Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Mar 21.
Published in final edited form as: Curr Neurol Neurosci Rep. 2017 Jul;17(7):52. doi: 10.1007/s11910-017-0762-x

Pathophysiology and Treatment of Memory Dysfunction after Traumatic Brain Injury

Rosalia Paterno 1,*, Kaitlin A Folweiler 2,*, Akiva S Cohen 2,3
PMCID: PMC5861722  NIHMSID: NIHMS949462  PMID: 28500417

Abstract

Memory is fundamental to everyday life, and cognitive impairments resulting from traumatic brain injury (TBI) have devastating effects on TBI survivors. A contributing component to memory impairments caused by TBI are alterations in the neural circuits associated with memory function. In this review, we aim to bring together experimental findings that characterize behavioral memory deficits and the underlying pathophysiology of memory-involved circuits after TBI. While there is little doubt that TBI causes memory and cognitive dysfunction, it is difficult to conclude which memory phase i.e., encoding, maintenance or retrieval is specifically altered by TBI. This is most likely due to variation in behavioral protocols and experimental models. Additionally we review a selection of experimental treatments that hold translational potential to mitigate memory dysfunction following injury.

Keywords: Traumatic brain injury, Memory, Hippocampus, Prefrontal cortex, Fluid percussion injury, Controlled cortical impact injury

Introduction

Traumatic brain injury (TBI) is defined as any force to the head that causes alteration in neurological function. TBI presents a significant health issue in the United States, with more than 2.5 million cases resulting in emergency department visits, hospitalizations or fatality [1]. Furthermore, memory impairment is one of the most common neurological manifestations of TBI [24]. Indeed, memory and TBI appear to be intrinsically linked, as the hippocampus and cortex, significant brain regions involved in the physiological circuits of memory, are often damaged after TBI [5,6]. In order to manage and mitigate memory dysfunction ascribed by TBI patients [7] it is imperative to determine the physiological mechanisms linking TBI and these neural substrates of memory.

In this review, we aim to present the current state of research linking memory and TBI by systematically describing the type of memory tested and the different animal models implemented. Additionally, we review studies addressing the underlying neural physiology of memory-associated circuits, predominately within the hippocampus and cortex, and how experimental models of TBI contribute to understanding amnesic pathophysiology. Lastly, we provide a brief overview of promising therapeutic strategies that have potential to target these physiological vulnerabilities within neuronal circuits of memory.

Experimental animal models of TBI

In order to study TBI pre-clinically, scientists have developed several animal models to mimic human pathophysiology. TBI animal models can be divided into closed and open head injury models. The closed head injury (CHI) models, such as Marmarou’s and Feeney’s weight-drop models, are characterized by the fact that the skull remains intact before the injury [810]. CHI causes blood brain barrier disruptions, edema permeability and transient alterations in neurological status [11]. Apart from the weight-drop models, another CHI model type is blast injury, which consists of projecting blast pressure waves from a compressed-gas driven shock tube onto the head of an anesthetized animal [12]. The blast model of TBI was developed to mimic the pressure waves from improvised explosive devices (IEDs) during combat warfare, and has also been shown to replicate human TBI cognitive symptoms and pathology [13].

In contrast to CHI models, open head injury models administer the injury through a craniectomy, directly onto the surface of the dura. The most common forms of these types of models are lateral fluid percussion injury (LFPI) and controlled cortical impact (CCI) [14,15]. LFPI induces injury by delivering a fluid pressure wave onto the exposed dura, and as a result, inducing a focal cortical contusion at the desired level of severity, as well as diffuse subcortical neuronal injury in the side ipsilateral to the injury [16,17]. CCI on the other hand, utilizes a pneumatically-driven impactor on the exposed dura to produce a precise injury, characterized by a focal cortical contusion. Both models replicate aspects of human TBI pathology as well as injury-induced cognitive deficits.

A hallmark of all the aforementioned injury models is that they can be adjusted to produce various levels of injury severity. While each of these models has certain strengths in replicating aspects of TBI, they each have limitations in that they cannot recapitulate all features of human TBI. Therefore, attention must be paid to the model selected in experimental studies in order to understand which aspect of injury the authors are best trying to replicate. Even with the individual limitations of animal models, they remain essential for studying the functional, biomechanical, cellular and molecular aspects of human TBI that are difficult to address in the clinical realm. Though we have highlighted a few common models here, please refer to additional reviews that further detail the diversity of animal models [1820•].

Types of memory and TBI

Memory is a dynamic neural and cognitive progression characterized by three separate processes: encoding, maintenance and retrieval [21]. Encoding is the transformation of an experience into a discrete neural representation, known as the memory engram. Maintenance refers to the endurance of the engram across time, and retrieval is the ability to voluntarily reinstate the engram into the forefront of consciousness. It is widely accepted that the hippocampus plays a major role in all three of these memory processes [22,23]. However, it is unknown how, and whether, TBI disproportionately alters one of these processes as compared to the others.

Previously, memory was categorized into short-term or long-term durations. However, this distinction is presently laden with misunderstanding and controversy [24]. Current research conceptualizes the time dependence of memory traces as working, episodic, and semantic memory. Working memory is defined as the cognitive ability to transiently hold, process, and manipulate information [25]. In regards to episodic and semantic memory, these types are actually less dependent on time, but rooted more in the amount of personal experience associated with the memory [26]. Episodic memory receives and stores information about temporally-dated episodes or events, and temporal-spatial relations among these events (e.g., “I remember that last Tuesday at 3pm I was sitting at my desk, talking to my co-worker, who was standing by the door”) [27]. Semantic memory, on the other hand, refers more to the retrieval of memorized facts or events, and their meanings, that one might not necessarily have had a personal experience with (e.g. reciting state capitals) [28]. Importantly, the transition from episodic to semantic memory is governed by a hypothetical cognitive process called consolidation. Studies in cognitive neuroscience have suggested that consolidation mediates the transition of the engram from the medial temporal lobe to the cerebral cortex [29]. Memory engrams that have presumably passed the consolidation phase and remain constant over time are often referred to as “reference memory” in rodent behavioral tasks.

In TBI animal models, it is common to use different type of behavioral tasks to study memory. The lack of standardized memory tasks in these studies makes it difficult to map any one result to specific types (working, episodic, and semantic) or processes (encoding, maintenance, and retrieval) of memory. To help clarify this, we have organized all of the animal studies linking TBI and memory based on the type of memory tested (Table 1). These studies are discussed in further detail below.

Table 1.

Summary of behavioral changes in experimental models of TBI

Table 1. Overview of the current literature organized by type of memory disrupted after experimental TBI. Abbreviations: d, day; CCI, controlled cortical impact; FPI, fluid percussion injury; ITI, inter-trial time; mod, moderate; LFPI, lateral fluid percussion injury; MWM, Morris water maze; PID, post injury days; rec, recent; ref, reference; rem, remote; sec, seconds; TOR, temporal order recognition memory.

Ref Injury Model Animal age and gender Animal group Memory Task Protocol Test performance PID Behavioral effects (TBI vs Sham) p value
Working memory
Kobori et al (2006) Moderate-severe CCI Adult male rats Sham: 8
TBI: 8
MWM working mem
Training: 5 trials with 5 sec delay between sample and choice phase and 4 min ITI
Test: the following day
13–14 and 27–28 ⇑ latency to reach the platform at 14 and 28 PID p<0.001
Hoskinson et al (2009) Moderate-severe CCI Adult male rats Sham: 6
TB:I 10
MWM working mem
Training 5 trials with 5 sec delay between sample and choice phase and 4 min ITI for 1 day followed by test
Swim T maze: 10 training trials/d for 7 d (10 s ITI) followed by 5 d testing
120 (3 months) ⇑ latency to reach the platform in MWM task p<0.05
⇓ correct choice percentage with the same amount of training in swim T maze p<0.001
Sebastian et al (2013) CCI Adult male rats Sham: 3–4
TBI: 3–4
Radial 8-arm maze
Training: 10 trials/d for 6 d starting at 21 PID
Test: 3–5 trials followed by test at 42 PID
42 (5 weeks) ⇑ number of working memory errs p<0.05
Both groups (Sham & TBI) improved performance over days p<0.001
Sebastian et al (2013) CCI Adult male rats Sham: 3–6
TBI: 3–6
Radial 8-arm maze
Training: 10 trials/d for 6 d starting at 42 PID
Test: 3–5 trials followed by test at 63 PID
63 (9 weeks) ⇑ number of working memory errs during training p<0.001
No difference during test NS
Lyeth et al (1990) Mild – Moderate LFPI Adult male rats Sham: 8
TBI mild: 7
TBI mod: 10
Radial 8-arm maze
Training: 1 session per day (10 sec ITI) until < 1 err/d for 3 consecutives days before TBI
5 – 25 Impaired working memory performance up to 5 PID in mild TBI p<0.05
⇑number of working memory err up to 15 PID in moderate TBI p<0.05
Intact ref memory NS
Whiting et al (2006) Moderate LFPI Adult male rats Sham: 9
TBI: 10
Swim T maze
Training: 15 trials/d with 7 sec delay between sample and choice phase (11–15 PID).
Test: 15, 30, 120 sec between sample and choice phase
16–18 No difference with 15 sec delay NS
Impaired working memory performance with delay 30 and 120 sec p<0.05
Eakin et al (2012) Mild LFPI Adult male rats Sham: 3
TBI: 3
Swim T maze
Training: 0 delay between sample and choice phase.
Test: 15, 30, 60, 120 delay between sample and choice phase
30–60 Impaired working memory performance with delays p<0.01
Sweet et al (2014) Moderate midline FPI Adult male rats Sham: 6
TBI: 6
Swim T maze
Training: 0 delay between sample and choice phase for 5 d.
Test: 15 trials/d with 60 sec delay between sample and choice phase for 12 d
3 weeks after electrodes implantation Impaired working memory performance during test p<0.01
Smith et al (2015) Mild/moderate LFPI Young adult mice Sham: 9
TBI: 7
T maze
Training: 10 trials/d with 0 delay per 7 d before TBI Test: immediately after TBI for 7 continuous days
7 Impaired working memory performance p<0.01
Spatial anterograde memory
Hamm et al (1992) Moderate CCI Adult male rats Sham: 8
TBI: 8
MWM ref
4 trials/d for 5 d (4 min ITI)
11–15 & 30–34 ⇑ latency to reach the platform at both time points p<0.05
⇑ time spent in the peripheral part p<0.05
Dash et al (1995) CCI Adult male rats Sham: 10
TBI: 10
MWM ref
4 trials/d for 5 d (4 min ITI)
14 ⇑ latency to reach the platform p<0.01
Fox et al (1998) CCI Adult male mice Sham: 11
TBI: 15
Barnes maze
1 session/d for 4 d
7–10 ⇑ escape latency p<0.05
⇓ spatial pattern strategy p<0.05
Pierce et al (1993) Moderate LFPI Adult male rats Sham: 15
TBI: 12
MWM ref
1992
7–8 ⇑ latency to reach the platform p<0.01
Whiting et al (2008) Moderate LFPI Adult male rats Sham: 8
TBI: 8
MWM ref
Training: blocks of 3 trials (30 min between blocks) until criteria reached
Test: after 4, 8, 24 h
11–12 ⇑ trials to reach the criteria p<0.01
No difference to reach the platform after 4, 8, 24h delay NS
Fedor et al (2010) Moderate LFPI Adult male rats Sham: 8
TBI: 12
Barnes maze
Training: 2 trials (2 min ITI) for 3 d
90 No difference in the escape latency NS
⇑ peripheral search strategy p<0.01
Hylin et al (2013) Mild LFPI (1 atm) & Mild LFPI (1.5atm) Adult male rats Sham: 17
TBI (1atm): 20
TBI(1.5 atm): 11
MWM ref
Training: 10 trials/d (4min ITI) for 1 d followed by a probe trial after 30 min
5 No difference in the mild (1 atm) group NS
⇑escape latency during training and probe trial in the mild (1.5 atm) group p<0.01
⇑ peripheral search strategy p=0.05
Lee at al (2013) Moderate LFPI Adult male rats Sham: 14
TBI: 13
Barnes maze
Training: 2 trials (2 min ITI) for 3 d
5–7 ⇑escape latency p<0.05
⇑ peripheral search strategy p<0.05
Zohar et al (2011) CHI Adult male mice Sham: 10
TBI: 10
MWM ref
Training: 6 trials/d for 4 d and test the following day
7, 30, 60, 90 ⇑escape latency p<0.05
Spatial retrograde memory
Smith et al (1991) Moderate and severe LFPI Adult male rats Sham: 12
TBI mod: 12
TBI sev: 13
MWM ref
Training: 10 trials/d for 2 days until 2.5 h before TBI
Test: 42 h post-injury
42 h Impaired task performance p<0.01
Okiama et al (1992) Moderate LFPI Adult male rats Sham: 11
TBI: 20
MWM ref
Training: 10 trials/d for 2 days before TBI
Test: 42 h post-injury
42 h ⇑escape latency p<0.01
Hicks et al (1993) Mild LFPI Adult male rats Sham: 10
TBI mild: 7
MWM ref
Training: 10 trials/d for 2 days until 2.5 h before TBI
Test: 42 h post-injury
42 h Impaired task performance p<0.05
Smith et al (1994) Moderate LFPI Adult male rats Sham: 11
TBI: 12
MWM ref
Training: 10 trials/d for 2 days until 2.5 h before TBI
Test: 7 PID
7 Impaired task performance p<0.05
Smith et al (1994) Moderate LFPI Adult male rats Sham: 15
TBI: 11
MWM ref
Training: 10 trials/d for 2 days until 2.5 h before TBI
Test: 14 PID
14 Impaired task performance p<0.05
Whiting et al (2008) Moderate LFPI Adult male rats Sham: 8
TBI rec: 8
TBI rem: 8
MWM ref
Training: 5 trials/d for 5 d (10 min ITI) before (1 or 13 days) TBI
Test: at 14 PID a reminder trial and a second probe trial
14 No difference with any delays NS
Sebastian et al (2013) CCI Adult male rats Sham: 4
TBI: 4
Radial 8-arm maze
Training: 10 trials/d for 6 d before TBI
Test: 3–5 trials followed by test at 14 PID
14 No impairment in ref memory NS
Impaired working memory performance p<0.05
Episodic memory
Gurkoff et al (2013) Moderate LFPI Adult male rats Sham: 12
TBI: 11
TOR
three times 5 min acquisition separated by 15 min interval followed by test
14 Impaired temporal order discrimination p<0.05
Gurkoff et al (2013) Moderate LFPI Adult male rats Sham: 11
TBI: 12
Topological task
Habituation for 15 min
Test after 5 min delay
14 No difference in the ability to discriminate a spatial change in the object NS
Gurkoff et al (2013) Moderate LFPI Adult male rats Sham: 9
TBI: 10
Metric tsk
Habituation for 15 min
Test after 5 min delay
14 Impaired ability to discriminate a distance change between objects p<0.05
Zhang et al (2015) Mild CHI Adult male rats Sham: 8
TBI: 11
TOR 45 and 90 Impaired temporal order discrimination p<0.05

Working memory

In the TBI animal literature, working memory is tested primarily using three different tasks: T-maze, radial arm maze, and a specific working memory protocol with the Morris water maze (MWM) [3032][32,31,30]. Each task consists of 3 trial phases: sample, delay, and choice phase. In the sample phase, animals must learn the path to the maze endpoint, whether that endpoint be a food reward at the end of a T-maze arm, or escape to the hidden platform. In the choice phase, depending on the learned rule, animals must choose the appropriate arm for T-maze and radial maze, or reach the hidden platform location in the MWM. A short delay with a duration lasting from seconds to minutes, separates the sample and choice phases to probe memory recall.

Following (moderate, severe or mild) CCI, working memory dysfunction persists for at least 16 weeks post-injury as assessed with all three of the previously described working memory tasks. [3335]. After LFPI, working memory dysfunction follows a temporal evolution that is dependent on injury severity. Acutely, after mild/moderate LFPI, animals demonstrated a deficit in the first 7 post-injury days (PID) with zero time delay between sample and choice phases, indicating a working memory deficit [36•]. When animals were allowed to recover for 15 to 60 PID, no deficits were seen when the delay time was less than 30 seconds, however working memory performance was impaired with a delay time above or equal to 30 seconds [3740]. Taken together, these data support the hypothesis that TBI diminishes the ability to encode new information in the acute post-injury phase and to maintain information chronically up to 60 days in rodents.

Spatial navigation memory

Spatial memory, a type of memory that records information about a subject’s environment and navigation within that environment, has representations in both working and episodic memory. In animal models, spatial memory is tested using different behavioral tasks. One of the most common spatial learning and memory tasks in rodents is the MWM [32,41,42]. In this task, animals are placed in a circular, cloudy water pool with a submerged platform. The goal of the task is for the animal to learn the fixed location of the submerged platform using visual cues located on the walls surrounding the pool, thereby developing a spatial map of the maze. Other widely used spatial tasks include the Barnes and radial arm mazes, in which animals have to find the escape hole after been placed either on an illuminated, circular platform or in an 8-arm radial maze [43,44]. In addition, in each of these tasks, reference memory can also be tested with the use of a probe trial. This trial is added after the testing period to determine if the animal can remember the spatial path. The reference trial is often referred to has an assessment of “long-term” memory, however as discussed above, the probe trial actually is examining the consolidation process.

With respect to spatial episodic memory, which includes information regarding a specific spatial episode/location, numerous experiments have been undertaken in order to test spatial memory after TBI. Specifically, two types of memory have been tested: retrograde and anterograde. In retrograde memory, animals are trained before injury and tested hours/days after injury. This procedure tests the ability of TBI animals to recall previously learned information. In the anterograde memory protocol, animals are robustly trained after injury for many days and then tested.

Anterograde spatial memory

When tested in Barnes and Morris water mazes after moderate CCI, animals demonstrated a significant impairment of spatial learning performance when examined at 10 PID and up to 1 month after injury [4548]. Specifically, TBI animals exhibited an increased escape latency in the ability to reach the platform or the dark hole, and a more peripheral search pattern compared to sham. Over time, TBI animal performance improved, reaching a performance plateau that could be maintained up to 15 days from the last training day [47,49]. The CHI experimental animal model, in a similar fashion, demonstrated an impaired spatial performance up to 3 months post-injury [50].

Using the LFPI animal model, the severity of the injury causes different behavioral changes. Specifically, a MWM study using two different levels of mild LFPI severity (1.0–1.5 atm) demonstrated an increased latency to escape during the learning trials. However, in testing trials, only the higher injury severity resulted in longer latency to reach the platform [51]. Moderate LFPI on the other hand was associated with significant impairments in spatial learning performance up to 15 PID. These animals recover with no deficit observed at 3 months post-injury [5254]. Of note, behavioral tasks examining spatial learning performance with TBI animals has demonstrated that enhanced training leads to better testing performance. Rigorous training results in TBI animals successfully completing the task when compared to cohorts with nominal training [55]. Further experiments are needed to specifically determine whether this effect is due to establishment of a neural spatial representation of the maze or, a non-spatial strategy.

Retrograde spatial memory

Animals trained before undergoing LFPI (i.e., retrograde) demonstrated an impairment in recalling the already learned information up to 14 PID [56]. A study with a radial arm maze showed impairment in memory retention after mild and moderate TBI up to 25 PID [35,37,5659]. Interesting, no deficits were observed when given a brief reminding procedural prompt [35,55]. These data suggest that information learned before TBI can be recalled with longer recovery time and robust reminder training.

Episodic memory

After TBI, there are few experiments focused on examining episodic memory. In order to test the time (or “when”) component of episodic memory, Gurkoff and colleagues used a temporal order task [60•]. Specifically, these animals were exposed to an odor sequence and after an hour delay or longer, were tested for their ability to discriminate the initial odor sequence versus a new sequence. Sham animals preferentially explored the initial odor, whereas injured animals demonstrated no preference. In similar fashion when using a different task, animals with mild CHI were unable to discriminate between odors up to 90 PID [61]. The propensity for control animals to prefer the initial sequence is analogous to the primacy effect in humans, which describes the tendency to more easily remember the first items presented in a sequence during a memory task [62,63].

To test the spatial (or “where”) component of episodic memory, topological tasks, such as object place recognition tasks, are typically used [64]. Here, animals are tested for their ability to distinguish when two objects have their locations transposed in space. Mild and moderate LFPI demonstrated no significant impairments when a short delay time window is used between familiarization and test phase [60•]. However, at a longer time window (1 hour), TBI animals were unable to discriminate that the object’s new location from the old location (unpublished data).

Pathophysiology of memory impairment after TBI

Underlying the deficits seen in experimental models of TBI, are alterations in the physiological circuitry of brain regions that confer the different types of memory we have detailed in this review. Disruption to the hippocampus and cortex—regions critical to memory function—are pathological features of both human and animal models of TBI [5,6,65]. The hippocampus is a key part of a large network of brain areas that interact to store and retrieve recent events, and guide memory-driven actions. It receives inputs from different cortical regions which are essential for episodic memory. Specifically, the hippocampus receives spatial and non-spatial information about the environment via projections from the medial entorhinal cortex and lateral entorhinal cortex, respectively. Interactions between the hippocampus and the prefrontal cortex are also involved in encoding, processing, and performance of working memory.

In order to review the current knowledge of TBI-induced pathophysiology within the hippocampus and cortex, we have chosen to view these structures and their sub-regions, with a “circuit-level” perspective. Therefore, we have narrowed our definition of pathophysiology to include functional changes in neuronal output (i.e., how likely it is to fire an action potential), examined by electrophysiological techniques. Additionally, we recognize that even though the hippocampus and prefrontal cortex have been physiologically well studied in the TBI literature, other brain regions that are important for memory, such as upstream cortical regions (e.g., entorhinal cortex) have not been as well characterized.

Hippocampus

As described earlier, experimental models of TBI show deficits in episodic memory. Physiological disruption of hippocampal circuitry, comprised of the dentate gyrus, areas CA3 and CA1, are thought to be largely responsible for disruption of episodic memory after TBI, including spatial memory [6671].

Dentate gyrus

The dentate gyrus (DG) is considered an important region in hippocampal memory processing. Specifically, it is involved in pattern separation of cortical inputs [7277]. Consequently, the DG is a crucial regulator of incoming excitability to the hippocampus, and acts as a filter or gatekeeper of cortical input to the hippocampus [78]. The filtering function of the DG is conferred by the low excitability of its principal cell type, the granule cell. Normally, granule cells have an extremely low propensity to fire action potentials due to a combination of their intrinsic properties and strong GABAAergic inhibition by diverse interneuron populations located around the granule cell layer and in the hilus [7983]. However, after TBI, the DG becomes hyperexcitable, thus disrupting its filtering capabilities [8486].

TBI has been shown to alter inhibitory transmission onto granule cells, comprised of phasic and tonic components. Phasic, synaptic GABAAergic transmission onto the granule cells is diminished by TBI [84,85,87]. One month after LFPI phasic inhibition recovers; however after CCI, inhibition remains diminished for several months post-injury [88,89]. Compromised phasic inhibition is associated with reduced expression of the potassium-chloride membrane transporter KCC2, thus decreasing the driving force of chloride through GABAA receptors [90]. Therefore, a reduction in phasic GABAAergic inhibition appears to be present across TBI models, but the duration of these changes varies with injury severity.

In contrast, tonic inhibition is altered in opposite directions depending on the cell type. For example, tonic inhibition is enhanced onto granule cells, while it is diminished onto a subpopulation of granule cells, known as semilunar granule cells [9195]. Therefore, alteration in inhibition leads to a complicated dysfunction of the DG.

In order to understand alterations in granule cell firing, it is imperative to examine alterations in DG interneurons. Of the diverse GABAAergic interneuron subtypes, only somatostatin-positive (SOM) interneurons, have been examined physiologically after TBI. A study by Hunt and colleagues demonstrated that SOM interneurons receive more glutamatergic synaptic input after injury, and fire more action potentials [89]. In addition, glutamatergic mossy cells have been shown to fire more action potentials, resulting in delayed excitatory postsynaptic currents in granule cells [87]. Increased mossy cell firing is due to a shift in enhanced afferent excitatory input as well as homeostatic compensation of intrinsic properties after TBI [96].

In summary, the current literature suggests that posttraumatic DG hyperexcitability is primarily due to changes in GABAAergic and glutamatergic synaptic transmission within this sub-region. Future studies are needed to explore the functionality of the other DG interneuron types.

Area CA3

Area CA3 also plays an important role in episodic memory encoding and retrieval [7274,97,98]. Despite the importance of this circuit in memory processing, few electrophysiological studies have examined area CA3 after TBI. It is known however, that area CA3 neurons are vulnerable to death after moderate TBI [86,99102]. One study using a CHI model showed no acute change in the intrinsic membrane properties of CA3 pyramidal neurons acutely after injury, but was not able to examine later time points because cell survival did not exceed 3 days after injury [103].

In addition to injury-induced cell loss, CA3 pyramidal cells may also be vulnerable to oxidative damage from in vitro slice preparation, if taken from mature, adult tissue [104]. Therefore, the high susceptibility to both intended and unintended tissue injury, may explain why so few studies have focused on area CA3 after experimental TBI.

Area CA1

Area CA1 has a distinct role in the encoding and retrieval of episodic memories [105108]. In contrast to the DG’s hyperexcitable response to injury, area CA1 circuit activity becomes hypoexcitable. One week after FPI, CA1 has demonstrated a decreased net response to afferent fiber stimulation, accompanied by a higher threshold to initiate population spikes [84,109]. Therefore, the output of the CA1 circuit, mediated by the firing of CA1 pyramidal neurons, is diminished after injury.

The hypoexcitable state of the CA1 circuit can partly be explained by changes in synaptic inputs onto pyramidal cells. A study utilizing a model of lateral cortical contusion injury showed a reduction in fiber volley amplitude of afferent Schaeffer collaterals two days after injury [110]. This same study also found that at 7 PID, fiber volley amplitudes were restored, yet synaptic strength remained depressed. One week following FPI, postsynaptic responses to evoked glutamatergic events from both AMPA- and NMDA-receptors have also been shown to be diminished [111]. This indicates that while glutamatergic afferent fibers can re-innervate their targets, the postsynaptic machinery may still be disrupted.

In addition to decreased glutamatergic excitation, pyramidal cells also receive increased GABAA-receptor mediated inhibition from local interneurons. A study from our laboratory using voltage-sensitive dye imaging revealed hyperpolarization in stratum oriens of area CA1 after LFPI, due to enhanced GABAergic inhibition from cannabinoid-sensitive cholecystokinin (CCK) interneurons [112]. Therefore, imbalances in excitatory and inhibitory synaptic inputs appear to underlie diminished CA1 circuit efficacy.

There are other studies from the TBI literature that instead report hyperexcitability in area CA1 after TBI. Two days after moderate LFPI, pyramidal cells showed hyperexcitability accompanied by transiently enhanced afferent input [113,114]. One study using a CHI model of injury, observed more frequent spontaneous action potentials in CA1 pyramidal cells [103]. Additionally, one week following CCI, CA1 pyramidal cells were shown to receive less inhibitory currents, as well as a selective loss of GABAAergic interneurons in stratum pyramidale [115]. Taken altogether, these differential results in CA1 may reflect experimental variations such as time points examined after injury, injury model, or injury severity, as well as slice preparation. Much is left to future studies to examine the effects of these factors on CA1 circuit function.

After TBI, only a few studies have attempted to correlate in vivo hippocampal activity with memory-associated behavior. Fedor and colleagues measured the correlation between hippocampal theta rhythm—a narrow neuronal oscillation associated with memory processing—and spatial memory performance in a Barnes Maze after moderate FPI [54]. Rats that had poor spatial strategies in the maze demonstrated a decrease in theta activity. Another study measuring single-neuron spiking activity found that FPI animals had decreased bursting activity in place cells, which was also associated with poor memory performance in the T-maze [116]. In contrast, in a mild FPI model, a study showed only a decrease in broadband activity (a measure of the overall multi-unit activity) and not theta oscillations [117] supporting the hypothesis that TBI severity leads to different pathologies.

Prefrontal cortex

The prefrontal cortex (PFC), a cortical region heavily connected to the hippocampus, is critically involved in working memory. Recently, we demonstrated that an impairment of working memory in a T-maze task is accompanied by reduced excitability in the medial PFC [36•]. One week after LFPI, layer II/III neurons received an imbalance of more frequent spontaneous and miniature EPSCs and smaller amplitude IPSCs. Changes in synaptic inputs were additionally accompanied by an increase in action potential threshold, as well as a decrease in principal neuron firing rate. Downstream in layer V of medial PFC, pyramidal neurons did not experience a change in afferent synaptic inputs, but rather had shifted intrinsic membrane properties, such as a decrease in input resistance. In contrast another study, using an acute slice injury model, reported increased excitatory, and decreased inhibitory, synaptic currents onto layer V pyramidal cells [118]. These results demonstrate that a combination of layer-specific synaptic and intrinsic alterations occur in the medial prefrontal cortex after injury.

In cortical slices, it has been demonstrated that there can be differential effects on the intrinsic properties of pyramidal neurons and their firing properties. Both axotomized (i.e., severed) and intact cortical pyramidal neurons maintain normal membrane properties, yet the axotomized cells have a lower propensity to fire action potentials compared with intact neurons [119]. An in vitro model of TBI mechanical with cultured cortical neurons, resulted in the membrane composition of glutamatergic receptors. That is, after stretch injury, calcium-permeable AMPA receptors were upregulated in plasma membranes [120,121].

In relation to brain macrocircuitry, the prefrontal cortex is involved in rhythmic neuronal oscillations, such as the thalamocortical relay. While no in vivo study has been conducted, in vitro brain slices reveal a significant decline in the presence of these oscillations after FPI [122]. In summary, there are significant functional shifts in the PFC circuit that diminish large-scale brain activity patterns and correlate with working memory impairments.

Potential therapeutic strategies to improve memory dysfunction after TBI

To date, the promising results from animal studies of potential TBI therapies (calcium-channel antagonist, steroids, glutamate agonists, NMDA-receptor antagonists, oxygen free-radical scavengers, immune system modulation, statins, progesterone, hypothermia, etc.) have not been translated into successful phase 3 clinical trials. The reasons for the failure to translate bench research to bedside clinical practice are multi-fold and a recent review authored by Chakraborty et al 2016 analyzes in detail potential different causes of this failure [123][123]. However, even with unsuccessful clinical trial results, we believe that some of those therapies can be useful in TBI treatment if they would be designed to guide the pathophysiology. Below, we will focus on new potential therapies that have demonstrated memory improvement, but have not been tested yet in the clinical setting with a clinical trial. All studies are described in the text and summarized in Table 2.

Table 2.

Summary of neurobehavioral changes induced by therapy in experimental models of TBI

Overview of promising TBI therapeutic strategies. Abbreviations: BCAA, branched chain amino acids; CCI, controlled cortical impact; d, day; DBS, deep brain stimulation; DR, dorsal raphe; EE, environmental enrichment; E/I, Excitatory/Inhibitory; FPI, fluid percussion injury; HFS, high frequency stimulation; hNSCs, fetal human neural stem cell; LFPI, lateral fluid percussion injury; LFS, low frequency stimulation; MR, Midbrain medial raphe; MSN, medial septal nucleus; MWM ref, Morris water maze reference protocol; NOR, novel object recognition task; PID, post injury days; sol, solution; STD, standard; TBS, theta burst stimulation; TOR, temporal order recognition task;

Ref Injury Model Animal age and gender Memory Task Brain target Therapy Behavioral effects (TBI + therapy vs TBI) p value
Deep brain stimulation (DBS)
Lee et al (2013) Moderate LFPI Adult male rats Barnes maze (5–7 PID) Medial septal nucleus (MSN) DBS: 7.7 Hz, 1 msec pulse width, 80 uA for 1′ before task performance ⇓ escape latency p=0.05
⇑ spatial strategy search --
Carballosa et al (2013) Moderate LFPI Adult male rats MWM ref (35–37 PID)
MWM working (38–39 PID)
Midbrain medial raphe (MR), dorsal raphe (DR) DBS MR: 8 or 24 Hz, 5′ trains alternate with 5′ break during daylight, starting 4–6 h post-injury for 7d
DBS DR: 8 Hz, starting 4–6 h post-injury for 7 d
DBS MR delay: starting 7 PID for 7 d
⇑ spatial learning in the group DBS MR or DR 8 Hz. p<0.05
Improved working memory in the group DBS MR 8 Hz p<0.05
Sweet et al (2014) Moderate midline FPI Adult male rats Swim T maze MWM ref Fornix LFS: 5 Hz
HFS: 130 Hz
TBS: 200 Hz in 50 ms trains (5 trains/sec)
Improved performance in both tasks with TBS stimulation p<0.05
Lee et al (2015) Moderate LFPI Adult male rats Object exploration task (5 PID)
Barnes maze (5–7 PID)
Medial septal nucleus (MSN) DBS LFS: 7.7 H z (20, 80, 200 uA)
DBS HFS: 100 Hz (80uA) (DBS started 1 min before task performance and terminated before to return to the home cage)
⇑ objects exploration and improved latency and search strategy in the group DBS 7.7 Hz, 80 uA p<0.05
Neural stem cell transplantation
Gao et al (2006) Moderate LFPI Adult male rats MWM ref (11–15 PID) Ipsilateral hippocampus Fetal human neural stem cell (hNSCs) transplantation (1PID) Improved escape latency p<0.05
⇑GNDF secretion p<0.05
Shear et al (2004) CCI Young adult male mice MWM ref (30, 90, 360 PID) Ipsilateral striatum Neural progenitor cell (NPC) transplantation (7 PID) Improved escape latency p<0.05
Bakshi et al (2006) Severe LFPI (3.1atm) Adult male rats MWM ref (42–45 PID) Perilesional region GDNF-expressing C17.2 cells transplantation (1PID) ⇑ spatial learning p<0.05
Lu et al (2007) CCI Adult male rats MWM ref (31–35PID) Cavity lesion Human marrow stromal cells transplantation (4 PID) ⇑ spatial learning p<0.05
Xiong et al (2009) CCI Young adult male rats MWM ref (31–35 PID) Cavity lesion Human marrow stromal cells transplantation (7 PID) ⇑ spatial learning p<0.05
Dietary therapy
Cole et al (2010) Mild-moderate LFPI Adult male mice Fear conditioning (6–7PID) Restore hippocampal E/I balance ad libitum BCAA (100 mM sol): starting 2 PID for 7 d ⇑ freezing percentage p<0.05
Lim et al (2013) Mild-moderate LFPI Adult male mice Sleep/wakefulness cycle Re-activation of orexin neurons ad libitum BCAA (100 mM sol): starting 2 PID for the entire experiment Mitigation of injury-induced inability to maintain wakefulness p<0.05
Environmental enrichment (EE)
Hamm et al (1996) Moderate midline FPI Adult male rats MWM ref (11–15 PID) ⇑ neural plasticity EE: immediately after TBI for 15 days ⇑ spatial learning p<0.01
Passineau et al (2001) Severe LFPI Adult male rats MWM ref (11–15 PID) ⇑ neural plasticity EE: immediately after TBI for 15 days Improved escape latency NS
⇓learning acquisition than Sham NS
Hicks et al (2002) Moderate LFPI Adult male rats MWM ref (15 PID) ⇑ neural plasticity EE + handling: immediately after TBI for 15 days Improved escape latency NS
⇓learning acquisition in TBI+EE vs Sham NS
No changes in neurotrophin/receptor mRNA --
Kline et al (2007) CCI Adult male rats MWM ref (14–18 PID) ⇑ neural plasticity EE: immediately after TBI for 21 days ⇑ spatial learning p<0.01
Hoffman et al (2008) CCI Adult male rats MWM ref (14–18 PID) ⇑ neural plasticity Early EE: immediately after TBI for 7 d followed by 2 weeks STD
Delay EE: 1 week STD followed by 2 weeks EE
Continuous EE: immediately after TBI for 3 weeks
⇑ spatial learning in continuous and delay group p<0.01
Matter et al (2011) CCI Adult male rats MWM ref (14–18 PID) ⇑ neural plasticity Continuous EE: immediately after TBI for 3 weeks
Early EE: immediately after TBI for 1–2 weeks
Early and late EE: 1 week EE+1 week STD+1 week EE
Early and late STD: 1 week STD +1 week EE+1 week STD
⇑ spatial learning in continuous and delayed (starting after 7 PID) p<0.01
de Witt et al (2011) CCI Adult male rats MWM ref (14–18 PID) ⇑ neural plasticity Early and continuous EE: immediately after TBI for 3 weeks
Abbreviated EE: 2, 4, 6 h per day
⇑ spatial learning in continuous and abbreviated 6h group p<0.01
Cheng et al (2012) CCI Adult male rats MWM ⇑ neural plasticity EE: immediately after TBI for 3 weeks followed by STD
EE continuous: immediately after TBI for 6 months
⇑ spatial learning up to 6 months post-rehabilitation in both groups (EE+STD and EE continuous) p<0.01
Darwish et al (2014) CCI Adult male rats NOR (7 and 14d) [test 1′ and 15′ delay] ⇑ neural plasticity EE: 2 h per d, beginning 1 PID and continuing for 14 d Improved recognition memory at 7 PID but not at 14 PID with 1′ delay NS
No improvement at 7 and 14 PID with 15′ delay
Darwish et al (2014) CCI Adult male rats TOR (60′ delay) ⇑ neural plasticity EE: 2 h per d, beginning 1 PID and continuing for 14 d No improvement at 7 and 14 PID NS

Deep brain stimulation

Electrical stimulus therapy has been used successfully to treat motor dysfunction in Parkinson disease, however only few studies tested this treatment in TBI. Only in the last few years, has deep brain stimulation been tested in TBI animal models with promising results. The basis of deep brain stimulation therapy is to improve abnormal synchrony between different brain regions [124]. Specifically, it was found that stimulation given within the theta frequency band to the medial septal nucleus transiently increased theta activity in the hippocampus and led to improved spatial search pattern and decreased escape latency during Barnes Maze performance 5–7 PID after moderate LFPI [125]. Another deep brain stimulation study showed an increased exploration time when animals were exposed to new objects [126]. Furthermore, theta stimulation of the midbrain medial raphe and dorsal raphe showed a decreased learning peak during reference memory acquisition, and theta-burst stimulation of the fornix demonstrated improved working memory performance [40,127]. Clinically, there have been a few studies where stimulation electrodes have been successfully implanted in severe TBI patients chronically [128132].

Neural Stem Cell Transplantation

In the last decade, the ability to repair and regenerate the injured brain has been used as potential therapeutic target for TBI. Different types of cells have been used for neural transplantation in TBI animal models. Interestingly, embryonic stem cell transplantation successfully improved cognitive dysfunction [133135]. However, some limitations on this technique have been raised due to a limited neural long-term survival and increased tumor risk [136]. Few studies investigated adult neural stem cell implantation showing interesting results but none of them tested memory performance [137,138]. Instead, as alternative strategy with less side effects, bone marrow stromal cells have shown therapeutic promise. This technique resulted in improved cognitive dysfunction, a decreased brain lesion volume and enhanced focal brain angiogenesis [139,140]. Despite the encouraging results of the neuronal transplantation from the animal setting, the clinical translation is still far off. Some issues such as generating sufficient neurons able to integrate in the existing neural network, controlling the hostile environment due to the injury, need to be solved before treating human brain to obtain a successful outcome.

Dietary therapy

Due to altered excitatory/inhibitory (E/I) balance caused by TBI, our laboratory sought to develop a dietary therapy based on precursors of the excitatory neurotransmitter glutamate. The inhibitory neurotransmitter GABA is synthesized from glutamate [141]. Branched chain amino acids (BCAAs) are key amino acids involved in de novo glutamate synthesis [142]. We have found that dietary BCAA therapy restores limbic E/I balance and ameliorates hippocampal-dependent contextual fear memory impairment in a mild/moderate FPI mouse model [143]. Furthermore, we have demonstrated that BCAA therapy mitigates injury-induced inability to maintain wakefulness [144••]. Specifically, BCAA therapy was shown to restore brain EEG activity during wake and sleep cycles and increases hypothalamic orexin neuronal firing, which are important in mediating wakefulness. We are currently investigating the efficacy of our dietary therapy on altered episodic-like and working memory tasks and the electrical brainwave activity that sub-serves these functions.

Environmental enrichment

As a non-invasive therapeutic approach, environmental enrichment has been demonstrated to robustly attenuate TBI-induced memory impairments [145149]. Environmental enrichment is a rodent housing condition which combines complex motor, sensory, and social stimuli within a large living space [150]. The most beneficial effects of this treatment have been seen when rodents are introduced to an enriched environment immediately after TBI and housed continuously for the duration of testing [151,152]. As a continuous-exposure model may not translate effectively in a clinical setting, other studies have demonstrated that environmental enrichment can still have cognitive benefits when delayed after injury, and also in abbreviated daily time periods [153,154]. Future studies optimizing the temporal effects of environmental enrichment will readily facilitate its therapeutic efficacy in human patients.

Conclusion

In consideration of the information summarized in this review, it is our opinion that the success of future TBI clinical trials will depend on a preclinical approach that incorporates both memory behavior and its underlying neural circuitry. While the current state of the literature reflects overall deficits in certain types of memory, future behavioral studies should expand on how the components of memory—encoding, maintenance, and retrieval—are affected after TBI. Specifically, behavioral assessment of these memory components will identify where TBI disrupts memory function.

To better understand alterations to memory components observed behaviorally, a circuit-level physiological approach can be of major benefit. By examining cellular and synaptic changes in the hippocampus and cortex, we can understand how behavioral memory deficits occur. New technologies such as optogentics or chemogenetics, could be utilized to substantiate the involvement of physiological mechanisms in behavioral outcome. In conclusion, a combination of behavior and circuit physiology in preclinical studies will aide in the discovery of specific therapeutic targets for clinical translation, and lead to meaningful recovery of memory function in TBI patients.

References

Papers of particular interest, published recently, have been highlighted as:

• Of importance

•• Of major importance

  • 1.Centers for Disease Control and Prevention. Report to Congress on Traumatic Brain Injury in the United States: Epidemiology and Rehabilitation; Atlanta, GA. 2015. [DOI] [PubMed] [Google Scholar]
  • 2.Nicholl J, LaFrance WC. Neuropsychiatric sequelae of traumatic brain injury. Semin Neurol. 2009:247–55. doi: 10.1055/s-0029-1223878. [DOI] [PubMed] [Google Scholar]
  • 3.Pierce JES, Smith DH, Trojanowski JQ, McIntosh TK. Enduring cognitive, neurobehavioral and histopathological changes persist for up to one year following severe experimental brain injury in rats. Neuroscience. 1998;87:359–69. doi: 10.1016/s0306-4522(98)00142-0. [DOI] [PubMed] [Google Scholar]
  • 4.Rutland-Brown W, Langlois JA, Thomas KE, Xi YL. Incidence of traumatic brain injury in the United States, 2003. J Head Trauma Rehabil. 2006;21:544–8. doi: 10.1097/00001199-200611000-00009. [DOI] [PubMed] [Google Scholar]
  • 5.Adams JH, Doyle D, Ford I, Gennarelli TA, Graham DI, McLellan DR. Diffuse axonal injury in head injury: definition, diagnosis and grading. Histopathology. 1989;15:49–59. doi: 10.1111/j.1365-2559.1989.tb03040.x. [DOI] [PubMed] [Google Scholar]
  • 6.Graham DI, Adams JH, Nicoll JAR, Maxwell WL, Gennarelli TA. The Nature, Distribution and Causes of Traumatic Brain Injury. Brain Pathol. 1995;5:397–406. doi: 10.1111/j.1750-3639.1995.tb00618.x. [DOI] [PubMed] [Google Scholar]
  • 7.McAllister TW, Sparling MB, Flashman La, Guerin SJ, Mamourian aC, Saykin aJ. Differential working memory load effects after mild traumatic brain injury. Neuroimage. 2001;14:1004–12. doi: 10.1006/nimg.2001.0899. [DOI] [PubMed] [Google Scholar]
  • 8.Marmarou CR, Prieto R, Taya K, Young HF, Marmarou A. Marmarou weight drop injury model. Anim Model Acute Neurol Inj. 2009:393–407. [Google Scholar]
  • 9.Feeney DM, Boyeson MG, Linn RT, Murray HM, Dail WG. Responses to cortical injury: I. Methodology and local effects of contusions in the rat. Brain Res. 1981;211:67–77. doi: 10.1016/0006-8993(81)90067-6. [DOI] [PubMed] [Google Scholar]
  • 10.Flierl Ma, Stahel PF, Beauchamp KM, Morgan SJ, Smith WR, Shohami E. Mouse closed head injury model induced by a weight-drop device. Nat Protoc. 2009;4:1328–37. doi: 10.1038/nprot.2009.148. [DOI] [PubMed] [Google Scholar]
  • 11.Shapira Y, Setton D, Artru AA, Shohami E. Blood-brain barrier permeability, cerebral edema, and neurologic function after closed head injury in rats. Anesth Analg. 1993;77:141–8. doi: 10.1213/00000539-199307000-00028. [DOI] [PubMed] [Google Scholar]
  • 12.Panzer MB, Matthews KA, Yu AW, Morrison B, Meaney DF, Bass CR. A multiscale approach to blast neurotrauma modeling: Part I - development of novel test devices for in vivo and in vitro blast injury models. Front Neurol. 2012 Mar; doi: 10.3389/fneur.2012.00046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Beamer M, Tummala SR, Gullotti D, Kopil K, Gorka S, Abel Ted, et al. Primary blast injury causes cognitive impairments and hippocampal circuit alterations. Exp Neurol. 2016;283:16–28. doi: 10.1016/j.expneurol.2016.05.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Dixon CE, Lyeth BG, Povlishock JT, Findling RL, Hamm RJ, Marmarou a, et al. A fluid percussion model of experimental brain injury in the rat. J Neurosurg. 1987;67:110–9. doi: 10.3171/jns.1987.67.1.0110. [DOI] [PubMed] [Google Scholar]
  • 15.Edward Dixon C, Clifton GL, Lighthall JW, Yaghmai AA, Hayes RL. A controlled cortical impact model of traumatic brain injury in the rat. J Neurosci Methods. 1991;39:253–62. doi: 10.1016/0165-0270(91)90104-8. [DOI] [PubMed] [Google Scholar]
  • 16.Cortez SC, McIntosh TK, Noble LJ. Experimental fluid percussion brain injury: vascular disruption and neuronal and glial alterations. Brain Res. 1989;482:271–82. doi: 10.1016/0006-8993(89)91190-6. [DOI] [PubMed] [Google Scholar]
  • 17.McIntosh TK, Vink R, Noble L, Yamakami I, Fernyak S, Soares H, et al. Traumatic brain injury in the rat: Characterization of a lateral fluid-percussion model. Neuroscience. 1989;28:233–44. doi: 10.1016/0306-4522(89)90247-9. [DOI] [PubMed] [Google Scholar]
  • 18.Johnson VE, Meaney DF, cullen DK, Smith DH. Animal models of traumatic brain injury. Handb Clin Neurol. 2015;127:115–28. doi: 10.1016/B978-0-444-52892-6.00008-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19•.Xiong Y, Mahmood A, Chopp M. Animal models of traumatic brain injury. Nat Rev Neurosci. 2013;14:128–42. doi: 10.1038/nrn3407. Xiong et al. extensively reviews the current animal models of traumatic brain injury. It is particularly useful to gain insight into the various aspects of experimental models used to study traumatic brain injury. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Namjoshi DR, Good C, Cheng WH, Panenka W, Richards D, Cripton Pa, et al. Towards clinical management of traumatic brain injury: a review of models and mechanisms from a biomechanical perspective. Dis Model Mech. 2013;0:1–14. doi: 10.1242/dmm.011320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Tulving E. How many memory systems are there? Am. Psychol. 1985;40:385–98. [Google Scholar]
  • 22.Josselyn SA, Köhler S, Frankland PW. Finding the engram. Nat Rev Neurosci. 2015;16:521–34. doi: 10.1038/nrn4000. [DOI] [PubMed] [Google Scholar]
  • 23.Tonegawa S, Pignatelli M, Roy DS, Ryan TJ. Memory engram storage and retrieval. Curr Opin Neurobiol. 2015:101–9. doi: 10.1016/j.conb.2015.07.009. [DOI] [PubMed] [Google Scholar]
  • 24.Cowan N. What are the differences between long-term, short-term, and working memory? Nelson. NIH Public Access. 2009;6123:323–38. doi: 10.1016/S0079-6123(07)00020-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Diamond A. Executive functions. Annu Rev Psychol. 2013;64:135–168. doi: 10.1146/annurev-psych-113011-143750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Tulving E, Markowitsch HJ. Episodic and declarative memory: role of the hippocampus. Hippocampus. 1998;8:198–204. doi: 10.1002/(SICI)1098-1063(1998)8:3<198::AID-HIPO2>3.0.CO;2-G. [DOI] [PubMed] [Google Scholar]
  • 27.Tulving E. Episodic and semantic memory. Organ Mem. 1972:381–403. [Google Scholar]
  • 28.Pause BM, Zlomuzica A, Kinugawa K, Mariani J, Pietrowsky R, Dere E. Perspectives on episodic-like and episodic memory. Front Behav Neurosci. 2013;7:33. doi: 10.3389/fnbeh.2013.00033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Alvarez P, Squire LR. Memory consolidation and the medial temporal lobe: A simple network model. Proc Natl Acad Sci USA. 1994;91:7041–5. doi: 10.1073/pnas.91.15.7041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Olton DS, Papas BC. Spatial memory and hippocampal function. Neuropsychologia. 1979;17:669–82. doi: 10.1016/0028-3932(79)90042-3. [DOI] [PubMed] [Google Scholar]
  • 31.Olton DS, Collison C, Werz MA. Spatial memory and radial arm maze performance of rats. Learn Motiv. 1977;8:289–314. [Google Scholar]
  • 32.Vorhees CV, Williams MT. Morris water maze: procedures for assessing spatial and related forms of learning and memory. Nat Protoc. 2006;1:848–58. doi: 10.1038/nprot.2006.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Hoskison MM, Moore AN, Hu B, Orsi S, Kobori N, Dash PK. Persistent working memory dysfunction following traumatic brain injury: Evidence for a time-dependent mechanism. Neuroscience IBRO. 2009;159:483–91. doi: 10.1016/j.neuroscience.2008.12.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Kobori N, Dash PK. Reversal of Brain Injury-Induced Prefrontal Glutamic Acid Decarboxylase Expression and Working Memory Deficits by D 1 Receptor Antagonism. J Neurosci. 2006;26:4236–46. doi: 10.1523/JNEUROSCI.4687-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Sebastian V, Diallo A, Ling DSF, Serrano PA. Robust training attenuates TBI-induced deficits in reference and working memory on the radial 8-arm maze. Front Behav Neurosci. 2013;7:38. doi: 10.3389/fnbeh.2013.00038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36•.Smith CJ, Xiong G, Elkind JA, Putnam B, Cohen AS. Brain injury impairs working memory and prefrontal circuit function. Front Neurol. 2015;6:1–13. doi: 10.3389/fneur.2015.00240. Smith et al demonstrate a working memory impairment caused by TBI and associated with circuit alteration in the medial prefrontal cortex. This is a good example of relating memory deficits to the underlying pathophysiology in the working memory circuit. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Lyeth BG, Jenkins LW, Hamm RJ, Dixon CE, Phillips LL, Clifton GL, et al. Prolonged Memory Impairment in the Absence of Hippocampal Cell-Death Following Traumatic Brain Injury in the Rat. Brain Res. 1990;526:249–58. doi: 10.1016/0006-8993(90)91229-a. [DOI] [PubMed] [Google Scholar]
  • 38.Whiting MD, Hamm RJ. Traumatic brain injury produces delay-dependent memory impairment in rats. J Neurotrauma. 2006;23:1529–34. doi: 10.1089/neu.2006.23.1529. [DOI] [PubMed] [Google Scholar]
  • 39.Eakin K, Miller JP. J Neurotrauma. Vol. 29. Atlanta, GA: Elsevier Inc; 2012. Mild traumatic brain injury is associated with impaired hippocampal spatiotemporal representation in the absence of histological changes; pp. 1180–7. [DOI] [PubMed] [Google Scholar]
  • 40.Sweet JA, Eakin KC, Munyon CN, Miller JP. Improved learning and memory with theta-burst stimulation of the fornix in rat model of traumatic brain injury. Hippocampus. 2014;24:1592–600. doi: 10.1002/hipo.22338. [DOI] [PubMed] [Google Scholar]
  • 41.Brandeis R, Brandys Y, Yehuda S. The use of the Morris Water Maze in the study of memory and learning. Int J Neurosci. 1989;48:29–69. doi: 10.3109/00207458909002151. [DOI] [PubMed] [Google Scholar]
  • 42.Morris R. Developments of a water-maze procedure for studying spatial learning in the rat. J Neurosci Methods. 1984;11:47–60. doi: 10.1016/0165-0270(84)90007-4. [DOI] [PubMed] [Google Scholar]
  • 43.Barnes CA. Memory deficits associated with senescence: a neurophysiological and behavioral study in the rat. J Comp Physiol Psychol. 1979;93:74–104. doi: 10.1037/h0077579. [DOI] [PubMed] [Google Scholar]
  • 44.Olton DS. The radial arm maze as a tool in behavioral pharmacology. Physiol Behav. 1987;40:793–7. doi: 10.1016/0031-9384(87)90286-1. [DOI] [PubMed] [Google Scholar]
  • 45.Fox GB, Fan L, LeVasseur Ra, Faden aI. Effect of traumatic brain injury on mouse spatial and nonspatial learning in the Barnes circular maze. J Neurotrauma. 1998;15:1037–46. doi: 10.1089/neu.1998.15.1037. [DOI] [PubMed] [Google Scholar]
  • 46.Dawish H, Mahmood A, Schallert T, Chopp M, Therrien B. Mild traumatic brain injury (MTBI) leads to spatial learning deficits. Brain Inj. 2012;26:151–65. doi: 10.3109/02699052.2011.635362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Hamm RJ. Cognitive deficits following traumatic brain injury produced by controlled cortical impact. J Neurotrauma. 1992;9:11–20. doi: 10.1089/neu.1992.9.11. [DOI] [PubMed] [Google Scholar]
  • 48.Kim D-K, Nishida H, An SY, Shetty AK, Bartosh TJ, Prockop DJ. Chromatographically isolated CD63 + CD81 + extracellular vesicles from mesenchymal stromal cells rescue cognitive impairments after TBI. Proc Natl Acad Sci. 2016;113:170–5. doi: 10.1073/pnas.1522297113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Dash PK, Moore AN, Dixon CE. Spatial Memory Deficits, Increased Phosphorylation of the Transcription Factor Creb, and Induction of the Ap-1 Complex Following Experimental Brain Injury. J Neurosci. 1995;15:2030–9. doi: 10.1523/JNEUROSCI.15-03-02030.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Zohar O, Rubovitch V, Milman A, Schreiber S, Pick CG. Behavioral consequences of minimal traumatic brain injury in mice. Acta Neurobiol Exp (Wars) 2011;71:36–45. doi: 10.55782/ane-2011-1821. [DOI] [PubMed] [Google Scholar]
  • 51.Hylin MJ, Orsi Sa, Zhao J, Bockhorst K, Perez A, Moore AN, et al. Behavioral and histopathological alterations resulting from mild fluid percussion injury. J Neurotrauma. 2013;30:702–15. doi: 10.1089/neu.2012.2630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Pierce JES, Smith DH, Eison MS, McIntosh TK. The nootropic compound BMY-21502 improves spatial learning ability in brain injured rats. Brain Res. 1993;624:199–208. doi: 10.1016/0006-8993(93)90078-2. [DOI] [PubMed] [Google Scholar]
  • 53.Lee DJ, Gurkoff GG, Izadi A, Berman RF, Ekstrom AD, Muizelaar JP, et al. Medial septal nucleus theta frequency deep brain stimulation improves spatial working memory after traumatic brain injury. J Neurotrauma. 2013;30:131–9. doi: 10.1089/neu.2012.2646. [DOI] [PubMed] [Google Scholar]
  • 54.Fedor M, Berman RF, Muizelaar JP, Lyeth BG. Hippocampal θ dysfunction after lateral fluid percussion injury. J Neurotrauma. 2010;27:1605–15. doi: 10.1089/neu.2010.1370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Whiting MD, Hamm RJ. Mechanisms of anterograde and retrograde memory impairment following experimental traumatic brain injury. Brain Res. 2008;1213:69–77. doi: 10.1016/j.brainres.2008.01.107. [DOI] [PubMed] [Google Scholar]
  • 56.Smith DH, Lowenstein DH, Gennarelli TA, Mcintosh K. NEUROSCI [ NC [ lETTERS. 1994;168:151–4. doi: 10.1016/0304-3940(94)90438-3. [DOI] [PubMed] [Google Scholar]
  • 57.Hicks RR, Smith DH, Lowenstein DH, Saint Marie R, McIntosh TK. Mild experimental brain injury in the rat induces cognitive deficits associated with regional neuronal loss in the hippocampus. J Neurotrauma. 1993;10:405–14. doi: 10.1089/neu.1993.10.405. [DOI] [PubMed] [Google Scholar]
  • 58.Okiyama K, Smith DH, Thomas MJ, McIntosh TK. Evaluation of a Novel Calcium-Channel Blocker, (S)-Emopamil, on Regional Cerebral Edema and Neurobehavioral Function After Experimental Brain Injury. J Neurosurg. 1992;77:607–15. doi: 10.3171/jns.1992.77.4.0607. [DOI] [PubMed] [Google Scholar]
  • 59.Smith DH, Okiyama K, Thomas MJ, Claussen B, McIntosh TK. Evaluation of memory dysfunction following experimental brain injury using the Morris water maze. J Neurotrauma. 1991;8:259–69. doi: 10.1089/neu.1991.8.259. [DOI] [PubMed] [Google Scholar]
  • 60•.Gurkoff GG, Gahan JD, Ghiasvand RT, Hunsaker MR, Van K, Feng J-F, et al. Evaluation of metric, topological, and temporal ordering memory tasks after lateral fluid percussion injury. J Neurotrauma. 2013;30:292–300. doi: 10.1089/neu.2012.2463. Gurkoff et al. tested various tasks to examine how TBI alters recognition memory. The found that the temporal aspect of episodic memory is impaired. [DOI] [PubMed] [Google Scholar]
  • 61.Zhang Y, Chopp M, Meng Y, Zhang ZG. Cerebrolysin improves cognitive performance in rats after mild traumatic brain injury. 2015;122:843–55. doi: 10.3171/2014.11.JNS14271. [DOI] [PubMed] [Google Scholar]
  • 62.Tulving E. Episodic Memory: From Mind to Brain. Annu Rev Psychol. 2002;53:1–25. doi: 10.1146/annurev.psych.53.100901.135114. [DOI] [PubMed] [Google Scholar]
  • 63.Serruya MD, Sederberg PB, Kahana MJ. Power shifts track serial position and modulate encoding in human episodic memory. Cereb Cortex. 2014;24:403–13. doi: 10.1093/cercor/bhs318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Oliveira AMM, Hawk JD, Abel T, Havekes R. Post-training reversible inactivation of the hippocampus enhances novel object recognition memory. Learn Mem. 2010;17:155–60. doi: 10.1101/lm.1625310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Carbonell WS, Grady MS. Regional and temporal characterization of neuronal, glial, and axonal response after traumatic brain injury in the mouse. Acta Neuropathol. 1999;98:396–406. doi: 10.1007/s004010051100. [DOI] [PubMed] [Google Scholar]
  • 66.Amaral DG, Rempel-Clower NL, Zola SM, Squire LR. Three cases of enduring memory impairment after bilateral damage limited to the hippocampal formation. J Neurosci. 1996;16:5233–55. doi: 10.1523/JNEUROSCI.16-16-05233.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Reed JM, Squire LR. Impaired recognition memory in patients with lesions limited to the hippocampal formation. Behav Neurosci. 1997;111:667–75. doi: 10.1037//0735-7044.111.4.667. [DOI] [PubMed] [Google Scholar]
  • 68.Cave CB, Squire LR. Equivalent impairment of spatial and nonspatial memory following damage to the human hippocampus. Hippocampus. 1991;1:329–40. doi: 10.1002/hipo.450010323. [DOI] [PubMed] [Google Scholar]
  • 69.Bohbot VD, Allen JJb, Nadel L. Memory Deficits Characterized by Patterns of Lesions to the Hippocampus and Parahippocampal Cortex. Ann N Y Acad Sci. 2000;911:355–68. doi: 10.1111/j.1749-6632.2000.tb06737.x. [DOI] [PubMed] [Google Scholar]
  • 70.Miller La, Lai R, Munoz DG. Contributions of the entorhinal cortex, amygdala and hippocampus to human memory. Neuropsychologia. 1998;36:1247–56. doi: 10.1016/s0028-3932(97)00139-5. [DOI] [PubMed] [Google Scholar]
  • 71.Deweer B, Pillon B, Pochon JB, Dubois B. Is the HM story only a “remote memory”? Some facts about hippocampus and memory in humans. Behav Brain Res. 2001:209–24. doi: 10.1016/s0166-4328(01)00366-7. [DOI] [PubMed] [Google Scholar]
  • 72.Marr D. Simple Memory: A Theory for Archicortex. Source Philos Trans R Soc London Ser B, Biol Sci. 1971;262:23–81. doi: 10.1098/rstb.1971.0078. [DOI] [PubMed] [Google Scholar]
  • 73.McNaughton BL, Morris RGM. Hippocampal synaptic enhancement and information storage within a distributed memory system. Trends Neurosci. 1987:408–15. [Google Scholar]
  • 74.O’Reilly RC, McClelland JL. Hippocampal conjunctive encoding, storage, and recall: Avoiding a trade-off. Hippocampus. 1994;4:661–82. doi: 10.1002/hipo.450040605. [DOI] [PubMed] [Google Scholar]
  • 75.Treves A, Rolls ET. Computational analysis of the role of the hippocampus in memory. Hippocampus. 1994;4:374–91. doi: 10.1002/hipo.450040319. [DOI] [PubMed] [Google Scholar]
  • 76.Leutgeb JK, Leutgeb S, Moser M-B, Moser EI. Pattern Separation in the Dentate Gyrus and CA3 of the Hippocampus. Science (80-) 2007;315:961–6. doi: 10.1126/science.1135801. [DOI] [PubMed] [Google Scholar]
  • 77.Bakker A, Kirwan CB, Miller M, Stark CEL. Pattern separation in the human hippocampal CA3 and dentate gyrus. Science. 2008;319:1640–2. doi: 10.1126/science.1152882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Hsu D. The dentate gyrus as a filter or gate: a look back and a look ahead. Prog Brain Res. 2007:601–13. doi: 10.1016/S0079-6123(07)63032-5. [DOI] [PubMed] [Google Scholar]
  • 79.Buzsáki G, Lai-Wo SL, Vanderwolf CH. Cellular bases of hippocampal EEG in the behaving rat. Brain Res Rev. 1983:139–71. doi: 10.1016/0165-0173(83)90037-1. [DOI] [PubMed] [Google Scholar]
  • 80.Halasy K, Somogyi P. Subdivisions in the multiple GABAergic innervation of granule cells in the dentate gyrus of the rat hippocampus. Eur J Neurosci. 1993;5:411–29. doi: 10.1111/j.1460-9568.1993.tb00508.x. [DOI] [PubMed] [Google Scholar]
  • 81.Han Z-S, Buhl EH, Lorinczi Z, Somogyi P. A High Degree of Spatial Selectivity in the Axonal and Dendritic Domains of Physiologically Identified Local-circuit Neurons in the Dentate Gyrus of the Rat Hippocampus. Eur J Neurosci. 1993;5:395–410. doi: 10.1111/j.1460-9568.1993.tb00507.x. [DOI] [PubMed] [Google Scholar]
  • 82.Soltesz I, Smetters DK, Mody I. Tonic Inhibition Originates from Synapses Close to the Soma. Neuron. 1995;14:1273–83. doi: 10.1016/0896-6273(95)90274-0. [DOI] [PubMed] [Google Scholar]
  • 83.Hollrigel GS, Toth K, Soltesz I. Neuroprotection by propofol in acute mechanical injury: role of GABAergic inhibition. J Neurophysiol. 1996;76:2412–22. doi: 10.1152/jn.1996.76.4.2412. [DOI] [PubMed] [Google Scholar]
  • 84.Witgen BM, Lifshitz J, Smith ML, Schwarzbach E, Liang SL, Grady MS, et al. Regional hippocampal alteration associated with cognitive deficit following experimental brain injury: A systems, network and cellular evaluation. Neuroscience. 2005;133:1–15. doi: 10.1016/j.neuroscience.2005.01.052. [DOI] [PubMed] [Google Scholar]
  • 85.Toth Z, Hollrigel GS, Gorcs T, Soltesz I. Instantaneous perturbation of dentate interneuronal networks by a pressure wave-transient delivered to the neocortex. J Neurosci. 1997;17:8106–17. doi: 10.1523/JNEUROSCI.17-21-08106.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Lowenstein DH, Thomas MJ, Smith DH, McIntosh TK. Selective vulnerability of dentate hilar neurons following traumatic brain injury: a potential mechanistic link between head trauma and disorders of the hippocampus. J Neurosci. 1992;12:4846–53. doi: 10.1523/JNEUROSCI.12-12-04846.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Santhakumar V, Bender R, Frotscher M, Ross ST, Hollrigel GS, Toth Z, et al. Granule cell hyperexcitability in the early post-traumatic rat dentate gyrus: the “irritable mossy cell” hypothesis. J Physiol. 2000;524(Pt 1):117–34. doi: 10.1111/j.1469-7793.2000.00117.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Santhakumar V, Ratzliff ADH, Jeng J, Toth Z, Soltesz I. Long-term hyperexcitability in the hippocampus after experimental head trauma. Ann Neurol. 2001;50:708–17. doi: 10.1002/ana.1230. [DOI] [PubMed] [Google Scholar]
  • 89.Hunt RF, Scheff SW, Smith BN. Synaptic reorganization of inhibitory hilar interneuron circuitry after traumatic brain injury in mice. J Neurosci. 2011;31:6880–90. doi: 10.1523/JNEUROSCI.0032-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Bonislawski DP, Schwarzbach EP, Cohen AS. Brain injury impairs dentate gyrus inhibitory efficacy. Neurobiol Dis. 2007;25:163–9. doi: 10.1016/j.nbd.2006.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Mtchedlishvili Z, Lepsveridze E, Xu H, Kharlamov EA, Lu B, Kelly KM. Increase of GABAA receptor-mediated tonic inhibition in dentate granule cells after traumatic brain injury. Neurobiol Dis Elsevier Inc. 2010;38:464–75. doi: 10.1016/j.nbd.2010.03.012. [DOI] [PubMed] [Google Scholar]
  • 92.Boychuk JA, Butler CR, Halmos KC, Smith BN. Enduring changes in tonic GABAA receptor signaling in dentate granule cells after controlled cortical impact brain injury in mice. Exp Neurol Elsevier BV. 2016;277:178–89. doi: 10.1016/j.expneurol.2016.01.005. [DOI] [PubMed] [Google Scholar]
  • 93.Pavlov I, Huusko N, Drexel M, Kirchmair E, Sperk G, Pitkänen A, et al. Neuroscience. Vol. 194. Elsevier Inc; 2011. Progressive loss of phasic, but not tonic, GABAA receptor-mediated inhibition in dentate granule cells in a model of post-traumatic epilepsy in rats; pp. 208–19. [DOI] [PubMed] [Google Scholar]
  • 94.Langlois JA, Rutland-Brown W, Wald MM, Li N, Yang Y, Glover DP, et al. J Neurotrauma. Vol. 32. Atlanta, GA: Elsevier Inc; 2015. Decrease in Tonic Inhibition Contributes to Increase in Dentate Semilunar Granule Cell Excitability after Brain Injury; pp. 1–13. [Google Scholar]
  • 95.Gupta A, Elgammal FS, Proddutur A, Shah S, Santhakumar V. Decrease in Tonic Inhibition Contributes to Increase in Dentate Semilunar Granule Cell Excitability after Brain Injury. J Neurosci. 2012;32:2523–37. doi: 10.1523/JNEUROSCI.4141-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Howard AL, Neu A, Morgan RJ, Echegoyen JC, Soltesz I. Opposing modifications in intrinsic currents and synaptic inputs in post-traumatic mossy cells: evidence for single-cell homeostasis in a hyperexcitable network. J Neurophysiol. 2007;97:2394–409. doi: 10.1152/jn.00509.2006. [DOI] [PubMed] [Google Scholar]
  • 97.Treves A, Rolls ET. Computational analysis of the role of the hippocampus in memory. Hippocampus. 1994;4:374–91. doi: 10.1002/hipo.450040319. [DOI] [PubMed] [Google Scholar]
  • 98.Hasselmo ME, Schnell E, Barkai E. Dynamics of learning and recall at excitatory recurrent synapses and cholinergic modulation in rat hippocampal region CA3. J Neurosci. 1995;15:5249–62. doi: 10.1523/JNEUROSCI.15-07-05249.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Chen Y, Constantini S, Trembovler V, Weinstock M, Shohami E. An experimental model of closed head injury in mice: pathophysiology, histopathology, and cognitive deficits. J Neurotrauma. 1996;13:557–68. doi: 10.1089/neu.1996.13.557. [DOI] [PubMed] [Google Scholar]
  • 100.Tang YP, Noda Y, Hasegawa T, Nabeshima T. A concussive-like brain injury model in mice (I): impairment in learning and memory. J Neurotrauma. 1997;14:851–62. doi: 10.1089/neu.1997.14.851. [DOI] [PubMed] [Google Scholar]
  • 101.Golarai G, Greenwood aC, Feeney DM, Connor Ja. Physiological and structural evidence for hippocampal involvement in persistent seizure susceptibility after traumatic brain injury. J Neurosci. 2001;21:8523–37. doi: 10.1523/JNEUROSCI.21-21-08523.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Baldwin SA, Gibson T, Callihan CT, Sullivan PG, Palmer E, Scheff SW. Neuronal cell loss in the CA3 subfield of the hippocampus following cortical contusion utilizing the optical disector method for cell counting. J Neurotrauma. 1997;14:385–98. doi: 10.1089/neu.1997.14.385. [DOI] [PubMed] [Google Scholar]
  • 103.Griesemer D, Mautes AM. Closed head injury causes hyperexcitability in rat hippocampal CA1 but not in CA3 pyramidal cells. J Neurotrauma. 2007;24:1823–32. doi: 10.1089/neu.2006.0237. [DOI] [PubMed] [Google Scholar]
  • 104.Ting JT, Daigle TL, Chen Q, Feng G. Acute brain slice methods for adult and aging animals: Application of targeted patch clamp analysis and optogenetics. Methods Mol Biol. 2014;1183:221–42. doi: 10.1007/978-1-4939-1096-0_14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Lee I, Kesner RP. Differential contributions of dorsal hippocampal subregions to memory acquisition and retrieval in contextual fear-conditioning. Hippocampus. 2004;14:301–10. doi: 10.1002/hipo.10177. [DOI] [PubMed] [Google Scholar]
  • 106.Lee I, Rao G, Knierim JJ. A double dissociation between hippocampal subfields: Differential time course of CA3 and CA1 place cells for processing changed environments. Neuron. 2004;42:803–15. doi: 10.1016/j.neuron.2004.05.010. [DOI] [PubMed] [Google Scholar]
  • 107.Leutgeb S, Leutgeb JK, Barnes CA, Moser EI, McNaughton BL, Moser M-B. Independent codes for spatial and episodic memory in hippocampal neuronal ensembles. Science. 2005;309:619–23. doi: 10.1126/science.1114037. [DOI] [PubMed] [Google Scholar]
  • 108.Leutgeb S, Leutgeb JK, Moser MB, Moser EI. Place cells, spatial maps and the population code for memory. Curr Opin Neurobiol. 2005:738–46. doi: 10.1016/j.conb.2005.10.002. [DOI] [PubMed] [Google Scholar]
  • 109.D’Ambrosio R, Maris DO, Grady MS, Winn HR, Janigro D. Selective loss of hippocampal long-term potentiation, but not depression, following fluid percussion injury. Brain Res. 1998;786:64–79. doi: 10.1016/s0006-8993(97)01412-1. [DOI] [PubMed] [Google Scholar]
  • 110.Norris CM, Scheff SW. Recovery of afferent function and synaptic strength in hippocampal CA1 following traumatic brain injury. J Neurotrauma. 2009;26:2269–78. doi: 10.1089/neu.2009.1029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Schwarzbach E, Bonislawski DP, Xiong G, Cohen AS. Mechanisms underlying the inability to induce area CA1 LTP in the mouse after traumatic brain injury. Hippocampus. 2006;16:541–50. doi: 10.1002/hipo.20183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Johnson BN, Palmer CP, Bourgeois EB, Elkind JA, Putnam BJ, Cohen AS. Augmented Inhibition from Cannabinoid-Sensitive Interneurons Diminishes CA1 Output after Traumatic Brain Injury. Front Cell Neurosci. 2014;8:435. doi: 10.3389/fncel.2014.00435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Reeves TM, Lyeth BG, Phillips LL, Hamm RJ, Povlishock JT. The effects of traumatic brain injury on inhibition in the hippocampus and dentate gyrus. Brain Res. 1997;757:119–32. doi: 10.1016/s0006-8993(97)00170-4. [DOI] [PubMed] [Google Scholar]
  • 114.Reeves TM, Kao CQ, Phillips LL, Bullock MR, Povlishock JT. Presynaptic excitability changes following traumatic brain injury in the rat. J Neurosci Res. 2000;60:370–9. doi: 10.1002/(SICI)1097-4547(20000501)60:3<370::AID-JNR12>3.0.CO;2-B. [DOI] [PubMed] [Google Scholar]
  • 115.Almeida-Suhett CP, Prager EM, Pidoplichko V, Figueiredo TH, Marini AM, Li Z, et al. GABAergic interneuronal loss and reduced inhibitory synaptic transmission in the hippocampal CA1 region after mild traumatic brain injury. Exp Neurol Elsevier BV. 2015;273:11–23. doi: 10.1016/j.expneurol.2015.07.028. [DOI] [PubMed] [Google Scholar]
  • 116.Eakin K, Miller JP. Mild traumatic brain injury is associated with impaired hippocampal spatiotemporal representation in the absence of histological changes. J Neurotrauma. 2012;29:1180–7. doi: 10.1089/neu.2011.2192. [DOI] [PubMed] [Google Scholar]
  • 117.Paterno R, Metheny H, Xiong G, Elkind J, Cohen AS. Mild Traumatic Brain Injury Decreases Broadband Power in Area CA1. J Neurotrauma. 2016;5 doi: 10.1089/neu.2015.4107. neu.2015.4107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Yang L, Benardo LS, Valsamis H, Ling DSF. Acute injury to superficial cortex leads to a decrease in synaptic inhibition and increase in excitation in neocortical layer V pyramidal cells. J Neurophysiol. 2007;97:178–87. doi: 10.1152/jn.01374.2005. [DOI] [PubMed] [Google Scholar]
  • 119.Greer JE, Povlishock JT, Jacobs KM. Electrophysiological Abnormalities in Both Axotomized and Nonaxotomized Pyramidal Neurons following Mild Traumatic Brain Injury. J Neurosci. 2012;32:6682–7. doi: 10.1523/JNEUROSCI.0881-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Goforth PB, Ellis EF, Satin LS. Enhancement of AMPA-mediated current after traumatic injury in cortical neurons. J Neurosci. 1999;19:7367–74. doi: 10.1523/JNEUROSCI.19-17-07367.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Goforth PB, Ren J, Schwartz BS, Satin LS, Ashman T, Gordon W, et al. Excitatory synaptic transmission and network activity are depressed following mechanical injury in cortical neurons. J Neurophysiol. 2011;105:2350–63. doi: 10.1152/jn.00467.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Kao C-Q, Goforth PB, Ellis EF, Satin LS. Potentiation of GABA(A) currents after mechanical injury of cortical neurons. J Neurotrauma. 2004;21:259–70. doi: 10.1089/089771504322972059. [DOI] [PubMed] [Google Scholar]
  • 123.Chakraborty S, Skolnick B, Narayan RK. Neuroprotection Trials in Traumatic Brain Injury. Curr Neurol Neurosci Rep. 2016:29. doi: 10.1007/s11910-016-0625-x. [DOI] [PubMed] [Google Scholar]
  • 124.de Hemptinne C, Swann NC, Ostrem JL, Ryapolova-Webb ES, San Luciano M, Galifianakis NB, et al. Therapeutic deep brain stimulation reduces cortical phase-amplitude coupling in Parkinson’s disease. Nat Neurosci. 2015;18:779–86. doi: 10.1038/nn.3997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Lee DJ, Gurkoff GG, Izadi A, Berman RF, Ekstrom AD, Muizelaar JP, et al. Medial septal nucleus theta frequency deep brain stimulation improves spatial working memory after traumatic brain injury. J Neurotrauma. 2013;30:131–9. doi: 10.1089/neu.2012.2646. [DOI] [PubMed] [Google Scholar]
  • 126.Lee DJ, Gurkoff GG, Izadi A, Seidl SE, Echeverri A, Melnik M, et al. Septohippocampal Neuromodulation Improves Cognition after Traumatic Brain Injury. J Neurotrauma. 2015;11:150902125930001. doi: 10.1089/neu.2014.3744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Carballosa Gonzalez MM, Blaya MO, Alonso OF, Bramlett HM, Hentall ID. Midbrain raphe stimulation improves behavioral and anatomical recovery from fluid-percussion brain injury. J Neurotrauma. 2013;30:119–30. doi: 10.1089/neu.2012.2499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Kang E-K, Kim D-Y, Paik N-J. Transcranial direct current stimulation of the left prefrontal cortex improves attention in patients with traumatic brain injury: a pilot study. J Rehabil Med. 2012;44:346–50. doi: 10.2340/16501977-0947. [DOI] [PubMed] [Google Scholar]
  • 129.Rezai AR, Sederberg PB, Bogner J, Nielson DM, Zhang J, Mysiw WJ, et al. Improved function after deep brain stimulation for chronic, severe traumatic brain injury. Neurosurgery. 2016;79:204–10. doi: 10.1227/NEU.0000000000001190. [DOI] [PubMed] [Google Scholar]
  • 130.Schiff ND, Giacino JT, Kalmar K, Victor JD, Baker K, Gerber M, et al. Behavioural improvements with thalamic stimulation after severe traumatic brain injury. Nature. 2007;448:600–3. doi: 10.1038/nature06041. [DOI] [PubMed] [Google Scholar]
  • 131.Schiff ND, Giacino JT, Kalmar K, Victor JD, Baker K, Gerber M, et al. Medial septal nucleus theta frequency deep brain stimulation improves spatial working memory after traumatic brain injury. J Neurotrauma. 2013;30:119–30. doi: 10.1089/neu.2012.2646. [DOI] [PubMed] [Google Scholar]
  • 132.Shin SS, Dixon CE, Okonkwo DO, Richardson RM. Neurostimulation for traumatic brain injury. J Neurosurg. 2014;121:1219–31. doi: 10.3171/2014.7.JNS131826. [DOI] [PubMed] [Google Scholar]
  • 133.Gao J, Prough DS, McAdoo DJ, Grady JJ, Parsley MO, Ma L, et al. Transplantation of primed human fetal neural stem cells improves cognitive function in rats after traumatic brain injury. Exp Neurol. 2006;201:281–92. doi: 10.1016/j.expneurol.2006.04.039. [DOI] [PubMed] [Google Scholar]
  • 134.Shear DA, Tate MC, Archer DR, Hoffman SW, Hulce VD, Laplaca MC, et al. Neural progenitor cell transplants promote long-term functional recovery after traumatic brain injury. Brain Res. 2004;1026:11–22. doi: 10.1016/j.brainres.2004.07.087. [DOI] [PubMed] [Google Scholar]
  • 135.Bakshi A, Shimizu S, Keck CA, Cho S, LeBold DG, Morales D, et al. Neural progenitor cells engineered to secrete GDNF show enhanced survival, neuronal differentiation and improve cognitive function following traumatic brain injury. Eur J Neurosci. 2006;23:2119–34. doi: 10.1111/j.1460-9568.2006.04743.x. [DOI] [PubMed] [Google Scholar]
  • 136.Riess P, Molcanyi M, Bentz K, Maegele M, Simanski C, Carlitscheck C, et al. Embryonic stem cell transplantation after experimental traumatic brain injury dramatically improves neurological outcome, but may cause tumors. J Neurotrauma. 2007;24:216–25. doi: 10.1089/neu.2006.0141. [DOI] [PubMed] [Google Scholar]
  • 137.Sun D, Gugliotta M, Rolfe A, Reid W, McQuiston AR, Hu W, et al. Sustained survival and maturation of adult neural stem/progenitor cells after transplantation into the injured brain. J Neurotrauma. 2011;28:961–72. doi: 10.1089/neu.2010.1697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Olstorn H, Moe MC, Røste GK, Bueters T, Langmoen IA. Transplantation of stem cells from the adult human brain to the adult rat brain. Neurosurgery. 2007;60:1089–98. doi: 10.1227/01.NEU.0000255461.91892.0D. [DOI] [PubMed] [Google Scholar]
  • 139.Lu D, Mahmood A, Qu C, Hong X, Kaplan D, Chopp M. Collagen scaffolds populated with human marrow stromal cells reduce lesion volume and improve functional outcome after traumatic brain injury. Neurosurgery. 2007;61:596–602. doi: 10.1227/01.NEU.0000290908.38438.B2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Xiong Y, Qu C, Mahmood A, Liu Z, Ning R, Li Y, et al. Brain Res. Vol. 1263. Elsevier B.V; 2009. Delayed transplantation of human marrow stromal cell-seeded scaffolds increases transcallosal neural fiber length, angiogenesis, and hippocampal neuronal survival and improves functional outcome after traumatic brain injury in rats; pp. 183–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Mathews GC, Diamond JS. Neuronal glutamate uptake Contributes to GABA synthesis and inhibitory synaptic strength. J Neurosci. 2003;23:2040–8. doi: 10.1523/JNEUROSCI.23-06-02040.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Kanamori K, Ross BD, Kondrat RW. Rate of glutamate synthesis from leucine in rat brain measured in vivo by 15N NMR. J Neurochem. 1998;70:1304–15. doi: 10.1046/j.1471-4159.1998.70031304.x. [DOI] [PubMed] [Google Scholar]
  • 143.Cole JT, Mitala CM, Kundu S, Verma A, Elkind JA, Nissim I, et al. Dietary branched chain amino acids ameliorate injury-induced cognitive impairment. Proc Natl Acad Sci U S A. 2010;107:366–71. doi: 10.1073/pnas.0910280107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144••.Lim MM, Elkind J, Xiong G, Galante R, Zhu J, Zhang L, et al. Dietary therapy mitigates persistent wake deficits caused by mild traumatic brain injury. Sci Transl Med. 2013;5:215ra173. doi: 10.1126/scitranslmed.3007092. Lim et al. demonstrate that lateral fluid percussion alters sleep by inducing an anability to maintain wakefulness. Furthermore, they demonstrate that branched chain amino acid dietary therapy restores sleep by increased activity of hypothalamic orexin neurons. The restoration of sleep activity may improve memory consolidation know to occur during sleep. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Bondi CO, Klitsch KC, Leary JB, Kline AE. Environmental enrichment as a viable neurorehabilitation strategy for experimental traumatic brain injury. J Neurotrauma. 2014;31:873–88. doi: 10.1089/neu.2014.3328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Darwish H, Mahmood A, Schallert T, Chopp M, Therrien B. Simvastatin and environmental enrichment effect on recognition and temporal order memory after mild-to-moderate traumatic brain injury. Brain Inj. 2014;28:1362–301. doi: 10.3109/02699052.2013.862737. [DOI] [PubMed] [Google Scholar]
  • 147.Hamm RJ, Temple MD, O’Dell DM, Pike BR, Lyeth BG. Exposure to environmental complexity promotes recovery of cognitive function after traumatic brain injury. J Neurotrauma. 1996;13:41–7. doi: 10.1089/neu.1996.13.41. [DOI] [PubMed] [Google Scholar]
  • 148.Hicks RR, Zhang L, Atkinson A, Stevenon M, Veneracion M, Seroogy KB. Environmental enrichment attenuates cognitive deficits, but does not alter neurotrophin gene expression in the hippocampus following lateral fluid percussion brain injury. Neuroscience. 2002;112:631–7. doi: 10.1016/s0306-4522(02)00104-5. [DOI] [PubMed] [Google Scholar]
  • 149.Passineau MJ, Green EJ, Dietrich WD. Therapeutic effects of environmental enrichment on cognitive function and tissue integrity following severe traumatic brain injury in rats. Exp Neurol. 2001;168:373–84. doi: 10.1006/exnr.2000.7623. [DOI] [PubMed] [Google Scholar]
  • 150.Sozda CN, Hoffman AN, Olsen AS, Cheng JP, Zafonte RD, Kline AE. Empirical comparison of typical and atypical environmental enrichment paradigms on functional and histological outcome after experimental traumatic brain injury. J Neurotrauma. 2010;27:1047–57. doi: 10.1089/neu.2010.1313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Hoffman AN, Malena RR, Westergom BP, Luthra P, Cheng JP, Aslam HA, et al. Environmental enrichment-mediated functional improvement after experimental traumatic brain injury is contingent on task-specific neurobehavioral experience. Neurosci Lett. 2008;431:226–30. doi: 10.1016/j.neulet.2007.11.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Kline AE, Wagner AK, Westergom BP, Malena RR, Zafonte RD, Olsen AS, et al. Acute treatment with the 5-HT1A receptor agonist 8-OH-DPAT and chronic environmental enrichment confer neurobehavioral benefit after experimental brain trauma. Behav Brain Res. 2007;177:186–94. doi: 10.1016/j.bbr.2006.11.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Matter AM, Folweiler KA, Curatolo LM, Kline AE. Temporal effects of environmental enrichment-mediated functional improvement after experimental traumatic brain injury in rats. Neurorehabil Neural Repair. 2011;25:558–64. doi: 10.1177/1545968310397206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.de Witt BW, Ehrenberg KM, McAloon RL, Panos AH, Shaw KE, Raghavan PV, et al. Abbreviated environmental enrichment enhances neurobehavioral recovery comparably to continuous exposure after traumatic brain injury. Neurorehabil Neural Repair. 2011;25:343–50. doi: 10.1177/1545968310390520. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES