins, S.C.; Marion, D.W. Alterations in Inducible 72-KDa Heat Shock Protein and the Chaperone Cofactor BAG-1 in Human Brain after Head Injury. J. Neurochem. 2003, 84, 514–521. [Google Scholar] [CrossRef] [PubMed]
  80. Kim, J.Y.; Kim, N.; Zheng, Z.; Lee, J.E.; Yenari, M.A. The 70 KDa Heat Shock Protein Protects against Experimental Traumatic Brain Injury. Neurobiol. Dis. 2013, 58, 289–295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Kim, N.; Kim, J.Y.; Yenari, M.A. Pharmacological Induction of the 70-KDa Heat Shock Protein Protects against Brain Injury. Neuroscience 2015, 284, 912–919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Petrov, T.; Underwood, B.D.; Braun, B.; Alousi, S.S.; Rafols, S.S. Upregulation of INOS Expression and Phosphorylation of EIF-2α Are Paralleled by Suppression of Protein Synthesis in Rat Hypothalamus in a Closed Head Trauma Model. J. Neurotrauma 2004, 18, 799–812. [Google Scholar] [CrossRef]
  83. Hylin, M.J.; Holden, R.C.; Smith, A.C.; Logsdon, A.F.; Qaiser, R.; Lucke-Wold, B.P. Juvenile Traumatic Brain Injury Results in Cognitive Deficits Associated with Impaired Endoplasmic Reticulum Stress and Early Tauopathy. Dev. Neurosci. 2018, 40, 175–188. [Google Scholar] [CrossRef] [PubMed]
  84. Sen, T.; Gupta, R.; Kaiser, H.; Sen, N. Activation of PERK Elicits Memory Impairment through Inactivation of CREB and Downregulation of PSD95 After Traumatic Brain Injury. J. Neurosci. 2017, 37, 5900–5911. [Google Scholar] [CrossRef]
  85. Logsdon, A.F.; Turner, R.C.; Lucke-Wold, B.P.; Robson, M.J.; Naser, Z.J.; Smith, K.E.; Matsumoto, R.R.; Huber, J.D.; Rosen, C.L. Altering Endoplasmic Reticulum Stress in a Model of Blast-Induced Traumatic Brain Injury Controls Cellular Fate and Ameliorates Neuropsychiatric Symptoms. Front. Cell. Neurosci. 2014, 8, 421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Tan, H.P.; Guo, Q.; Hua, G.; Chen, J.X.; Liang, J.C. Inhibition of Endoplasmic Reticulum Stress Alleviates Secondary Injury after Traumatic Brain Injury. Neural Regen. Res. 2018, 13, 827. [Google Scholar] [CrossRef] [PubMed]
  87. Staal, J.A.; Dickson, T.C.; Chung, R.S.; Vickers, J.C. Disruption of the Ubiquitin Proteasome System Following Axonal Stretch Injury Accelerates Progression to Secondary Axotomy. J. Neurotrauma 2009, 26, 781–788. [Google Scholar] [CrossRef]
  88. Yao, X.; Liu, J.; McCabe, J.T. Alterations of Cerebral Cortex and Hippocampal Proteasome Subunit Expression and Function in a Traumatic Brain Injury Rat Model. J. Neurochem. 2008, 104, 353–363. [Google Scholar] [CrossRef]
  89. Sakai, K.; Fukuda, T.; Iwadate, K. Immunohistochemical Analysis of the Ubiquitin Proteasome System and Autophagy Lysosome System Induced after Traumatic Intracranial Injury: Association with Time between the Injury and Death. Am. J. Forensic Med. Pathol. 2014, 35, 38–44. [Google Scholar] [CrossRef]
  90. Clark, R.S.B.; Bayir, H.; Chu, C.T.; Alber, S.M.; Kochanek, P.M.; Watkins, S.C. Autophagy Is Increased in Mice after Traumatic Brain Injury and Is Detectable in Human Brain after Trauma and Critical Illness. Autophagy 2007, 4, 88–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Liu, C.L.; Chen, S.; Dietrich, D.; Hu, B.R. Rapid Communication Changes in Autophagy after Traumatic Brain Injury. J. Cereb. Blood Flow Metab. 2008, 28, 674–683. [Google Scholar] [CrossRef]
  92. Tansey, M.G.; Wallings, R.L.; Houser, M.C.; Herrick, M.K.; Keating, C.E.; Joers, V. Inflammation and Immune Dysfunction in Parkinson Disease. Nat. Rev. Immunol. 2022, 22, 657–673. [Google Scholar] [CrossRef]
  93. Leng, F.; Edison, P. Neuroinflammation and Microglial Activation in Alzheimer Disease: Where Do We Go from Here? Nat. Rev. Neurol. 2021, 17, 157–172. [Google Scholar] [CrossRef]
  94. Gilhus, N.E.; Deuschl, G. Neuroinflammation—A Common Thread in Neurological Disorders. Nat. Rev. Neurol. 2019, 15, 429–430. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, Q.; Zheng, J.; Pettersson, S.; Reynolds, R.; Tan, E.K. The Link between Neuroinflammation and the Neurovascular Unit in Synucleinopathies. Sci. Adv. 2023, 9, eabq1141. [Google Scholar] [CrossRef] [PubMed]
  96. Ransohoff, R.M. How Neuroinflammation Contributes to Neurodegeneration. Science 2016, 353, 777–783. [Google Scholar] [CrossRef] [PubMed]
  97. Simon, D.W.; McGeachy, M.J.; Baylr, H.; Clark, R.S.B.; Loane, D.J.; Kochanek, P.M. The Far-Reaching Scope of Neuroinflammation after Traumatic Brain Injury. Nat. Rev. Neurol. 2017, 13, 171–191. [Google Scholar] [CrossRef] [Green Version]
  98. Lu, Y.; Jarrahi, A.; Moore, N.; Bartoli, M.; Brann, D.W.; Baban, B.; Dhandapani, K.M. Inflammaging, Cellular Senescence, and Cognitive Aging after Traumatic Brain Injury. Neurobiol. Dis. 2023, 180, 106090. [Google Scholar] [CrossRef] [PubMed]
  99. Newell, E.A.; Todd, B.P.; Mahoney, J.; Pieper, A.A.; Ferguson, P.J.; Bassuk, A.G. Combined Blockade of Interleukin-1α and -1β Signaling Protects Mice from Cognitive Dysfunction after Traumatic Brain Injury. eNeuro 2018, 5, e0385-17.2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Burda, J.E.; Bernstein, A.M.; Sofroniew, M.V. Astrocyte Roles in Traumatic Brain Injury. Exp. Neurol. 2016, 275, 305–315. [Google Scholar] [CrossRef] [Green Version]
  101. Paul, B.D.; Snyder, S.H.; Bohr, V.A. Signaling by CGAS–STING in Neurodegeneration, Neuroinflammation, and Aging. Trends Neurosci. 2021, 44, 83–96. [Google Scholar] [CrossRef] [PubMed]
  102. Li, Q.; Cao, Y.; Dang, C.; Han, B.; Han, R.; Ma, H.; Hao, J.; Wang, L. Inhibition of Double-Strand DNA-Sensing CGAS Ameliorates Brain Injury after Ischemic Stroke. EMBO Mol. Med. 2020, 12, e11002. [Google Scholar] [CrossRef]
  103. Hu, X.; Zhang, H.; Zhang, Q.; Yao, X.; Ni, W.; Zhou, K. Emerging Role of STING Signalling in CNS Injury: Inflammation, Autophagy, Necroptosis, Ferroptosis and Pyroptosis. J. Neuroinflamm. 2022, 19, 242. [Google Scholar] [CrossRef] [PubMed]
  104. Abdullah, A.; Zhang, M.; Frugier, T.; Bedoui, S.; Taylor, J.M.; Crack, P.J. STING-Mediated Type-I Interferons Contribute to the Neuroinflammatory Process and Detrimental Effects Following Traumatic Brain Injury. J. Neuroinflamm. 2018, 15, 323. [Google Scholar] [CrossRef] [PubMed]
  105. Barrett, J.P.; Knoblach, S.M.; Bhattacharya, S.; Gordish-Dressman, H.; Stoica, B.A.; Loane, D.J. Traumatic Brain Injury Induces CGAS Activation and Type I Interferon Signaling in Aged Mice. Front. Immunol. 2021, 12, 710608. [Google Scholar] [CrossRef] [PubMed]
  106. Jin, X.; Ishii, H.; Bai, Z.; Itokazu, T.; Yamashita, T. Temporal Changes in Cell Marker Expression and Cellular Infiltration in a Controlled Cortical Impact Model in Adult Male C57BL/6 Mice. PLoS ONE 2012, 7, e41892. [Google Scholar] [CrossRef] [Green Version]
  107. Turtzo, L.C.; Lescher, J.; Janes, L.; Dean, D.D.; Budde, M.D.; Frank, J.A. Macrophagic and Microglial Responses after Focal Traumatic Brain Injury in the Female Rat. J. Neuroinflamm. 2014, 11, 82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Loane, D.J.; Kumar, A.; Stoica, B.A.; Cabatbat, R.; Faden, A.I. Progressive Neurodegeneration After Experimental Brain Trauma: Association with Chronic Microglial Activation. J. Neuropathol. Exp. Neurol. 2014, 73, 14–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Chiò, A.; Benzi, G.; Dossena, M.; Mutani, R.; Mora, G. Severely Increased Risk of Amyotrophic Lateral Sclerosis among Italian Professional Football Players. Brain 2005, 128, 472–476. [Google Scholar] [CrossRef] [Green Version]
  110. Lehman, E.J.; Hein, M.J.; Baron, S.L.; Gersic, C.M. Neurodegenerative Causes of Death among Retired National Football League Players. Neurology 2012, 79, 1970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Russell, E.R.; MacKay, D.F.; Lyall, D.; Stewart, K.; MacLean, J.A.; Robson, J.; Pell, J.P.; Stewart, W. Neurodegenerative Disease Risk among Former International Rugby Union Players. J. Neurol. Neurosurg. Psychiatry 2022, 93, 1262–1268. [Google Scholar] [CrossRef]
  112. Kurtzke, J.F.; Beebe, G.W. Epidemiology of Amyotrophic Lateral Sclerosis. Neurology 1980, 30, 453. [Google Scholar] [CrossRef]
  113. Turner, M.R.; Abisgold, J.; Yeates, D.G.R.; Talbot, K.; Goldacre, M.J. Head and Other Physical Trauma Requiring Hospitalisation Is Not a Significant Risk Factor in the Development of ALS. J. Neurol. Sci. 2010, 288, 45–48. [Google Scholar] [CrossRef]
  114. Chen, H.; Richard, M.; Sandler, D.P.; Umbach, D.M.; Kamel, F. Head Injury and Amyotrophic Lateral Sclerosis. Am. J. Epidemiol. 2007, 166, 810–816. [Google Scholar] [CrossRef] [Green Version]
  115. Armon, C.; Nelson, L.M. Is Head Trauma a Risk Factor for Amyotrophic Lateral Sclerosis? An Evidence Based Review. Amyotroph. Lateral Scler. 2012, 13, 351–356. [Google Scholar] [CrossRef] [PubMed]
  116. Fournier, C.N.; Gearing, M.; Upadhyayula, S.R.; Klein, M.; Glass, J.D. Head Injury Does Not Alter Disease Progression or Neuropathologic Outcomes in ALS. Neurology 2015, 84, 1788–1795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Deutsch, M.B.; Mendez, M.F.; Teng, E. Interactions between Traumatic Brain Injury and Frontotemporal Degeneration. Dement. Geriatr. Cogn. Disord. 2015, 39, 143–153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Kalkonde, Y.v.; Jawaid, A.; Qureshi, S.U.; Shirani, P.; Wheaton, M.; Pinto-Patarroyo, G.P.; Schulz, P.E. Medical and Environmental Risk Factors Associated with Frontotemporal Dementia: A Case-Control Study in a Veteran Population. Alzheimers Dement. 2012, 8, 204–210. [Google Scholar] [CrossRef]
  119. Rosso, S.M.; Landweer, E.J.; Houterman, M.; Donker Kaat, L.; van Duijn, C.M.; van Swieten, J.C. Medical and Environmental Risk Factors for Sporadic Frontotemporal Dementia: A Retrospective Case-Control Study. J. Neurol. Neurosurg. Psychiatry 2003, 74, 1574–1576. [Google Scholar] [CrossRef] [Green Version]
  120. DeJesus-Hernandez, M.; Mackenzie, I.R.; Boeve, B.F.; Boxer, A.L.; Baker, M.; Rutherford, N.J.; Nicholson, A.M.; Finch, N.C.A.; Flynn, H.; Adamson, J.; et al. Expanded GGGGCC Hexanucleotide Repeat in Noncoding Region of C9ORF72 Causes Chromosome 9p-Linked FTD and ALS. Neuron 2011, 72, 245–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Renton, A.E.; Majounie, E.; Waite, A.; Simón-Sánchez, J.; Rollinson, S.; Gibbs, J.R.; Schymick, J.C.; Laaksovirta, H.; van Swieten, J.C.; Myllykangas, L.; et al. A Hexanucleotide Repeat Expansion in C9ORF72 Is the Cause of Chromosome 9p21-Linked ALS-FTD. Neuron 2011, 72, 257–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Soppela, H.; Krüger, J.; Hartikainen, P.; Koivisto, A.; Haapasalo, A.; Borroni, B.; Remes, A.M.; Katisko, K.; Solje, E. Traumatic Brain Injury Associates with an Earlier Onset in Sporadic Frontotemporal Dementia. J. Alzheimer’s Dis. 2023, 91, 225–232. [Google Scholar] [CrossRef] [PubMed]
  123. Thomsen, G.M.; Vit, J.P.; Lamb, A.; Gowing, G.; Shelest, O.; Alkaslasi, M.; Ley, E.J.; Svendsen, C.N. Acute Traumatic Brain Injury Does Not Exacerbate Amyotrophic Lateral Sclerosis in the SOD1G93A Rat Model. eNeuro 2015, 2, e0059-14.2015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Thomsen, G.M.; Ma, A.M.; Ko, A.; Harada, M.Y.; Wyss, L.; Haro, P.S.; Vit, J.P.; Shelest, O.; Rhee, P.; Svendsen, C.N.; et al. A Model of Recurrent Concussion That Leads to Long-Term Motor Deficits, CTE-like Tauopathy and Exacerbation of an ALS Phenotype. J. Trauma. Acute Care Surg. 2016, 81, 1070–1078. [Google Scholar] [CrossRef] [PubMed]
  125. Alkaslasi, M.R.; Cho, N.E.; Dhillon, N.K.; Shelest, O.; Haro-Lopez, P.S.; Linaval, N.T.; Ghoulian, J.; Yang, A.R.; Vit, J.P.; Avalos, P.; et al. Poor Corticospinal Motor Neuron Health Is Associated with Increased Symptom Severity in the Acute Phase Following Repetitive Mild TBI and Predicts Early ALS Onset in Genetically Predisposed Rodents. Brain Sci. 2021, 11, 160. [Google Scholar] [CrossRef]
  126. Wiesner, D.; Tar, L.; Linkus, B.; Chandrasekar, A.; olde Heuvel, F.; Dupuis, L.; Tsao, W.; Wong, P.C.; Ludolph, A.; Roselli, F. Reversible Induction of TDP-43 Granules in Cortical Neurons after Traumatic Injury. Exp. Neurol. 2018, 299, 15–25. [Google Scholar] [CrossRef] [PubMed]
  127. Evans, T.M.; Jaramillo, C.A.; Sataranatarajan, K.; Watts, L.; Sabia, M.; Qi, W.; van Remmen, H. The Effect of Mild Traumatic Brain Injury on Peripheral Nervous System Pathology in Wild-Type Mice and the G93A Mutant Mouse Model of Motor Neuron Disease. Neuroscience 2015, 298, 410–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Mackenzie, I.R.A.; Rademakers, R. The Role of TDP-43 in Amyotrophic Lateral Sclerosis and Frontotemporal Dementia. Curr. Opin. Neurol. 2008, 21, 693. [Google Scholar] [CrossRef] [Green Version]
  129. Gao, F.; Hu, M.; Zhang, J.; Hashem, J.; Chen, C. TDP-43 Drives Synaptic and Cognitive Deterioration Following Traumatic Brain Injury. Acta Neuropathol. 2022, 144, 187–210. [Google Scholar] [CrossRef]
  130. Wang, H.K.; Lee, Y.C.; Huang, C.Y.; Liliang, P.C.; Lu, K.; Chen, H.J.; Li, Y.C.; Tsai, K.J. Traumatic Brain Injury Causes Frontotemporal Dementia and TDP-43 Proteolysis. Neuroscience 2015, 300, 94–103. [Google Scholar] [CrossRef] [PubMed]
  131. Yang, Z.; Lin, F.; Robertson, C.S.; Wang, K.K. Dual Vulnerability of TDP-43 to Calpain and Caspase-3 Proteolysis after Neurotoxic Conditions and Traumatic Brain Injury. J. Cereb. Blood Flow. Metab. 2014, 34, 1444–1452. [Google Scholar] [CrossRef] [Green Version]
  132. Wright, D.K.; Liu, S.; van der Poel, C.; McDonald, S.J.; Brady, R.D.; Taylor, L.; Yang, L.; Gardner, A.J.; Ordidge, R.; O’Brien, T.J.; et al. Traumatic Brain Injury Results in Cellular, Structural and Functional Changes Resembling Motor Neuron Disease. Cereb. Cortex 2017, 27, 4503–4515. [Google Scholar] [CrossRef] [Green Version]
  133. Johnson, V.E.; Stewart, W.; Trojanowski, J.Q.; Smith, D.H. Acute and Chronically Increased Immunoreactivity to Phosphorylation- Independent but Not Pathological TDP-43 after a Single Traumatic Brain Injury in Humans. Acta Neuropathol. 2011, 122, 715–726. [Google Scholar] [CrossRef] [Green Version]
  134. Rosenbohm, A.; Kassubek, J.; Weydt, P.; Marroquin, N.; Volk, A.E.; Kubisch, C.; Huppertz, H.J.; Weber, M.; Andersen, P.M.; Weishaupt, J.H.; et al. Can Lesions to the Motor Cortex Induce Amyotrophic Lateral Sclerosis? J. Neurol. 2014, 261, 283–290. [Google Scholar] [CrossRef] [PubMed]
  135. Rugbjerg, K.; Ritz, B.; Korbo, L.; Martinussen, N.; Olsen, J.H. Risk of Parkinson’s Disease after Hospital Contact for Head Injury: Population Based Case-Control Study. BMJ 2008, 337, 34–36. [Google Scholar] [CrossRef] [Green Version]
  136. Kenborg, L.; Rugbjerg, K.; Lee, P.C.; Ravnskjær, L.; Christensen, J.; Ritz, B.; Lassen, C.F. Head Injury and Risk for Parkinson Disease: Results from a Danish Case-Control Study. Neurology 2015, 84, 1098–1103. [Google Scholar] [CrossRef] [Green Version]
  137. Spangenberg, S.; Hannerz, H.; Tüchsen, F.; Mikkelsen, K.L. A Nationwide Population Study of Severe Head Injury and Parkinson’s Disease. Park. Relat. Disord. 2009, 15, 12–14. [Google Scholar] [CrossRef]
  138. Jafari, S.; Etminan, M.; Aminzadeh, F.; Samii, A. Head Injury and Risk of Parkinson Disease: A Systematic Review and Meta-Analysis. Mov. Disord. 2013, 28, 1222–1229. [Google Scholar] [CrossRef] [PubMed]
  139. Balabandian, M.; Noori, M.; Lak, B.; Karimizadeh, Z.; Nabizadeh, F. Traumatic Brain Injury and Risk of Parkinson’s Disease: A Meta-Analysis. Acta Neurol. Belg. 2023; ahead-of-print. [Google Scholar] [CrossRef]
  140. Gardner, R.C.; Byers, A.L.; Barnes, D.E.; Li, Y.; Boscardin, J.; Yaffe, K. Mild TBI and Risk of Parkinson Disease. Neurology 2018, 90, e1771–e1779. [Google Scholar] [CrossRef]
  141. Gardner, R.C.; Burke, J.F.; Nettiksimmons, J.; Goldman, S.; Tanner, C.M.; Yaffe, K. Traumatic Brain Injury in Later Life Increases Risk for Parkinson’s Disease. Ann. Neurol. 2015, 77, 987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Bower, J.H.; Maraganore, D.M.; Peterson, B.J.; McDonnell, S.K.; Ahlskog, J.E.; Rocca, W.A. Head Trauma Preceding PD. Neurology 2003, 60, 1610–1615. [Google Scholar] [CrossRef] [PubMed]
  143. Allen, N.E.; Schwarzel, A.K.; Canning, C.G. Recurrent Falls in Parkinson’s Disease: A Systematic Review. Park. Dis. 2013, 2013, 906274. [Google Scholar] [CrossRef] [Green Version]
  144. Chandra, S.; Chen, X.; Rizo, J.; Jahn, R.; Südhof, T.C. A Broken α-Helix in Folded α-Synuclein. J. Biol. Chem. 2003, 278, 15313–15318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Sun, J.; Wang, L.; Bao, H.; Premi, S.; Das, U.; Chapman, E.R.; Roy, S. Functional Cooperation of α-Synuclein and VAMP2 in Synaptic Vesicle Recycling. Proc. Natl. Acad. Sci. USA 2019, 166, 11113–11115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Glenner, G.G.; Wong, C.W. Alzheimer’s Disease: Initial Report of the Purification and Characterization of a Novel Cerebrovascular Amyloid Protein. Biochem. Biophys. Res. Commun. 1984, 120, 885–890. [Google Scholar] [CrossRef] [PubMed]
  147. Masters, C.L.; Multhaup, G.; Simms, G.; Pottgiesser, J.; Martins, R.N.; Beyreuther, K. Neuronal Origin of a Cerebral Amyloid: Neurofibrillary Tangles of Alzheimer’s Disease Contain the Same Protein as the Amyloid of Plaque Cores and Blood Vessels. EMBO J. 1985, 4, 2757–2763. [Google Scholar] [CrossRef] [PubMed]
  148. Mori, H.; Takio, K.; Ogawarag, M.; Selkoen, D.J. Mass Spectrometry of Purified Amyloid Beta Protein in Alzheimer’s Disease. J. Biol. Chem. 1992, 267, 17062–17086. [Google Scholar] [CrossRef]
  149. Iwai, A.; Masliah, E.; Yoshimoto, M.; Ge, N.; Fianagan, L.; Rohan De Silva, H.A.; Kittei, A.; Saitoh, T. The Precursor Protein of Non-Ap Component of Alzheimer’s Disease Amyloid Is a Presynaptic Protein of the Central Nervous System. Neuron 1995, 14, 467–475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Hasan, S.; Mielke, M.M.; Turcano, P.; Ahlskog, J.E.; Bower, J.H.; Savica, R. Traumatic Brain Injury Preceding Clinically Diagnosed α-Synucleinopathies: A Case-Control Study. Neurology 2020, 94, e764–e773. [Google Scholar] [CrossRef]
  151. Nguyen, T.P.; Schaffert, J.; Lobue, C.; Womack, K.B.; Hart, J.; Cullum, C.M. Traumatic Brain Injury and Age of Onset of Dementia with Lewy Bodies. J. Alzheimer’s Dis. 2018, 66, 717–723. [Google Scholar] [CrossRef] [PubMed]
  152. Spillantini, M.G.; Schmidt, M.L.; Lee, V.M.Y.; Trojanowski, J.Q.; Jakes, R.; Goedert, M. α-Synuclein in Lewy Bodies. Nature 1997, 388, 839–840. [Google Scholar] [CrossRef]
  153. Mondello, S.; Buki, A.; Italiano, D.; Jeromin, A. α-Synuclein in CSF of Patients with Severe Traumatic Brain Injury. Neurology 2013, 80, 1662–1668. [Google Scholar] [CrossRef]
  154. Su, E.; Bell, M.J.; Wisniewski, S.R.; Adelson, P.D.; Janesko-Feldman, K.L.; Salonia, R.; Clark, R.S.B.; Kochanek, P.M.; Kagan, V.E.; Bayir, H. α-Synuclein Levels Are Elevated in Cerebrospinal Fluid Following Traumatic Brain Injury in Infants and Children: The Effect of Therapeutic Hypothermia. Dev. Neurosci. 2010, 32, 385–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Acosta, S.A.; Tajiri, N.; de la Pena, I.; Bastawrous, M.; Sanberg, P.R.; Kaneko, Y.; Borlongan, C.V. Alpha-Synuclein as a Pathological Link Between Chronic Traumatic Brain Injury and Parkinson’s Disease. J. Cell. Physiol. 2015, 230, 1024–1032. [Google Scholar] [CrossRef] [Green Version]
  156. Uryu, K.; Giasson, B.I.; Longhi, L.; Martinez, D.; Murray, I.; Conte, V.; Nakamura, M.; Saatman, K.; Talbot, K.; Horiguchi, T.; et al. Age-Dependent Synuclein Pathology Following Traumatic Brain Injury in Mice. Exp. Neurol. 2003, 184, 214–224. [Google Scholar] [CrossRef]
  157. Impellizzeri, D.; Campolo, M.; Bruschetta, G.; Crupi, R.; Cordaro, M.; Paterniti, I.; Cuzzocrea, S.; Esposito, E. Traumatic Brain Injury Leads to Development of Parkinson’s Disease Related Pathology in Mice. Front. Neurosci. 2016, 10, 458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Braak, H.; Sandmann-Keil, D.; Gai, W.; Braak, E. Extensive Axonal Lewy Neurites in Parkinson’s Disease: A Novel Pathological Feature Revealed by α-Synuclein Immunocytochemistry. Neurosci. Lett. 1999, 265, 67–69. [Google Scholar] [CrossRef] [PubMed]
  159. Crane, P.K.; Gibbons, L.E.; Dams-O’Connor, K.; Trittschuh, E.; Leverenz, J.B.; Dirk Keene, C.; Sonnen, J.; Montine, T.J.; Bennett, D.A.; Leurgans, S.; et al. Association between Traumatic Brain Injury and Late Life Neurodegenerative Conditions and Neuropathological Findings. JAMA Neurol. 2016, 73, 1062. [Google Scholar] [CrossRef] [PubMed]
  160. Kanaan, N.M.; Manfredsson, F.P. Loss of Functional Alpha-Synuclein: A Toxic Event in Parkinson’s Disease? J. Park. Dis. 2012, 2, 249–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Carlson, S.W.; Yan, H.Q.; Li, Y.; Henchir, J.; Ma, X.; Young, M.S.; Ikonomovic, M.D.; Dixon, C.E. Differential Regional Responses in Soluble Monomeric Alpha Synuclein Abundance Following Traumatic Brain Injury. Mol. Neurobiol. 2021, 58, 362. [Google Scholar] [CrossRef] [PubMed]
  162. Jenkins, P.O.; Roussakis, A.A.; de Simoni, S.; Bourke, N.; Fleminger, J.; Cole, J.; Piccini, P.; Sharp, D. Distinct Dopaminergic Abnormalities in Traumatic Brain Injury and Parkinson’s Disease. J. Neurol. Neurosurg. Psychiatry 2020, 91, 631–637. [Google Scholar] [CrossRef] [PubMed]
  163. Yan, H.Q.; Kline, A.E.; Ma, X.; Li, Y.; Dixon, C.E. Traumatic Brain Injury Reduces Dopamine Transporter Protein Expression in the Rat Frontal Cortex. Neuroreport 2002, 13, 1899–1901. [Google Scholar] [CrossRef] [PubMed]
  164. Massucci, J.L.; Kline, A.E.; Ma, X.; Zafonte, R.D.; Dixon, C.E. Time Dependent Alterations in Dopamine Tissue Levels and Metabolism after Experimental Traumatic Brain Injury in Rats. Neurosci. Lett. 2004, 372, 127–131. [Google Scholar] [CrossRef]
  165. Yan, H.Q.; Kline, A.E.; Ma, X.; Hooghe-Peters, E.L.; Marion, D.W.; Dixon, C.E. Tyrosine Hydroxylase, but Not Dopamine Beta-Hydroxylase, Is Increased in Rat Frontal Cortex after Traumatic Brain Injury. Neuroreport 2001, 12, 2323–2327. [Google Scholar] [CrossRef] [PubMed]
  166. Rudelli, R.; Strom, J.O.; Welch, P.T.; Ambler, M.W. Posttraumatic Premature Alzheimer’s Disease. Neuropathologic Findings and Pathogenetic Considerations. Arch. Neurol. 1982, 39, 570–575. [Google Scholar] [CrossRef]
  167. Claude, H.; Cuel, J. Démence Pré-Sénile Post-Traumatique Après Fracture Du Crâne. Considérations Médico-Légales. Ann. Med. -Leg. 1939, 19, 173–184. [Google Scholar]
  168. Gu, D.; Ou, S.; Liu, G. Traumatic Brain Injury and Risk of Dementia and Alzheimer’s Disease: A Systematic Review and Meta-Analysis. Neuroepidemiology 2022, 56, 4–16. [Google Scholar] [CrossRef] [PubMed]
  169. Fleminger, S.; Oliver, D.L.; Lovestone, S.; Rabe-Hesketh, S.; Giora, A. Head Injury as a Risk Factor for Alzheimer’s Disease: The Evidence 10 Years on; a Partial Replication. J. Neurol. Neurosurg. Psychiatry 2003, 74, 857–862. [Google Scholar] [CrossRef] [PubMed]
  170. Perry, D.C.; Sturm, V.E.; Peterson, M.J.; Pieper, C.F.; Bullock, T.; Boeve, B.F.; Miller, B.L.; Guskiewicz, K.M.; Berger, M.S.; Kramer, J.H.; et al. Association of Traumatic Brain Injury with Subsequent Neurological and Psychiatric Disease: A Meta-Analysis. J. Neurosurg. 2016, 124, 511–526. [Google Scholar] [CrossRef] [Green Version]
  171. Nemetz, P.N.; Leibson, C.; Naessens, J.M.; Beard, M.; Kokmen, E.; Annegers, J.F.; Kurland, L.T. Traumatic Brain Injury and Time to Onset of Alzheimer’s Disease: A Population-Based Study. Am. J. Epidemiol. 1999, 149, 32–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Mohamed, A.Z.; Nestor, P.J.; Cumming, P.; Nasrallah, F.A. Traumatic Brain Injury Fast-Forwards Alzheimer’s Pathology: Evidence from Amyloid Positron Emission Tomorgraphy Imaging. J. Neurol. 2021, 1, 3. [Google Scholar] [CrossRef]
  173. Tran, H.T.; LaFerla, F.M.; Holtzman, D.M.; Brody, D.L. Controlled Cortical Impact Traumatic Brain Injury in 3xTg-AD Mice Causes Acute Intra-Axonal Amyloid-β Accumulation and Independently Accelerates the Development of Tau Abnormalities. J. Neurosci. 2011, 31, 9513–9525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Tajiri, N.; Kellogg, S.L.; Shimizu, T.; Arendash, G.W.; Borlongan, C.V. Traumatic Brain Injury Precipitates Cognitive Impairment and Extracellular Aβ Aggregation in Alzheimer’s Disease Transgenic Mice. PLoS ONE 2013, 8, e78851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Zyśk, M.; Clausen, F.; Aguilar, X.; Sehlin, D.; Syvänen, S.; Erlandsson, A. Long-Term Effects of Traumatic Brain Injury in a Mouse Model of Alzheimer’s Disease. J. Alzheimer’s Dis. 2019, 72, 161–180. [Google Scholar] [CrossRef] [Green Version]
  176. Washington, P.M.; Morffy, N.; Parsadanian, M.; Zapple, D.N.; Burns, M.P. Experimental Traumatic Brain Injury Induces Rapid Aggregation and Oligomerization of Amyloid-Beta in an Alzheimer’s Disease Mouse Model. J. Neurotrauma 2014, 31, 125–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Shishido, H.; Kishimoto, Y.; Kawai, N.; Toyota, Y.; Ueno, M.; Kubota, T.; Kirino, Y.; Tamiya, T. Traumatic Brain Injury Accelerates Amyloid-β Deposition and Impairs Spatial Learning in the Triple-Transgenic Mouse Model of Alzheimer’s Disease. Neurosci. Lett. 2016, 629, 62–67. [Google Scholar] [CrossRef] [PubMed]
  178. Sawmiller, D.; Li, S.; Shahaduzzaman, M.; Smith, A.J.; Obregon, D.; Giunta, B.; Borlongan, C.V.; Sanberg, P.R.; Tan, J. Luteolin Reduces Alzheimer’s Disease Pathologies Induced by Traumatic Brain Injury. Int. J. Mol. Sci. 2014, 15, 895–904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Lecca, D.; Bader, M.; Tweedie, D.; Hoffman, A.F.; Jung, Y.J.; Hsueh, S.C.; Hoffer, B.J.; Becker, R.E.; Pick, C.G.; Lupica, C.R.; et al. (-)-Phenserine and the Prevention of Pre-Programmed Cell Death and Neuroinflammation in Mild Traumatic Brain Injury and Alzheimer’s Disease Challenged Mice. Neurobiol. Dis. 2019, 130, 104528. [Google Scholar] [CrossRef]
  180. Crawford, F.C.; Wood, M.; Ferguson, S.; Mathura, V.S.; Faza, B.; Wilson, S.; Fan, T.; O’Steen, B.; Ait-Ghezala, G.; Hayes, R.; et al. Genomic Analysis of Response to Traumatic Brain Injury in a Mouse Model of Alzheimer’s Disease (APPsw). Brain Res. 2007, 1185, 45–58. [Google Scholar] [CrossRef]
  181. Wu, Y.; Wu, H.; Zeng, J.; Pluimer, B.; Dong, S.; Xie, X.; Guo, X.; Ge, T.; Liang, X.; Feng, S.; et al. Mild Traumatic Brain Injury Induces Microvascular Injury and Accelerates Alzheimer-like Pathogenesis in Mice. Acta Neuropathol. Commun. 2021, 9, 74. [Google Scholar] [CrossRef] [PubMed]
  182. Lou, D.; Du, Y.; Huang, D.; Cai, F.; Zhang, Y.; Li, T.; Zhou, W.; Gao, H.; Song, W. Traumatic Brain Injury Alters the Metabolism and Facilitates Alzheimer’s Disease in a Murine Model. Mol. Neurobiol. 2018, 55, 4928–4939. [Google Scholar] [CrossRef]
  183. Ott, A.; Stolk, R.P.; van Harskamp, F.; Pols, H.A.P.; Hofman, A.; Breteler, M.M.B. Diabetes Mellitus and the Risk of Dementia: The Rotterdam Study. Neurology 1999, 53, 1937–1942. [Google Scholar] [CrossRef] [PubMed]
  184. Rivera, E.J.; Goldin, A.; Fulmer, N.; Tavares, R.; Wands, J.R.; de La Monte, S.M. Insulin and Insulin-like Growth Factor Expression and Function Deteriorate with Progression of Alzheimer’s Disease: Link to Brain Reductions in Acetylcholine. J. Alzheimer’s Dis. 2005, 8, 247–268. [Google Scholar] [CrossRef]
  185. Craft, S.; Baker, L.D.; Montine, T.J.; Minoshima, S.; Watson, G.S.; Claxton, A.; Arbuckle, M.; Callaghan, M.; Tsai, E.; Plymate, S.R.; et al. Intranasal Insulin Therapy for Alzheimer Disease and Amnestic Mild Cognitive Impairment: A Pilot Clinical Trial. Arch. Neurol. 2012, 69, 29–38. [Google Scholar] [CrossRef] [Green Version]
  186. Reger, M.A.; Watson, G.S.; Green, P.S.; Wilkinson, C.W.; Baker, L.D.; Cholerton, B.; Fishel, M.A.; Plymate, S.R.; Breitner, J.C.S.; DeGroodt, W.; et al. Intranasal Insulin Improves Cognition and Modulates β-Amyloid in Early AD. Neurology 2008, 70, 440–448. [Google Scholar] [CrossRef]
  187. Franklin, W.; Krishnan, B.; Taglialatela, G. Chronic Synaptic Insulin Resistance after Traumatic Brain Injury Abolishes Insulin Protection from Amyloid Beta and Tau Oligomer-Induced Synaptic Dysfunction. Sci. Rep. 2019, 9, 8228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Korkmaz, N.; Kesikburun, S.; Atar, M.Ö.; Sabuncu, T. Insulin Resistance and Related Factors in Patients with Moderate and Severe Traumatic Brain Injury. Ir. J. Med. Sci. 2022; ahead-of-print. [Google Scholar] [CrossRef]
  189. Sekar, S.; Viswas, R.S.; Mahabadi, H.M.; Alizadeh, E.; Fonge, H.; Taghibiglou, C. Concussion/Mild Traumatic Brain Injury (TBI) Induces Brain Insulin Resistance: A Positron Emission Tomography (PET) Scanning Study. Int. J. Mol. Sci. 2021, 22, 9005. [Google Scholar] [CrossRef] [PubMed]
  190. de Felice, F.G.; Vieira, M.N.N.; Bomfim, T.R.; Decker, H.; Velasco, P.T.; Lambert, M.P.; Viola, K.L.; Zhao, W.Q.; Ferreira, S.T.; Klein, W.L. Protection of Synapses against Alzheimer’s-Linked Toxins: Insulin Signaling Prevents the Pathogenic Binding of Aβ Oligomers. Proc. Natl. Acad. Sci. USA 2009, 106, 1971–1976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Karelina, K.; Sarac, B.; Freeman, L.M.; Gaier, K.R.; Weil, Z.M. Traumatic Brain Injury and Obesity Induce Persistent Central Insulin Resistance. Eur. J. Neurosci. 2016, 43, 1034–1043. [Google Scholar] [CrossRef] [PubMed]
  192. Bird, S.M.; Sohrabi, H.R.; Sutton, T.A.; Weinborn, M.; Rainey-Smith, S.R.; Brown, B.; Patterson, L.; Taddei, K.; Gupta, V.; Carruthers, M.; et al. Cerebral Amyloid-β Accumulation and Deposition Following Traumatic Brain Injury—A Narrative Review and Meta-Analysis of Animal Studies. Neurosci. Biobehav. Rev. 2016, 64, 215–228. [Google Scholar] [CrossRef] [PubMed]
  193. Johnson, V.E.; Stewart, W.; Smith, D.H. Widespread Tau and Amyloid-Beta Pathology Many Years After a Single Traumatic Brain Injury in Humans. Brain Pathol. 2012, 22, 142–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Ikonomovic, M.D.; Uryu, K.; Abrahamson, E.E.; Ciallella, J.R.; Trojanowski, J.Q.; Lee, V.M.Y.; Clark, R.S.; Marion, D.W.; Wisniewski, S.R.; DeKosky, S.T. Alzheimer’s Pathology in Human Temporal Cortex Surgically Excised after Severe Brain Injury. Exp. Neurol. 2004, 190, 192–203. [Google Scholar] [CrossRef] [PubMed]
  195. Nakagawa, Y.; Reed, L.; Nakamura, M.; McIntosh, T.K.; Smith, D.H.; Saatman, K.E.; Raghupathi, R.; Clemens, J.; Saido, T.C.; Lee, V.M.Y.; et al. Brain Trauma in Aged Transgenic Mice Induces Regression of Established Aβ Deposits. Exp. Neurol. 2000, 163, 244–252. [Google Scholar] [CrossRef] [PubMed]
  196. Nakagawa, Y.; Nakamura, M.; Mcintosh, T.K.; Rodriguez, A.; Berlin, J.A.; Smith, D.H.; Saatman, K.E.; Raghupathi, R.; Clemens, J.; Saido, T.C.; et al. Traumatic Brain Injury in Young, Amyloid-β Peptide Overexpressing Transgenic Mice Induces Marked Ipsilateral Hippocampal Atrophy and Diminished Aβ Deposition during Aging. J. Comp. Neurol. 1999, 411, 390–398. [Google Scholar] [CrossRef]
  197. Cummins, T.L.; Doré, V.; Feizpour, A.; Krishnadas, N.; Bourgeat, P.; Elias, A.; Lamb, F.; Williams, R.; Hopwood, M.; Landau, S.; et al. Tau, β-Amyloid, and Glucose Metabolism Following Service-Related Traumatic Brain Injury in Vietnam War Veterans: The Australian Imaging Biomarkers and Lifestyle Study of Aging-Veterans Study (AIBL-VETS). J. Neurotrauma 2023. ahead-of-print. [Google Scholar] [CrossRef] [PubMed]
  198. VE, J.; Stewart, W.; DH, S. Traumatic Brain Injury and Amyloid-β Pathology: A Link to Alzheimer’s Disease? Nat. Rev. Neurosci. 2010, 11, 361–370. [Google Scholar] [CrossRef] [Green Version]
  199. Uryu, K.; Chen, X.H.; Martinez, D.; Browne, K.D.; Johnson, V.E.; Graham, D.I.; Lee, V.M.Y.; Trojanowski, J.Q.; Smith, D.H. Multiple Proteins Implicated in Neurodegenerative Diseases Accumulate in Axons after Brain Trauma in Humans. Exp. Neurol. 2007, 208, 185–192. [Google Scholar] [CrossRef] [Green Version]
  200. XH, C.; Siman, R.; Iwata, A.; Meaney, D.F.; Trojanowski, J.Q.; Smith, D.H. Long-Term Accumulation of Amyloid-Beta, Beta-Secretase, Presenilin-1, and Caspase-3 in Damaged Axons Following Brain Trauma. Am. J. Pathol. 2004, 165, 357–371. [Google Scholar] [CrossRef]
  201. Smith, D.H.; Chen, X.H.; Nonaka, M.; Trojanowski, J.Q.; Lee, V.M.Y.; Saatman, K.E.; Leoni, M.J.; Xu, B.N.; Wolf, J.A.; Meaney, D.F. Accumulation of Amyloid Beta and Tau and the Formation of Neurofilament Inclusions Following Diffuse Brain Injury in the Pig. J. Neuropathol. Exp. Neurol. 1999, 58, 982–992. [Google Scholar] [CrossRef] [PubMed]
  202. Itoh, T.; Satou, T.; Nishida, S.; Tsubaki, M.; Hashimoto, S.; Ito, H. Expression of Amyloid Precursor Protein after Rat Traumatic Brain Injury. Neurol. Res. 2013, 31, 103–109. [Google Scholar] [CrossRef] [PubMed]
  203. Smith, D.H.; Chen, X.H.; Iwata, A.; Graham, D.I. Amyloid β Accumulation in Axons after Traumatic Brain Injury in Humans. J. Neurosurg. 2003, 98, 1072–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Knoblach, S.M.; Nikolaeva, M.; Huang, X.; Fan, L.; Krajewski, S.; Reed, J.C.; Faden, A.I. Multiple Caspases Are Activated after Traumatic Brain Injury: Evidence for Involvement in Functional Outcome. J. Neurotrauma 2004, 19, 1155–1170. [Google Scholar] [CrossRef] [PubMed]
  205. Blasko, I.; Beer, R.; Bigl, M.; Apelt, J.; Franz, G.; Rudzki, D.; Ransmayr, G.; Kampfl, A.; Schliebs, R. Experimental Traumatic Brain Injury in Rats Stimulates the Expression, Production and Activity of Alzheimer’s Disease β-Secretase (BACE-1). J. Neural Transm. 2004, 111, 523–536. [Google Scholar] [CrossRef]
  206. Stone, J.R.; Okonkwo, D.O.; Singleton, R.H.; Mutlu, L.K.; Helm, G.A.; Povlishock, J.T. Caspase-3-Mediated Cleavage of Amyloid Precursor Protein and Formation of Amyloid β Peptide in Traumatic Axonal Injury. J. Neurotrauma 2004, 19, 601–614. [Google Scholar] [CrossRef]
  207. Yang, W.J.; Chen, W.; Chen, L.; Guo, Y.J.; Zeng, J.S.; Li, G.Y.; Tong, W.S. Involvement of Tau Phosphorylation in Traumatic Brain Injury Patients. Acta Neurol. Scand. 2017, 135, 622–627. [Google Scholar] [CrossRef] [PubMed]
  208. Hawkins, B.E.; Krishnamurthy, S.; Castillo-Carranza, D.L.; Sengupta, U.; Prough, D.S.; Jackson, G.R.; DeWitt, D.S.; Kayed, R. Rapid Accumulation of Endogenous Tau Oligomers in a Rat Model of Traumatic Brain Injury. J. Biol. Chem. 2013, 288, 17042–17050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Edwards, G.; Zhao, J.; Dash, P.K.; Soto, C.; Moreno-Gonzalez, I. Traumatic Brain Injury Induces Tau Aggregation and Spreading. J. Neurotrauma 2020, 37, 80–92. [Google Scholar] [CrossRef] [Green Version]
  210. Zanier, E.R.; Bertani, I.; Sammali, E.; Pischiutta, F.; Chiaravalloti, M.A.; Vegliante, G.; Masone, A.; Corbelli, A.; Smith, D.H.; Menon, D.K.; et al. Induction of a Transmissible Tau Pathology by Traumatic Brain Injury. Brain 2018, 141, 2685–2699. [Google Scholar] [CrossRef] [PubMed]
  211. Cook, C.; Carlomagno, Y.; Gendron, T.F.; Dunmore, J.; Scheffel, K.; Stetler, C.; Davis, M.; Dickson, D.; Jarpe, M.; DeTure, M.; et al. Acetylation of the KXGS Motifs in Tau Is a Critical Determinant in Modulation of Tau Aggregation and Clearance. Hum. Mol. Genet. 2014, 23, 104–116. [Google Scholar] [CrossRef] [PubMed]
  212. Sohn, P.D.; Tracy, T.E.; Son, H.I.; Zhou, Y.; Leite, R.E.P.; Miller, B.L.; Seeley, W.W.; Grinberg, L.T.; Gan, L. Acetylated Tau Destabilizes the Cytoskeleton in the Axon Initial Segment and Is Mislocalized to the Somatodendritic Compartment. Mol. Neurodegener. 2016, 11, 47. [Google Scholar] [CrossRef] [Green Version]
  213. Irwin, D.J.; Cohen, T.J.; Grossman, M.; Arnold, S.E.; Xie, S.X.; Lee, V.M.Y.M.-Y.; Trojanowski, J.Q. Acetylated Tau, a Novel Pathological Signature in Alzheimer’s Disease and Other Tauopathies. Brain 2012, 135, 807–818. [Google Scholar] [CrossRef] [Green Version]
  214. Min, S.W.; Cho, S.H.; Zhou, Y.; Schroeder, S.; Haroutunian, V.; Seeley, W.W.; Huang, E.J.; Shen, Y.; Masliah, E.; Mukherjee, C.; et al. Acetylation of Tau Inhibits Its Degradation and Contributes to Tauopathy. Neuron 2010, 67, 953–966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Min, S.-W.W.; Chen, X.; Tracy, T.E.; Li, Y.; Zhou, Y.; Wang, C.; Shirakawa, K.; Minami, S.S.; Defensor, E.; Mok, S.A.; et al. Critical Role of Acetylation in Tau-Mediated Neurodegeneration and Cognitive Deficits. Nat. Med. 2015, 21, 1154–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Tracy, T.E.; Sohn, P.D.; Minami, S.S.; Wang, C.; Min, S.W.; Li, Y.; Zhou, Y.; Le, D.; Lo, I.; Ponnusamy, R.; et al. Acetylated Tau Obstructs KIBRA-Mediated Signaling in Synaptic Plasticity and Promotes Tauopathy-Related Memory Loss. Neuron 2016, 90, 245–260. [Google Scholar] [CrossRef] [Green Version]
  217. Yan, Y.; Wang, X.; Chaput, D.; Shin, M.K.; Koh, Y.; Gan, L.; Pieper, A.A.; Woo, J.A.A.; Kang, D.E. X-Linked Ubiquitin-Specific Peptidase 11 Increases Tauopathy Vulnerability in Women. Cell 2022, 185, 3913–3930.e19. [Google Scholar] [CrossRef] [PubMed]
  218. Paul, B.D. DUB’ling down Uncovers an X-Linked Vulnerability in Alzheimer’s Disease. Cell 2022, 185, 3854–3856. [Google Scholar] [CrossRef] [PubMed]
  219. Shultz, S.R.; Wright, D.K.; Zheng, P.; Stuchbery, R.; Liu, S.J.; Sashindranath, M.; Medcalf, R.L.; Johnston, L.A.; Hovens, C.M.; Jones, N.C.; et al. Sodium Selenate Reduces Hyperphosphorylated Tau and Improves Outcomes after Traumatic Brain Injury. Brain 2015, 138, 1297–1313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Begum, G.; Yan, H.Q.; Li, L.; Singh, A.; Edward Dixon, C.; Sun, D. Docosahexaenoic Acid Reduces ER Stress and Abnormal Protein Accumulation and Improves Neuronal Function Following Traumatic Brain Injury. J. Neurosci. 2014, 34, 3743–3755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Saletti, P.G.; Casillas-Espinosa, P.M.; Lisgaras, C.P.; Mowrey, W.B.; Li, Q.; Liu, W.; Brady, R.D.; Ali, I.; Silva, J.; Yamakawa, G.; et al. Tau Phosphorylation Patterns in the Rat Cerebral Cortex after Traumatic Brain Injury and Sodium Selenate Effects: An EpiBioS4Rx Project 2 Study. J. Neurotrauma, 2023; ahead-of-print. [Google Scholar] [CrossRef]
  222. Wang, B.; Han, S. Exosome-Associated Tau Exacerbates Brain Functional Impairments Induced by Traumatic Brain Injury in Mice. Mol. Cell. Neurosci. 2018, 88, 158–166. [Google Scholar] [CrossRef]
  223. Wang, Y.; Hall, R.A.; Lee, M.; Kamgar-Parsi, A.; Bi, X.; Baudry, M. The Tyrosine Phosphatase PTPN13/ FAP-1 Links Calpain-2, TBI and Tau Tyrosine Phosphorylation OPEN. Sci. Rep. 2017, 7, 11771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Wu, Z.; Wang, Z.H.; Liu, X.; Zhang, Z.; Gu, X.; Yu, S.P.; Keene, C.D.; Cheng, L.; Ye, K. Traumatic Brain Injury Triggers APP and Tau Cleavage by Delta-Secretase, Mediating Alzheimer’s Disease Pathology. Prog. Neurobiol. 2020, 185, 101730. [Google Scholar] [CrossRef]
  225. Rubenstein, R.; Chang, B.; Grinkina, N.; Drummond, E.; Davies, P.; Ruditzky, M.; Sharma, D.; Wang, K.; Wisniewski, T. Tau Phosphorylation Induced by Severe Closed Head Traumatic Brain Injury Is Linked to the Cellular Prion Protein. Acta Neuropathol. Commun. 2017, 5, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Albayram, O.; Kondo, A.; Mannix, R.; Smith, C.; Tsai, C.Y.; Li, C.; Herbert, M.K.; Qiu, J.; Monuteaux, M.; Driver, J.; et al. Cis P-Tau Is Induced in Clinical and Preclinical Brain Injury and Contributes to Post-Injury Sequelae. Nat. Commun. 2017, 8, 1000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Kondo, A.; Shahpasand, K.; Mannix, R.; Qiu, J.; Moncaster, J.; Chen, C.H.; Yao, Y.; Lin, Y.M.; Driver, J.A.; Sun, Y.; et al. Antibody against Early Driver of Neurodegeneration Cis P-Tau Blocks Brain Injury and Tauopathy. Nature 2015, 523, 431–436. [Google Scholar] [CrossRef] [Green Version]
  228. Wang, C.; Shao, C.; Zhang, L.; Siedlak, S.L.; Meabon, J.S.; Peskind, E.R.; Lu, Y.; Wang, W.; Perry, G.; Cook, D.G.; et al. Oxidative Stress Signaling in Blast Tbi-Induced Tau Phosphorylation. Antioxidants 2021, 10, 955. [Google Scholar] [CrossRef] [PubMed]
  229. Huber, B.R.; Meabon, J.S.; Martin, T.J.; Mourad, P.D.; Bennett, R.; Kraemer, B.C.; Cernak, I.; Petrie, E.C.; Emery, M.J.; Swenson, E.R.; et al. Blast Exposure Causes Early and Persistent Aberrant Phospho- and Cleaved-Tau Expression in a Murine Model of Mild Blast-Induced Traumatic Brain Injury. J. Alzheimer’s Dis. 2013, 37, 309–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Goldstein, L.E.; Fisher, A.M.; Tagge, C.A.; Zhang, X.L.; Velisek, L.; Sullivan, J.A.; Upreti, C.; Kracht, J.M.; Ericsson, M.; Wojnarowicz, M.W.; et al. Chronic Traumatic Encephalopathy in Blast-Exposed Military Veterans and a Blast Neurotrauma Mouse Model. Sci. Transl. Med. 2012, 4, 212–213. [Google Scholar] [CrossRef]
  231. Acosta, S.A.; Tajiri, N.; Sanberg, P.R.; Kaneko, Y.; Borlongan, C.V. Increased Amyloid Precursor Protein and Tau Expression Manifests as Key Secondary Cell Death in Chronic Traumatic Brain Injury. J. Cell. Physiol. 2017, 232, 665–677. [Google Scholar] [CrossRef] [PubMed]
  232. Hoshino, S.; Tamaoka, A.; Takahashi, M.; Kobayashi, S.; Furukawa, T.; Oaki, Y.; Mori, O.; Matsuno, S.; Shoji, S.; Inomata, M. Emergence of Immunoreactivities for Phosphorylated Tau and Amyloid-β Protein in Chronic Stage of Fluid Percussion Injury in Rat Brain. NeuroReport 1998, 9, 1879–1883. [Google Scholar] [CrossRef] [PubMed]
  233. Shi, Q.-X.; Chen, B.; Nie, C.; Zhao, Z.-P.; Zhang, J.-H.; Si, S.-Y.; Cui, S.-J.; Gu, J. wen A Novel Model of Blast Induced Traumatic Brain Injury Caused by Compressed Gas Produced Sustained Cognitive Deficits in Rats: Involvement of Phosphorylation of Tau at the Thr205 Epitope. Brain Res. Bull. 2020, 157, 149–161. [Google Scholar] [CrossRef]
  234. Lv, Q.; Lan, W.; Sun, W.; Ye, R.; Fan, X.; Ma, M.; Yin, Q.; Jiang, Y.; Xu, G.; Dai, J.; et al. Intranasal Nerve Growth Factor Attenuates Tau Phosphorylation in Brain after Traumatic Brain Injury in Rats. J. Neurol. Sci. 2014, 345, 48–55. [Google Scholar] [CrossRef] [PubMed]
  235. Collins-Praino, L.; Gutschmidt, D.; Sharkey, J.; Arulsamy, A.; Corrigan, F. Temporal Changes in Tau Phosphorylation and Related Kinase and Phosphatases Following Two Models of Traumatic Brain Injury. J. Neuroinflamm. Neurodegener. Dis. 2018, 2, 100007. [Google Scholar]
  236. Zhu, W.; Zhao, L.; Li, T.; Xu, H.; Ding, Y.; Cui, G. Docosahexaenoic Acid Ameliorates Traumatic Brain Injury Involving JNK-Mediated Tau Phosphorylation Signaling. Neurosci. Res. 2020, 157, 44–50. [Google Scholar] [CrossRef] [PubMed]
  237. Arun, P.; Oguntayo, S.; Van Albert, S.; Gist, I.; Wang, Y.; Nambiar, M.P.; Long, J.B. Acute Decrease in Alkaline Phosphatase after Brain Injury: A Potential Mechanism for Tauopathy. Neurosci. Lett. 2015, 609, 152–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Zhao, Z.A.; Zhao, Y.; Ning, Y.L.; Yang, N.; Peng, Y.; Li, P.; Chen, X.Y.; Liu, D.; Wang, H.; Chen, X.; et al. Adenosine A2A Receptor Inactivation Alleviates Early-Onset Cognitive Dysfunction after Traumatic Brain Injury Involving an Inhibition of Tau Hyperphosphorylation. Transl. Psychiatry 2017, 7, e1123. [Google Scholar] [CrossRef] [Green Version]
  239. Schafer, D.P.; Jha, S.; Liu, F.; Akella, T.; McCullough, L.D.; Rasband, M.N. Disruption of the Axon Initial Segment Cytoskeleton Is a New Mechanism for Neuronal Injury. J. Neurosci. 2009, 29, 13242–13254. [Google Scholar] [CrossRef] [Green Version]
  240. Dams-O’Connor, K.; Awwad, H.O.; Hoffman, S.W.; Pugh, M.J.; Johnson, V.; Keene, C.D.; McGavern, L.; Mukherjee, P.; Opanashuk, L.; Umoh, N.; et al. Alzheimer’s Disease-Related Dementias Summit 2022: National Research Priorities for the Investigation of Post-Traumatic Brain Injury Alzheimer’s Disease and Related Dementias. J. Neurotrauma 2023. ahead-of-print. [Google Scholar] [CrossRef] [PubMed]
Figure 1. TBI leads to an increase in ROS, which chemically modify proteins, lipids, and nucleic acids, deplete endogenous antioxidants, and activate cellular responses via Nrf2. Oxidative stress causes mitochondrial dysfunction, inflammation, ferroptosis, and ultimately neurodegeneration.
Figure 1. TBI leads to an increase in ROS, which chemically modify proteins, lipids, and nucleic acids, deplete endogenous antioxidants, and activate cellular responses via Nrf2. Oxidative stress causes mitochondrial dysfunction, inflammation, ferroptosis, and ultimately neurodegeneration.
Biomedicines 11 01154 g001
Figure 2. Potential mechanisms by which TBI increases the risk of aging-related neurodegenerative disease, both generally (dotted arrow in top panel) as well as specifically (dotted arrows in bottom panels) in ALS, FTD, PD, and AD. The red arrows indicate an increased risk of the specified condition.
Figure 2. Potential mechanisms by which TBI increases the risk of aging-related neurodegenerative disease, both generally (dotted arrow in top panel) as well as specifically (dotted arrows in bottom panels) in ALS, FTD, PD, and AD. The red arrows indicate an increased risk of the specified condition.
Biomedicines 11 01154 g002
Table 1. Meta-analyses assessing the association between TBI and PD.
Table 1. Meta-analyses assessing the association between TBI and PD.
First AuthorOdds/Risk Ratio (*)
Jafari (2013) [138]1.57 (1.35–1.83)
Balabandian (2023) [139]1.48 (1.22–1.74)
* (95% confidence interval).
Table 2. Meta-analyses assessing the association between TBI and AD.
Table 2. Meta-analyses assessing the association between TBI and AD.
First AuthorOdds Ratio (*)Notes
Fleminger (2003) [169]1.58 (1.21–2.06)Male-specific effect
Perry (2016) [170]1.40 (1.02–1.90)Only mild TBI included
Li (2017) [13]1.52 (1.26–1.80)Male-specific effect
* Odds ratio (95% confidence interval).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Barker, S.; Paul, B.D.; Pieper, A.A. Increased Risk of Aging-Related Neurodegenerative Disease after Traumatic Brain Injury. Biomedicines 2023, 11, 1154. https://doi.org/10.3390/biomedicines11041154

AMA Style

Barker S, Paul BD, Pieper AA. Increased Risk of Aging-Related Neurodegenerative Disease after Traumatic Brain Injury. Biomedicines. 2023; 11(4):1154. https://doi.org/10.3390/biomedicines11041154

Chicago/Turabian Style

Barker, Sarah, Bindu D. Paul, and Andrew A. Pieper. 2023. "Increased Risk of Aging-Related Neurodegenerative Disease after Traumatic Brain Injury" Biomedicines 11, no. 4: 1154. https://doi.org/10.3390/biomedicines11041154

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