U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Laskowitz D, Grant G, editors. Translational Research in Traumatic Brain Injury. Boca Raton (FL): CRC Press/Taylor and Francis Group; 2016.

Cover of Translational Research in Traumatic Brain Injury

Translational Research in Traumatic Brain Injury.

Show details

Chapter 3Diffuse Axonal Injury

and .

Traumatic brain injury (TBI) is the leading cause of death in the United States in people between the ages of 1 and 44 years and occurs in hundreds of thousands of subjects yearly. Recently, the importance of apparently mild injuries has been recognized as a public health crisis for soldiers in the combat theater, children and young adults in sport activities, and others throughout their normal life. Our understanding of the pathology of TBI is rudimentary despite years of study. This chapter will summarize an important aspect of TBI pathology—diffuse axonal injury (DAI)—that is increasingly recognized as an important cause of long-term disability and mortality.

DAI describes a process of widespread axonal damage in the aftermath of acute or repetitive TBI, leading to deficits in cerebral connectivity that may or may not recover over time. It is a component of injury in 40% to 50% of hospital admissions for traumatic brain injury (TBI)1 and one of the most common pathologies in all closed-head trauma.25 DAI is typically characterized by coma without focal lesion6,7 on presentation and pathologically defined by axonal damage in multiple regions of the brain parenchyma, often causing impairments in cognitive, autonomic motor, and sensory function by virtue of disrupted neuronal connectivity. Areas commonly affected include axons in the brainstem, parasagittal white matter near the cerebral cortex, and corpus callosum.1 Though DAI is often characterized as a structural disease, where the fundamental brain architecture is irreparably disrupted on a microscopic scale, many aspects of DAI are influenced not only by physical factors but by perturbations in any number of pathways including metabolism, electrochemistry, and inflammation among others. DAI preceded by a period of relative lucidity diagnosed with imaging findings has also been described.8,9 This may correlate with animal models of DAI demonstrating direction dependence of trauma influencing coma, however, lower extremity long bone orthopedic intervention factored heavily in these clinical cases, and imaging findings in these patients’ brains may have been confounded by emboli mimicking DAI findings.

DAI was first described in comatose trauma subjects who demonstrated scattered axonal injury in the cerebrum, cerebellum, and brainstem on postmortem examination.10 Advances in histopathology during the 1950s further revealed the extent of axonal injury associated with DAI at the cellular level. Through the 1960s additional studies would describe an early histopathological pattern of scattered axonal injury characterized by worsening swelling and distortion of normal architecture into retraction bulbs and helical structures. Extensive characterization of these histologic findings in trauma victims subject to large acceleration/deceleration forces led to the initial definitions of “diffuse axonal injury” described in 1982 by Adams and colleagues.2 The Adams classification is used to grade DAI from a pathology standpoint and is described (Table 3.1).11

TABLE 3.1

TABLE 3.1

Adams Classification of DAI

With increasingly sophisticated methods for neurological evaluation, the understanding of DAI as a diffuse process has evolved. It is now instead considered a regional process influencing multiple disparate brain regions, initiated by a global and significant high-energy traumatic insult.1

CLINICAL FEATURES

Some degree of DAI is likely in patients suffering moderate to severe TBI with loss of consciousness,1 with initial Glasgow Coma Scale (GCS) assessment perhaps reflecting functional impairment of the brainstem and the reticular activating system within the midbrain. Seminal work by Genarelli et al. indicates that DAI can be the sole factor in causing coma after TBI.12 In such circumstances, improvements in responsiveness and alertness may develop slowly a protracted course over weeks to months of intensive rehabilitation. DAI has been historically defined using histopathological methods,1 though recent advances in neuroimaging have considerably improved in vivo diagnosis. In terms of functional outcome DAI is likely the most common cause of severe impairment after TBI. Disruptions in consciousness were initially attributed specifically to brainstem injury, however, coma after DAI is also frequently associated with axonal damage in cerebral white matter as well.13 Persistent cognitive and memory deficits, seen in TBI in general, are prominent in these patients with deficits in information processing.14,15 Magnetic resonance imaging (MRI) data concurrently demonstrate a dose-dependent effect of DAI lesions in the brain on cognitive impairment.1619

Hypothalamic injury and panhypopituitarism have been associated with DAI, possibly due to shear injury across the pituitary stalk from the same high kinetic energy forces that cause DAI.2023 Additionally dopaminergic pathways in the anteroventral third ventricular region mediate arginine vasopressin release and may be disrupted in DAI. This may contribute to sodium and free water derangements seen complicating post-TBI management.23

INITIAL INJURY AND PRIMARY AXOTOMY

The initial traumatic insult causes dynamic deformation of the brain parenchyma, thereby putting long-tract structures such as axons and blood vessels at risk for stretch and shear injuries. Distortion of the axonal cytoskeleton subsequently disrupts normal axonal transport mechanisms, leading to accumulation of transport products in injured regions24 and alterations in neuronal homeostasis. Until recently, the axonopathy that occurs subsequent to trauma has been characterized as the progressive formation of axonal varicosities within 2 to 3 hours after injury with disconnection by 6 to 12 hours.25 More recently, another form of axonal injury has been described characterized by increased axolemmal permeability, mitochondrial swelling, and cytoskeletal compaction damaging microtubule and neurofilament structures.26,27 Notably this form of injury is not associated with the characteristic axonal swelling associated histologically with injury, and reversed axonal transport has been considered as a possible mechanism for the observed disturbances in axon function without swelling.28 Shearing of axonal fibers leading to complete disconnection after trauma, or primary axotomy, can cause a more pronounced accumulation of transport products in the injury referred to as an “axonal bulb” (also referred to as a “retraction ball”) (Figure 3.1a). The appearance of these bulbs can appear in transected tissue for years following injury reflecting complete and persistent axonal separation,29 even when varicosities of presumably intact axons resolve, though a full understanding of the temporal course of this process remains to be done. Despite the apparent correlation of traumatic force magnitude with degree of axonal damage, primary axotomy is considered a minor contributor compared to secondary mechanisms in the overall axonopathy seen after TBI.6,30

FIGURE 3.1. (See color insert.

FIGURE 3.1

(See color insert.) (a) Silver-stained tissue of juvenile rat brain tissue at 40x following impact-acceleration TBI revealing axonal retraction bulb at center. (b) Silver-stained tissue of juvenile rat brain tissue at 40x following impact-acceleration (more...)

Stretch injury without complete axotomy is considered a greater contributor to the pathology seen in DAI than injury with axotomy. In vitro studies of neuronal preparations reveal that immediately subsequent to stretch injury axonal arrangement is distorted and some axons become undulated and twisted due to cytoskeletal damage (Figure 3.1b). Internodal regions of the axon appear particularly vulnerable, whether due to specific mechanical features of this region, a lack of association with supportive oligodendrocytes,30 or perhaps the density of transmembrane ion channels in this area.31 Changes to the cytoskeleton are prominent in the wake of trauma, as neurofilament morphology in some structures changes within 15 minutes after injury25 with subsequent disruption of axonal transport and accumulation of transport products. Though axonal transport is commonly mediated by the microtubule cytoskeleton as opposed to neurofilaments, neurofilament morphology is considered a mirror of concurrent microtubular integrity.25 Supporting this, a loss of microtubules following trauma has also been noted in models of TBI and axonal injury.32 Ultimately, the use of the microtubule stabilizing medication taxol ameliorates microtubule damage after experimental TBI,33 but cytotoxic effects of the drug preclude its use as a therapy for DAI at this time.34 TBI damage decreases axonal elasticity and prevents recovery of the axon to its original structural conformation.33 The individual axon response to injury appears idiosyncratic to some degree and governed by multiple factors. Though large-scale traumatic forces would likely render roughly uniform mechanical stresses across microscopic axon bundles, initial histological examination reveals scattered axonopathy in affected tissue. One possible explanation is that myelination varies between neighboring axons, and finely myelinated small axons appear more vulnerable to injury.25 Histologically visible findings likely underestimate injury in neighboring fibers with varying degrees of injury, as the proportion of histologically abnormal axons seen in trauma models and patients is incongruent with the magnitude of DAI symptoms seen after TBI. Also suggests some axons may appear morphologically normal but are functionally incapacitated after trauma.35 Due to the progressive nature of DAI early after trauma, imaging and histological findings are also significantly affected by time of assessment. Staining techniques such as hematoxylin and eosin and silver staining delineating structural features are of limited benefit as they may underestimate the burden of gross axonal pathology.10 Rather, immunohistochemical stains identifying axonal transport products such as amyloid precursor protein (APP), which can be identified in damaged axons within 2 hours after injury, may be more useful. Commonly seen in neurodegenerative diseases in older patients,36 APP positive damaged axons appear in high frequency in TBI patients who die soon after injury, even in children.37,38 Therefore importantly, other clinical pathologies affecting axonal transport should be considered in addition to injury mechanisms in the histological evaluation of DAI.39

SECONDARY AXOTOMY AND DISRUPTED NEURONAL HOMEOSTASIS

Traumatic brain injury is intrinsically heterogeneous and therefore axonal injury does not occur in a vacuum, rather multiple injury cascades from oxidative, excitotoxic, and inflammatory pathways occur as well and affect the evolution of axonal pathology. Following trauma, axons that have not been ruptured by primary axotomy may have been injured from physical stretch due to deformation of the brain. Axonal stretch injures cells through several mechanisms. In addition to direct cytoskeletal damage, stretch disrupts membrane permeability and precipitates depolarization.40 This in turn alters the electrochemistry of the damaged axon and triggers the release of excitatory neurotransmitters such as glutamate to concentrations as much as 50 times normal in TBI.41,42 This surge is worsened by dysfunction of normal glutamate reuptake processes by neighboring astrocytes in injury. Glutamate acting on NMDA and AMPA receptors increases cytosolic calcium influx and concentration. This in turn engages secondary messenger systems involving calcium, alters transmembrane ion gradients, and the osmotic load of the neuronal cytosol among other mechanisms that precipitate additional neuronal injury. Calcium ion increases seem dependent on two sources: (1) entry of sodium secondary to depolarization and reversal of the Na+/Ca2+ membrane antiporter31,43,44 and (2) release of calcium from the endoplasmic reticulum.45 Other possible sources of altered axonal electrophysiology include increased sodium ion channel permeability,31,44,46 decreased Na-K ATPase activity resulting in ATP starvation of the neuron,4749 and ectopic distribution of channels.50 Calcium induces changes in mitochondria leading to the opening of the mitochondrial permeability transition pore. This ultimately leads to neutralization of the electrochemical gradient essential for mitochondrial function, permits water influx, and instigates mitochondrial demise.51 Calcium-mediated proteolysis by calpain and caspases in turn leads to further damage to the axonal cytoskeletal as well as ion channel structures.43,5153 The effects of these calcium-dependent enzymes are demonstrated in the lysis of the cytoskeletal protein spectrin in disparate products of proteolysis characteristic for each enzyme. Calcineurin inhibition with cyclosporine A and tacrolimus and calpain inactivation has been demonstrated to mitigate axonal injury in vivo5358 and may offer potential therapies in the future. Oxidative stress and disrupted neuronal energy metabolism secondary to mitochondrial injury and eventual destruction also plays an important role in neuronal cell death. This is due to cytoskeletal injury, protein alteration, and metabolic failure, a process partially ameliorated by free-radical scavengers.59 Neuroinflammatory processes and microglial activation also contribute to local injury processes and endure long after the initial insult.29,60,61

Axons not severed in primary axotomy may proceed to breakage via secondary axotomy. Secondary axotomy is a process of rapidly progressive axonal deterioration, breakage, and retraction occurring following but not at the time of injury. Axonal bulbs form and the axon assumes a helical appearance as the proximal damaged axon retracts. Concurrently the severed distal axonal fragment begins Wallerian degeneration by 24 hours, a process of progressive lysis and disintegration.25 By 72 hours the damaged, separated segment is consumed by activated glia. Recent evidence suggests that unmyelinated axons are also at higher risk for secondary disconnection compared to myelinated axons as demonstrated in in vitro models where these smaller fibers are more prone to secondary disconnection than larger myelinated axons.62,63 Myelin sheath integrity changes during the evolution of DAI after injury, with progressive demyelination occurring in affected areas and white matter atrophy up to 1 year after injury.64 Of note, oligodendrocytes in apoptosis have been observed in affected tissue following TBI and may reflect myelin degeneration in progress. Appearance of persistent myelin basic protein (MBP) in the cerebrospinal fluid (CSF) of patients following trauma65 suggests an ongoing sheath degeneration consistent with the appearance of sheath globoids seen in conjunction with axonal bulbs after DAI.66

NEUROINFLAMMATION

The primary inflammatory response seen in the brain after DAI is mediated by microglia, as it is in other TBI. In a study by Oehmichen et al.67 immunohistochemical labeling with β-APP for axonal damage and CD68 for microglia identified colocalization of the labels in half of the patients who survived 5 to 15 days in areas of the brain—thereby demonstrating moderate microglial infiltration in areas of axonal injury.68 Microglia migrate rapidly into injured areas, and activated microglia extend cytoplasmic processes toward injured axons so as to isolate damaged structures. Thalamic infiltration with astrocytes has been noted at 4 to 8 hours after injury in animal DAI models, with maximal injury markers evident at 48 hours to 2 weeks after injury. Concomitant microglia activation has also been observed in the cortex and hippocampus at 4 hours, with upregulation of MHC Class II epitopes in white matter 24 hours following injury. Macrophages localize to the meninges and perivascular spaces within 24 to 48 hours and persist for up to 2 weeks, however, infiltration into the parenchyma is more limited.69

Cytokines involved in neuroinflammation include mediators commonly seen in TBI such as the IL-1 family, IL-6, IL-10, and TNF-α. The IL-1 family includes IL-1α, IL-1β, and IL-18, and is a well-studied constellation of cytokines that stimulate lymphocytes and macrophages, as well as trigger further inflammatory mediators. A study by Lu et al.70 demonstrates an increase in cortical IL-1α and β following an impact-acceleration TBI in rats. Hans et al.71 also using an impact acceleration model noted IL-6 activity increases 1 hour following histologically confirmed DAI peaking at 2 to 4 hours then receding to normal in 24 hours. It was noted that IL-6 mRNA and expression was highest in regions of axonal damage. IL-6 in particular is important in regulation of inflammation and activity of granulocytes, lymphocytes, and NK cells as well as inducing release of soluble TNFR and IL-1 receptor antagonists. In a fluid-percussion model of DAI, Kita and colleagues72 identified increasing TNF-α concentrations in the brainstem and corpus callosum over the first 3 hours after trauma. TNF-α is well known as a pro-inflammatory, proapoptotic cytokine with effects on macrophage/monocyte/NK cell stimulation, as well as secretion of platelet activation factor, ICAM, thromboxane A2, prostaglandin E2 as well as endogenous nitric oxide.69 TNF-α has been detected in lysosomes of microglia, astrocytes, and oligodendroglia in other models of DAI as well.7374 A factor that has been demonstrated to exacerbate the inflammatory response is hypoxia.74 Given that respiratory compromise can occur through impact apnea, mechanical asphyxiation, inhalation injury or other mechanisms associated with TBI, hypoxia compounding DAI likely influences injury evolution following moderate-severe TBI.

The leukocyte adhesion molecule ICAM-1 is another potential mediator of post-DAI secondary injury in that it is upregulated following DAI75,76 16 hours after injury, peaking 4 days after injury. Lymphocyte adhesion associated chemokines MIP-2 and MCP-1 are also elevated after focal TBI, however, only MCP-1 was increased after DAI in this experimental DAI model. MCP-1 is associated with monocyte recruitment from the bloodstream and interestingly MIP-2 is also a neutrophil attractant.

EVOLUTION OF DAI AND PLASTICITY

Historically, DAI has been described as a process that evolved in days to weeks following trauma, with peak neuropathology occurring at 1 to 2 days after injury.25 Histopathological and imaging evidence,77 however, suggests that trauma incites a protracted process of axonal degeneration and impaired transport over months in a progressive, Alzheimer’s-like neurodegenerative process.29,78 Likewise, the processes of axonal recovery, regrowth, and neuronal plasticity are poorly understood in the healing patient. After experimental DAI in mice NG2 positive oligodendrocyte precursors initially decrease, but then proliferate by three days after injury in the damaged corpus callosum and subventricular zone.79 Mature oligodendrocytes demonstrate a similar tendency in total numbers, which decrease during the first 3 days following injury but appear higher at 7 days after injury. Despite diffuse axonal damage, frank demyelination is not seen in the first week after injury, and redundant or disordered myelin may persist.80 It can be difficult to assess myelin-sheath-associated proteins in the tissue as a marker of injury due to retention and ongoing breakdown of damaged sheath.80 Regrowth of myelin with oligodendrocyte recovery likely influences other recovery and plasticity processes, however, the extent of this interaction is not well characterized.79 Basic recovery processes such as restoration of cytoskeletal structures with microtubule and neurofilament turnover and recovery of normal ionic homeostasis in the injured axon have been postulated,10 and it is possible that axonal varicosities seen with disrupted axonal transport following TBI can resolve with restoration of normal axonal function. The utility of biomarkers to predict recovery is unclear, as the thresholds for marker detection may involve injury beyond the possibility for axon recovery.40 Axonal regrowth is normally inhibited by resident myelin sheath proteins such as Nogo-A, oligodendrocyte-myelin glycoprotein (OMgp), and myelin-associated protein (MAP).68,81,82 Binding of these mediators to Nogo receptors (NgR) results in Rho A mediated inhibition of axonal growth and prevents axonal elongation. Though this is likely protective in normal neuronal architecture, it has been postulated that this process impedes axonal regeneration after trauma. Interference with the Nogo signaling pathway using a NgR inhibitor and other medications in in vitro and in vivo models releases axonal growth inhibition and has been shown to restore some recovery of spinal cord function after hemisection in a rat model.68 Further exploration in cerebral axonal disconnection is necessary to determine the manipulability of this pathway for treating DAI.77 Altogether parameters governing the extent and degree of these recovery processes, determinants of injured axonal survival or death, as well as their direct effect on outcome are important subjects for further investigation.83

SPECIAL CONSIDERATIONS: BLAST INJURY, CHRONIC TRAUMATIC ENCEPHALOPATHY, AND ABUSIVE HEAD TRAUMA

Interest in blast injury has increased markedly in recent times due to use of explosives as weapons and for breaching structures in theater. Blast injury causes a unique constellation of symptoms, including DAI, due to the high amount of kinetic energy transferred to patients after injury. Multisystem injury can be expected in addition to direct blast injury, including penetrating injury due to shrapnel, and respiratory injury from thermal and toxic exposure. Explosive blasts release energy in the form of acoustic, light, thermal, and electromagnetic energy, which can all potentially interact with a patient.84 A mathematically idealized blast wave initially develops with a near instantaneous increase in local air pressure to peak levels, followed by an exponential decrease in pressure that approaches nadir below baseline atmospheric pressure, and then recovers to baseline over roughly twice the time required to reach peak pressure.84 Blast forces can reflect off surfaces and redirect as compound waves toward the patient in a closed space, and lead to successive buffeting of the patient from multiple directions.84 The primary blast passes on the order of milliseconds through tissue and loses less energy perpetuating through noncompressible fluids and body tissue in comparison to air, which dampens blast forces. Additional injury results from physical displacement of the victim and subsequent contact with other structures as the force of the explosive accelerates the body. This type of additional trauma has been described as a “blast plus” component of injury, possibly incurring additional acceleration-deceleration injuries to the brain or other organs. Though helmets reduce direct deceleration injury in military and civilian trauma, they do not usually stop rotational acceleration/deceleration that is particularly injurious to axons.85,86 Chronic low level repetitive blast events may also cause some level of cognitive disturbance, as has been demonstrated in military door breachers.87 Progressive advances in military blast injury management highlight inefficiencies in civilian blast TBI care88 and have contributed to improvements in postblast DAI management.89,90

Chronic traumatic encephalopathy (CTE) was first described in boxers in 192891 and has received increasing attention due to increased recognition and association with contact sports prone to repeated head injury. Though predominantly described in American football and boxing, it has also been described in soccer, hockey,92,93 and mixed martial arts94 though the true burden of CTE is unknown in multiple sports. Because its pathophysiology is a long and usually slow progressive course evolving from frequent repetitive head injuries, it is frequently diagnosed well into its chronic phase when persistent long-term neurological and behavioral effects predominate. Brain histopathology performed on CTE-symptomatic ex-boxers postmortem reveals scattered intracellular accumulation of microtubule associated tau protein and neuropil threads9597 and similar findings have been found in postmortem examinations of professional American football players.98,99 Clinical findings include expected cognitive and neuropsychiatric disturbances with disinhibition,100 and focal neurological signs may include speech problems, ataxia, spasticity, and extrapyramidal symptoms.98 Recent advancements in imaging present an ominous picture of boxing, with 76% of boxers having imaging abnormalities consistent with DAI occurring in a dose-dependent fashion that correlates with career length and number of bouts.101 CTE is also unique as a TBI syndrome in that it rarely involves penetrating injury or skull deformation, and the pathology arises primarily from global acceleration-deceleration forces causing neuropathology primarily through DAI. Consequently, CTE is not well modeled by most current preclinical TBI models. CTE and Alzheimer’s disease (AD) bear similarities histopathologically in terms of progressive axonal degeneration. A number of studies appear to link the history of a single reported TBI with eventual AD, as well as acceleration of dementia onset.102112 Given that the apolipoprotein E4 (ApoE4) phenotype is associated with increased severity of AD pathology, TBI patients manifesting ApoE4 are also predisposed to future axonopathic neurodegeneration demonstrated in boxers with severe CTE.113

Abusive head trauma (AHT) in infants and young children has also been a topic of increasing interest due to the injury pattern and advances in fathoming its unique characteristics. AHT victims frequently suffer significant repetitive injury potentially over time, and a multitude of injury mechanisms in addition to axonal shear and focal lesions. Over 90% of patients who die of AHT also develop subdural hematoma.114 Mechanistically effects of a proportionally larger head with less neck control than an adult would seem to predispose children to greater acceleration-deceleration forces115 reinforced by adult data that brain injury is worse in adult patients when the neck is limp.116 This implicates axonal injury as an important component in the pathology observed. Stresses on the axon likely entail direct shear as well as internal stresses from centripetally displaced axonal cytoplasm within the axon course. This differs from direct contact injury such as that seen in contusion.1 Frequently a period of medical neglect follows injury, during which the AHT victim may become hypopneic from multiple mechanisms including depressed mental status and direct injury to medullary respiratory centers. Moreover, the physical exam may be complicated by a varied, underreported history of partially healed repeated injuries of the brain and other organs. A series by Geddes et al.96 examined 53 cases of inflicted head injury and found that whereas children over a year of age demonstrated DAI patterns similar to adults, infants under the age of 1 year manifested infrequent axonal pathology, however, vascular injury and cerebral edema was more prominent concomitant with anoxic injury. Thirty-one percent of patients do demonstrate axonal injury in the cervical cord or the craniocervical junction consistent with shaking, however, providing a causal relationship for medullary damage in AHT and associated respiratory failure.40 The overall low amount of cerebral axonal damage is puzzling given the magnitude of functional deficits seen, and some authors115 propose that myelination protects the mature axon from injury since most axonal damage when seen is in intermodal areas. Several have also postulated that the immature axon may be less susceptible to axonal damage or that shaking injuries mechanistically were insufficient to cause axonal injury.117,118

MECHANISTICS

Animal, fabricated plastic, and in silico models have all been instrumental in characterizing the mechanical forces that cause DAI pathology. As can be expected there are a multitude of external forces imposed on the brain in TBI leading to rotational, tensile, and compressive strains on the tissues. Moreover the size and complexity of the human brain are important factors as well, due to its inertia, relative freedom of movement on the cervical spine, and variable density of the grey and white matter resulting in separation at significant shear stress. Consequently, only an exact model of the human brain’s unique geometry and environment adequately models shear forces important for human TBI. The falx separating the cerebral hemisphere and tentorium in particular alters shear wave propagation in the brain and can amplify local forces on the brain near the falx’s insertion points.119121 Even at the cellular level cytoplasm is displaced by traumatic forces within the axon and can directly damage the cytoskeleton.25,122 In particular the threadlike structure of axons and small blood vessels in the brain expectedly places them at exquisite risk for shear injury in the brain under deformational stress. Though DAI is attributed to white matter injury, myelination increases the dynamic modulus of white matter axons and makes them stiffer. Per unit length, they are therefore, less susceptible to shear stress than corresponding grey matter structures, a factor balanced by white matter axons being longer and therefore more prone to DAI.119,123 Evidence suggests normal compliance of the cellular membrane and cytoskeleton permits some pliability in low velocity brain deformation. Under higher, rapidly applied stresses, however, axons become stiffer, brittle, and susceptible to shear.44,124127

The substance of the brain is analogous to a semisolid of heterogeneous densities and composition. Therefore, any model must account for this in simulating the shear forces that characterize DAI.1 When force is applied to a semisolid it is deformed in a manner consistent with its elasticity, viscosity, and plasticity. The tissue response to deformation subsequently depends on the duration and amplitude of the force application. When this force is distributed through the tissue by cells in contact with each other, shear forces develop within that cause permanent damage when the magnitude of shear exceeds the tissues’ elastic threshold. Longer durations of force application increase the amount of energy transferred to the brain and maintain a degree of physical displacement that rents the tissue, therefore causing more damage. Acceleration-deceleration at high speed or in repetitive shaking injuries can result in significant directional forces lasting over 1 second in duration and place the tissue in peril. Moreover high kinetic energy head injury frequently involves multiple directions of acceleration-deceleration that compounds shear forces on the brain. Given that the dynamic modulus, or tolerance of physical stress, of threadlike axons and vascular structures is direction dependent, compounded shear directions in the brain can potentially injure multiple vulnerable long axon pathways. In particular, angular movement causes greater injury than translational movement, and angular head motion in the coronal plane causes greater injury than sagittal movement.12 In repetitive injury, initial damage caused by DAI may also alter the compliance of affected tissue. In a study of rodent impact-acceleration DAI the modulus of injured axons decreased after initial injury, implying that on successive impacts greater deformation of axons occurs, thereby causing axonopathy.128

ANIMAL MODELS

The structural complexity of the human brain contributes to difficulty in modeling of the multitude of dynamic forces acting on it in DAI. Gyrencephalic models best represent the complex dynamics of TBI from an anatomical standpoint; apparatuses that consistently cause reproducible injury with high kinetic stresses are resource intensive and limited by the types of injury caused. Presently, lissencephalic models in smaller animals are more accessible to investigators due to their economy, reproducibility, and capacity for titration of injury magnitude.129

GYRENCEPHALIC MODELS

A nonhuman primate model of DAI was first described by Genarelli in 1982 utilizing high velocity acceleration-deceleration kinetics.12,130 Notably, the investigation also indicated DAI’s strong association with post-traumatic coma independent of mass effect from intracranial hemorrhage. In this inertial acceleration study apparatus, the animal body was restrained in a frame, whereas the head was secured to a rotating armature that caused rapid acceleration and sudden deceleration about the cervical axis through a designated plane. Though primates are reasonably analogous to human injury by virtue of similar brain structure and size (approximately 95 g), significantly higher rotational acceleration was required to cause the same expected mechanical shear seen in human TBI121,131 as the lighter brains of smaller animals cannot be accelerated by this apparatus to velocities necessary to yield axonal injury.132 Subsequent investigations in this laboratory have been performed on miniature swine, which have similarly sized gyrencephalic brains and are more accessible than primate animals.129 This model brings the head through a biphasic centroidal rotation of 100 degrees over 20 ms and yields diffuse lesions consistent with DAI. Even with extensive white matter injury in these models, ultimately consciousness was largely dependent on brainstem pathology suggesting location of insult is more important than degree of total injury.133

BLAST INJURY

Blast injury research in animal models has advanced considerably in the past decade pursuant to the extensive use of explosive devices in warfare. A number of models have been described, including open air blast where animals are arrayed around a central explosive,134 and models of enclosed vehicle spaces to simulate explosive attacks.135 Open field models, especially those that simulate vehicle or confined space situations where the blast environment is simulated, provide realistic situation-specific injury models.136 The most straightforward and commonly used apparatuses for blast injury, however, particularly for small animals, are blast tube models.137 Blast tubes produce calibrated blast peak pressure levels of 20 kPa-350 kPa and replicate the kinetics of blast overpressure to the brain.138 A compressed gas shock tube used to produce this type of injury (Figure 3.2d) consists of a compression chamber separated from an expansion chamber by Mylar sheets, which rupture at a predetermined pressure.139 Within the expansion apparatus an animal can be positioned at different angles to the shock wave direction to determine directional effects. Compressed air or helium is pressurized behind the Mylar membrane, and upon rupture the shock wave progresses down the expansion chamber toward the animal. Altogether this model is reproducible and relatively safe to investigators compared to explosive-driven models, though it does not replicate effects of heat and toxic organic compound containing exhaust on the subject. Explosive-driven blast tubes are more commonly used with larger animals models partly due to the amount of energy needed to replicate a larger blast.140 The explosive charge is placed in the tube prior to the expansion area naturally without a separation membrane.

FIGURE 3.2. (See color insert.

FIGURE 3.2

(See color insert.) Major experimental models of traumatic brain injury used in simulating DAI in lissencephalic animals: (a) controlled cortical impact, (b) impact acceleration, (c) fluid percussion, (d) compressed gas blast tube.

LISSENCEPHALIC MODELS

Lissencephalic models, primarily rodent models, benefit from simple mechanical designs, economy of scale, relative safety profile for investigators, and the freedom to work with genetically manipulated murine colonies. A disadvantage is that they do not adequately model complex human neuropathology after injury due to the lower cortical complexity seen in rodents.141 Certainly human craniospinal angle, brain and skull geometry, brain functional topography, gray/white matter mass ratio, and junctional surfaces, as well as cortical complexity are not adequately replicated in these models.

Controlled Cortical Impact Model

The Controlled Cortical Impact (CCI) model (Figure 3.2a) uses a pneumatic or electromagnetically driven impactor to deliver a direct percussive injury traditionally to an animal’s exposed dura or brain.132,142 The animal’s skull position in three axes is fixed by mouth and ear bars, and the subject can remain intubated through the injury to ameliorate trauma-induced apnea. However, this means the brain is not subject to global acceleration-deceleration forces seen in DAI permitted by free head movement. Injury magnitude can be titrated with adjustments in impactor velocity, displacement, and duration of deformation. Pitfalls of this injury include that it causes a focal injury primarily, with marked contusion and hemorrhage. It also violates the skull integrity, alters skull compliance, and likely changes intracranial pressure dynamics. Settings needed for large white matter injury also causes significant contusion and tissue destruction in overlying cortex. Modifications of this model to simulate diffuse injury have been employed using a rubber tipped impactor against a unrestrained, intact skull,143,144 as well as a flat impactor against the bregma of a partially restrained skull with blunt, rubber covered ear bars, though this results in skull fractures.79

Impact Acceleration Model

The impact acceleration or weight drop model (Figure 3.2b) involves letting a weight fall freely within a low friction guide tube to strike an animal’s head, typically a rodent, either midline between the lambda and bregma or laterally.141,145,146 In the midline approach there is a risk of skull fracture and concurrent contusion injury, and it is generally recommended that in this approach the skull is exposed and a metal helmet disc be affixed to the skull to distribute the impact. There are a variety of surfaces on which the rodent rests. Classically the device places the animal restrained on a foam surface of standard elastance. This permits free movement of the head with the weight drop in the direction of the weight’s fall. The animal should be restrained, as the weight impact can eject the animal from the device. Mass of the weight and distance of weight fall can be adjusted to vary degree of injury. Additionally the surface on which the rodent rests has been changed in some models to tissue paper or aluminum foil so as to alter recoil characteristics after impact to simulate other types of injury.73 Axonal injury is induced in the brainstem with this injury model using the midline approach, however, in distributions different from human DAI in part due to rodent brain and skull geometry. Pitfalls of this model include the potential for multiple strikes by the weight after initial contact, which can be avoided by pulling a brake line attached to the upper surface of the weight immediately after impact. Additionally, because the animal’s head is not secured to the surface there is the potential for inconsistency between trials if the head is not placed in the same location or if the weight trajectory is not consistent either due to poor fit within the guide tube or excessive distance between the guide tube and the skull.

Fluid Percussion Model

In the fluid percussion model (Figure 3.2c) the animal’s skull is exposed and trephinated for access for the injury device. Through a surgically inserted saline-filled reservoir continuous with the rodent’s CSF space, a percussive shock wave is directed through the closed system that distributes force over the cortical surface.147 This access point can enter the otherwise closed cranium either at midline or laterally.148,149 The magnitude of injury is titratable by altering the volume of saline displacement, though causes of error may include air in the fluid system or animal, or compliant structural weaknesses in the fluid system that diffuse impact energy.150 It is mechanistically more similar to contusion or extraaxial hemorrhage as opposed to DAI, though the model transfers kinetic energy to a broad area of brain and some axonal injury does occur.150 The fact that generalized acceleration-deceleration of the brain does not occur in this model is a limitation in replicating DAI, and axonal injury in this model occurs in peripheral areas, as opposed to the brainstem and other central areas.141

Linear and Angular Acceleration-Deceleration

Li and colleagues have described an adult rat model of combined linear (transverse) and angular acceleration-deceleration closed-head DAI.151 This novel device is composed of a rotating head helmet affixed to the animal using ear bars on the moving portion of the device. The helmet rotates through 75 degrees coronally as it is pneumatically driven over a geared track over 4.68 ms, also producing 1.57 cm of translational lateral motion. Following injury with this model, rats demonstrated APP positive staining in the corpus callosum, brainstem, subcortical and periventricular white matter, hippocampus, and thalamus, and notable subarachnoid hemorrhages were present.151 Electron microscopy and diffusion tensor imaging152,153 have also confirmed changes consistent with DAI in this model. At the time of this publication this device had only been described in use by a single group of investigators.

In Vitro DAI Models

Consistent replication of trauma forces on isolated neurons is also important in DAI research, and the majority of mechanistic findings in axonal injury have been elucidated in isolated cell preparations.150 In vitro preparations of neurons and associated glia may be prepared on elastic membranes that simulate primary injury via stretching of the elastic in one or more directions. A single axis stretch can be used to injure cell preparations and organotypic tissues such as prepared hippocampal slices154,155 in a pattern consistent with pure shear. A biaxial stretch pattern is more commonly described with equal stretch in two dimensions without isolating direction of stretch or shear. In these models, the elastic membrane can be placed over a percussion chamber that uses fluid or gas to rapidly displace the membrane in direction perpendicular to the plane of the membrane. A nonequal biaxial stretch model that stretches in one direction and shears in another can also be used.90 In these models combined cell culture preparations with neurons and glia, as well as organotypic slices, have been used.

Optic Nerve Stretch Model

Basic molecular mechanisms are readily simplified using the optic nerve stretch model, which mechanistically allows reproducible stretch injury in vivo along the major axis of parallel axon bundles in the second cranial nerve as a central nervous system structure. Histological analysis is facilitated by the straightforward structure of the large nerve, and maintenance of the animal postinjury permits long duration studies (as long as 3 months has been described).156 Much of the current understanding in the basic subcellular mechanisms of traumatic axonal stretch has been elucidated using this model in rodents.31,125 Pitfalls include anatomical differences between nerve and parenchymal white matter, differences in neighboring cell populations from the parenchyma, and the mechanistically simple single-direction stretch injury that does not incorporate shear across axons or oscillations.

BIOMARKER ASSAYS

The various pathological processes described earlier characterizing DAI evolution also produce an array of potential biomarkers for monitoring, some of which are temporally useful for monitoring specific processes. An advantage of using biological samples for assessment of brain injury is sample portability and typically a low risk to the patient. Characteristics of an optimal biomarker include temporal relevance to injury state, and a short half-life and processing time are important such that current status is adequately reflected. Another benefit of timely injury state reporting by a biomarker is that it can assess therapy effect. Other optimal characteristics importantly include specificity to the injury of interest.

Calcium-Dependent Proteolysis and αII Spectrin Breakdown Products

Calpain and caspase are calcium-dependent enzymes involved in cytoskeletal breakdown following TBI. With significant increases in intracellular calcium immediately following injury, calpain activity rises early after TBI. A second peak occurs with progression of secondary axonal injury processes to necrosis and apoptosis.157159 Some evidence suggests necrotic cell death may be related to glutamate exposure as well as environmental calcium concentrations with elevated glutamate and calcium concentrations favoring necrosis.160162 Caspases are also active following TBI reflecting apoptotic mechanisms initiated by intrinsic and extrinsic pathway mechanisms, as well as by oxidative stress. Caspase-3 activity rises between 8 and 12 hours after injury peaking at 24 hours before decreasing to lower levels by 3 days.

Axonal damage mediated by calpain and caspases is reflected in patterned breakdown of all spectrin breakdown characterized by known protease cleavage points, and calpain and caspase activity can be attributed based on predictable lengths of αII spectrin breakdown product (SBDP) produced after cleavage. αII spectrin is a component of the cytoskeleton and connects axolemmal components to the presynaptic terminal, stabilizing the nodal structure of myelinated axons. Calpain lyses the spectrin cytoskeleton into SBDP 145 and SBDP 150, and caspase-3 renders SBDP120. Consequently, the relative contribution of each of these calcium-dependent proteases makes cytoskeletal breakdown potentially distinguishable. SBDP increases in mice as quickly as within 15 minutes of insult163 and in rats peaks at 24 hours after injury with a progressive decline through 3 days after TBI.164,165 Following injury most calpain-generated SBDP is found associated with axonal bulbs, and can geographically be found in the neocortex and subcortical white matter at the gray–white matter junction. Concomitant decreases in intact αII spectrin can be seen in the corpus callosum at this time, recovering to normal titers by 7 days after experimental injury.166 In humans, SBDP is seen to rise within 6 hours postinjury and peak concentrations can be seen by 2 to 3 days after injury.167 A comparative study of SBDP concentrations after adult TBI in 38 patients demonstrated predominant elevations in SBDP 145 and 150 3 hours after injury, whereas SBDP 120 did not manifest as high an elevation after injury and suggested that calpain-mediated cytoskeletal breakdown predominates over caspase-3 processes after TBI.168 Further studies assessing other markers of comparative caspase and calpain activation are necessary to determine whether relative SBDP concentrations reveal differences in the proteases’ relative activity.

Neurofilament Markers

As neurofilament protein changes feature prominently in histological changes seen after TBI, the protein subunits that comprise these fibers are useful cytoskeletal injury markers. The NF heteropolymer is comprised of light chain (NF-L, 68 kDa), medium chain (NF-M, 160 kDa), and heavy chain (NF-H, 200 kDa) subunits.163,169,170 The subunits are transported within the axon, and in axonal injury accumulate in areas of disrupted transport.171173 Phosphorylation of the subunits influences the speed of transport and controls axonal structure and organization.174 Following TBI, NF sidearms compact progressively and weaken structurally as they undergo proteolysis by caspase and calpain.26 NF-L is one of the first subunits to be degraded within a few hours after injury. In pig and rat experimental models of TBI, phosphorylated NF-L decreases between 0.5 and 6 hours, as well as in the corpus callosum of TBI patients.175 Immunohistochemistry labeling NF-L demonstrates accumulation of NF-L in axonal bulbs and varicosities, becoming progressively more conspicuous between 1 and 3 days postinjury.176 Phosphorylated NF-H also decreases 1 day after injury and up to 80% in the corpus callosum of TBI patients.175 NF-H demonstrates a bimodal peak pattern at both 12 and 48 hours, likely reflecting primary and secondary injury effects on the cytoskeleton. Comparatively NF-M has not been demonstrated to be a TBI biomarker, demonstrating modest immunohistochemical detectability at 24 hours after injury, though this increased by 3 days.177

Amyloid Precursor Protein, Amyloid β, and Microtubule-Associated Protein Tau

Beta amyloid precursor protein (APP) is a transmembrane glycoprotein synthesized in neurons and plays a role in cell adhesion, growth, and response to injury.178 It has neurotrophic functions including promotion of axonal sprouting, neurite outgrowth, and synaptogenesis essential for repair after injury.16 Because it is transported in the axon via fast axonal transport mechanisms, it accumulates in areas of disrupted transport rapidly after injury.179 In animal models it shows up as early as 15 minutes after the initial insult. Consequently, it has been well described in surgical pathology specimens as a marker of axonal damage in TBI patients.36 Factors that may affect APP accumulation include rate of APP production24,180 and whether axonal transport is functional between the injury and the neuronal body such that accumulation is possible. Additionally, axonal transport velocity decreases with age and relative tissue concentrations of previously accumulated APP may vary.178 Even in trauma, APP only appears in affected tissue 30% to 50% of the time compared to aII spectrin breakdown products.181 Furthermore, it is somewhat nonspecific for DAI as it appears in patients with diffuse axonopathy from other causes such as metabolic182 and ischemic causes.179,183 The burden of APP positive lesions seen in the brain does not necessarily predict outcome as well, as a large prospective series of trauma patients did not find an association between lesions and 6-month outcome after TBI except in children and young adults.184 Therefore, an adequate clinical history is necessary to interpret APP if it is to be used as a TBI biomarker in infrequently obtained ex vivo tissue.

Amyloid β is a peptide that normally exists extracellularly in monomeric form. Following injury it polymerizes into plaques neurotoxic to neighboring tissue. Rapid escalation of the marker is seen in the brain within the first day following injury in experimental animals, and may continue to be seen through day 14.176 These plaques are readily identifiable on immunohistology and can last for years following insult, though regression is possible secondary to enzyme mechanisms for catabolism. A factor affecting appearance of the lesions is their perseverance, as amyloid β positive lesions that appear in areas of axonal injury after severe TBI, may regress over the first few years after injury.29 It is primarily used in TBI as an ex vivo histological marker of axonal injury, as serum concentrations are roughly 100-fold lower than in the CSF185 and do not change with injury.186

Microtubule-associated protein tau (MAP-tau) is a microtubule bound protein that plays a role in cytoskeletal organization and axonal transport. MAP-tau exists in different states of phosphorylation, and hyperphosphorylated tau is commonly associated with Alzheimer’s disease.16 After axonal injury and the initiation of microtubular cytoskeletal breakdown, MAP-tau is cleaved by activated calpain and caspase-3 into cleaved tau (c-tau) fragments that subsequently mediate amyloid β-induced tau hyperphosphorylation. These hyperphosphorylated tau complexes begin to accumulate in the tissue as neurofibrillary tangles and subsequently trigger apoptotic cell death.30 Rapidly after injury, soluble c-tau fragments can be found in the CSF in 1 hour postinjury and rise precipitously at 24 hours, with subsequent decline over 3 days.187189 CSF concentrations of c-tau may also be useful as an outcome marker, as higher quantities are negatively associated with functional outcome.187,188 In a comparison study of 28 severe TBI patients to medical neurology and nonneurology patients,187 TBI patients manifested c-tau increases of ~40,000-fold over concentrations in controls and correlated with Glasgow Outcome Score at discharge. Moreover, CSF concentrations of c-tau demonstrate axonal damage in patients with DAI with more sensitivity than a CT scan,187,188 however the biomarker in serum has not been shown to be similarly selective173,174 likely due to the blood-brain barrier (BBB).

Glial Fibrillary Acidic Protein

Glial fibrillary acidic protein (GFAP) is an acidic filament protein located within astrocytes and otherwise not found outside the central nervous system (CNS). Following experimental DAI in rats, increased GFAP mRNA expression has been noted for at least 11 days following injury.190 Following initial injury, it increases markedly to peak within the first 24 hours after injury and then progressively declines to baseline by approximately a week, though patients with unfavorable outcome still demonstrated higher GFAP concentrations than favorable outcome patients at 11 to 14 days.191 In several series, serum concentrations have been demonstrated to be predictive of ultimate survival,191194 as well as correlate with imaging findings. A comparison study of GFAP and S-100β in 92 severe TBI patients demonstrated that though both markers were comparable in their correlation mortality, GFAP was better able to discriminate between severe disability and persistent vegetative state than S-100β.194,195 However, a recent study assessing day 1 GFAP among other markers after all TBI requiring computed tomography (CT) imaging of the head did not find similar outcome discrimination.193 This may suggest that GFAP elevation may only be most informative in patients with sufficient brain injury as to have astrocyte injury from secondary hypoxia-ischemia or other mechanisms.

Myelin Basic Protein

Myelin basic protein (MBP) is a major constituent of the white matter and comprises approximately 30% of it. Closely associated in axonal sheath layers, it is readily released into the CSF following injury and can appear as early as 2 hours after TBI in rats.196 Following injury it is lysed by calpain, matrix metalloproteinases, and lysosomal proteases and by-products of its dissolution can be detected as well. Though MBP is readily found in the CSF, the BBB appears to impede its translocation into the serum and therefore it may be difficult to access as an acute indicator of axonal injury.65

S-100β

S-100β is a low molecular weight (10.5 kDa) acidic calcium binding protein found in glial and Schwann cells of white matter.197 It is unclear whether S-100β may be protective or detrimental, as it may stimulate glial repair processes but may also be linked to calcium influx.198 As a biomarker, clinical studies have found elevations in S-100β early in injury are tied to eventual functional outcome.199201 S-100β also cannot diffuse across intact BBB, however, and serum concentrations are heavily dependent on BBB permeability.202,203 Importantly, there is a potential for confounding in multisystem trauma with musculoskeletal injury as S-100β is also found in adipocytes and chondrocytes. Some trauma literature has yielded conflictual results regarding the utility of the marker in predicting outcome204207 particularly in severe extracranial injury.208 Therefore, careful interpretation is likely indicated in trauma patients with multisystem injury.

Neuron-Specific Enolase

Neuron specific enolase is housed in the cytoplasm of neurons and can also be found in circulating erythrocytes and platelets. Within the axon it travels via the slow axonal transport system. It is not normally seen extracellularly and can be found in injured tissue as soon as 1.5 hours after injury. Though hemolysis will contaminate serum and possibly hemorrhagic CSF samples with hematogenous NSE, corrections exist for correcting hemolysis effect on NSE in serum samples.209 As a biomarker NSE has correlates with Glasgow Outcome Score within the first year following pediatric TBI.210 NSE outcome correlations appear to best match early peak values, and therefore the value of NSE might be an assessment of magnitude of early injury in the patient. However, the marker’s half-life is 20 hours, and over the course of early injury it is likely not useful for revealing acute changes in status.

Ubiquitin Carboxy-Terminal Hydroxylase L1

Ubiquitin carboxy-terminal hydroxylase L1 (UCH-L1) is expressed in neurons as a small cysteine protease that hydrolyzes the C-terminal bond between ubiquitin and polypeptides. Several studies have demonstrated significant UCH-L1 increases in the CSF of severely injured patients within an hour after injury.193,211,212 Moreover, the magnitude of UCH-L1 increase was seen to be larger in patients with worse 6-month outcome and postinjury complications.211 The BBB also restricts translocation of UCH-L1 and therefore serum concentrations are highly dependent on BBB integrity, though serum concentrations are proportional to increases in CSF concentrations.197

NEUROIMAGING IN DIFFUSE AXONAL INJURY

With improvements in imaging technologies, radiographic analysis of the brain following TBI is gaining importance in outcome prediction, as has been demonstrated in predicative models based on head computed tomography.213,214 Neuroradiology data in the clinical setting is typically evaluated in qualitative terms,215 though this is fallible as critical pathology indicative of outcome is missed to more than an insignificant degree even among blinded expert neuroradiologists interpreting CT and magnetic resonance (MR) data in TBI patients.216,217

Computed Tomography of the Brain

Head CT frequently does not identify pathology associated with DAI.218 Only 10% of patients with DAI demonstrate hemorrhagic punctate lesions of the corpus callosum and gray–white matter junctions of the cerebrum and pontine–mesencephalic junction near the cerebellar peduncles. Weeks after injury, atrophic changes may occur in the white matter, dependent on the degree and topography of injury. Areas where atrophy may appear more prominent include areas of high white matter concentration and periventricular parenchyma recedes and ventriculomegaly may develop219 without sulcal effacement or other evidence of hydrocephalus. Clear advantages of neuroimaging with CT remain, however, particularly in terms of image acquisition speed and scanner accessibility in most acute trauma environments.

Conventional Magnetic Resonance Imaging (MRI)

Like CT, MR neuroimaging is based on tomographic reconstruction of the brain in imaging slices acquired sequentially, but in the case of MR, data is reconstructed from the magnetic resonance properties of hydrogen atoms in a powerful magnetic field. MRI is more sensitive for the diffuse and physically small pathology found in DAI as opposed to head CT.216,220,221 Shearing forces associated with DAI transect small blood vessels running in parallel with axons, resulting in detectable microscopic hemorrhages forming in these areas. Gradient echo and Image T2star.jpg (T2 star) weighted MR imaging detects signal dropout caused by iron-containing heme groups in slow-moving blood. Distributions of these microhemorrhages in areas associated with axonal injury such as the corpus callosum, brainstem, and other white matter tracts strongly suggest an imaging diagnosis of DAI. DAI microhemorrhages typically appear as punctate signal-free lesions in the white matter that “bloom” or appear slightly larger than their true anatomic size due to iron-induced magnetic field distortion. Consequently, signal loss caused by punctate hemorrhages from DAI can be visualized for years after injury though lesions fade over time.222 Confounders that may appear to be microhemorrhages include air and calcium deposits; however, these are uncommon in areas affected by DAI. Density of Image T2star.jpg lesions have been associated with severity of injury in terms of admission GCS223225 as well as maximum ICP during admission and 3 month Glasgow Outcome Score.226 An additional factor is magnet field strength, as 3 Tesla MRI demonstrates practically twice the sensitivity of 1.5 Tesla devices for microhemorrhages,227 and using different MRI devices utilizing different pulse sequence methods, even of the same field strength between manufacturers, can complicate interpretation between serial exams.

Punctate areas of T2 signal hyperintensity in the white matter and at gray-white matter junctions in the frontal or occipital lobes are also suggestive of DAI (Figure 3.3).217,228 These nonhemorrhagic lesions are optimally visualized in high-resolution T2-weighted sequences with or without fluid attenuated inversion recovery (FLAIR) suppressing CSF signals. Even in the absence of microhemorrhages conventional MRI has been demonstrated superior to CT in evaluating DAI.229 With regard to outcome, nonhemorrhagic lesions may not necessarily be associated with worsening outcome suggesting different mechanisms for their appearance.229 In terms of longterm follow-up imaging, atrophy and gliosis as a consequence of DAI may be also be evaluated with conventional MRI.

FIGURE 3.3. MRI imaging of a 2-year-old patient with DAI after being ejected from a motor vehicle.

FIGURE 3.3

MRI imaging of a 2-year-old patient with DAI after being ejected from a motor vehicle. T2-weighted FLAIR imaging appears on the left, with corresponding susceptibility weighted imaging (SWI) slice appearing on right. Punctate areas of increased T2 signal (more...)

Susceptibility Weighted Imaging

Similarly to Image T2star.jpg weighted imaging, susceptibility weighted imaging (SWI) is a highresolution three-dimensional imaging sequence that produces images based on local magnetic field aberrations.230 However, it is much more sensitive than Image T2star.jpg, detecting traumatic microhemorrhages at roughly 6 times the sensitivity of Image T2star.jpg, and renders hemorrhage volume approximately twice the size seen on conventional MRI gradient echo imaging.230 Though the physics of the sequence are different, it similarly accentuates hemorrhage, calcification, air, or other features causing aberrations in magnetic susceptibility. In patients with severe DAI and resultant cerebral edema, venous stasis in SWI is an ominous sign and prominently appears as dark areas of susceptibility artifacts outlining engorged veins coursing through the brain parenchyma. In severe cerebral edema and associated venous stasis, delayed transit of venous blood causes pronounced SWI signal decrease in the distribution of medullary cerebral veins, appearing as irregular radially distributed rays of signal loss through the white matter, and is described as diffuse vascular injury (Figure 3.4).231

FIGURE 3.4. SWI revealing diffuse vascular injury in an infant after abusive head trauma, appearing as serpentine areas of dark absent signal in the right occipital lobe.

FIGURE 3.4

SWI revealing diffuse vascular injury in an infant after abusive head trauma, appearing as serpentine areas of dark absent signal in the right occipital lobe.

Diffusion Weighted Imaging (DWI) and Diffusion Tensor Imaging (DTI)

Diffusion imaging is performed with serial acquisition of signal data after an initial coding from radiofrequency pulse, thereby identifying fluid movement within a magnetic gradient. DWI has demonstrated improved detection of nonhemorrhagic lesions over conventional MRI in areas of cytotoxic or vasogenic edema. In a series of 74 adult DAI patients, DWI-derived apparent diffusion coefficient (ADC) imaging revealed a higher burden of high intensity ADC signals in the corpus callosum and brainstem, and the magnitude of abnormal signal correlated with duration of coma.232 In children, identification of deep frontal and temporal white matter and basal ganglia diffusion abnormality was associated with poorer outcome in a subset of severe pediatric TBI patients.233 DTI is a further modification of this imaging technique where multiple diffusion vectors are measured in three dimensions, and diffusion magnitude can be determined in multiple directions for a given cube-shaped voxel of space in the brain. Data is aggregated from individual voxels arranged in a threedimensional grid representing the brain in imaged space, with each voxel typically measuring approximately 2 to 3 mm on a side and between 6 and 64 or more diffusion directions measured. Each voxel is assigned vectors measuring a magnitude of diffusion in three perpendicular cardinal directions, with the principal vector being the major axis of diffusion within the voxel space (Figure 3.5). Regarding utility of DTI, data continues to emerge234,235 on its utility in TBI and it is more sensitive to change following TBI than high-resolution conventional MRI.236 Multiple metrics have been derived for interpreting DTI data. Axial diffusivity is the magnitude of the major axis of diffusion inside the voxel and may indicate relative integrity of fluid filled tubular structures such as axons.218,237 Radial diffusivity is the mean magnitude of the two minor axes perpendicular to the major axis, and some authors suggest it may reflect integrity of tubular structures, such as the myelin sheath of axons.236 Fractional anisotropy (FA) is an approximation of the overall directionality of flow within a voxel, and a value between 0 and 1 is derived from the three eigenvectors of a voxel, with higher values reflecting greater anisotropy or directionality. Given that a standard clinical voxel averages a volume of approximately 8 mm3 and may contain several hundreds of thousands of neurons, these presumptions are potentially prone to inaccuracy particularly in regions with large numbers of crossing structures or varying axon density.238,239 Despite data suggesting the utility of DTI in describing anatomical changes associated with outcome in significant DAI, DTI have not consistently shown similar significance in mild TBI.240,241 Tractography is an additional postprocessing technique that utilizes DTI data and “seeds” placed in white matter regions of interest in the brain to track approximated white matter tracts. From these seeds and user-determined tracing thresholds, diffusion tensors in each voxel are used to trace out estimated diffusion tracts. When seeds are placed within white matter, these tracts are presumed to represent bundles of continuous axons. DTI and tractography data must be interpreted considering operator seed placement as well as the effects of multiple cellular entities in addition to axons on diffusion, such as astrocytes, crossing fibers, and vascular structures among others.242 It is worth considering that DTI metrics generalize the myriad diffusivities of hundreds of thousands of cells in a single voxel into just three eigenvectors.

FIGURE 3.5. (See color insert.

FIGURE 3.5

(See color insert.) Diffusion tensor imaging images of relative fractional anisotropy in an 8-year-old patient after TBI from being struck by a motor vehicle. The left grayscale image reflects FA magnitude and the right image incorporates direction data, (more...)

Strategies have emerged for improving DTI data resolution, including adjustments for complex fluid dynamics in biological systems, such as in diffusion kurtosis imaging (DKI).243,244 Other methodologies involve increased sampling of diffusion directions such that multiple principal vectors can be derived, as seen in high angular resolution diffusion imaging (HARDI)245 and diffusion spectrum imaging (DSI),246 which permit high-resolution tractography as seen in high-definition fiber tracking (HDFT).247 The utility of these methods in assessing DAI in heterogeneous TBI patients will rely on future imaging investigations. Variability between patients is important in clinical practice given accurate postinjury DTI interpretation requires some estimate of preinjury DTI conditions. A number of studies have indicated reduced anisotropy (evaluated with FA) in DAI-vulnerable white matter voxels (genu and splenium of the corpus callosum, internal capsule) is associated with postinjury cognitive deficits,248254 however, significant sample and methodology variance between these studies preclude their generalizability.255,256 Deciphering the predictive accuracy of DTI on TBI patient outcome therefore is still an important topic of ongoing research. This may be more challenging in DAI than in focal injury patients, since an unaffected hemisphere can often serve as a control in focal injury patients.247 In contrast, DAI patients will likely require aggregate DTI atlas data from large-scale population studies to generalize DTI findings.257 Presently, this information is in development,257,258 particularly in children where myelination steadily progresses throughout neurodevelopment259 and DTI metrics vary over brain maturation. Moreover the effects of age, subacute or chronic medical conditions on DTI findings over the human lifespan are still being elucidated. Therefore, interpreting abnormal findings in trauma patients across the spectrum of age will require a much more complete picture of normal DTI than is currently available.260

Magnetization Transfer Imaging

Magnetization transfer imaging (MTI) identifies chemical alterations in tissue caused by injury, based on associated changes in tissue magnetic moment. The protons of large organic macromolecules are selectively saturated by a radiofrequency pulse that is off-resonance from free water protons,261 and the magnetic moments of these macromolecule protons are “transferred” to water surrounding the tissue. These free water protons are interrogated thereby producing the MTI data.261 The methodology appears to be useful in TBI, as areas of decreased magnetization transfer ratio (MTR) have been identified in the white matter of trauma patients in the absence of findings on conventional MRI.262267 MTI in TBI patients also demonstrates abnormal findings in trauma patients with ongoing cognitive deficits despite having “normal” conventional MRI sequences.262,264 Histological confirmation of MTI-related lesions in animal models has been performed,263 though the exact physiology underlying these changes is not clear and likely multifactorial.261 A study describing MTI and MR spectroscopy of the corpus callosum for long-term follow-up of TBI patients has also been performed, revealing a lack of association between MTI and outcome in this investigation, though proton MRS demonstrated that decreased N-acetylaspartate to creatinine ratio (as an indicator of decreased mitochondria from neuronal loss) might suggest poor outcome. Pitfalls of MTI include lengthy study time and susceptibility to motion artifact, as well as qualitative interpretation susceptible to variability between interpreters.261

Morphometric and Volumetric Analysis

Morphometric analysis entails isolation and rendering of specific brain structure volumes by virtue of their contrast from surrounding tissue, and can be based on automatic parcellation or manual three-dimensional tracing based on obtained MR images. Morphology and volumes of these structures in turn reveal disease effects, and is ideally suited to analyzing both the direct physical injuries incurred by TBI as well as evolution of brain appearance over time. Morphometric analysis has associated anatomic injury with specific deficits268271 and readily captures volume changes in atrophic structures as they evolve after injury. A study examining volumetric analysis based on T2-weighted FLAIR images in 24 adults with severe DAI found volume of affected white matter correlated significantly with eventual GOSE (Glasgow Outcome Scale-Extended) at 6 months.272 Strangman and colleagues273 also found that longitudinal changes in the volume of high connectivity areas such as the hippocampus, thalamus, lateral prefrontal cortex, and cingulate cortex are predictive of neurorehabilitation outcome. In a more topology oriented approach, Kinnunen and colleagues274 utilized tract-based spatial statistics (TBSS) to characterize risk of specific deficits after TBI based on white matter changes. Given that morphometry is highly dependent on accurate tissue definition, it is easily confounded by measurement error. Edema in particular affects the appearance of structures in both CT and MRI, and the heterogeneous nature of TBI and DAI makes parcellation of structures difficult for both manual and automated systems especially when the brain is severely deformed from pathology.275

Functional MRI

Functional MRI (fMRI) findings are based on the detection of elevated blood oxygen level dependent (BOLD) signals presumably in regions of the brain parenchyma with increased metabolic demand.276,277 Determination of where the brain is most active is based on relative oxygen content differences detected in blood delivery and return. In trauma, frontoparietal activity in the brain is attenuated in DAI survivors compared to controls in working memory tasks, particularly in the right middle and superior frontal lobe cortex,278 and fMRI may identify functional compensation and forecast response to learning strategies in patients rehabilitating from DAI.144,279 Over time, fMRI has been demonstrated to reveal compensatory regional activation well into recovery after TBI.280,281 There is enthusiasm among imaging specialists275 that identification of active brain combined with DTI structural data will permit connectivity mapping of the brain as it recovers from injury. In the acute phase, issues with fMRI include that BOLD signals can be confounded by blood pooling from other sources in trauma and also perfusion effects caused by increased ICP. With careful consideration of these effects fMRI may prove a useful methodology for profiling outcomes of brain recovery after DAI in the future.

ELECTROPHYSIOLOGY AND MAGNETOENCEPHALOGRAPHY

DAI’s effects on long axons and cerebral connectivity affects neural electrophysiology after injury, manifest in evoked potentials, electroencephalography, as well as magnetoencephalography (MEG). As a functional test of brain activity, MEG records magnetic flux at the head surface caused by neuronal depolarization in shallow cortical structures below the scalp.261 The methodology has high temporal and spatial resolution and is a promising means of identifying abnormal cortical areas. In TBI patients, MEG demonstrates changes such as reduced background activity persistently after injury.282,283 Given that abnormal EEG activity (and therefore MEG as well) in gray matter may be a consequence of white matter de-afferentation by axonal damage,284 regionally abnormal MEG findings may identify affected cortical areas after DAI. In a case series of 10 mild TBI patients with no CT or MR radiographic evidence of injury in nine, combined analysis using DTI and MEG revealed gray matter MEG lesions topographically near areas of reduced anisotropy identified with DTI.283 MEG machines are not common even in large medical centers, however, and clinical expertise in interpretation, as well as clinical access for patients who cannot easily get into the device, may be challenging. Appropriate MEG helmet sizing for pediatric patients may also be challenging as there will be increased distance between a child’s skull and biomagnetometer coils compared to an adult’s head.285

Electrophysiologic methods identifying altered evoked potentials in DAI patients may also provide insight into patterns of disrupted cerebral connectivity following injury. DAI survivors may demonstrate increased resting corticospinal motor thresholds necessary to initiate MEPs, and reduced area under the MEP waveform indicating damaged cerebral connectivity.286 In a study of 52 adult survivors of TBI associated with motor vehicle accidents and chronic functional impairment from DAI, DTI of the corticospinal tract, and motor evoked potential (MEP) evaluation with transcranial magnetic stimulation,287 revealed that patients who did not generate MEP readings in response to stimulation had lower corticospinal tract FA than DAI patients who did generate MEPs as well as uninjured controls. Though electrophysiology is informative, imaging may provide better lesion localization not captured using these electrophysiological methods. Further research may suggest whether serial evoked potential data is informative regarding functional recovery after TBI.

THERAPEUTICS

Calcineurin Modulators

As mentioned earlier, calcineurin (CN) modulators are a potential therapy for axonal injury by virtue of CN’s responsiveness to calcium as a calcium/calmodulin phosphatase, and possible role in calcium-induced axonal damage following TBI and excitotoxic injury. Cyclosporin A (CsA) initially received attention as a axonal injury modulator in part because its capacity suppresses calcineurin-mediated formation of the mitochondrial transition pore, averting mitochondrial failure and subsequent apoptotic cell death. In a rodent model of impact acceleration DAI CsA administration 30 minutes after injury resulted in decreased loss of mitochondrial N-acetylaspartate, and treated animals had higher ATP production than untreated controls signifying mitochondrial preservation.288 Histologically CsA pretreatment decreases APP positive axonal swellings in rats following impact acceleration DAI.56,57 Time dependency of CsA dosing in relationship to TBI seems important as well, given that administration of CsA 1 hour after midline fluid percussion injury in rodents worsens axon function determined by compound action potential (CAP), and in comparison when CsA was given 15 minutes postinjury, CAP is preserved.289 Toxicity of CsA has also been a concern as a DAI therapy since it can cause life-threatening seizures, and drug delivery to the brain is restricted by intact BBB thereby making brain concentrations in TBI unpredictable.

Tacrolimus (FK506) has also been investigated in multiple preclinical series as a potential TBI and DAI therapy, and histopathological evidence postinjury suggests a beneficial role for the agent in preserving axonal transport and reducing APP accumulation.55,290,291 Functionally, however, the role of tacrolimus does not affect neurofilament sidearm compaction in the progression to axonal disconnection54 and therefore does not protect against all histopathological effects of axonal injury. Additionally, despite beneficial effects of tacrolimus in juvenile animal DAI models, reducing histological axonal damage, it functionally did not facilitate recovery of CAP deficits in a modified CCI model for juvenile rat DAI.292

Stem Cell Therapy

The limited regenerative capacity of the brain makes the potential for stem cell therapy very attractive and the concept of introducing pluripotential cells into the CNS to regenerate brain has been explored sporadically since the end of the nineteenth century.293 In addition to possibly reconstituting damaged neurons, astrocytes, and oligodendrocytes among other brain cells, stem cells could also potentially be neuroprotective by maintaining the penumbral cell milieu through trophic, anti-inflammatory, direct chemotactic, or other mechanisms, thereby limiting secondary injury.293 Stem cell delivery routes vary, as intracerebral administration optimizes transplant localization but poses an invasive risk, intravenous delivery suffers from pulmonary first-pass sequestration and poor target specificity, and intraarterial delivery may result in pathologic emboli.294 Multiple studies have been performed demonstrating disparate results in animal models utilizing varied methodologies, but stem cell differentiation into neurons appears rare compared to other cell types.295 Recently, Riess and colleagues intracerebrally transplanted murine embryonic stem cells in rats following a fluid-percussion model of DAI, and found that functional outcome appeared improved, but only 1 of 10 animals at 7 weeks retained transplanted cells, while 2 animals developed tumors.296 Stem cells from more restricted lineages may also be practical and have less tumorigenic potential.7 Examples include neurogenic stem cells derived from remnants of primitive embryonic rests in the walls of the ventricular system (subventricular zone295) and hippocampus among other structures,293 or bone-marrow-derived stem cells.294 Another potential strategy is the use of oligodendrocyte precursor cells to target remyelination of injured axons.297 In humans, a clinical trial administered human embryonic stem cells intracranially to a cohort of comatose patients and reported some improved functional outcomes compared to a matched group of comatose controls.298 In this study, coma etiologies included 15 DAI patients of the total 25 treated patients, some with signs of frontal lobe atrophy on MR imaging. Further clinical investigations into human patient safety and functional outcomes after neurological stem cell therapies are necessary to determine the scope of utility of stem cell therapy in the clinical setting.

Recombinant Human Erythropoetin

Erythropoetin (EPO) as an erythrogenesis stimulator is found in the CSF throughout the life cycle and likely has a role in the basic homeostasis of the CNS.299 The role of EPO in the CNS is unclear, but it appears to reduce neuronal apoptosis possibly by EPO-receptor initiated autophosphorylation of Jak-2 and subsequent activation of antiapoptotic pathways through PI3K-Akt/protein kinase B/bcl-2, NF-KB, and RAS mechanisms among others.300 Animal models have demonstrated improved preservation of neuronal architecture in experimental TBI treated with EPO,301,302 and improved in-hospital mortality has been demonstrated in a retrospective clinical TBI study303 and trauma subset of a Phase II trial on critically ill adults.304 Following this Phase II trial, a Phase III trial of EPO administration in ICU patients also demonstrated a significant reduction in mortality (adjusted hazard ratio 0.40 at day 140, 95% CI 0.23–0.69), however, a significant increase in incidence of thrombotic complications accompanied this.305 An additional issue potentiating side effects is that the intact BBB impedes EPO diffusion into the brain and it is difficult to predict optimal drug dosing in TBI.299 Despite animal TBI model evidence of axonal preservation after injury with EPO, EPO has been implicated in inducing neuronal calcium ion influx, and therefore could theoretically potentiate calcium-mediated neuronal injury.306,307

Docosahexanoic Acid

Docosahexanoic acid (DHA) is well known as an omega-3 fatty acid, major constituent of the brain and neuronal phospholipid membranes, and potential anti-inflammatory specifically decreasing cyclooxygenase activity.308,309 In an impact acceleration rodent model of DAI, rats supplemented with 10 or 40 mg/kg/d DHA containing feed on postinjury day 1 manifested significantly lower APP positive stained corticospinal tract axons and decreased caspase-3 activation than untreated TBI animals, with observed concentrations in treated animals similar to those seen in uninjured animals.308,309 Functional and outcome studies have not been performed in DAI models, but in a clinical trial could be readily implemented in compatible TBI management protocols permitting early feeds.

SUMMARY

In total DAI is a complex process that merely begins with stretching and shearing of axons, at times proceeding to a persistent syndrome of cerebral disconnection and longstanding functional impairment. Research has improved our understanding of the mechanisms of axonopathy in DAI, its biomarkers, and potential therapeutic targets to treat it. Rapid technological innovations in biomarkers and neuroimaging, especially in MRI, are increasing our understanding of DAI’s evolution following TBI, however, there remain significant hurdles with regard to balancing data granularity and generalizable interpretation methods given the heterogeneous nature of TBI. With increasing scientific attention on brain connectivity, a working knowledge of DAI and its potential future therapies will also likely see significant advances in the near future.

REFERENCES

1.
Meythaler J.M. et al. Current concepts: Diffuse axonal injury-associated traumatic brain injury. Arch Phys Med Rehabil. 2001;82(10):1461–1471. [PubMed: 11588754]
2.
Adams J.H. et al. Diffuse axonal injury due to nonmissile head injury in humans: An analysis of 45 cases. Ann Neurol. 1982;12(6):557–563. [PubMed: 7159059]
3.
Povlishock J.T. et al. Axonal change in minor head injury. J Neuropathol Exp Neurol. 1983;42(3):225–242. [PubMed: 6188807]
4.
Povlishock J.T, Katz D.I. Update of neuropathology and neurological recovery after traumatic brain injury. J Head Trauma Rehabil. 2005;20(1):76–94. [PubMed: 15668572]
5.
Smith D.H, Meaney D.F. Axonal damage in traumatic brain injury. Neuroscientist. 2000;6(6):483–495.
6.
Christman C.W. et al. Ultrastructural studies of diffuse axonal injury in humans. J Neurotrauma. 1994;11(2):173–186. [PubMed: 7523685]
7.
Li X.-Y, Feng D.-F. Diffuse axonal injury: Novel insights into detection and treatment. J Clin Neurosci. 2009;16(5):614–619. [PubMed: 19285410]
8.
Corbo J, Tripathi P. Delayed presentation of diffuse axonal injury: A case report. Ann Emerg Med. 2004;44(1):57–60. [PubMed: 15226709]
9.
Gieron M.A, Korthals J.K, Riggs C.D. Diffuse axonal injury without direct head trauma and with delayed onset of coma. Pediatr Neurol. 1998;19(5):382–384. [PubMed: 9880145]
10.
Johnson V.E, Stewart W, Smith D.H. Axonal pathology in traumatic brain injury. Exp Neurol. 2013;246(C):35–43. [PMC free article: PMC3979341] [PubMed: 22285252]
11.
Adams H. et al. Diffuse brain damage of immediate impact type. Its relationship to “primary brain-stem damage” in head injury. Brain. 1977;100(3):489–502. [PubMed: 589428]
12.
Gennarelli T.A. et al. Diffuse axonal injury and traumatic coma in the primate. Ann Neurol. 1982;12(6):564–574. [PubMed: 7159060]
13.
Gaetz M. The neurophysiology of brain injury. Clin Neurophysiol. 2004;115(1):4–18. [PubMed: 14706464]
14.
Meyers C.A. et al. Early versus late lateral ventricular enlargement following closed head injury. J Neurol Neurosurg Psychiatry. 1983;46(12):1092–1097. [PMC free article: PMC491773] [PubMed: 6607318]
15.
Wilson J.T. et al. Early and late magnetic resonance imaging and neuropsychological outcome after head injury. J Neurol Neurosurg Psychiatry. 1988;51(3):391–396. [PMC free article: PMC1032866] [PubMed: 3361330]
16.
Blennow K, Hardy J, Zetterberg H. The neuropathology and neurobiology of traumatic brain injury. Neuron. 2012;76(5):886–899. [PubMed: 23217738]
17.
Lipton M.L. et al. Multifocal white matter ultrastructural abnormalities in mild traumatic brain injury with cognitive disability: A voxel-wise analysis of diffusion tensor imaging. J Neurotrauma. 2008;25(11):1335–1342. [PubMed: 19061376]
18.
Niogi S.N. et al. Extent of microstructural white matter injury in postcon-cussive syndrome correlates with impaired cognitive reaction time: A 3T diffusion tensor imaging study of mild traumatic brain injury. AJNR Am J Neuroradiol. 2008;29(5):967–973. [PMC free article: PMC8128563] [PubMed: 18272556]
19.
Wilde E.A. et al. Diffusion tensor imaging of acute mild traumatic brain injury in adolescents. Neurology. 2008;70(12):948–955. [PubMed: 18347317]
20.
Baxter D. et al. Pituitary dysfunction after blast traumatic brain injury. Ann Neurol. 2013;74(4):527–536. [PMC free article: PMC4223931] [PubMed: 23794460]
21.
Jeong J.H. et al. Negative effect of hypopituitarism following brain trauma in patients with diffuse axonal injury. J Neurosurg. 2010;113(3):532–538. [PubMed: 19943735]
22.
Sundaram N.K, Geer E.B, Greenwald B.D. The impact of traumatic brain injury on pituitary function. Endocrinol Metab Clin North Am. 2013;42(3):565–583. [PubMed: 24011887]
23.
Richmond E, Rogol A.D. Endocrine. 2013. Traumatic brain injury: Endocrine consequences in children and adults. [PubMed: 24030696]
24.
Singleton R.H. et al. Traumatically induced axotomy adjacent to the soma does not result in acute neuronal death. J Neurosci. 2002;22(3):791–802. [PMC free article: PMC6758486] [PubMed: 11826109]
25.
Povlishock J.T, Christman C.W. The pathobiology of traumatically induced axonal injury in animals and humans: A review of current thoughts. J Neurotrauma. 1995;12(4):555–564. [PubMed: 8683606]
26.
Pettus E.H. et al. Traumatically induced altered membrane permeability: Its relationship to traumatically induced reactive axonal change. J Neurotrauma. 1994;11(5):507–522. [PubMed: 7861444]
27.
Farkas O, Povlishock J.T. Cellular and subcellular change evoked by diffuse traumatic brain injury: A complex web of change extending far beyond focal damage. Prog Brain Res. 2007;161:43–59. [PubMed: 17618969]
28.
Stone J.R. et al. Impaired axonal transport and altered axolemmal permeability occur in distinct populations of damaged axons following traumatic brain injury. Exp Neurol. 2004;190(1):59–69. [PubMed: 15473980]
29.
Chen X.-H. et al. A lack of amyloid beta plaques despite persistent accumulation of amyloid beta in axons of long-term survivors of traumatic brain injury. Brain Pathol. 2009;19(2):214–223. [PMC free article: PMC3014260] [PubMed: 18492093]
30.
Li J. et al. Biomarkers associated with diffuse traumatic axonal injury: Exploring pathogenesis, early diagnosis, and prognosis. J Trauma. 2010;69(6):1610–1618. [PubMed: 21150538]
31.
Stys P.K. Anoxic and ischemic injury of myelinated axons in CNS white matter: From mechanistic concepts to therapeutics. J Cereb Blood Flow Metab. 1998;18(1):2–25. [PubMed: 9428302]
32.
Maxwell W.L, Povlishock J.T, Graham D.L. A mechanistic analysis of nondisruptive axonal injury: A review. J Neurotrauma. 1997;14(7):419–440. [PubMed: 9257661]
33.
Tang-Schomer M.D. et al. Mechanical breaking of microtubules in axons during dynamic stretch injury underlies delayed elasticity, microtubule disassembly, and axon degeneration. FASEB J. 2010;24(5):1401–1410. [PMC free article: PMC2879950] [PubMed: 20019243]
34.
Adlard P.A, King C.E, Vickers J.C. The effects of taxol on the central nervous system response to physical injury. Acta Neuropathol. 2000;100(2):183–188. [PubMed: 10963366]
35.
Tomei G. et al. Morphology and neurophysiology of focal axonal injury experimentally induced in the guinea pig optic nerve. Acta Neuropathol. 1990;80(5):506–513. [PubMed: 2251908]
36.
Ikonomovic M.D. et al. Alzheimer’s pathology in human temporal cortex surgically excised after severe brain injury. Exp Neurol. 2004;190(1):192–203. [PubMed: 15473992]
37.
Roberts G.W. et al. Beta A4 amyloid protein deposition in brain after head trauma. Lancet. 1991;338(8780):1422–1423. [PubMed: 1683421]
38.
Roberts G.W. et al. Beta amyloid protein deposition in the brain after severe head injury: Implications for the pathogenesis of Alzheimer’s disease. J Neurol Neurosurg Psychiatry. 1994;57(4):419–425. [PMC free article: PMC1072869] [PubMed: 8163989]
39.
Ommaya A.K. Head injury mechanisms and the concept of preventive management: A review and critical synthesis. J Neurotrauma. 1995;12(4):527–546. [PubMed: 8683604]
40.
Medana I.M. Axonal damage: A key predictor of outcome in human CNS diseases. Brain. 2003;126(3):515–530. [PubMed: 12566274]
41.
Bullock R. et al. Evidence for prolonged release of excitatory amino acids in severe human head trauma. Relationship to clinical events. Ann N Y Acad Sci. 1995;765:290–297. discussion 298. [PubMed: 7486616]
42.
Bullock R. et al. Factors affecting excitatory amino acid release following severe human head injury. J Neurosurg. 1998;89(4):507–518. [PubMed: 9761042]
43.
Iwata A. et al. Traumatic axonal injury induces proteolytic cleavage of the voltage-gated sodium channels modulated by tetrodotoxin and protease inhibitors. J Neurosci. 2004;24(19):4605–4613. [PMC free article: PMC6729402] [PubMed: 15140932]
44.
Wolf J.A. et al. Traumatic axonal injury induces calcium influx modulated by tetrodotoxin-sensitive sodium channels. J Neurosci. 2001;21(6):1923–1930. [PMC free article: PMC6762603] [PubMed: 11245677]
45.
Staal J.A. et al. Initial calcium release from intracellular stores followed by calcium dysregulation is linked to secondary axotomy following transient axonal stretch injury. J Neurochem. 2010;112(5):1147–1155. [PubMed: 19968758]
46.
Rosenberg L.J, Teng Y.D, Wrathall J.R. Effects of the sodium channel blocker tetrodotoxin on acute white matter pathology after experimental contusive spinal cord injury. J Neurosci. 1999;19(14):6122–6133. [PMC free article: PMC6783088] [PubMed: 10407048]
47.
Fink D.J, Datta S, Mata M. Isoform specific reductions in Na+,K(+)-ATPase catalytic (alpha) subunits in the nerve of rats with streptozotocin-induced diabetes. J Neurochem. 1994;63(5):1782–1786. [PubMed: 7931333]
48.
Tavalin S.J, Ellis E.F, Satin L.S. Inhibition of the electrogenic Na pump underlies delayed depolarization of cortical neurons after mechanical injury or glutamate. J Neurophysiol. 1997;77(2):632–638. [PubMed: 9065836]
49.
Ahmed S.M. et al. Stretch-induced injury alters mitochondrial membrane potential and cellular ATP in cultured astrocytes and neurons. J Neurochem. 2000;74(5):1951–1960. [PubMed: 10800938]
50.
Kornek B. et al. Distribution of a calcium channel subunit in dystrophic axons in multiple sclerosis and experimental autoimmune encephalomyelitis. Brain. 2001;124(Pt 6):1114–1124. [PubMed: 11353727]
51.
Büki A. et al. Cytochrome c release and caspase activation in traumatic axonal injury. J Neurosci. 2000;20(8):2825–2834. [PMC free article: PMC6772193] [PubMed: 10751434]
52.
Büki A, Koizumi H, Povlishock J.T. Moderate posttraumatic hypothermia decreases early calpain-mediated proteolysis and concomitant cytoskeletal compromise in traumatic axonal injury. Exp Neurol. 1999;159(1):319–328. [PubMed: 10486200]
53.
Saatman K.E. et al. Behavioral efficacy of posttraumatic calpain inhibition is not accompanied by reduced spectrin proteolysis, cortical lesion, or apoptosis. J Cereb Blood Flow Metab. 2000;20(1):66–73. [PubMed: 10616794]
54.
Marmarou C.R, Povlishock J.T. Administration of the immunophilin ligand FK506 differentially attenuates neurofilament compaction and impaired axonal transport in injured axons following diffuse traumatic brain injury. Exp Neurol. 2006;197(2):353–362. [PubMed: 16297913]
55.
Singleton R.H. et al. The immunophilin ligand FK506 attenuates axonal injury in an impact-acceleration model of traumatic brain injury. J Neurotrauma. 2001;18(6):607–614. [PubMed: 11437083]
56.
Okonkwo D.O. et al. Cyclosporin A limits calcium-induced axonal damage following traumatic brain injury. Neuroreport. 1999;10(2):353–358. [PubMed: 10203334]
57.
Okonkwo D.O, Povlishock J.T. An intrathecal bolus of cyclosporin A before injury preserves mitochondrial integrity and attenuates axonal disruption in traumatic brain injury. J Cereb Blood Flow Metab. 1999;19(4):443–451. [PubMed: 10197514]
58.
Staal J.A. et al. Cyclosporin-A treatment attenuates delayed cytoskeletal alterations and secondary axotomy following mild axonal stretch injury. Devel Neurobio. 2007;67(14):1831–1842. [PubMed: 17702000]
59.
Deng-Bryant Y. et al. Neuroprotective effects of tempol, a catalytic scavenger of peroxynitrite-derived free radicals, in a mouse traumatic brain injury model. J Cereb Blood Flow Metab. 2008;28(6):1114–1126. [PubMed: 18319733]
60.
Loane D.J, Byrnes K.R. Role of microglia in neurotrauma. Neurotherapeutics. 2010;7(4):366–377. [PMC free article: PMC2948548] [PubMed: 20880501]
61.
Gentleman S.M. et al. Long-term intracerebral inflammatory response after traumatic brain injury. Forensic Sci Int. 2004;146(2–3):97–104. [PubMed: 15542269]
62.
Reeves T.M, Phillips L.L, Povlishock J.T. Myelinated and unmyelinated axons of the corpus callosum differ in vulnerability and functional recovery following traumatic brain injury. Exp Neurol. 2005;196(1):126–137. [PubMed: 16109409]
63.
Staal J.A, Vickers J.C. Selective vulnerability of non-myelinated axons to stretch injury in an in vitro co-culture system. J Neurotrauma. 2011;28(5):841–847. [PubMed: 21235329]
64.
Bramlett H.M, Dietrich W.D. Synuclein aggregation: Possible role in traumatic brain injury. Exp Neurol. 2003;184(1):27–30. [PubMed: 14637075]
65.
Su E. et al. Increased CSF concentrations of myelin basic protein after TBI in infants and children: Absence of significant effect of therapeutic hypothermia. Neurocrit Care. 2012;17(3):401–407. [PMC free article: PMC3836276] [PubMed: 22890910]
66.
Ng H.K, Mahaliyana R.D, Poon W.S. The pathological spectrum of diffuse axonal injury in blunt head trauma: Assessment with axon and myelin strains. Clin Neurol Neurosurg. 1994;96(1):24–31. [PubMed: 8187378]
67.
Oehmichen M, Theuerkauf I, Meissner C. Is traumatic axonal injury (AI) associated with an early microglial activation? Application of a double-labeling technique for simultaneous detection of microglia and AI. Acta Neuropathol. 1999;97(5):491–494. [PubMed: 10334486]
68.
GrandPré T, Li S, Strittmatter S.M. Nogo-66 receptor antagonist peptide promotes axonal regeneration. Nature. 2002;417(6888):547–551. [PubMed: 12037567]
69.
Lin Y, Wen L. Inflammatory response following diffuse axonal injury. Int J Med Sci. 2013;10(5):515–521. [PMC free article: PMC3607236] [PubMed: 23532682]
70.
Lu K.-T. et al. Extracellular signal-regulated kinase-mediated IL-1-induced cortical neuron damage during traumatic brain injury. Neurosci Lett. 2005;386(1):40–45. [PubMed: 16024175]
71.
Hans V.H. et al. Experimental axonal injury triggers interleukin-6 mRNA, protein synthesis and release into cerebrospinal fluid. J Cereb Blood Flow Metab. 1999;19(2):184–194. [PubMed: 10027774]
72.
Kita T. et al. The role of tumor necrosis factor-alpha in diffuse axonal injury following fluid-percussive brain injury in rats. Int J Legal Med. 2000;113(4):221–228. [PubMed: 10929238]
73.
Khuman J. et al. Tumor necrosis factor alpha and Fas receptor contribute to cognitive deficits independent of cell death after concussive traumatic brain injury in mice. J Cereb Blood Flow Metab. 2011;31(2):778–789. [PMC free article: PMC3049532] [PubMed: 20940727]
74.
Yan E.B. et al. Post-traumatic hypoxia exacerbates neurological deficit, neuroinflammation and cerebral metabolism in rats with diffuse traumatic brain injury. J Neuroinflammation. 2011;8(1):147. [PMC free article: PMC3215944] [PubMed: 22034986]
75.
Rancan M. et al. Upregulation of ICAM-1 and MCP-1 but not of MIP-2 and sensorimotor deficit in response to traumatic axonal injury in rats. J Neurosci Res. 2001;63(5):438–446. [PubMed: 11223919]
76.
Rhodes J.K.J, Sharkey J, Andrews P.J.D. The temporal expression, cellular localization, and inhibition of the chemokines MIP-2 and MCP-1 after traumatic brain injury in the rat. J Neurotrauma. 2009;26(4):507–525. [PubMed: 19210118]
77.
Büki A, Povlishock J.T. All roads lead to disconnection?—Traumatic axonal injury revisited. Acta Neurochir (Wien). 2005;148(2):181–194. [PubMed: 16362181]
78.
Blumbergs P.C, Jones N.R, North J.B. Diffuse axonal injury in head trauma. J Neurol Neurosurg Psychiatry. 1989;52(7):838–841. [PMC free article: PMC1031929] [PubMed: 2769276]
79.
Sullivan G.M. et al. Oligodendrocyte lineage and subventricular zone response to traumatic axonal injury in the corpus callosum. J Neuropathol Exp Neurol. 2013;72(12):1106–1125. [PMC free article: PMC4130339] [PubMed: 24226267]
80.
Ludwin S.K. Oligodendrocyte survival in Wallerian degeneration. Acta Neuropathol. 1990;80(2):184–191. [PubMed: 1697140]
81.
Grandpré T, Strittmatter S.M. Nogo: A molecular determinant of axonal growth and regeneration. Neuroscientist. 2001;7(5):377–386. [PubMed: 11597097]
82.
Wang T. et al. The role of Nogo-A in neuroregeneration: A review. Brain Res Bull. 2012;87(6):499–503. [PubMed: 22414960]
83.
Keyvani K, Schallert T. Plasticity-associated molecular and structural events in the injured brain. J Neuropathol Exp Neurol. 2002;61(10):831–840. [PubMed: 12387449]
84.
Magnuson J, Leonessa F, Ling G.S.F. Neuropathology of explosive blast traumatic brain injury. Curr Neurol Neurosci Rep. 2012;12(5):570–579. [PubMed: 22836523]
85.
Richter M. et al. Head injury mechanisms in helmet-protected motorcyclists: Prospective multicenter study. J Trauma. 2001;51(5):949–958. [PubMed: 11706346]
86.
Pellman E.J. et al. Concussion in professional football: Reconstruction of game impacts and injuries. Neurosurgery. 2003;53(4):799–812. discussion 812-814. [PubMed: 14519212]
87.
Tate C.M. et al. Serum brain biomarker level, neurocognitive performance, and self-reported symptom changes in soldiers repeatedly exposed to low-level blast: A breacher pilot study. J Neurotrauma. 2013;30(19):1620–1630. [PubMed: 23687938]
88.
Bochicchio G.V. et al. Blast injury in a civilian trauma setting is associated with a delay in diagnosis of traumatic brain injury. Am Surg. 2008;74(3):267–270. [PubMed: 18376697]
89.
DuBose J.J. et al. Isolated severe traumatic brain injuries sustained during combat operations: Demographics, mortality outcomes, and lessons to be learned from contrasts to civilian counterparts. J Trauma. 2011;70(1):11–16. discussion 16–18. [PubMed: 21217475]
90.
Chen Y.C, Smith D.H, Meaney D.F. In-vitro approaches for studying blast-induced traumatic brain injury. J Neurotrauma. 2009;26(6):861–876. [PMC free article: PMC2803321] [PubMed: 19397424]
91.
Martland H.S. Punch drunk. JAMA. 1928;91(15):1103–1107.
92.
McKee A.C. et al. The neuropathology of sport. Acta Neuropathol. 2014;127(1):29–51. [PMC free article: PMC4255282] [PubMed: 24366527]
93.
Powell J.W, Barber-Foss K.D. Traumatic brain injury in high school athletes. JAMA. 1999;282(10):958–963. [PubMed: 10485681]
94.
Shin W. et al. Diffusion measures indicate fight exposure-related damage to cerebral white matter in boxers and mixed martial arts fighters. Am J Neuroradiol. 2013 [PMC free article: PMC7965766] [PubMed: 23928146]
95.
Dale G.E. et al. Neurofibrillary tangles in dementia pugilistica are ubiquitinated. J Neurol Neurosurg Psychiatry. 1991;54(2):116–118. [PMC free article: PMC1014344] [PubMed: 1850450]
96.
Geddes J.F. et al. Neuronal cytoskeletal changes are an early consequence of repetitive head injury. Acta Neuropathol. 1999;98(2):171–178. [PubMed: 10442557]
97.
Tokuda T. et al. Re-examination of ex-boxers’ brains using immunohisto-chemistry with antibodies to amyloid beta-protein and tau protein. Acta Neuropathol. 1991;82(4):280–285. [PubMed: 1759560]
98.
McKee A.C. et al. Chronic traumatic encephalopathy in athletes: Progressive tauopathy after repetitive head injury. J Neuropathol Exp Neurol. 2009;68(7):709–735. [PMC free article: PMC2945234] [PubMed: 19535999]
99.
Omalu B. et al. Emerging histomorphologic phenotypes of chronic traumatic encephalopathy in American athletes. Neurosurgery. 2011;69(1):173–183. [PubMed: 21358359]
100.
Jordan B.D. Chronic traumatic brain injury associated with boxing. Semin Neurol. 2000;20(2):179–185. [PubMed: 10946737]
101.
Orrison W.W. et al. Traumatic brain injury: A review and high-field MRI findings in 100 unarmed combatants using a literature-based checklist approach. J Neurotrauma. 2009;26(5):689–701. [PubMed: 19335205]
102.
Fleminger S. et al. Head injury as a risk factor for Alzheimer’s disease: The evidence 10 years on; A partial replication. J Neurol Neurosurg Psychiatry. 2003;74(7):857–862. [PMC free article: PMC1738550] [PubMed: 12810767]
103.
Graves A.B. et al. The association between head trauma and Alzheimer’s disease. Am J Epidemiol. 1990;131(3):491–501. [PubMed: 2405648]
104.
Guo Z. et al. Head injury and the risk of AD in the MIRAGE study. Neurology. 2000;54(6):1316–1323. [PubMed: 10746604]
105.
Mortimer J.A. et al. Head injury as a risk factor for Alzheimer’s disease. Neurology. 1985;35(2):264–267. [PubMed: 3969219]
106.
Mortimer J.A. et al. Head trauma as a risk factor for Alzheimer’s disease: A collaborative re-analysis of case-control studies. EURODEM Risk Factors Research Group. Int J Epidemiol. 1991;20(Suppl 2):S28–S35. [PubMed: 1833351]
107.
O’Meara E.S. et al. Head injury and risk of Alzheimer’s disease by apolipoprotein E genotype. Am J Epidemiol. 1997;146(5):373–384. [PubMed: 9290497]
108.
Plassman B.L. et al. Documented head injury in early adulthood and risk of Alzheimer’s disease and other dementias. Neurology. 2000;55(8):1158–1166. [PubMed: 11071494]
109.
Salib E, Hillier V. Head injury and the risk of Alzheimer’s disease: A case control study. Int J Geriatr Psychiatry. 1997;12(3):363–368. [PubMed: 9152722]
110.
Nemetz P.N. et al. Traumatic brain injury and time to onset of Alzheimer’s disease: A population-based study. Am J Epidemiol. 1999;149(1):32–40. [PubMed: 9883791]
111.
Sullivan P, Petitti D, Barbaccia J. Head trauma and age of onset of dementia of the Alzheimer type. JAMA. 1987;257(17):2289–2290. [PubMed: 3573227]
112.
Schofield P.W. et al. Alzheimer’s disease after remote head injury: An incidence study. J Neurol Neurosurg Psychiatry. 1997;62(2):119–124. [PMC free article: PMC486721] [PubMed: 9048710]
113.
Jordan B.D. et al. Apolipoprotein E epsilon4 associated with chronic traumatic brain injury in boxing. JAMA. 1997;278(2):136–140. [PubMed: 9214529]
114.
Case M.E. Inflicted traumatic brain injury in infants and young children. Brain Pathol. 2008;18(4):571–582. [PMC free article: PMC8095515] [PubMed: 18782169]
115.
Case M.E. Accidental traumatic head injury in infants and young children. Brain Pathol. 2008;18(4):583–589. [PMC free article: PMC8095612] [PubMed: 18782170]
116.
Cantu R.C, Gean A.D. Second-impact syndrome and a small subdural hematoma: An uncommon catastrophic result of repetitive head injury with a characteristic imaging appearance. J Neurotrauma. 2010;27(9):1557–1564. [PMC free article: PMC2956379] [PubMed: 20536318]
117.
Geddes J.F. et al. Neuropathology of inflicted head injury in children. II. Microscopic brain injury in infants. Brain. 2001;124(Pt 7):1299–1306. [PubMed: 11408325]
118.
Duhaime A.C. et al. Nonaccidental head injury in infants—The “shaken-baby syndrome.” N Engl J Med. 1998;338(25):1822–1829. [PubMed: 9632450]
119.
Clayton E.H, Genin G.M, Bayly P.V. Transmission, attenuation and reflection of shear waves in the human brain. J R Soc Interface. 2012;9(76):2899–2910. [PMC free article: PMC3479912] [PubMed: 22675163]
120.
Meaney D.F. et al. Biomechanical analysis of experimental diffuse axonal injury. J Neurotrauma. 1995;12(4):689–694. [PubMed: 8683620]
121.
Margulies S.S, Thibault L.E, Gennarelli T.A. Physical model simulations of brain injury in the primate. J Biomech. 1990;23(8):823–836. [PubMed: 2384494]
122.
Povlishock J.T. Pathobiology of traumatically induced axonal injury in animals and man. Ann Emerg Med. 2013;22(6):980–986. [PubMed: 8503536]
123.
Wright R.M. et al. A multiscale computational approach to estimating axonal damage under inertial loading of the head. J Neurotrauma. 2013;30(2):102–118. [PubMed: 22992118]
124.
Smith D.H. et al. High tolerance and delayed elastic response of cultured axons to dynamic stretch injury. J Neurosci. 1999;19(11):4263–4269. [PMC free article: PMC6782601] [PubMed: 10341230]
125.
Gennarelli T.A. et al. Axonal injury in the optic nerve: A model simulating diffuse axonal injury in the brain. J Neurosurg. 1989;71(2):244–253. [PubMed: 2746348]
126.
Metz H, McElhaney J, Ommaya A.K. A comparison of the elasticity of live, dead, and fixed brain tissue. J Biomech. 1970;3(4):453–458. [PubMed: 5000415]
127.
Laksari K, Shafieian M, Darvish K. Constitutive model for brain tissue under finite compression. J Biomech. 2012;45(4):642–646. [PubMed: 22281404]
128.
Shafieian M, Darvish K.K, Stone J.R. Changes to the viscoelastic properties of brain tissue after traumatic axonal injury. J Biomech. 2009;42(13):2136–2142. [PubMed: 19698945]
129.
O’Connor W.T, Smyth A, Gilchrist M.D. Animal models of traumatic brain injury: A critical evaluation. Pharmacol Ther. 2011;130(2):106–113. [PubMed: 21256863]
130.
Graham D.I, Adams J.H, Gennarelli T.A. Mechanisms of non-penetrating head injury. Prog Clin Biol Res. 1988;264:159–168. [PubMed: 3289020]
131.
Holbourn A. Mechanics of head injuries. Lancet. 1943;242(6267):438–441.
132.
Lighthall J.W. Controlled cortical impact: A new experimental brain injury model. J Neurotrauma. 1988;5(1):1–15. [PubMed: 3193461]
133.
Browne K.D. et al. Mild traumatic brain injury and diffuse axonal injury in swine. J Neurotrauma. 2011;28(9):1747–1755. [PMC free article: PMC3172883] [PubMed: 21740133]
134.
Rubovitch V. et al. A mouse model of blast-induced mild traumatic brain injury. Exp Neurol. 2011;232(2):280–289. [PMC free article: PMC3202080] [PubMed: 21946269]
135.
Bauman R.A. et al. An introductory characterization of a combat-casualty-care relevant swine model of closed head injury resulting from exposure to explosive blast. J Neurotrauma. 2009;26(6):841–860. [PubMed: 19215189]
136.
Säljö A. et al. Neuropathology and pressure in the pig brain resulting from low-impulse noise exposure. J Neurotrauma. 2008;25(12):1397–1406. [PubMed: 19146459]
137.
Ling G. et al. Explosive blast neurotrauma. J Neurotrauma. 2009;26(6):815–825. [PubMed: 19397423]
138.
Hicks R.R. et al. Neurological effects of blast injury. J Trauma. 2010;68(5):1257–1263. [PMC free article: PMC2958428] [PubMed: 20453776]
139.
Long J.B. et al. Blast overpressure in rats: Recreating a battlefield injury in the laboratory. J Neurotrauma. 2009;26(6):827–840. [PubMed: 19397422]
140.
Gyorgy A. et al. Time-dependent changes in serum biomarker levels after blast traumatic brain injury. J Neurotrauma. 2011;28(6):1121–1126. [PubMed: 21428721]
141.
Xiong Y, Mahmood A, Chopp M. Animal models of traumatic brain injury. Nat Rev Neurosci. 2013;14(2):128–142. [PMC free article: PMC3951995] [PubMed: 23329160]
142.
Lighthall J.W, Goshgarian H.G, Pinderski C.R. Characterization of axonal injury produced by controlled cortical impact. J Neurotrauma. 1990;7(2):65–76. [PubMed: 2376865]
143.
Shitaka Y. et al. Repetitive closed-skull traumatic brain injury in mice causes persistent multifocal axonal injury and microglial reactivity. J Neuropathol Exp Neurol. 2011;70(7):551–567. [PMC free article: PMC3118973] [PubMed: 21666502]
144.
Maruichi K. et al. Graded model of diffuse axonal injury for studying head injury-induced cognitive dysfunction in rats. Neuropathology. 2009;29(2):132–139. [PubMed: 18702633]
145.
Marmarou A. et al. A new model of diffuse brain injury in rats. Part I: Pathophysiology and biomechanics. J Neurosurg. 1994;80(2):291–300. [PubMed: 8283269]
146.
Foda M.A, Marmarou A. A new model of diffuse brain injury in rats. Part II: Morphological characterization. J Neurosurg. 1994;80(2):301–313. [PubMed: 8283270]
147.
Thibault L.E. et al. Biomechanical aspects of a fluid percussion model of brain injury. J Neurotrauma. 2013;9(4):311–322. [PubMed: 1291691]
148.
McIntosh T.K. et al. Traumatic brain injury in the rat: Characterization of a midline fluid-percussion model. Cent Nerv Syst Trauma. 1987;4(2):119–134. [PubMed: 3690695]
149.
Thompson H.J. et al. Lateral fluid percussion brain injury: A 15-year review and evaluation. J Neurotrauma. 2005;22(1):42–75. [PubMed: 15665602]
150.
Wang H.-C, Ma Y.-B. Experimental models of traumatic axonal injury. J Clin Neurosci. 2010;17(2):157–162. [PubMed: 20042337]
151.
Li X.Y. et al. Diffuse axonal injury induced by simultaneous moderate linear and angular head accelerations in rats. Neuroscience. 2010;169(1):357–369. [PubMed: 20451584]
152.
Li J. et al. Exploring temporospatial changes in glucose metabolic disorder, learning, and memory dysfunction in a rat model of diffuse axonal injury. J Neurotrauma. 2012;29(17):2635–2646. [PMC free article: PMC3510449] [PubMed: 22880625]
153.
Li J. et al. Quantitative evaluation of microscopic injury with diffusion tensor imaging in a rat model of diffuse axonal injury. Eur J Neurosci. 2011;33(5):933–945. [PubMed: 21385236]
154.
Lusardi T.A. et al. Effect of acute calcium influx after mechanical stretch injury in vitro on the viability of hippocampal neurons. J Neurotrauma. 2004;21(1):61–72. [PubMed: 14987466]
155.
Pfister B.J. et al. An in vitro uniaxial stretch model for axonal injury. Ann Biomed Eng. 2003;31(5):589–598. [PubMed: 12757202]
156.
Mohammed Sulaiman A. et al. Stereology and ultrastructure of chronic phase axonal and cell soma pathology in stretch-injured central nerve fibers. J Neurotrauma. 2011;28(3):383–400. [PubMed: 21190396]
157.
Saatman K.E, Creed J, Raghupathi R. Calpain as a therapeutic target in traumatic brain injury. Neurotherapeutics. 2010;7(1):31–42. [PMC free article: PMC2842949] [PubMed: 20129495]
158.
Saatman K.E. et al. Traumatic axonal injury results in biphasic calpain activation and retrograde transport impairment in mice. J Cereb Blood Flow Metab. 2003;23(1):34–42. [PubMed: 12500089]
159.
Agoston D.V. et al. Proteomic biomarkers for blast neurotrauma: Targeting cerebral edema, inflammation, and neuronal death cascades. J Neurotrauma. 2009;26(6):901–911. [PubMed: 19397421]
160.
Ankarcrona M. et al. Glutamate-induced neuronal death: A succession of necrosis or apoptosis depending on mitochondrial function. Neuron. 1995;15(4):961–973. [PubMed: 7576644]
161.
Zipfel G.J. et al. Neuronal apoptosis after CNS injury: The roles of glutamate and calcium. J Neurotrauma. 2000;17(10):857–869. [PubMed: 11063053]
162.
Choi D.W. Calcium: Still center-stage in hypoxic-ischemic neuronal death. Trends Neurosci. 1995;18(2):58–60. [PubMed: 7537408]
163.
Serbest G. et al. Temporal profiles of cytoskeletal protein loss following traumatic axonal injury in mice. Neurochem Res. 2007;32(12):2006–2014. [PubMed: 17401646]
164.
Huh J.W. et al. Regionally distinct patterns of calpain activation and traumatic axonal injury following contusive brain injury in immature rats. Dev Neurosci. 2006;28(4–5):466–476. [PubMed: 16943669]
165.
Thompson S.N. et al. Relationship of calpain-mediated proteolysis to the expression of axonal and synaptic plasticity markers following traumatic brain injury in mice. Exp Neurol. 2006;201(1):253–265. [PubMed: 16814284]
166.
Reeves T.M. et al. Proteolysis of submembrane cytoskeletal proteins ankyrin-G and αII-spectrin following diffuse brain injury: A role in white matter vulnerability at nodes of Ranvier. Brain Pathol. 2010;20(6):1055–1068. [PMC free article: PMC3265329] [PubMed: 20557305]
167.
Farkas O. et al. Spectrin breakdown products in the cerebrospinal fluid in severe head injury—Preliminary observations. Acta Neurochir (Wien). 2005;147(8):855–861. [PubMed: 15924207]
168.
Brophy G.M. et al. alphaII-Spectrin breakdown product cerebrospinal fluid exposure metrics suggest differences in cellular injury mechanisms after severe traumatic brain injury. J Neurotrauma. 2009;26(4):471–479. [PMC free article: PMC2848834] [PubMed: 19206997]
169.
Anderson K.J. et al. The phosphorylated axonal form of the neurofilament subunit NF-H (pNF-H) as a blood biomarker of traumatic brain injury. J Neurotrauma. 2008;25(9):1079–1085. [PMC free article: PMC2820728] [PubMed: 18729720]
170.
Van Geel W.J.A, Rosengren L.E, Verbeek M.M. An enzyme immunoassay to quantify neurofilament light chain in cerebrospinal fluid. J Immunol Methods. 2005;296(1–2):179–185. [PubMed: 15680162]
171.
Park E. et al. Heavy neurofilament accumulation and α-spectrin degradation accompany cerebellar white matter functional deficits following forebrain fluid percussion injury. Exp Neurol. 2007;204(1):49–57. [PubMed: 17070521]
172.
Letournel F. et al. Neurofilament high molecular weight–green fluorescent protein fusion is normally expressed in neurons and transported in axons: A neuronal marker to investigate the biology of neurofilaments. Neuroscience. 2006;137(1):103–111. [PubMed: 16289584]
173.
Teunissen C.E, Dijkstra C, Polman C. Biological markers in CSF and blood for axonal degeneration in multiple sclerosis. Lancet Neurol. 2005;4(1):32–41. [PubMed: 15620855]
174.
Sihag R.K. et al. Role of phosphorylation on the structural dynamics and function of types III and IV intermediate filaments. Exp Cell Res. 2007;313(10):2098–2109. [PMC free article: PMC2570114] [PubMed: 17498690]
175.
McCracken E. et al. Calpain activation and cytoskeletal protein breakdown in the corpus callosum of head-injured patients. J Neurotrauma. 1999;16(9):749–761. [PubMed: 10521135]
176.
Hamberger A. et al. Redistribution of neurofilaments and accumulation of betaamyloid protein after brain injury by rotational acceleration of the head. J Neurotrauma. 2003;20(2):169–178. [PubMed: 12675970]
177.
Chen X.H. et al. Evolution of neurofilament subtype accumulation in axons following diffuse brain injury in the pig. J Neuropathol Exp Neurol. 1999;58(6):588–596. [PubMed: 10374749]
178.
Smith D.H, Meaney D.F, Shull W.H. Diffuse axonal injury in head trauma. J Head Trauma Rehabil. 2003;18(4):307–316. [PubMed: 16222127]
179.
Kaur B, Rutty G.N, Timperley W.R. The possible role of hypoxia in the formation of axonal bulbs. J Clin Pathol. 1999;52(3):203–209. [PMC free article: PMC501080] [PubMed: 10450180]
180.
Chen Y, Tang B.L. The amyloid precursor protein and postnatal neurogenesis/neuroregeneration. Biochem Biophys Res Commun. 2006;341(1):1–5. [PubMed: 16406235]
181.
Stone J.R, Singleton R.H, Povlishock J.T. Intra-axonal neurofilament compaction does not evoke local axonal swelling in all traumatically injured axons. Exp Neurol. 2001;172(2):320–331. [PubMed: 11716556]
182.
Reichard R.R, Smith C, Graham D.I. The significance of beta-APP immuno-reactivity in forensic practice. Neuropathol Appl Neurobiol. 2005;31(3):304–313. [PubMed: 15885067]
183.
Reichard R.R. et al. Beta-amyloid precursor protein staining in nonhomicidal pediatric medicolegal autopsies. J Neuropathol Exp Neurol. 2003;62(3):237–247. [PubMed: 12638728]
184.
Teasdale G.M. et al. Association of apolipoprotein E polymorphism with outcome after head injury. Lancet. 1997;350(9084):1069–1071. [PubMed: 10213549]
185.
Raby C.A. et al. Traumatic brain injury increases beta-amyloid peptide 1-42 in cerebrospinal fluid. J Neurochem. 1998;71(6):2505–2509. [PubMed: 9832149]
186.
Olsson A. et al. Marked increase of beta-amyloid(1-42) and amyloid precursor protein in ventricular cerebrospinal fluid after severe traumatic brain injury. J Neurol. 2004;251(7):870–876. [PubMed: 15258792]
187.
Zemlan F.P. et al. C-tau biomarker of neuronal damage in severe brain injured patients: Association with elevated intracranial pressure and clinical outcome. Brain Res. 2002;947(1):131–139. [PubMed: 12144861]
188.
Zemlan F.P. et al. Quantification of axonal damage in traumatic brain injury: Affinity purification and characterization of cerebrospinal fluid tau proteins. J Neurochem. 1999;72(2):741–750. [PubMed: 9930748]
189.
Cengiz P. et al. Cerebrospinal fluid cleaved-tau protein and 9-hydroxyoctadeca-dienoic acid concentrations in pediatric patients with hydrocephalus. Pediatr Crit Care Med. 2008;9(5):524–529. [PubMed: 18679140]
190.
Bodjarian N. et al. Strong expression of GFAP mRNA in rat hippocampus after a closed-head injury. Neuroreport. 1997;8(18):3951–3956. [PubMed: 9462472]
191.
Nylén K. et al. Increased serum-GFAP in patients with severe traumatic brain injury is related to outcome. J Neurol Sci. 2006;240(1–2):85–91. [PubMed: 16266720]
192.
Wiesmann M. et al. Outcome prediction in traumatic brain injury: Comparison of neurological status, CT findings, and blood levels of S100B and GFAP. Acta Neurol Scand. 2010;121(3):178–185. [PubMed: 19804476]
193.
Diaz-Arrastia R. et al. Acute biomarkers of traumatic brain injury: Relationship between plasma levels of ubiquitin C-terminal hydrolase-L1 and glial fibrillary acidic protein. J Neurotrauma. 2014;31(1):19–25. [PMC free article: PMC3880090] [PubMed: 23865516]
194.
Pelinka L.E. et al. GFAP versus S100B in serum after traumatic brain injury: Relationship to brain damage and outcome. J Neurotrauma. 2004;21(11):1553–1561. [PubMed: 15684648]
195.
Svetlov S.I. et al. Biomarkers of blast-induced neurotrauma: Profiling molecular and cellular mechanisms of blast brain injury. J Neurotrauma. 2013;26(6):913–921. [PMC free article: PMC6469534] [PubMed: 19422293]
196.
Ottens A.K. et al. Proteolysis of multiple myelin basic protein isoforms after neurotrauma: Characterization by mass spectrometry. J Neurochem. 2008;104(5):1404–1414. [PMC free article: PMC3584683] [PubMed: 18036155]
197.
Jeter C.B. et al. Biomarkers for the diagnosis and prognosis of mild traumatic brain injury/concussion. J Neurotrauma. 2013;30(8):657–670. [PubMed: 23062081]
198.
Hurley R.A. et al. Traumatic axonal injury: Novel insights into evolution and identification. J Neuropsychiatry Clin Neurosci. 2004;16(1):1–7. [PubMed: 14990753]
199.
de Kruijk J.R. et al. S-100B and neuron-specific enolase in serum of mild traumatic brain injury patients. A comparison with health controls. Acta Neurol Scand. 2001;103(3):175–179. [PubMed: 11240565]
200.
Savola O, Hillbom M. Early predictors of post-concussion symptoms in patients with mild head injury. Eur J Neurol. 2003;10(2):175–181. [PubMed: 12603294]
201.
Undén J, Romner B. A new objective method for CT triage after minor head injury—Serum S100B. Scand J Clin Lab Invest. 2009;69(1):13–17. [PubMed: 19199125]
202.
Bazarian J.J. et al. Serum S-100B and cleaved-tau are poor predictors of longterm outcome after mild traumatic brain injury. Brain Inj. 2006;20(7):759–765. [PubMed: 16809208]
203.
Ingebrigtsen T, Romner B. Biochemical serum markers for brain damage: A short review with emphasis on clinical utility in mild head injury. Restor Neurol Neurosci. 2003;21(3–4):171–176. [PubMed: 14530579]
204.
Kleindienst A. et al. The neurotrophic protein S100B: Value as a marker of brain damage and possible therapeutic implications. Prog Brain Res. 2007;161:317–325. [PubMed: 17618987]
205.
Kleindienst A., Ross Bullock, M. A critical analysis of the role of the neurotrophic protein S100B in acute brain injury. J Neurotrauma. 2006;23(8):1185–1200. [PubMed: 16928177]
206.
Piazza O. et al. S100B is not a reliable prognostic index in paediatric TBI. Pediatr Neurosurg. 2007;43(4):258–264. [PubMed: 17627141]
207.
Nygren de boussard C. et al. S100 in mild traumatic brain injury. Brain Inj. 2004;18(7):671–683. [PubMed: 15204328]
208.
Ohrt-Nissen S. et al. How does extracerebral trauma affect the clinical value of S100B measurements? Emerg Med J. 2011;28(11):941–944. [PMC free article: PMC3198008] [PubMed: 20947920]
209.
Berger R, Richichi R. Derivation and validation of an equation for adjustment of neuron-specific enolase concentrations in hemolyzed serum. Pediatr Crit Care Med. 2009;10(2):260–263. [PubMed: 19188872]
210.
Berger R.P. et al. Serum biomarker concentrations and outcome after pediatric traumatic brain injury. J Neurotrauma. 2007;24(12):1793–1801. [PubMed: 18159990]
211.
Papa L. et al. Ubiquitin C-terminal hydrolase is a novel biomarker in humans for severe traumatic brain injury. Crit Care Med. 2010;38(1):138–144. [PMC free article: PMC3445330] [PubMed: 19726976]
212.
Berger R.P. et al. Serum concentrations of ubiquitin C-terminal hydrolase-L1 and αII-spectrin breakdown product 145 kDa correlate with outcome after pediatric TBI. J Neurotrauma. 2012;29(1):162–167. [PMC free article: PMC3253308] [PubMed: 22022780]
213.
Murray G.D. et al. Multivariable prognostic analysis in traumatic brain injury: Results from the IMPACT study. J Neurotrauma. 2007;24(2):329–337. [PubMed: 17375997]
214.
Predicting outcome after traumatic brain injury: Practical prognostic models based on large cohort of international patients. BMJ. 336(7641):425–429. MRC CRASH Trial Collaborators et al. 2008. [PMC free article: PMC2249681] [PubMed: 18270239]
215.
Tollard E. et al. Experience of diffusion tensor imaging and 1H spectroscopy for outcome prediction in severe traumatic brain injury: Preliminary results. Crit Care Med. 2009;37(4):1448–1455. [PubMed: 19242330]
216.
Lee H. et al. Focal lesions in acute mild traumatic brain injury and neurocognitive outcome: CT versus 3T MRI. J Neurotrauma. 2008;25(9):1049–1056. [PubMed: 18707244]
217.
Laalo J.P, Kurki T.J, Tenovuo O.S. Interpretation of magnetic resonance imaging in the chronic phase of traumatic brain injury: What is missed in the original reports? Brain Inj. 2014;28(1):66–70. [PubMed: 24328801]
218.
Provenzale J.M. Imaging of traumatic brain injury: A review of the recent medical literature. AJR Am J Roentgenol. 2010;194(1):16–19. [PubMed: 20028899]
219.
Reider-Groswasser I. et al. CT findings in persistent vegetative state following blunt traumatic brain injury. Brain Inj. 1997;11(12):865–870. [PubMed: 9413620]
220.
Jones N.R. et al. Correlation of postmortem MRI and CT appearances with neuropathology in brain trauma: A comparison of two methods. J Clin Neurosci. 1998;5(1):73–79. [PubMed: 18644293]
221.
Kim J.J, Gean A.D. Imaging for the diagnosis and management of traumatic brain injury. Neurotherapeutics. 2011;8(1):39–53. [PMC free article: PMC3026928] [PubMed: 21274684]
222.
Imaizumi T. et al. Dynamics of dotlike hemosiderin spots associated with intracerebral hemorrhage. J Neuroimaging. 2003;13(2):155–157. [PubMed: 12722499]
223.
Tong K.A. et al. Hemorrhagic shearing lesions in children and adolescents with posttraumatic diffuse axonal injury: Improved detection and initial results. Radiology. 2003;227(2):332–339. [PubMed: 12732694]
224.
Yanagawa Y. et al. A quantitative analysis of head injury using T2*-weighted gradient-echo imaging. J Trauma. 2000;49(2):272–277. [PubMed: 10963538]
225.
Scheid R. et al. Diffuse axonal injury associated with chronic traumatic brain injury: Evidence from T2*-weighted gradient-echo imaging at 3 T. AJNR Am J Neuroradiol. 2003;24(6):1049–1056. [PMC free article: PMC8149043] [PubMed: 12812926]
226.
Yanagawa Y. et al. Relationship between maximum intracranial pressure and traumatic lesions detected by T2*-weighted imaging in diffuse axonal injury. J Trauma. 2009;66(1):162–165. [PubMed: 19131819]
227.
Luccichenti G. et al. 3 Tesla is twice as sensitive as 1.5 Tesla magnetic resonance imaging in the assessment of diffuse axonal injury in traumatic brain injury patients. Funct Neurol. 2010;25(2):109–114. [PubMed: 20923609]
228.
Viswanathan A. Cerebral microhemorrhage. Stroke. 2006;37(2):550–555. [PubMed: 16397165]
229.
Paterakis K. et al. Outcome of patients with diffuse axonal injury: The significance and prognostic value of MRI in the acute phase. J Trauma. 2000;49(6):1071–1075. [PubMed: 11130491]
230.
Tong K.A. et al. Susceptibility-weighted MR imaging: A review of clinical applications in children. AJNR Am J Neuroradiol. 2008;29(1):9–17. [PMC free article: PMC8119104] [PubMed: 17925363]
231.
Iwamura A. et al. Diffuse vascular injury: Convergent-type hemorrhage in the supratentorial white matter on susceptibility-weighted image in cases of severe traumatic brain damage. Neuroradiology. 2012;54(4):335–343. [PubMed: 21611726]
232.
Zheng W.B. et al. Coma duration prediction in diffuse axonal injury: Analyses of apparent diffusion coefficient and clinical prognostic factors. Conf Proc IEEE Eng Med Biol Soc. 2006;1:1052–1055. [PubMed: 17946873]
233.
Galloway N.R. et al. Diffusion-weighted imaging improves outcome prediction in pediatric traumatic brain injury. J Neurotrauma. 2008;25(10):1153–1162. [PubMed: 18842104]
234.
Mac Donald C.L. et al. Detection of blast-related traumatic brain injury in U.S. military personnel. N Engl J Med. 2011;364(22):2091–2100. [PMC free article: PMC3146351] [PubMed: 21631321]
235.
Wang J.Y. et al. Diffusion tensor tractography of traumatic diffuse axonal injury. Arch Neurol. 2008;65(5):619–626. [PubMed: 18474737]
236.
Bendlin B.B. et al. Longitudinal changes in patients with traumatic brain injury assessed with diffusion-tensor and volumetric imaging. Neuroimage. 2008;42(2):503–514. [PMC free article: PMC2613482] [PubMed: 18556217]
237.
Kou Z. et al. The role of advanced MR imaging findings as biomarkers of traumatic brain injury. J Head Trauma Rehabil. 2010;25(4):267–282. [PubMed: 20611045]
238.
Oouchi H. et al. Diffusion anisotropy measurement of brain white matter is affected by voxel size: Underestimation occurs in areas with crossing fibers. AJNR Am J Neuroradiol. 2007;28(6):1102–1106. [PMC free article: PMC8134156] [PubMed: 17569968]
239.
Sarlls J.E, Pierpaoli C. In vivo diffusion tensor imaging of the human optic chiasm at sub-millimeter resolution. Neuroimage. 2009;47(4):1244–1251. [PMC free article: PMC2735123] [PubMed: 19520170]
240.
Tang C.Y. et al. Diffuse disconnectivity in TBI: A resting state fMRI and DTI study. Transl Neurosci. 2012;3(1):9–14. [PMC free article: PMC3583347] [PubMed: 23459252]
241.
Zhang K. et al. Are functional deficits in concussed individuals consistent with white matter structural alterations: Combined FMRI & DTI study. Exp Brain Res. 2010;204(1):57–70. [PMC free article: PMC2919274] [PubMed: 20496060]
242.
Zhang Y. et al. Atlas-guided tract reconstruction for automated and comprehensive examination of the white matter anatomy. Neuroimage. 2010;52(4):1289–1301. [PMC free article: PMC2910162] [PubMed: 20570617]
243.
Jensen J.H, Helpern J.A. MRI quantification of non-Gaussian water diffusion by kurtosis analysis. NMR Biomed. 2010;23(7):698–710. [PMC free article: PMC2997680] [PubMed: 20632416]
244.
Lu H. et al. Three-dimensional characterization of non-Gaussian water diffusion in humans using diffusion kurtosis imaging. NMR Biomed. 2006;19(2):236–247. [PubMed: 16521095]
245.
Tuch D.S. et al. High angular resolution diffusion imaging reveals intravoxel white matter fiber heterogeneity. Magn Reson Med. 2002;48(4):577–582. [PubMed: 12353272]
246.
Wedeen V.J. et al. Mapping complex tissue architecture with diffusion spectrum magnetic resonance imaging. Magn Reson Med. 2005;54(6):1377–1386. [PubMed: 16247738]
247.
Shin S.S. et al. High-definition fiber tracking for assessment of neurological deficit in a case of traumatic brain injury: Finding, visualizing, and interpreting small sites of damage. J Neurosurg. 2012;116(5):1062–1069. [PubMed: 22381003]
248.
Chan J.H.M. et al. Diffuse axonal injury: Detection of changes in anisotropy of water diffusion by diffusion-weighted imaging. Neuroradiology. 2003;45(1):34–38. [PubMed: 12525952]
249.
Lee Z.I. et al. Diffusion tensor magnetic resonance imaging of microstructural abnormalities in children with brain injury. Am J Phys Med Rehabil. 2003;82(7):556–559. [PubMed: 12819543]
250.
Jang S.H. Diffusion tensor imaging studies on corticospinal tract injury following traumatic brain injury: A review. NeuroRehabilitation. 2011;29(4):339–345. [PubMed: 22207060]
251.
Arfanakis K. et al. Diffusion tensor MR imaging in diffuse axonal injury. AJNR Am J Neuroradiol. 2002;23(5):794–802. [PMC free article: PMC7974716] [PubMed: 12006280]
252.
Bazarian J.J. et al. Diffusion tensor imaging detects clinically important axonal damage after mild traumatic brain injury: A pilot study. J Neurotrauma. 2007;24(9):1447–1459. [PubMed: 17892407]
253.
Miles L. et al. Short-term DTI predictors of cognitive dysfunction in mild traumatic brain injury. Brain Inj. 2008;22(2):115–122. [PubMed: 18240040]
254.
Mayer A.R. et al. A prospective diffusion tensor imaging study in mild traumatic brain injury. Neurology. 2010;74(8):643–650. [PMC free article: PMC2830922] [PubMed: 20089939]
255.
Hulkower M.B. et al. A decade of DTI in traumatic brain injury: 10 Years and 100 articles later. AJNR Am J Neuroradiol. 2013;34(11):2064–2074. [PMC free article: PMC7964847] [PubMed: 23306011]
256.
Marquez de la Plata C.D. et al. Diffusion tensor imaging biomarkers for traumatic axonal injury: Analysis of three analytic methods. J Int Neuropsychol Soc. 2010;17(1):24–35. [PMC free article: PMC3097093] [PubMed: 21070694]
257.
Mori S. et al. Atlas-based neuroinformatics via MRI: Harnessing information from past clinical cases and quantitative image analysis for patient care. Annu Rev Biomed Eng. 2013;15(1):71–92. [PMC free article: PMC3719383] [PubMed: 23642246]
258.
Faria A.V. et al. Atlas-based analysis of neurodevelopment from infancy to adulthood using diffusion tensor imaging and applications for automated abnormality detection. Neuroimage. 2010;52(2):415–428. [PMC free article: PMC2886186] [PubMed: 20420929]
259.
Paus T. et al. Maturation of white matter in the human brain: A review of magnetic resonance studies. Brain Res Bull. 2001;54(3):255–266. [PubMed: 11287130]
260.
Metting Z. et al. Structural and functional neuroimaging in mild-to-moderate head injury. Lancet Neurol. 2007;6(8):699–710. [PubMed: 17638611]
261.
Hunter J.V. et al. Emerging imaging tools for use with traumatic brain injury research. J Neurotrauma. 2012;29(4):654–671. [PMC free article: PMC3289847] [PubMed: 21787167]
262.
Bagley L.J. et al. Magnetization transfer imaging of traumatic brain injury. J Magn Reson Imaging. 2000;11(1):1–8. [PubMed: 10676614]
263.
McGowan J.C. et al. Diffuse axonal pathology detected with magnetization transfer imaging following brain injury in the pig. Magn Reson Med. 1999;41(4):727–733. [PubMed: 10332848]
264.
McGowan J.C. et al. Magnetization transfer imaging in the detection of injury associated with mild head trauma. AJNR Am J Neuroradiol. 2000;21(5):875–880. [PMC free article: PMC7976760] [PubMed: 10815663]
265.
Duckworth J.L, Stevens R.D. Imaging brain trauma. Curr Opin Crit Care. 2010;16(2):92–97. [PubMed: 20160645]
266.
Mamere A.E. et al. Evaluation of delayed neuronal and axonal damage secondary to moderate and severe traumatic brain injury using quantitative MR imaging techniques. AJNR Am J Neuroradiol. 2009;30(5):947–952. [PMC free article: PMC7051670] [PubMed: 19193759]
267.
Sinson G. et al. Magnetization transfer imaging and proton MR spectroscopy in the evaluation of axonal injury: Correlation with clinical outcome after traumatic brain injury. AJNR Am J Neuroradiol. 2001;22(1):143–151. [PMC free article: PMC7975548] [PubMed: 11158900]
268.
Asikainen I, Kaste M, Sarna S. Early and late posttraumatic seizures in traumatic brain injury rehabilitation patients: Brain injury factors causing late seizures and influence of seizures on long-term outcome. Epilepsia. 1999;40(5):584–589. [PubMed: 10386527]
269.
Englander J. et al. The association of early computed tomography scan findings and ambulation, self-care, and supervision needs at rehabilitation discharge and at 1 year after traumatic brain injury. Arch Phys Med Rehabil. 2003;84(2):214–220. [PubMed: 12601652]
270.
Sidaros A. et al. Diffusion tensor imaging during recovery from severe traumatic brain injury and relation to clinical outcome: A longitudinal study. Brain. 2008;131(2):559–572. [PubMed: 18083753]
271.
Porto L. et al. Morphometry and diffusion MR imaging years after childhood traumatic brain injury. Eur J Paediatr Neurol. 2011;15(6):493–501. [PubMed: 21783392]
272.
Marquez de la Plata C. et al. Magnetic resonance imaging of diffuse axonal injury: Quantitative assessment of white matter lesion volume. J Neurotrauma. 2007;24(4):591–598. [PubMed: 17439343]
273.
Strangman G.E. et al. Regional brain morphometry predicts memory rehabilitation outcome after traumatic brain injury. Front Hum Neurosci. 2010;4(182) [PMC free article: PMC2967347] [PubMed: 21048895]
274.
Kinnunen K.M. et al. White matter damage and cognitive impairment after traumatic brain injury. Brain. 2011;134(2):449–463. [PMC free article: PMC3030764] [PubMed: 21193486]
275.
Irimia A. et al. Neuroimaging of structural pathology and connectomics in traumatic brain injury: Toward personalized outcome prediction. Neuroimage Clin. 2012;1(1):1–17. [PMC free article: PMC3757727] [PubMed: 24179732]
276.
Belanger H.G. et al. Recent neuroimaging techniques in mild traumatic brain injury. J Neuropsychiatry Clin Neurosci. 2007;19(1):5–20. [PubMed: 17308222]
277.
Hillary F.G. et al. Functional magnetic resonance imaging technology and traumatic brain injury rehabilitation: Guidelines for methodological and conceptual pitfalls. J Head Trauma Rehabil. 2002;17(5):411–430. [PubMed: 12802252]
278.
Sanchez-Carrion R. et al. Frontal hypoactivation on functional magnetic resonance imaging in working memory after severe diffuse traumatic brain injury. J Neurotrauma. 2008;25(5):479–494. [PubMed: 18363509]
279.
Ueno H. et al. Brain activations in errorless and errorful learning in patients with diffuse axonal injury: A functional MRI study. Brain Inj. 2009;23(4):291–298. [PubMed: 19330592]
280.
Palacios E.M. et al. Resting-state functional magnetic resonance imaging activity and connectivity and cognitive outcome in traumatic brain injury. JAMA Neurol. 2013;70(7):845. [PubMed: 23689958]
281.
Christodoulou C. et al. Functional magnetic resonance imaging of working memory impairment after traumatic brain injury. J Neurol Neurosurg Psychiatry. 2001;71(2):161–168. [PMC free article: PMC1737512] [PubMed: 11459886]
282.
Luo Q. et al. Complexity analysis of resting state magnetoencephalography activity in traumatic brain injury patients. J Neurotrauma. 2013;30(20):1702–1709. [PMC free article: PMC3796321] [PubMed: 23692211]
283.
Huang M.-X. et al. Integrated imaging approach with MEG and DTI to detect mild traumatic brain injury in military and civilian patients. J Neurotrauma. 2009;26(8):1213–1226. [PubMed: 19385722]
284.
Gloor P, Ball G, Schaul N. Brain lesions that produce delta waves in the EEG. Neurology. 1977;27(4):326–333. [PubMed: 557774]
285.
Gaetz W, Otsubo H, Pang E.W. Magnetoencephalography for clinical pediatrics: The effect of head positioning on measurement of somatosensory-evoked fields. Clin Neurophysiol. 2008;119(8):1923–1933. [PubMed: 18579439]
286.
Bernabeu M. et al. Abnormal corticospinal excitability in traumatic diffuse axonal brain injury. J Neurotrauma. 2009;26(12):2185–2193. [PMC free article: PMC2824228] [PubMed: 19604100]
287.
Yasokawa Y.-T. et al. Correlation between diffusion-tensor magnetic resonance imaging and motor-evoked potential in chronic severe diffuse axonal injury. J Neurotrauma. 2007;24(1):163–173. [PubMed: 17263680]
288.
Signoretti S. et al. The protective effect of cyclosporin A upon N-acetylaspartate and mitochondrial dysfunction following experimental diffuse traumatic brain injury. J Neurotrauma. 2004;21(9):1154–1167. [PubMed: 15453986]
289.
Colley B.S, Phillips L.L, Reeves T.M. The effects of cyclosporin-A on axonal conduction deficits following traumatic brain injury in adult rats. Exp Neurol. 2010;224(1):241–251. [PMC free article: PMC2885519] [PubMed: 20362574]
290.
Suehiro E. et al. The immunophilin ligand FK506 attenuates the axonal damage associated with rapid rewarming following posttraumatic hypothermia. Exp Neurol. 2001;172(1):199–210. [PubMed: 11681852]
291.
Fujita M. et al. The combination of either tempol or FK506 with delayed hypothermia: Implications for traumatically induced microvascular and axonal protection. J Neurotrauma. 2011 [PMC free article: PMC3136741] [PubMed: 21521034]
292.
DiLeonardi A.M, Huh J.W, Raghupathi R. Differential effects of FK506 on structural and functional axonal deficits after diffuse brain injury in the immature rat. J Neuropathol Exp Neurol. 2012;71(11):959–972. [PMC free article: PMC3495060] [PubMed: 23095847]
293.
Schouten J.W. et al. A review and rationale for the use of cellular transplantation as a therapeutic strategy for traumatic brain injury. J Neurotrauma. 2004;21(11):1501–1538. [PubMed: 15684646]
294.
Harting M.T. et al. Cell therapies for traumatic brain injury. Neurosurg Focus. 2008;24(3–4):E18. [PMC free article: PMC3721356] [PubMed: 18341394]
295.
Salman H, Ghosh P, Kernie S.G. Subventricular zone neural stem cells remodel the brain following traumatic injury in adult mice. J Neurotrauma. 2004;21(3):283–292. [PubMed: 15115603]
296.
Riess P. et al. Embryonic stem cell transplantation after experimental traumatic brain injury dramatically improves neurological outcome, but may cause tumors. J Neurotrauma. 2007;24(1):216–225. [PubMed: 17263685]
297.
Kulbatski I. et al. Glial precursor cell transplantation therapy for neurotrauma and multiple sclerosis. Prog Histochem Cytochem. 2008;43(3):123–176. [PubMed: 18706353]
298.
Seledtsov V.I. et al. Cell therapy of comatose states. Bull Exp Biol Med. 2006;142(1):129–132. [PubMed: 17369922]
299.
Velly L. et al. Erythropoietin 2nd cerebral protection after acute injuries: A double-edged sword? Pharmacol Ther. 2010;128(3):445–459. [PubMed: 20732352]
300.
Mammis A, McIntosh T.K, Maniker A.H. Erythropoietin as a neuroprotective agent in traumatic brain injury: Review. Surg Neurol. 2009;71(5):527–531. discussion 531. [PubMed: 18789503]
301.
Hellewell S.C. et al. Erythropoietin improves motor and cognitive deficit, axonal pathology, and neuroinflammation in a combined model of diffuse traumatic brain injury and hypoxia, in association with upregulation of the erythropoietin receptor. J Neuroinflammation. 2013;10(156) [PMC free article: PMC3896698] [PubMed: 24344874]
302.
Yatsiv I. et al. Erythropoietin is neuroprotective, improves functional recovery, and reduces neuronal apoptosis and inflammation in a rodent model of experimental closed head injury. FASEB J. 2005;19(12):1701–1703. [PubMed: 16099948]
303.
Talving P. et al. Erythropoiesis stimulating agent administration improves survival after severe traumatic brain injury: A matched case control study. Ann Surg. 2010;251(1):1–4. [PubMed: 19779323]
304.
Corwin H.L. et al. Efficacy of recombinant human erythropoietin in critically ill patients: A randomized controlled trial. JAMA. 2002;288(22):2827–2835. [PubMed: 12472324]
305.
Corwin H.L. et al. Efficacy and safety of epoetin alfa in critically ill patients. N Engl J Med. 2007;357(10):965–976. [PubMed: 17804841]
306.
Yamamoto M. et al. Effect of erythropoietin on nitric oxide production in the rat hippocampus using in vivo brain microdialysis. Neuroscience. 2004;128(1):163–168. [PubMed: 15450363]
307.
Tubbs R.S. et al. Does the neuroprotective agent erythropoietin amplify diffuse axonal injury in its early stages? Med Hypotheses. 2007;69(6):1385–1386. [PubMed: 17449190]
308.
Mills J.D. et al. Omega-3 fatty acid supplementation and reduction of traumatic axonal injury in a rodent head injury model. J Neurosurg. 2011;114(1):77–84. [PubMed: 20635852]
309.
Bailes J.E, Mills J.D. Docosahexaenoic acid reduces traumatic axonal injury in a rodent head injury model. J Neurotrauma. 2010;27(9):1617–1624. [PubMed: 20597639]
© 2016 by Taylor & Francis Group, LLC.
Bookshelf ID: NBK326722PMID: 26583181

Views

  • PubReader
  • Print View
  • Cite this Page

Other titles in this collection

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...