Skip to main content
Frontiers in Psychiatry logoLink to Frontiers in Psychiatry
. 2018 May 15;9:196. doi: 10.3389/fpsyt.2018.00196

Single-Prolonged Stress: A Review of Two Decades of Progress in a Rodent Model of Post-traumatic Stress Disorder

Michael J Lisieski 1, Andrew L Eagle 2, Alana C Conti 3,4, Israel Liberzon 5,6, Shane A Perrine 1,*
PMCID: PMC5962709  PMID: 29867615

Abstract

Post-traumatic stress disorder (PTSD) is a common, costly, and often debilitating psychiatric condition. However, the biological mechanisms underlying this disease are still largely unknown or poorly understood. Considerable evidence indicates that PTSD results from dysfunction in highly-conserved brain systems involved in stress, anxiety, fear, and reward. Pre-clinical models of traumatic stress exposure are critical in defining the neurobiological mechanisms of PTSD, which will ultimately aid in the development of new treatments for PTSD. Single prolonged stress (SPS) is a pre-clinical model that displays behavioral, molecular, and physiological alterations that recapitulate many of the same alterations observed in PTSD, illustrating its validity and giving it utility as a model for investigating post-traumatic adaptations and pre-trauma risk and protective factors. In this manuscript, we review the present state of research using the SPS model, with the goals of (1) describing the utility of the SPS model as a tool for investigating post-trauma adaptations, (2) relating findings using the SPS model to findings in patients with PTSD, and (3) indicating research gaps and strategies to address them in order to improve our understanding of the pathophysiology of PTSD.

Keywords: PTSD, single prolonged stress, anxiety, amygdala, hippocampus, prefrontal cortex, HPA axis, pre-clinical models

Introduction

Post-traumatic stress disorder (PTSD) is a psychiatric disorder that develops following direct or indirect exposure to an extremely stressful (traumatic) event or series of events. Symptoms of PTSD include intrusive memories related to the traumatic event, distress in response to trauma-related cues and avoidance of those cues, negative alterations in cognition and mood, and increased arousal and reactivity (1). PTSD is diagnosed when these symptoms last longer than 1 month and cause functional impairment and distress. It's estimated that 3.5% of the US population (more than 11 million Americans) suffer from PTSD in a given year, but less than half of these patients are in treatment, and less than half of those in treatment receive minimally adequate care (2) receiving treatment may not experience remission; the two pharmacotherapeutics currently approved by the US Food and Drug Administration for treatment of PTSD have relatively limited efficacy (3) and focus on reducing symptoms rather aimed at the underlying pathophysiology of PTSD. Given the high prevalence and significant health burden of PTSD, research aimed at better understanding its biological basis and using this knowledge to develop efficacious treatments is absolutely essential.

Because our ability to examine the neurobiology of PTSD in humans is limited, pre-clinical models provide a valuable opportunity to investigate the effects of traumatic stress exposure. The goal of this review is to provide an overview of research conducted using one pre-clinical model of PTSD, the single prolonged stress (SPS) model, and to examine its validity, utility, and potential for contributing to new and improved therapies for PTSD. While we focus on the SPS model, several other pre-clinical models have been proposed as relevant for PTSD. These include Pavlovian fear conditioning (4, 5), stress-enhanced fear learning (6, 7), exposure to predators or predator scent (810), physiological stressors such as underwater trauma (11), or restraint stress (12, 13), and protocols that (like SPS) combine multiple stressors, for example the combination of social instability with predator exposure (14) or social isolation with foot shock exposure (15). Several recent reviews have evaluated and compared these models (1622); however, making comparisons among them is not the primary aim of this review. Rather, after briefly describing the symptomatology of PTSD and key brain systems involved in this disorder, we review studies that demonstrate the utility of the SPS model to expand our understanding of clinical PTSD and the molecular mechanisms that may underlie it, and discuss potential avenues for research and novel treatment targets. In addition, a condensed summary of findings regarding the behavioral and neurobiological effects of SPS is given in Supplementary Tables 1, 2.

Symptomatology of PTSD

The American Psychiatric Association's Diagnostic and Statistical Manual of Mental Disorders (DSM 5), identifies four clusters of PTSD symptoms: intrusive re-experiencing of the traumatic event(s), avoidance of trauma-related thoughts and reminders, negative alterations in cognition and mood, and alterations in arousal (1). This clustering of PTSD symptomatology has empirical support; studies analyzing the factor structure of PTSD symptoms among diverse clinical populations such as victims of intimate partner violence (23), refugees (24), and earthquake survivors (25) provide evidence for factor structures in PTSD symptomatology which are broadly consistent with (if not identical to) the DSM 5 model.

Translating clinical diagnostic criteria and patient characteristics to constructs that can be operationalized and used in pre-clinical behavioral research is a perennial challenge. The Research Domain Criteria developed by the National Institute of Mental Health (26) attempts to provide parsimonious definitions of neurobehavioral phenotypes that are relevant to psychiatric disorders, each of which can be linked to specific behaviors, cellular, and molecular processes. We take a similar approach, grouping findings in SPS research by behavioral domain.

Only subsets of PTSD symptoms readily map onto behavioral domains that can be modeled in non-humans. For example, assessing nightmares, self-blame, and the presence of trauma-related thoughts rely on self-report, and therefore, cannot be modeled in non-humans. However, the majority of behavioral domains relevant to PTSD symptoms can be studied in non-humans in some form; these include exaggerated startle response, disruption of sleep and circadian cycle, and avoidance or fear of trauma-related cues. Increased generalization and persistence of conditioned fear, while not part of the DSM5 diagnostic criteria, has been repeatedly observed in humans with PTSD (2729), presumably involves mechanisms underlying the persistence of traumatic memories, and is readily studied in non-humans (30). Also experimentally accessible are characteristics of PTSD which are usually self-reported in humans but can also inferred through behavioral testing, such as changes in anxiety-like behavior and threat assessment, constriction of affect and reward, and disruption of attention and cognitive performance. Therefore, pre-clinical neurobehavioral analyses provide powerful tools for studying most, though not all, key behavioral processes disrupted in PTSD.

Neuroanatomy of PTSD

Neuroimaging studies have identified a set of brain regions that likely contribute to the behavioral abnormalities in PTSD, including the hippocampus (HC), amygdala (AMY), and prefrontal cortex (PFC) (3133). This triad of brain regions is also central to the brain circuit implicated in fear/safety learning (34). The function of the hypothalamic-pituitary-adrenal (HPA) axis also plays a key role in stress response (35) and has repeatedly been demonstrated to be altered in PTSD (36). As many clinical and pre-clinical studies on PTSD investigate these systems in some way, we will briefly describe their structure and function and review evidence of their involvement in PTSD from clinical and human laboratory studies.

Hippocampus

The HC is a subcortical structure that plays a critical role in learning and memory (37), and integrates contextual information to regulate behavior (38). The HC therefore stands out as a likely substrate for memory-related problems in PTSD, which include persistent re-experiencing of traumatic events, increased salience of negative emotional memories (39), deficits in working and verbal memory (4042), and impaired context-dependent modulation of memory (31, 43). Subregions of the HC are functionally heterogenious; studies in rat show that the dorsal HC (homologous to the human posterior HC) is primarily involved in information processing and cognitive performance whereas the ventral HC (homologous to the human anterior HC) is important in regulation of stress response and affect (44).

Hippocampal volume is reduced in individuals with PTSD compared to controls; however, reduced HC size is also seen in individuals exposed to severe stressors with or without PTSD (45, 46) and in non-trauma exposed twins of individuals with PTSD (47). Additionally, larger HC size is related to a positive response to PTSD treatment (48). These findings suggest that the link between HC size and PTSD diagnosis results both from vulnerability conferred by having a small HC and from a reduction in HC size caused by traumatic stress exposure. Corroborating structural studies, functional neuroimaging studies have shown altered HC function in individuals with PTSD (49). The HC has reciprocal connections to many other brain areas; importantly, the HC interacts with the PFC to regulate memory (50, 51) and with the AMY to regulate emotional arousal (52) and consolidation of fear memories (53). HC connectivity is uniquely altered in PTSD (54) and HC connectivity can predict symptom severity in patients with PTSD (55), suggesting that these connections may underlie PTSD-related deficits and emphasizing the importance of the HC in PTSD pathophysiology (54, 55).

Prefrontal cortex

The PFC encompasses a large area in the frontal lobe of the brain and has roles in decision-making, “executive” functions such as attention and working memory, and regulation of emotion (56, 57). Importantly for PTSD, the PFC plays a role in the regulation of fear learning, expression, and extinction (58). The function of two PFC subregions, the anterior cingulate cortex (ACC) and the medial PFC (mPFC), has been found to be altered in PTSD (59, 60). These brain regions regulate affective responses generated by other brain regions. Of great importance for understanding PTSD and its treatment, the human ventromedial PFC (vmPFC; analogous to the infralimbic cortex, IL, in rodents) plays a critical role in the extinction of fearful memories by processing safety signals (61) and interacting with the AMY to inhibit fear expression (62). Correspondingly, disruption of vmPFC function appears to contribute to altered emotional processing (63) and impaired retention of fear extinction learning (28, 29, 64) in PTSD.

Amygdala

The AMY is often considered “a hub” of emotional processing. Of particular relevance to PTSD, the AMY plays an important role in fear learning and extinction (65). Sensory information primarily flows into the basolateral AMY (BLA), where long-term potentiation key to associative fear learning takes place, and the signal is then conveyed to the central AMY (CeA) which regulates the output of fear behavior (65). The AMY receives input from brain regions such as the IL and sensory areas, which may act as “gates” that inhibit expression of learned fears (66). Studies in clinical populations with PTSD have found that AMY response to emotional stimuli is exaggerated in individuals with PTSD (67) and that AMY responses to fearful stimuli predict treatment response (68). These results and associated pre-clinical results indicate that dysfunction of the AMY and its connections with other brain regions may underlie the excessive persistence of fearful memories and other emotional symptoms of PTSD.

Other brain regions

Brain regions outside of the HC-AMY-PFC circuit described above have also been implicated in PTSD. Insular cortex function, for example, is altered in PTSD and anxiety-related disorders (69) and has roles in processes relevant to PTSD such as affective processing and interoceptive awareness (70). Despite this role, no studies using the SPS model have yet directly investigated insula structure or function.

Evidence that individuals with PTSD have increased startle response (71) and reduced magnitude of P3 (72), an event-related potential that may reflect increased noradrenergic output (73), suggests that locus coeruleus (LC) function might be altered in PTSD. Building on this work, several studies have investigated the effects of SPS on LC function.

Finally, evidence for dysfunction of reward and hedonic processes in PTSD (74) converges with neuroimaging evidence that striatum dysfunction is related to the affective and cognitive symptoms of PTSD (75, 76). Changes in striatal function may also explain the links between PTSD and substance use disorders (SUD) (77), compulsive behaviors (7880), and risk-taking (81, 82). Several studies have investigated striatal function following SPS, largely in relation to its role in responses to drugs of abuse and other extrinsic rewards.

HPA axis

A great deal of research has focused on the HPA axis in PTSD. The HPA axis coordinates stress responses through a highly regulated neurohormonal cascade. Corticotrophin releasing hormone (CRH) is released from hypothalamic neurons, liberating adrenocorticotropic hormone (ACTH) from pituitary cells, which in turn stimulates the secretion of cortisol (CORT) into the bloodstream by the adrenal cortex. CORT, in turn, has effects on a wide variety of tissues, including activating glucocorticoid receptors (GR) and mineralocorticoid receptors (MR) in the HC, AMY, hypothalamus, pituitary, and other regions of the brain. Activation of hypothalamic GR receptors decreases HPA axis activity, creating a negative feedback loop that limits the persistence of stress responses. The HPA axis also interacts with the brain non-hormonally; the paraventricular nucleus of the hypothalamus, where CRH-secreting cells are located, has connections with, and is regulated by, other brain regions implicated in PTSD, such as the HC (83) and AMY (84, 85). These brain regions participate in both feedback inhibition and feed-forward stimulation of the HPA axis to regulate stress responses (86). Particularly severe or persistent stressors can cause long-lasting changes in the function of HPA axis and brain responses to CORT and CRH, and this “stress programming” may contribute to the variety of maladaptive effects that severe stress can have on individuals (85, 87, 88).

Single prolonged stress (SPS)

SPS is a multimodal traumatic stress exposure protocol including sequential exposure to three stressors (2 h of restraint, a 20-min group swim, and exposure to ether until loss of consciousness) during a single continuous session. This protocol was originally designed to cause a robust stress response through three different pathways—psychological (restraint), physiological (forced group swim), and pharmacological (ether). SPS was originally described by Liberzon and Young (89) as “time-dependent [stress] sensitization,” and was observed to cause an abnormal phenotype characterized by enhanced fast negative feedback of the HPA axis which was emerging 7 days, but not 1 day, after stress exposure. This incubation period, which has also been called a sensitization or consolidation period, is usually 7 days in length. While some studies have investigated post-SPS neurobiological and behavioral changes before and after this time point, more studies are needed to define the temporal development and persistence of PTSD-like characteristics following SPS. Later work revealed similar behavioral and neuroendocrine effects in mice exposed to a similar procedure (90), demonstrating the viability of this approach in multiple rodent species. The specificity of the phenotype and the time-dependent nature of the change corresponds to the altered neuroendocrine response to stress in individuals with PTSD (9193) and the dynamic course of PTSD symptoms, which tend to escalate over time to produce long-term functional impairment (94, 95). Further studies demonstrated that combined exposure to all three components of SPS is required for the expression of PTSD-like phenotypes (96), suggesting that interaction among the multiple components of SPS induces abnormal neurobiological phenotypes associated with PTSD differently than any component stressor alone (2325, 97). Over the past 20 years, SPS had been widely used to study adaptations following a severe traumatic stressor, with great emphasis placed on characterizing behavioral changes following SPS and their relationship to underlying neurobiological processes.

In order to provide a useful platform for studying post-trauma adaptations relevant to PTSD, the SPS model (and any pre-clinical model) must satisfy few basic criteria. In particular, a model of PTSD should recapitulate the phenomenology of the disorder by producing behavioral changes that resemble behavioral changes seen in PTSD, and which rely on similar mechanisms (18). Furthermore, interventions that improve PTSD symptoms should also improve corresponding behavioral changes in the model, and vice versa. SPS has been demonstrated to satisfy these conditions of validity (97), as it uses a single episode of severe stress to generate persistent behavioral changes in PTSD-relevant behavioral domains which are responsive to interventions used to treat PTSD.

The remainder of this review describes recent progress that has been made in using SPS to understand the pathophysiology of PTSD. Where appropriate, we include brief comparisons to research on human patients to emphasize the validity of SPS as a model of traumatic stress.

Behavioral effects of SPS

Abnormal fear learning

PTSD has long been proposed to involve aberrant function of normal fear learning processes (98) leading to abnormal responses to potential threats and non-threatening (but trauma-associated) stimuli (99). This has led researchers using SPS to focus on behaviors and brain mechanisms involved in fear memory formation and retrieval (100). Several phases of fear conditioning can be studied: the acquisition of a fear memory, the expression of conditioned fear after a delay, the extinction of the conditioned fear response, the retention (recall) of this extinction, and the reactivation of an extinguished fear response through spontaneous recovery, reinstatement, or renewal (101). While fear conditioning alone is not a model of PTSD, it is used to assess abnormalities in fear learning that are associated with PTSD. In experimental studies that use PTSD models, fear conditioning can be broadly categorized into two paradigms: trauma cue-specific fear conditioning wherein a neutral stimulus that was paired with the trauma exposure is used to reactivate memory of the traumatic event, and de novo fear conditioning in which a neutral stimulus (cue or context) is paired with a novel aversive stimulus, with this whole process taking place after exposure to a traumatic event.

Trauma-cue responses

One of the defining characteristics of PTSD is increased reactivity to and avoidance of cues associated with the traumatic stressor (1). This has been investigated in a number of human laboratory studies (102), it was even proposed as potential biomarker to predict (103) and index (104) treatment response. There are relatively few reports on the effects of SPS on trauma cue-specific fear despite the fact that SPS, as a procedure consisting of a single exposure to a strongly aversive event, is amenable to this type of experimental design.

Defensive reactions to and avoidance of trauma cues have been demonstrated for up to 43 days following SPS exposure in rats (105) and up to 7 days following SPS exposure in mice (90). This has been observed to occur to both direct trauma-associated cues (e.g., restraint apparatus, swim tank, ether chamber) and peripherally-associated cues (e.g., holding chambers, tones, scents) (90, 105). Additionally, while the fear response appears to extinguish following repeated exposure to a trauma-associated scent cue, it can be reinstated by exposure to the anxiogenic drug yohimbine (106), suggesting that the extinguished response remains sensitive to reactivation. Further studies are needed to determine the neuronal underpinnings of trauma cue reactivity in SPS and how this differs from standard fear conditioning protocols.

De novo fear conditioning

When individuals with PTSD undergo extinction learning, this extinction is not as well-retained as in control subjects; thus, it is said that these individuals with PTSD have impaired extinction retention (28, 29, 107). This could be due to trauma-induced neurobiological changes in PTSD (28), or may reflect a pre-existing impairment in memory that predisposes susceptible individuals to PTSD following trauma exposure (108). Mirroring clinical findings, SPS produces robust extinction retention deficits in rats trained to associate cues (auditory tones) with foot shock (96, 109114), and in mice exposed to SPS (90). A similar pattern of delayed or poorly-retained extinction is also demonstrated by some studies of contextual fear conditioning following SPS, in which an environmental context serves to predict the shock rather than a discrete cue (115–,117). The effect of SPS on extinction retention is time-dependent, requiring an incubation period to emerge (111). SPS-induced extinction retention deficits have been associated with enhanced GR expression in the HC (96) and decreased activity in the IL during fear recall (118).

SPS also enhances the magnitude of conditioned fear response to contexts associated with aversive stimuli such as foot shock in rats (114, 119128) and mice (129). This enhanced contextual fear conditioning has been observed to recover by 4 weeks following SPS (130), in line with the clinical finding that populations exposed to some kinds of traumatic events (non-intentional traumatic events, in particular) tend to express symptoms initially but gradually recover spontaneously without treatment (131). Conditioned taste aversion, while not a type fear learning, is also enhanced following SPS (132), indicating that memory for aversive events may be abnormally high following trauma across modalities and tasks.

Finally, PTSD may produce specific dysregulation of contextual processing of aversive stimuli, leading to overgeneralization of learned fear and impaired ability to learn or respond to safety signals. Clinical research using fear conditioning paradigms demonstrates that PTSD patients do not adequately use context or cues to distinguish between threatening signals (those which indicate that an aversive stimulus is likely to occur) and safety signals (those which indicate that an aversive stimulus is unlikely to appear) (31, 133). This has been associated with the tendency to over-generalize negative contextual associations (43). Furthermore, clinical studies suggest that overgeneralization of conditioned fear may predict treatment responsiveness (27), indicating the importance of understanding the effects of traumatic stress on fear generalization.

In contrast to studies in humans showing a reduction of discrimination between fearful and non-fearful contexts in humans with PTSD (31), contextual fear discrimination was intact in SPS-exposed rats in some studies (122). However, we have recently shown that a subset of susceptible mice exposed to SPS have impaired fear discrimination in a cue-conditioning task, indicating that they generalize conditioned fear to non-threatening stimuli more readily than control animals (30).

Studies using de novo fear conditioning provide ample evidence that SPS mirrors fear deficits observed in PTSD, providing support for the model's applicability and translational relevance. Further research is necessary to determine what factors predispose trauma-exposed individuals to maladaptive disruptions of fear learning both in regard to trauma-cue reactivity and abnormalities in de novo fear conditioning. SPS, as a well-validated pre-clinical model of post-traumatic alterations in fear learning, can serve as a powerful platform to study the mechanisms underlying these changes, which can inform translational work aimed at improving treatments for PTSD.

Hyperarousal

Hyperarousal is one of the criteria of PTSD as defined by the DSM 5 (1). Symptoms that indicate hyperarousal include exaggerated startle response, hypervigilance, difficulty concentrating, and disruption of sleep. Several behaviors related to these symptoms are amenable to testing in non-human models. Exaggerated startle response, in particular, has been studied relatively extensively following SPS. Indices of cognitive performance that rely on attention and memory have also been studied following SPS, as have neurophysiological aspects of sleep; additionally, anxiety-like defensive behavior observed in rodents may correspond to the hypervigilance reported by patients with PTSD, as factor analyses of PTSD symptoms have identified an “anxious hyperarousal”-like factor that includes endorsement of both anxiety and arousal items (2325). Effects of SPS on cognitive performance, sleep, and anxiety-like behavior are discussed in subsequent sections.

The startle response is used to assess anxiety and hypervigilance across many species and reflects sensitivity to potential threat (134). PTSD commonly enhances startle responses (71, 135, 136), indicating that this may be a useful translational tool to investigate the hyperarousal that often characterizes PTSD (137). SPS has been repeatedly confirmed to increase startle response (105140), and specifically only after an incubation period (122). Exaggerated acoustic startle has been used as tool for the evaluation of PTSD treatments in the SPS model; for example, the anti-epileptic drug, topiramate, which may be useful in treating PTSD (141), was shown to attenuate SPS effects on startle response (140). Other classes of drugs including cannabinoid agonists (138), neuropeptide Y (NPY) (139), and psychostimulants (105) also blocked SPS effects on startle, indicating that drugs that target these neurotransmitter systems may be efficacious treatments for post-traumatic hyperarousal.

Cognitive dysfunction

PTSD is often associated with more general cognitive dysfunction, including memory impairment (40), deficits in mental processing speed (142), and inattention (143). Cognitive impairment may be a result of PTSD pathophysiology, and/or it may be a risk factor for PTSD, as clinical evidence suggests that deficits in executive function may be detected in individuals who will go on to develop PTSD following trauma exposure (144146). Similarly, SPS induces impairments in attention and non-fear memory performance. For example, SPS decreases learning performance in a spatial water maze (122, 147). However, the interpretation of these results is confounded by re-exposure to water, which is a reminder of the stressor in SPS-exposed rats but not controls (106) and so may affect performance in this task. We and others have also observed impairment in object recognition (148, 149) and social discrimination (148). This impairment in social discrimination may be attributed to decreased behavioral flexibility, a notion corroborated by evidence that SPS decreases performance in set shifting tasks (150152). These studies are in line with evidence that SPS reduced spontaneous activity in the LC (153) and altered mPFC function (118, 154), as both the LC (155) and mPFC (156) are important in maintaining attention and cognitive flexibility.

Sleep

Sleep disruption likely contributes to the development and maintenance of PTSD in humans (157), and is associated with diverse aspects of PTSD pathology, including sustained fear responses to trauma cues (158), hyperarousal (159), increased use of cannabis (160), and suicidality (161). Therefore, understanding the role of sleep in PTSD is necessary to identify post-trauma adaptations leading to novel targets for PTSD treatment. Mirroring findings on disrupted sleep architecture in patients with PTSD (162), SPS causes increased wakefulness and REM sleep (163165). Importantly, the degree of sleep disruption predicted SPS-induced extinction retention deficits (113), which suggests that traumatic stress-induced sleep alterations likely contribute to other PTSD symptoms.

Anxiety

While general anxiety is not a defining characteristic of PTSD in the DSM5, increased anxiety is often reported by individuals with PTSD (166). Sustained defensive (unconditioned anxiety-like) behaviors are reliably measured in rodents using standard behavioral assays, including the elevated plus maze, open field, and light-dark box. A large number of reports show that SPS produces or increases anxiety-like behavior in these established behavioral models of anxiety (120, 124, 130, 167173). Studies using modifications to SPS, such as re-exposure to stressors (147) or the addition of other stressors to the SPS paradigm (129, 139, 174179), also report increased anxiety-like behavior. On the other hand, other studies fail to find a consistent anxiety-like phenotype produced by SPS (180184). One powerful strategy to determine the relationship between anxiety-like behavior and more canonical PTSD-like phenotypes is to identify subpopulations of individuals that express anxiety-like behavior following traumatic stress and quantify other stress-related behaviors in these individuals to determine whether anxiety-like behavior changes in concert with other behaviors affected by SPS. Employing this approach, recent studies have identified that animals with high post-SPS anxiety also show exaggerated response to trauma cues (105). Furthermore, the time course of anxiety following SPS may be different from other phenotypes, as it has been found that anxiety-like behavior recovers toward baseline within 4 weeks following SPS (130).

Despite the wide use of anxiety-like behavior as a marker of “PTSD-like” changes following SPS, there are no published reports of the effects of SPS on some brain regions important in anxiety-like behavior (for example, the bed nucleus of the stria terminalis and insula), even though these brain regions have been previously implicated in the pathology of PTSD, for example in human neuroimaging studies (185) and in rodent studies on stress sensitization using other animal models (186, 187). Future research is necessary to relate specific post-SPS behavioral changes to neuronal systems outside of the HC-mPFC-AMY circuit and HPA axis involved in PTSD, including studies that differentiate populations which are uniquely affected by stress. This approach may be productive in identifying specific treatment targets for subgroups of PTSD patients, such as those who show treatment resistance to standard medications, those with high anxiety, and those with persistent (rather than spontaneously regressing) symptoms of PTSD (19).

Disruptions of affect and reward

Negative changes in affect comprise part of the current diagnostic criteria for PTSD (1); this symptom cluster is sometimes referred to as “emotional numbing” and includes amnesia, anhedonia, and detachment (188). Reactivity to natural rewards and motivation to pursue them are both readily measured in pre-clinical rodent models, making these behavioral domains amenable to testing following SPS.

Negative affective disturbances including anhedonia and decreased motivation are commonly reported by PTSD patients (189). We have repeatedly observed decreased sucrose preference following SPS (169, 190). While these findings are suggestive of anhedonia, they do not distinguish between affective response and motivation, as changes in either of these processes could lead to reduced consumption of a palatable substance. Demonstration of reduced positive affective responses awaits future research using more specific analyses such as facial reactivity analysis (191) or measurements of reward-related ultrasonic vocalization (192).

Learned helplessness, a putatively affective disruption typically associated with depression (193), is increased following SPS in the water maze (194) and forced swim test (97, 139, 177, 178, 195). Importantly, increased learned helpless in the forced swim test models may model learned helplessness associated with depression (196) and is reversed by antidepressants (197). However, the use of forced-swim tests following SPS is likely confounded in that water acts as a potent trauma-associated cue (106) and stressor (132, 183, 198) in its own right. Supporting this, SPS did not affect a similar measure of learned helplessness that involves no water exposure, the tail suspension test (97). Additionally, forced swim test behavior changes over multiple exposures to water (199), and so post-SPS changes may simply reflect adaptive habituation rather than a maladaptive change (200, 201). Therefore, future studies into the effects of SPS on learned helplessness-like behavior should include test procedures that do not themselves serve as trauma cues, such as tail suspension (202), and exposure to inescapable shock (203, 204).

Modeling comorbid conditions using SPS

While PTSD can be debilitating on its own, it is also highly comorbid with other psychiatric disorders. For example, SUD are disproportionately common among patients with PTSD, and this comorbidity is associated with increased mortality and poor treatment response (205, 206). Individuals with PTSD are also at increased risk for chronic pain (207, 208) and somatic disorders such as cardiovascular disease, respiratory disease, gastrointestinal disorders, and a variety of autoimmune disorders (208, 209). While a growing number of studies have investigated inflammation following SPS (see section Inflammation and Glia), none have yet directly modeled comorbid immune disorders. However, SPS has been used to investigate mechanisms related to substance use and pain perception following traumatic stress.

Substance use

Individuals with PTSD are prone to increased use of drugs including alcohol, nicotine, opioids, anxiolytics, marijuana, and psychostimulants (77, 205, 210212). This may be due to a number of factors, including self-medication for PTSD symptoms, increased vulnerability to stress following substance use, or the presence of shared risk factors for PTSD and SUD (213). SPS, when combined with models of substance use, provides a powerful pre-clinical platform to investigate these processes and their neural underpinnings. Our initial studies utilized drug sensitization models to probe neurochemical changes following SPS, and more recent studies have combined self-administration models with SPS to gain insight into the biological relationships between SUD and PTSD.

Drug sensitization is a process by which the brain becomes more responsive to the presence of a drug over repeated use, which is a useful model for probing mechanisms regulating addiction (214). Drug sensitization is altered following SPS, although findings vary by the substance investigated. For example, we have shown that SPS blunts ethanol-induced behavioral sensitization (215), but enhances behavioral sensitization to repeated cocaine (216) and methamphetamine (217). Another group reported that SPS produced noradrenergic-dependent cross-sensitization to acute amphetamine (218). These interactions between SPS and drugs of abuse have been attributed to alterations in the norepinephrine system (218), dopaminergic (D2) receptors (190, 215), CB type 1 (CB1) receptors (215), and a marker of glutamatergic synapses (PSD-95) (215) in the striatum, indicating that they may result from diverse and widespread post-SPS changes in neurobiology.

In line with the hypothesis that traumatic stress exposure augments drug seeking and taking, SPS enhances ethanol reward (176). However, SPS, somewhat surprisingly, has no effect on short-access self-administration, extinction, and reinstatement of cocaine, and decreases cocaine conditioned place preference and long-access self-administration (190, 216). Experiments investigating other aspects of self-administration behavior, including motivation and drug sensitivity, need to be considered in future studies looking at drug seeking and taking behaviors in order to better understand the deficits caused by traumatic stress exposure on the reward system.

While findings that SPS alters drug sensitization and self-administration are broadly aligned with clinical findings of altered substance use in PTSD, it is as yet unclear how the mixed results could be synthesized. In resolving this issue, it will be imperative to consider individual differences in substance use responses after trauma exposure (219). An approach that identifies SPS-susceptible and resistant subpopulations has been used to show that a single injection of amphetamine can block expression of PTSD-like behaviors in PTSD-susceptible rats (218). Such heterogeneity is clinically meaningful; for example, PTSD patients with alcohol use disorder tend to have greater hyperarousal symptoms than those using other drugs (220), indicating that subpopulations of trauma-exposed individuals may have a higher propensity to self-administer specific drugs of abuse and may respond differentially to treatments. Further investigations of a wider variety of substances of abuse in SPS-exposed animals using sophisticated approaches to analyze heterogeneity among these animals provides an opportunity to study the converging effects of PTSD and SUD on brain function rigorously and robustly.

While several studies have investigated how SPS exposure affects responses to drugs of abuse, only one study has investigated the effects of exposure to an addictive drug on later responses to SPS, finding that cocaine did not enhance vulnerability to anxiety-like behavior after SPS (180). Since clinical findings indicate that many patients with both PTSD and SUD acquire SUD first (221), further studies investigating how a history of exposure to drugs of abuse may modulate the effects of SPS on other behaviors, and how pre-existing traits may influence the effects of both SPS and drug exposure, are important to thoroughly model the causal interactions between SUD exposure and the development of PTSD.

Enhanced pain sensitivity

Chronic pain syndromes are often comorbid with PTSD (222). However, clinical studies have fallen short of demonstrating the effects of PTSD on pain thresholds, emotional response to pain, and a potential shared vulnerability to chronic pain (223). This highlights the need for a pre-clinical model like SPS to investigate these phenotypes. Indeed, while there is some evidence to suggest that SPS decreases pain sensitivity (120), others have reported long-lasting increases in sensitivity using the same assay (224) and increased visceral hyperalgesia (225). Pre-clinical studies using the SPS model have also suggested mechanisms by which traumatic stress may enhance pain sensitivity, such as altered PKCγ in the spinal cord (225), enhanced reactivity of LC to noxious stimuli (153), and modulation via nociceptin/orphanin FQ signaling (224). These, and potentially other novel pathways, are promising targets for the treatment of chronic pain in patients with PTSD which continue to be investigated using SPS.

Individual differences in vulnerability to SPS

As only a subset of individuals exposed to trauma develops PTSD, an important goal of research is to identify risk factors, including pre-existing or post-trauma factors (including age, sex, genetic, and personality differences, and early post-trauma response), that confer susceptibility or resilience to PTSD. To model pre-existing differences, such factors must be identified by manipulating or measuring relevant phenotypes before and after SPS exposure in longitudinal studies. Initial studies using such an approach have shown that greater pre-SPS anxiety predicts post-SPS enhanced reactivity to trauma-related cues (105, 106). In addition to considering pre-existing risk factors, individual variability in responsivity to trauma can be accounted for by differentiating SPS-exposed animals into susceptible and resilient populations using well-validated, replicable behavioral tests. For example, rats categorized as “vulnerable” to SPS using measures of enhanced anxiety and arousal also exhibit anhedonia, locomotor depression, and avoidance of trauma-related cues in a separate set of behavioral tests, whereas animals who are “resistant” to the effects of SPS on anxiety and arousal do not (106, 226). Future research using SPS should aim to take advantage of such longitudinal approaches to identify possible risk factors for PTSD, examine variability in responses to traumatic stress between “resilient” and “vulnerable” individuals, and identify the neural mechanisms that underlie these individual differences in traumatic stress responses.

In addition to identifying pre-existing or early-post-trauma traits that predict the enduring behavioral effects of traumatic stress, the effects of sex, age, and life history can be investigated using SPS. There is strong evidence that women are at increased risk for PTSD (206, 227) and experience symptoms differently than men, reporting greater distress, sleep disturbance, and diminished interest in daily activities (228230), and exhibiting less maladaptive coping behavior (231). Evidence from human studies suggests that sex hormones such as estrogen may affect stress reactivity (232); however, many questions remain unanswered, particularly in the SPS model. The only report of sex differences following SPS found that SPS did not induce fear extinction retention deficits in female rats as it did in male rats, though it did alter GR expression in the dorsal HC of female rats (109). While two studies have shown that estradiol administration mitigates the behavioral effects of SPS, one study used only male rats (125) and the other did not report the sex of its subjects (126). These findings are therefore difficult to interpret, and further research on the role of sex and sex hormone in the SPS model are important in establishing its generalizability across sexes.

Epidemiological evidence indicates that age of trauma exposure interacts with other characteristics to determine likelihood of developing PTSD following a traumatic experience (233). Therefore, it is important to determine the role of age in the development of SPS. However, to date, no peer-reviewed publications have evaluated age as a factor influencing susceptibility to SPS. Clinical evidence also indicates that early life stress and environmental adversity contribute to the vulnerability to later life PTSD (234). Supporting this, one study found that neonatal isolation interacts with adult SPS by promoting greater SPS-induced increases in anxiety and contextual fear (120). The lack of published studies regarding the effects of age and sex on post-SPS behavioral and neurobiological changes highlights the need for research investigating age, sex, and previous life history on post-SPS outcomes, which could define neurobiological mechanisms which control susceptibility to PTSD.

Neural and molecular mechanisms in SPS

Neuroendocrine responses and corticosteroid receptors

HPA axis dysregulation is observed in many cases of clinical PTSD, making it an attractive potential biomarker for PTSD (137). While some studies indicate that basal CORT levels are low in PTSD (235, 236), differences in basal CORT levels between PTSD patients and controls are not consistently seen (237239). A more reliable measure of HPA axis dysregulation is increased suppression of ACTH (91) and CORT (92) secretion in response to central stimulation of GR (a normal mechanism of negative feedback). This can be tested using the dexamethasone suppression test, and results consistently show an enhanced negative feedback response across many different populations with PTSD (240) and in animals exposed to SPS (90, 122, 241). One potential mechanism for this altered HPA axis activity is altered function and/or expression of steroid receptors, specifically GR and MR. The expression of GR and MR is programmed by stress and determines the responsiveness of neurons to circulating CORT; a high GR/MR ratio leads to increased sensitivity to the stress-related effects of increases in circulating CORT, including negative feedback of the HPA axis (242, 243). A large and growing body of evidence implicates altered expression and sensitivity of GR and MR in PTSD (88), and these targets, particularly an increased GR/MR ratio, may serve as useful biomarkers for PTSD phenotypes (36). Many studies have used the SPS model to investigate changes in HPA axis function and brain expression of GR and MR and to link these neurobiological changes to aberrant post-stress behaviors.

Mirroring clinical findings, SPS enhances rapid negative feedback of corticosterone (CORT; analogous to cortisol in humans) release (122, 138, 183, 241), and we have recently replicated this finding in mice (90). SPS also increases GR expression in many regions important for emotion, arousal, and memory, including HC (112, 184, 244), PFC (112, 245), and LC (246). Interestingly, no changes in GR as a result of SPS have been observed in AMY (184, 247). The effects of SPS on MR levels appear to be more complex. MR has been shown to decrease after SPS in HC (248) and LC (246), and MR expression in the mPFC appears to be biphasic, increasingly shortly after SPS, decreasing over the next 14 days, and then rebounding toward baseline (249).

Together these findings suggest that altered GR/MR expression may contribute to altered neuroendocrine response following traumatic stress and that this may underlie the phenotype seen following SPS. Supporting this, GR antagonists block SPS-enhanced contextual fear (122). Increased glucocorticoid signaling may contribute to PTSD-related memory impairment by promoting over-consolidation of aversive memories (250) and contributing to hypervigilance, anxiety, and emotional distress (251). This makes steroid receptor expression in the HPA axis and limbic brain regions a promising target for further investigation using the SPS model.

Neurotransmitter and neuropeptide systems

Excitatory and inhibitory neurotransmission

Several behavioral domains disrupted by PTSD depend on the balance of inhibitory (e.g., glycine, GLY, and gamma-aminobutyric acid, GABA) and excitatory (e.g., glutamate; GLU) neurotransmission, including dissociation (252), sleep (253), and fear learning and extinction (254, 255); therefore, disruptions in these excitatory and inhibitory systems may underlie many of the behavioral changes seen in PTSD. Approaches using 1H-MRS to probe the GLU and GABA systems in vivo have identified differences between individuals with PTSD and controls. Notably, PTSD is associated with increased GABA (256) and decreased GLU (257) in ACC, which may predict hyperarousal symptoms (253). In contrast, temporal, parietal, occipital, and insular cortices tend to show decreased GABA and increased GLU in PTSD, and peripheral measures suggest an overall decrease in GABA and an overall increase in GLU (258). In addition, a PET study showed that GABAA binding is decreased in the cortex, HC, and thalamus (259), indicating that PTSD also dysregulates GABA neurotransmission at the level of GABA receptors. These studies demonstrate that PTSD is associated with region-specific changes in excitatory and inhibitory neurotransmission which may contribute to PTSD symptomatology, making these systems important targets of pre-clinical study.

Several studies using SPS have investigated markers of GLU neurotransmission. SPS decreases mPFC, but not HC or AMY, GLU and glutamine 7 d after exposure (154, 260) and has also been observed to increase striatal expression of PSD-95, a glutamatergic synaptic marker (215). Additionally, increased phosphorylation of type 1A AMPA receptors in mPFC has been associated with cognitive inflexibility following SPS (151). In addition to these changes in mPFC GLU function, GLU levels in cerebral spinal fluid (CSF) increased 14 d after SPS exposure (130), suggesting brain-wide changes in GLU-based excitatory neurotransmission or metabolism. Finally, there is evidence that some drugs block the behavioral effects of SPS by regulating GLU neurotransmission. For example, the HDAC inhibitor vorinostat blocked SPS-induced impairments in fear extinction and increased expression of type 2B NMDA receptors (261), and the anticonvulsant and AMPA receptor antagonist topiramate attenuated SPS-induced exaggeration of acoustic startle response (140). This evidence suggests that altered GLU signaling may be driving SPS-induced memory impairments.

Less is known about the effects of SPS on inhibitory neurotransmission. While one report showed decreased GABA levels in HC after SPS (198), another report found no changes in GABA in HC, AMY, or mPFC (154). HC glycine transporter was persistently increased following contextual fear conditioning in SPS-exposed, but not control, rats (121), and this upregulation was associated with decreased extracellular GLY (262). How this may contribute to PTSD-associated behavior is unclear and likely complex, especially as GLY has a role in regulating excitatory neurotransmission. Nonetheless, this system may be a useful target for treatment. For example, D-cycloserine, a partial agonist of GLY binding sites on NMDA receptors which can ameliorate SPS-induced impairments in fear extinction and alter GLU receptor expression in the HC (117), has been proposed as a potential therapeutic for treating anxiety and PTSD in humans (263). More research is necessary to determine how the closely-connected systems of GLU, GABA, and glycine signaling are related to maladaptive post-traumatic behaviors. Notably, GLU and CORT signaling converge onto intracellular pathways that are altered following SPS, suggesting that the interactions between GLU and CORT may be involved in some post-traumatic adaptations (264).

Neuromodulators

The monoamines, serotonin (5-hydroxytryptamine; 5-HT), dopamine (DA), and norepinephrine (NE), likely play a significant role in the neurobiology of PTSD (265). Indeed, the only FDA-approved medications for PTSD, sertraline (Zoloft) and paroxetine (Paxil), are selective 5-HT reuptake inhibitors (SSRIs). However, these drugs have only moderate efficacy (3), underscoring the need to better understand the contribution of monoamines to PTSD symptomatology and neuropathology. Much attention has been paid to 5-HT in PTSD research as it is the main target for SSRIs using in the treatment of PTSD.

The main source of 5-HT in the forebrain is the dorsal raphe nuclei. Serotoninergic projections are distributed widely throughout the brain, play roles in many behaviors including emotional, social, and reward-related behaviors, and likely interact with CRF and CORT signaling to regulate stress coping behavior (266) and contribute to stress-related disorders (267). As SSRIs improve some PTSD symptoms (3), dysregulated 5-HT release or signaling may underlie post-traumatic behavioral changes. Our own and others' findings show that the behavioral effects of SPS can be ameliorated by a variety of SSRIs, including escitalopram (114), fluoxetine (268), and paroxetine (90, 127, 269). SPS also dysregulates 5-HT receptor expression and signaling across diverse brain regions. For example, SPS increases expression of 5-HT2C receptors in the AMY (119), where these receptors regulate contextual fear and responses to anxiogenic stimuli (270), and increases 5HT1A receptors in the oculomotor nucleus (271). In contrast, SPS decreases HC expression of 5-HT3A receptors, which are thought to be involved in neurogenesis and antidepressant effects (272). This change has been indirectly linked to PTSD symptomology by the finding that an intrahippocampal 5-HT3A receptor agonist infused immediately after SPS blocked fear extinction deficits (181). SPS might also affect 5-HT release and/or reuptake, as 5-HT concentrations are increased in the HC following SPS (183) and 5-HT reuptake transporter (SERT) knockdown in dorsal raphe projection neurons prevented SPS-induced fear extinction deficits (182). Alterations in 5-HT have been linked to behavioral changes seen following SPS; for example, markers of 5-HT turnover in the dorsal HC has been correlated with increased fear generalization in mice exposed to SPS (30). Changes in 5-HT appear to be region-specific, as SPS has been found to cause no change in 5-HT release in the IL, although this structure regulates fear extinction and retention of extinguished fear response (114).

NE is released into the forebrain from the LC and plays a critical regulatory role in arousal and attention (155). Its dysfunction likely contributes to PTSD symptoms, particularly those involving perturbations of arousal such as hypervigilance, startle, sleep disruptions, and trauma-cue reactivity (273). Notably, individuals with PTSD show increased CSF concentrations of NE (274), suggesting increased NE release, and NE dysfunction predicts specific PTSD symptoms, such as disrupted sleep (275) and startle (276). Correspondingly, NE function appears to be disrupted by SPS. SPS decreases NE release in the IL (114) and increases tyrosine hydroxylase and dopamine β-hydroxylase levels, enzymes responsible for the synthesis of catecholamines, in the LC (277). Additionally, SPS decreases tonic firing but increases phasic firing in LC neurons (153), which may contribute to a hypervigilant state (155) contributing to the enhanced startle and fear extinction deficits seen following SPS. To date, no study has investigated this causal relationship; therefore, LC function and NE signaling are promising areas for further investigation and potential development of novel treatment targets.

DA is broadly implicated in a variety of adaptive functions, including movement, learning, emotion, reward, and motivation (278). Its dysregulation has been implicated in multiple psychiatric disorders including schizophrenia (279), depression (280), and addiction (281). Recent evidence suggests that DA might play a unique role in behaviors associated with PTSD, including fear learning and extinction (282), addiction (283), and anhedonia (74). Multiple clinical findings link DA dysfunction to PTSD: DA metabolites in CSF decrease following traumatic reminders (284), striatal DA transporter (DAT)expression is increased in PTSD patients (285), and polymorphisms in genes encoding DAT and DA receptors are associated with PTSD (286). Mirroring these findings, we have observed SPS-related decreases in DA levels and turnover (i.e., DOPAC/DA) (190), D2 receptor availability (190) and concentration (215), corresponding with increased DAT concentrations in the striatum (190). SPS also decreases DA release in the PFC (114), which may promote extinction learning deficits (58). These effects may also be related to changes in reward function, affect, and sensitization to drugs of abuse described previously. Thus, future studies into the mechanisms for these changes in monoamines by SPS has the potential to provide great insight into affective and motivational symptoms of PTSD, including the increased risk of substance use in PTSD (287), and facilitate the development of targeted therapeutics.

Finally, acetylcholine (ACh), which signals through a variety of receptor subtypes to regulate both excitatory and inhibitory neurotransmission centrally and in the parasympathetic nervous system, may contribute to SPS phenotypes. Initial evidence, though scant, indicates that SPS increases muscarinic ACh receptor binding in HC and mPFC (132) and also that activation of spinal α7 nicotinic ACh receptors can reverse SPS-induced hyperalgesia (288). Especially in light of clinical findings that PTSD symptoms are associated with parasympathetic dysregulation (289) this suggests that the ACh system may be a promising target for treating both the central and peripheral pathophysiology of PTSD.

Cannabinoids

Cannabinoids (CBs) are lipid-based neuromodulators that largely inhibit neural function throughout the brain by acting as a retrograde neurotransmitter (290). The CB system appears to be altered in PTSD, and the recent, rapid development of CB research and pharmacology makes this a potentially fertile field for new interventions for PTSD (291). For example, THC (292) and the synthetic CB nabilone (293, 294) have been investigated in early clinical trials for PTSD-related nightmares and sleep disruption, and oral THC can facilitate successful recall of extinction memory in humans (295). Pre-clinical research suggests a role for CB signaling in other symptoms of PTSD, informed by the common use of cannabis as a coping behavior following the onset of PTSD (160, 296). The CB receptor agonist, Win55,212-2, can block SPS-induced impairment of contextual fear extinction when administered systemically or infused directly into BLA, subiculum, or vmPFC (116, 138). This effect appears to rely on an interaction with GR signaling, as GR blockade in the subiculum or BLA eliminates this effect (116). This is in line with evidence indicating that CB receptors in BLA modulate stress effects on affective memories (297). Additionally, mice exposed to SPS followed by ethanol administration have decreased CB1 receptor expression in the striatum (215). Further investigation of CB mechanisms using the SPS model could help explain the widespread use and abuse of cannabis in PTSD patients, as well as yield additional targeted treatments for PTSD.

Neuropeptides

A variety of neuropeptides have been implicated in stress responses and many are known to regulate HPA axis function. We focus here on the neuropeptides that have been shown to be affected both in PTSD and SPS, including oxytocin and NPY.

Oxytocin has important roles in social behavior and emotional reactivity (298). Variation in oxytocin/vasopressin genes might be related to the risk of PTSD development (299, 300), and oxytocin may facilitate effective psychotherapy for PTSD (301) and even prevent the development of PTSD in some individuals (302). Studies using SPS have corroborated these findings, showing SPS-induced increases in oxytocin receptor (OXTR) availability in the AMY, hypothalamus, and HC (89). Despite these findings, oxytocin administration does not reverse SPS-induced deficits in extinction learning (115). Nonetheless, a recent clinical trial found that oxytocin had significant effects in preventing the development of PTSD in a subgroup of patients showing high acute symptoms when given shortly following trauma (303). Further work to evaluate the mechanisms by which SPS affects oxytocin signaling may be fruitful in determining the potential of the oxytocin system as a treatment target for PTSD.

NPY is widely distributed in the brain and plays a role in fear (304), arousal (305), and traumatic stress responses (306). Clinical studies have linked variations in plasma NPY and polymorphisms in genes for NPY and its receptors with PTSD (307), making this system an attractive pharmacotherapeutic target for stress-related disorders (306). Several studies using SPS indicate that NPY interacts with brain monoamine systems and the HPA axis to orchestrate post-SPS behavioral changes. SPS decreases NPY receptor subtype 2, while at the same time increasing CRH and CRH receptor subtype 1 expression in the LC (277). Intranasal NPY administration shortly after SPS can prevent these changes as well as rescue SPS-induced increases in hippocampal GR and SPS-induced behavioral abnormalities (139, 178, 308). SPS also increases the number of NPY-reactive fibers in the BLA, which are positioned in contact with GR/MR-containing pyramidal neurons; this change is associated with an increase in morphological complexity in BLA neurons, but it is as yet unclear what role NPY plays in this change (247). These studies suggest that the role of the NPY system in regulating post-SPS changes in neurophysiology and behavior may be a promising field of study using the SPS model.

Cellular adaptations in SPS

A unique advantage of pre-clinical models is that they allow insights into cellular processes underlying disease development and progression. Recent studies have begun to interrogate the effects of SPS on cellular and intracellular adaptations in the brain. Neuronal morphology can adapt in response to various stimuli, including stress (309), and determines the functional properties of neurons and their synaptic connections (310, 311). Studies of neural morphological changes following SPS have identified increased dendritic branching in the AMY (247) which may be related to altered fear conditioning, and somatic changes suggesting neuronal damage and apoptosis in the AMY (122) and HC (312, 313). Studies using SPS have not yet replicated findings from other pre-clinical models of traumatic stress showing reduced dendritic branching in the HC (314), which may explain the observation in rodents (315) and humans (45, 46) that HC volume is decreased following traumatic stress. While expanding and clarifying research on the effects of SPS on neuronal morphology remains an important goal, several laboratories have developed research programs aimed at unraveling the intracellular mechanisms that regulate responses to SPS, with a focus on mitogen activated-protein kinase (MAPK), protein kinase B (also known as Akt), and apoptotic signaling pathways.

Signal transduction pathways

Signaling from G-protein coupled receptors (including many neuropeptide receptors, muscarinic receptors, and nearly all monoamine receptors) is propagated through multiple intracellular signaling pathways, including the MAPK pathway (316) and the PI3K/Akt/mTOR pathway (317). The MAPK pathway responds to a wide variety of extracellular signals to regulate many cellular processes such as differentiation, growth, and apoptosis (318). In neurons, the MAPK pathway appears to be an important regulator of neuroplasticity (319). Early investigations into the effects of SPS on the MAPK pathway found that SPS increases the phosphorylation (activation) of MAPK in the mPFC (171, 320, 321) and AMY (147, 322, 323). These studies suggest that MAPK signaling plays a role in coordinating transcriptional responses to SPS via activation of cFos (320), in regulating neuronal apoptosis (323), and in the behavioral effects of SPS including hyperalgesia (171) and increased anxiety-like behavior (147).

The PI3K/Akt/mTOR pathway is activated by a wide variety of receptors, and coordinates cellular responses such as cell growth, proliferation, and survival via signaling through the serine/threonine protein kinases Akt and mTOR (317). We reported that SPS increased levels of phosphorylation (activation) of Akt in the HC concurrently with increases in GR (184). Increased Akt phosphorylation has been linked to GR activation (324) which may contribute to cell death (325). While this coincides with research using other stress models that suggest that Akt/mTOR activation is involved in responses to traumatic stress (326328) another group has reported phasic decreases in Akt and mTOR phosphorylation following SPS (329). Further research is required to resolve this discrepancy and clarify the role of the PI3K/Akt/mTOR signaling pathway in adaptations following traumatic stress.

Both the MAPK and PI3K/Akt/mTOR pathways are involved in regulating synaptic plasticity and cellular morphology, particularly changes underlying fear learning (255). While relatively few studies using SPS have examined these effects, there is some indication that SPS produces morphological changes in neurons, specifically effects on soma size and dendrites. SPS has been observed to induce dendritic arborization in the BLA (247), changes which resemble changes in HC morphology observed following predator scent exposure, another traumatic stress model (314). Dendritic arborization may underlie neuronal plasticity and is controlled by diverse processes, including CORT response (330). Soma size and dendritic density were increased following SPS in vasopressin neurons in the hypothalamus, and these neurons had a blunted morphological response to an acute stressor (331). Neurotrophic signaling (e.g., through brain-derived neurotropic factor; BDNF) (332) has also been found to be altered by SPS, although results are inconsistent: some studies have shown that BDNF and components of its signaling pathway were upregulated following SPS (128, 147) while in another SPS appeared to decrease BDNF mRNA (168). There is also direct evidence that synaptic plasticity is altered by SPS. Specifically, hippocampal LTD is enhanced 1 day after SPS, and LTP in the HC and AMY is impaired 7 day after SPS (122). Additionally, SPS impacts cellular morphology. These findings suggest that changes in cellular physiology may contribute to the symptoms of PTSD, but do not provide a direct link between changes in signaling factors, cellular function, and behavioral changes. Further work is necessary to address this research gap, which will allow the investigation of interventions to prevent or reverse such changes in cellular plasticity following traumatic stress exposure.

Apoptotic pathways

In addition to changes in cellular morphology and electrophysiology, several studies have reported changes in histological and biochemical markers of apoptosis in brain tissue following SPS. Post-SPS apoptosis has been reported in up to 20% of AMY neurons (194, 333), in up to 45% of LC neurons (334), and in up to 20% of HC neurons (335). Multiple mechanisms to account for this observed loss of neurons have been proposed; these include oxidative-stress-induced loss of interneurons (123) and mitochondria-mediated apoptosis in the HC (313), multiple endoplasmic reticulum-based signaling pathways in the mPFC (321, 336), and caspase activation in the AMY (337). Some results disagree with these. For example, one study investigating changes in HC following SPS revealed no signs of neuronal death or morphological changes (122). Additionally, studies on post-SPS apoptosis so far do not routinely include measurements in brain areas that are not expected to show changes based on the specific pattern of post-SPS behavioral changes. If SPS (and potentially traumatic stress in general) causes apoptosis specifically in areas related to fear, anxiety, and stress-related behaviors, this may account for many behavioral changes seen in PTSD. On the other hand, these data may indicate that SPS produces widespread, non-specific apoptosis, which is unlikely to be mirrored in the human condition. While it is presently impossible to directly assess apoptosis in living patients with PTSD, post-mortem PFC samples from PTSD patients show dysregulation of cell survival and apoptosis-related genes (338), suggesting that apoptosis is abnormal in some way in people with PTSD. Stress-induced apoptosis may underlie differences in hippocampal volume seen in PTSD, but other phenomena could be at play; a study using mice showed that SPS causes HC volume loss along with a decrease in markers of axonal growth, indicating that reduced axonal size may underlie the change in HC size rather than changes in cell number or soma size (315). Further research on cell cycle regulation, neurogenesis, and apoptosis following SPS is important in resolving whether neuronal apoptosis following SPS is regionally specific or widespread, what its role in post-SPS behavioral changes is, and how it relates to human neurobiology.

Inflammation and glia

Neuroinflammation has been proposed to play a part in a diverse set of psychiatric conditions, including major depressive disorder, bipolar disorder, schizophrenia, and PTSD (339, 340). Results from studies investigating inflammatory biomarkers have found that C-reactive protein and NF-kB pathway components are related to PTSD, and can even predict symptom severity (341343). Inflammation may be related to the temporal course of PTSD, as inflammatory markers are only increased in those with current PTSD and return to baseline in individuals who have recovered (344). This is consistent with a wide body of research showing that GR signaling, which is altered in PTSD and following SPS, regulates immune function (345).

Following SPS, inflammation appears to be increased in the HC. SPS causes a consistent increase in IL-6 (123, 346) and a less consistent effect on TNFα, with one study showing increased TNFα after SPS (168) and another showing no change (123). An “enhanced SPS” protocol that includes an inescapable foot shock has been shown to increase IL-6, IL-2B, and iNOS levels in the HC, and these changes can be mitigated by administration of the NOS and COX-2 inhibitor gastrodin (346). Further supporting the hypothesis that inflammation plays a role in post-SPS behavioral changes, a 14-day treatment of the non-steroidal anti-inflammatory drug ibuprofen reversed SPS-induced increases in circulating CORT, anxiety-like behavior, and HC markers of inflammation (168).

There is also evidence that glial processes may contribute to post-SPS dysfunction and poor behavioral recovery. For example, microglial activation caused by glucocorticoid signaling in the spinal cord (288, 347) and HC (348) appears to contribute to post-SPS hyperalgesia. Immunohistochemical data indicate that glial cell number is decreased in the HC in a time-dependent manner following SPS, and magnetic resonance spectroscopy findings of decreased choline in the HC (312) and creatine in the HC (312) and mPFC (154) are consistent with a glial-related mechanism (312). Especially given the clinical association between PTSD and inflammatory disease, further research on the effects of SPS on inflammatory mediators and their relationship to post-SPS behavioral aberrations is needed in our understanding of the pathogenesis of PTSD.

Glia may participate in non-inflammatory changes following SPS, as well. Astroglia regulate processes important to neuronal function including synapse formation and GLU processing (349) and exhibit changes following other laboratory stressors (350). A post-SPS loss of cells expressing glial fibrillary acidic protein in the HC suggests a decrease in astroglial number(312), while administration of fibroblast growth factor 2, which regulates glial differentiation, has been found to mitigate the effects of SPS on contextual fear conditioning and unconditioned anxiety, possibly by restoring altered astrocytic GLU processing (130, 351). As astroglia actively regulate neuronal signaling, further research aimed at understanding changes in astroglial number, structure, and function following SPS may uncover mechanisms by which SPS alters neurotransmission.

It is unlikely that the intracellular factors heretofore investigated represent all factors involved in post-SPS adaptations. Discovery-focused approaches such as genomic and proteomic analyses have the potential to identify new molecules important in post-SPS adaptations. To date, only two studies report large-scale analyses of gene expression changes following SPS, including in the HC (119, 121) and ACC (119). While one study (121) found a set of differentially-expressed genes that were related to neurotransmission, the other (119) found that a diverse set of genes were affected, and the sets of genes reported to be differentially expressed following SPS in these two studies were non-overlapping. Further discovery-oriented research is necessary to clarify these results and identify other factors that may be involved in brain responses to traumatic stress.

Conclusion

The SPS model of exposure to traumatic stress demonstrates significant validity as an animal model of PTSD. Its content validity is demonstrated by its ability to produce a behavioral and neurohormonal syndrome that corresponds to many characteristics of PTSD, particularly reactivity to trauma cues, impairment of fear extinction and extinction retention, hyperarousal, changes in neurotransmitter and neuropeptide systems, and altered responsivity of the HPA axis. Its criterion validity is demonstrated by the efficacy of many clinically utilized drugs in ameliorating post-SPS behavioral changes, and in the general agreement of findings in SPS with findings from other traumatic stress models and human laboratory research.

There are of course limits to the scope and validity of the SPS model. Like all animal models, it cannot be used straightforwardly to answer hypotheses about experiential (rather than behavioral) symptoms such as nightmares, intrusive thoughts, or trauma-related guilt. Although it is a complex, multimodal stressor, SPS does not account for many of the complex features of trauma experienced by humans that may influence the development of PTSD, including varying characteristics of traumatic events, historical and social context, and post-trauma cognitive and emotional processing. Species-level differences between rodents and humans in neuroanatomy and the neurobiology of stress must be considered when translating findings from model to clinic. Additionally, there are many things that can be measured in laboratory animals that are unlikely to ever be corroborated with data from human patients; for example, invasive measurements of brain function and measurements taken immediately pre- and post-trauma are difficult, if not impossible, to obtain in humans. Like pre-clinical research on traumatic stress as a whole, research using the SPS model has not yet led to the development of new treatments for PTSD, and so it has not yet achieved the strongest demonstration of predictive validity. Finally, SPS (like any models of traumatic stress) is likely to correspond closely to only a subset of individuals with PTSD—particularly those whose trauma was isolated to a single event—and may not correspond closely to adaptations seen in individuals who have survived multiple traumatic events, or particularly complex traumas such as those involving interpersonal interactions or bodily injury.

Despite these limitations, the SPS model stands out as a well-characterized, translationally tractable paradigm to study of the neurobiology of traumatic stress-induced adaptations in brain and behavior. In concert with other pre-clinical and clinical approaches to the study of traumatic stress and PTSD, SPS holds great potential to expand both our basic knowledge of the mechanisms by which traumatic stress affects behavior and the clinical armamentarium with which we can help individuals with PTSD achieve a greater quality of life.

Author contributions

ML conducted the literature review and wrote the initial draft of this manuscript. AE, AC, IL, and SP advised ML in this process and provided revisions and critical feedback.

Conflict of interest statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Glossary

Abbreviations

5-HT

5-hydroxytryptamine (serotonin)

ACC

anterior cingulate

ACh

acetylcholine

ACTH

adrenocorticotropic hormone

Akt

protein kinase B

AMY

amygdala

AMPA

α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid

CB1

cannabinoid receptor, subtype 1

CeA

central amygdala

CSF

cerebrospinal fluid

CORT

corticosterone/cortisol

COX-2

cyclooxygenase-2

CRH

corticotropin releasing hormone

D2, dopamine receptor

subtype 2

DA

dopamine

DAT

dopamine transporter

DSM-5

Diagnostic and Statistical Manual of Mental Disorders, Fifth Edition

FDA

US Food and Drug Administration

GABA

gamma-aminobutyric acid

GAP43

growth associated protein 43

GLU

glutamate

GLY

glycine

GR

glucocorticoid receptor

HC

hippocampus

HPA

hypothalamus-pituitary-adrenal

IL-2B

interleukin 2B

IL-6

interleukin 6

ILC

infralimbic cortex

iNOS

calcium-insensitive nitic oxide synthase

LC

locus coeruleus

LTP

long-term potentiation

MAPK

mitogen activated protein kinase

mPFC

medial prefrontal cortex

MR

mineralocorticoid receptor

mTOR

mammalian target of rapamycin

NE

norepinephrine

NF-κB

nuclear factor κ-light-chain-enhancer of activated B cells

NOS

nitric oxide synthase

NPY

neuropeptide Y

PFC

prefrontal cortex

PI3K

phosphatidylinositide 3-kinase

PKC

protein kinase C

PSD-95

post-synaptic density 95

PTSD

post-traumatic stress disorder

SPS

single prolonged stress

SERT

5-HT transporter

SSRI

selective serotonin reuptake inhibitor

THC

Tetrahydrocannabinol

TNFα

tumor necrosis factor α

vmPFC

ventromedial prefrontal cortex.

Footnotes

Funding. This work was supported by the Wayne State MD/Ph.D. program (ML), the Wayne State Department of Psychiatry and Behavioral Neurosciences (ML, SP), NIDA R01 DA-042057 (SP), VA Merit Award RX001511 (ML, SP, and AC), and Department of Defense Award W81XWH-13-1-0377 (IL).

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpsyt.2018.00196/full#supplementary-material

References

  • 1.APA Diagnostic and Statistical Manual of Mental Disorders. DSM 5. Arlington: American Psychiatric Association; (2013). [Google Scholar]
  • 2.Wang PS, Lane M, Olfson M, Pincus HA, Wells KB, Kessler RC. Twelve-month use of mental health services in the United States: results from the National Comorbidity Survey Replication. Arch Gen Psychiatry (2005) 62:629–40. 10.1001/archpsyc.62.6.629 [DOI] [PubMed] [Google Scholar]
  • 3.Watts BV, Schnurr PP, Mayo L, Young-Xu Y, Weeks WB, Friedman MJ. Meta-analysis of the efficacy of treatments for posttraumatic stress disorder. J Clin Psychiatry (2013) 74:541–50. 10.4088/JCP.12r08225 [DOI] [PubMed] [Google Scholar]
  • 4.Foa EB, Zinbarg R, Rothbaum BO. Uncontrollability and unpredictability in post-traumatic stress disorder: an animal model. Psychol Bull. (1992) 112:218. 10.1037/0033-2909.112.2.218 [DOI] [PubMed] [Google Scholar]
  • 5.Johnson LR, McGuire J, Lazarus R, Palmer AA. Pavlovian fear memory circuits and phenotype models of PTSD. Neuropharmacology (2012) 62:638–46. 10.1016/j.neuropharm.2011.07.004 [DOI] [PubMed] [Google Scholar]
  • 6.Rau V, Fanselow MS. Exposure to a stressor produces a long lasting enhancement of fear learning in rats. Stress (2009) 12:125–33. 10.1080/10253890802137320 [DOI] [PubMed] [Google Scholar]
  • 7.Rau V, DeCola JP, Fanselow MS. Stress-induced enhancement of fear learning: an animal model of posttraumatic stress disorder. Neurosci Biobehav Rev. (2005) 29:1207–23. 10.1016/j.neubiorev.2005.04.010 [DOI] [PubMed] [Google Scholar]
  • 8.Clinchy M, Sheriff MJ, Zanette LY. Predator-induced stress and the ecology of fear. Funct Ecol. (2013) 27:56–65. 10.1111/1365-2435.12007 [DOI] [Google Scholar]
  • 9.Cohen H, Matar MA, Richter-Levin G, Zohar J. The contribution of an animal model toward uncovering biological risk factors for PTSD. Ann N Y Acad Sci. (2006) 1071:335–50. 10.1196/annals.1364.026 [DOI] [PubMed] [Google Scholar]
  • 10.Cohen H, Zohar J. An animal model of posttraumatic stress disorder: the use of cut-off behavioral criteria. Ann N Y Acad Sci. (2004) 1032:167–78. 10.1196/annals.1314.014 [DOI] [PubMed] [Google Scholar]
  • 11.Richter-Levin G. Acute and long-term behavioral correlates of underwater trauma—potential relevance to stress and post-stress syndromes. Psychiatry Res. (1998) 79:73–83. 10.1016/S0165-1781(98)00030-4 [DOI] [PubMed] [Google Scholar]
  • 12.Dal-Zotto S, Martí O, Armario A. Glucocorticoids are involved in the long-term effects of a single immobilization stress on the hypothalamic–pituitary–adrenal axis. Psychoneuroendocrinology (2003) 28:992–1009. 10.1016/S0306-4530(02)00120-8 [DOI] [PubMed] [Google Scholar]
  • 13.Armario A, Escorihuela RM, Nadal R. Long-term neuroendocrine and behavioural effects of a single exposure to stress in adult animals. Neurosci Biobehav Rev. (2008) 32:1121–35. 10.1016/j.neubiorev.2008.04.003 [DOI] [PubMed] [Google Scholar]
  • 14.Zoladz PR, Diamond DM. Predator-based psychosocial stress animal model of PTSD: Preclinical assessment of traumatic stress at cognitive, hormonal, pharmacological, cardiovascular and epigenetic levels of analysis. Exp Neurol. (2016) 284:211–9. 10.1016/j.expneurol.2016.06.003 [DOI] [PubMed] [Google Scholar]
  • 15.Berardi A, Trezza V, Palmery M, Trabace L, Cuomo V, Campolongo P. An updated animal model capturing both the cognitive and emotional features of post-traumatic stress disorder (PTSD). Front Behav Neurosci. (2014) 8:142. 10.3389/fnbeh.2014.00142 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Borghans B, Homberg JR. Animal models for posttraumatic stress disorder: an overview of what is used in research. World J Psychiatry (2015) 5:387–96. 10.5498/wjp.v5.i4.387 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Goswami S, Rodríguez-Sierra O, Cascardi M, Paré D. Animal models of post-traumatic stress disorder: face validity. Front Neurosci. (2013) 7:89. 10.3389/fnins.2013.00089 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Daskalakis NP, Yehuda R, Diamond DM. Animal models in translational studies of PTSD. Psychoneuroendocrinology (2013) 38:1895–911. 10.1016/j.psyneuen.2013.06.006 [DOI] [PubMed] [Google Scholar]
  • 19.Daskalakis NP, Yehuda R. Principles for developing animal models of military PTSD. Eur J Psychotraumatol. (2014) 5. 10.3402/ejpt.v5.23825 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Matar MA, Zohar J, Cohen H. Translationally relevant modeling of PTSD in rodents. Cell Tissue Res. (2013) 354:127–39. 10.1007/s00441-013-1687-6 [DOI] [PubMed] [Google Scholar]
  • 21.Whitaker AM, Gilpin NW, Edwards S. Animal models of post-traumatic stress disorder and recent neurobiological insights. Behav Pharmacol. (2014) 25:398–408. 10.1097/FBP.0000000000000069 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Berardi A, Trezza V, Campolongo P. Modeling specific phobias and posttraumatic stress disorder in rodents: the challenge to convey both cognitive and emotional features. Rev Neurosci. (2012) 23:645–57. 10.1515/revneuro-2012-0054 [DOI] [PubMed] [Google Scholar]
  • 23.Krause ED, Kaltman S, Goodman LA, Dutton MA. Longitudinal factor structure of posttraumatic stress symptoms related to intimate partner violence. Psychol Assess. (2007) 19:165. 10.1037/1040-3590.19.2.165 [DOI] [PubMed] [Google Scholar]
  • 24.Rasmussen A, Smith H, Keller AS. Factor structure of PTSD symptoms among west and central African refugees. J Traum Stress. (2007) 20:271–80. 10.1002/jts.20208 [DOI] [PubMed] [Google Scholar]
  • 25.Demirchyan A, Goenjian AK, Khachadourian V. Factor Structure and psychometric properties of the Posttraumatic Stress Disorder (PTSD) checklist and DSM-5 PTSD symptom set in a long-term postearthquake cohort in armenia. Assessment (2015) 22:594–606. 10.1177/1073191114555523 [DOI] [PubMed] [Google Scholar]
  • 26.NIMH RDoC Matrix 2017. Available online at: https://www.nimh.nih.gov/research-priorities/rdoc/constructs/rdoc-matrix.shtml (Accessed May 13, 2017).
  • 27.Aikins DE, Jackson ED, Christensen A, Walderhaug E, Afroz S, Neumeister A. Differential conditioned fear response predicts duloxetine treatment outcome in male veterans with PTSD: a pilot study. Psychiatry Res. (2011) 188:453–5. 10.1016/j.psychres.2011.05.009 [DOI] [PubMed] [Google Scholar]
  • 28.Milad MR, Orr SP, Lasko NB, Chang Y, Rauch SL, Pitman RK. Presence and acquired origin of reduced recall for fear extinction in PTSD: results of a twin study. J Psychiatr Res. (2008) 42:515–20. 10.1016/j.jpsychires.2008.01.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Milad MR, Pitman RK, Ellis CB, Gold AL, Shin LM, Lasko NB, et al. Neurobiological basis of failure to recall extinction memory in posttraumatic stress disorder. Biol Psychiatry (2009) 66:1075–82. 10.1016/j.biopsych.2009.06.026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Aikins DE, Strader JA, Kohler RJ, Bihani N, Perrine SA. Differences in hippocampal serotonergic activity in a mouse single prolonged stress paradigm impact discriminant fear acquisition and retention. Neurosci Lett. (2017) 639:162–6. 10.1016/j.neulet.2016.12.056 [DOI] [PubMed] [Google Scholar]
  • 31.Garfinkel SN, Abelson JL, King AP, Sripada RK, Wang X, Gaines LM, et al. Impaired contextual modulation of memories in PTSD: an fMRI and psychophysiological study of extinction retention and fear renewal. J Neurosci. (2014) 34:13435–43. 10.1523/JNEUROSCI.4287-13.2014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Karl A, Schaefer M, Malta LS, Dorfel D, Rohleder N, Werner A. A meta-analysis of structural brain abnormalities in PTSD. Neurosci Biobehav Rev. (2006) 30:1004–31. 10.1016/j.neubiorev.2006.03.004 [DOI] [PubMed] [Google Scholar]
  • 33.Campanella C, Bremner JD. Neuroimaging of PTSD. In: Posttraumatic Stress Disorder, ed Bremner JD. Hoboken, NJ: John Wiley & Sons, Inc. (2016). p. 291–318. [Google Scholar]
  • 34.Tovote P, Fadok JP, Luthi A. Neuronal circuits for fear and anxiety. Nat Rev Neurosci. (2015) 16:317–31. 10.1038/nrn3945 [DOI] [PubMed] [Google Scholar]
  • 35.Packard AE, Egan AE, Ulrich-Lai YM. HPA axis interactions with behavioral systems. Comprehens Physiol. (2016) 6:1897–934. 10.1002/cphy.c150042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Daskalakis NP, Cohen H, Nievergelt CM, Baker DG, Buxbaum JD, Russo SJ, et al. New translational perspectives for blood-based biomarkers of PTSD: from glucocorticoid to immune mediators of stress susceptibility. Exp Neurol. (2016) 284:133–40. 10.1016/j.expneurol.2016.07.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Squire LR. Memory and the hippocampus: a synthesis from findings with rats, monkeys, and humans. Psychol Rev. (1992) 99:195. 10.1037/0033-295X.99.2.195 [DOI] [PubMed] [Google Scholar]
  • 38.Smith DM, Bulkin DA. The form and function of hippocampal context representations. Neurosci Biobehav Rev. (2014) 40:52–61. 10.1016/j.neubiorev.2014.01.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Megías JL, Ryan E, Vaquero JM, Frese B. Comparisons of traumatic and positive memories in people with and without PTSD profile. Appl Cognit Psychol. (2007) 21:117–30. 10.1002/acp.1282 [DOI] [Google Scholar]
  • 40.Johnsen GE, Asbjørnsen AE. Consistent impaired verbal memory in PTSD: a meta-analysis. J Affect Disord. (2008) 111:74–82. 10.1016/j.jad.2008.02.007 [DOI] [PubMed] [Google Scholar]
  • 41.Johnsen GE, Asbjørnsen AE. Verbal learning and memory impairments in posttraumatic stress disorder: the role of encoding strategies. Psychiatry Res. (2009) 165:68–77. 10.1016/j.psychres.2008.01.001 [DOI] [PubMed] [Google Scholar]
  • 42.Scott JC, Matt GE, Wrocklage KM, Crnich C, Jordan J, Southwick SM, et al. A quantitative meta-analysis of neurocognitive functioning in posttraumatic stress disorder. Psychol Bull. (2015) 141:105–40. 10.1037/a0038039 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Levy-Gigi E, Szabo C, Richter-Levin G, Kéri S. Reduced hippocampal volume is associated with overgeneralization of negative context in individuals with PTSD. Neuropsychology (2015) 29:151. 10.1037/neu0000131 [DOI] [PubMed] [Google Scholar]
  • 44.Fanselow MS, Dong HW. Are the dorsal and ventral hippocampus functionally distinct structures? Neuron (2010) 65:7–19. 10.1016/j.neuron.2009.11.031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Woon FL, Sood S, Hedges DW. Hippocampal volume deficits associated with exposure to psychological trauma and posttraumatic stress disorder in adults: a meta-analysis. Progress Neuro Psychopharmacol Biol Psychiatry (2010) 34:1181–8. 10.1016/j.pnpbp.2010.06.016 [DOI] [PubMed] [Google Scholar]
  • 46.Bootsman F, Brouwer RM, Schnack HG, Kemner SM, Hillegers MH, Sarkisyan G, et al. A study of genetic and environmental contributions to structural brain changes over time in twins concordant and discordant for bipolar disorder. J Psychiatric Res. (2016) 79:116–24. 10.1016/j.jpsychires.2016.04.011 [DOI] [PubMed] [Google Scholar]
  • 47.Gilbertson MW, Shenton ME, Ciszewski A, Kasai K, Lasko NB, Orr SP, et al. Smaller hippocampal volume predicts pathologic vulnerability to psychological trauma. Nat Neurosci. (2002) 5:1242–7. 10.1038/nn958 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Levy-Gigi E, Szabó C, Kelemen O, Kéri S. Association among clinical response, hippocampal volume, and FKBP5 gene expression in individuals with posttraumatic stress disorder receiving cognitive behavioral therapy. Biol Psychiatry (2013) 74:793–800. 10.1016/j.biopsych.2013.05.017 [DOI] [PubMed] [Google Scholar]
  • 49.Patel R, Spreng RN, Shin LM, Girard TA. Neurocircuitry models of posttraumatic stress disorder and beyond: a meta-analysis of functional neuroimaging studies. Neurosci Biobehav Rev. (2012) 36:2130–42. 10.1016/j.neubiorev.2012.06.003 [DOI] [PubMed] [Google Scholar]
  • 50.Croxson PL, Johansen-Berg H, Behrens TEJ, Robson MD, Pinsk MA, Gross CG, et al. Quantitative investigation of connections of the prefrontal cortex in the human and macaque using probabilistic diffusion tractography. J Neurosci. (2005) 25:8854–66. 10.1523/JNEUROSCI.1311-05.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Cohen MX. Hippocampal-prefrontal connectivity predicts midfrontal oscillations and long-term memory performance. Curr Biol. (2011) 21:1900–5. 10.1016/j.cub.2011.09.036 [DOI] [PubMed] [Google Scholar]
  • 52.Fastenrath M, Coynel D, Spalek K, Milnik A. Dynamic modulation of amygdala-hippocampal connectivity by emotional arousal. J Neurosci (2014) 34:13935–47. 10.1523/JNEUROSCI.0786-14.2014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Hermans EJ, Kanen JW, Tambini A, Fernández G, Davachi L, Phelps EA. Persistence of amygdala–hippocampal connectivity and multi-voxel correlation structures during awake rest after fear learning predicts long-term expression of fear. Cereb Cortex (2017) 27:3028–41. 10.1093/cercor/bhw145 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Chen AC, Etkin A. Hippocampal network connectivity and activation differentiates post-traumatic stress disorder from generalized anxiety disorder. Neuropsychopharmacology (2013) 38:1889–98. 10.1038/npp.2013.122 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Dunkley BT, Doesburg SM, Sedge PA, Grodecki RJ, Shek PN, Pang EW, et al. Resting-state hippocampal connectivity correlates with symptom severity in post-traumatic stress disorder. NeuroImage Clin. (2014) 5:377–84. 10.1016/j.nicl.2014.07.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Dixon ML, Thiruchselvam R, Todd R, Christoff K. Emotion and the prefrontal cortex: an integrative review. Psychol Bull. (2017) 143:1033–81. 10.1037/bul0000096 [DOI] [PubMed] [Google Scholar]
  • 57.Yuan P, Raz N. Prefrontal cortex and executive functions in healthy adults: a meta-analysis of structural neuroimaging studies. Neurosci Biobehav Rev. (2014) 42:180–92. 10.1016/j.neubiorev.2014.02.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Gilmartin MR, Balderston NL, Helmstetter FJ. Prefrontal cortical regulation of fear learning. Trends Neurosci. (2014) 37:455–64. 10.1016/j.tins.2014.05.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Pitman RK, Rasmusson AM, Koenen KC, Shin LM, Orr SP, Gilbertson MW, et al. Biological studies of post-traumatic stress disorder. Nat Rev Neurosci. (2012) 13:769–87. 10.1038/nrn3339 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Garfinkel SN, Liberzon I. Neurobiology of PTSD: a review of neuroimaging findings. Psychiatric Ann. (2009) 39:370–81. 10.3928/00485713-20090527-01 [DOI] [Google Scholar]
  • 61.Harrison BJ, Fullana MA, Via E, Soriano-Mas C, Vervliet B, Martínez-Zalacaín I, et al. Human ventromedial prefrontal cortex and the positive affective processing of safety signals. NeuroImage (2017) 152:12–8. 10.1016/j.neuroimage.2017.02.080 [DOI] [PubMed] [Google Scholar]
  • 62.Giustino TF, Maren S. The role of the medial prefrontal cortex in the conditioning and extinction of fear. Front Behav Neurosci. (2015) 9:298. 10.3389/fnbeh.2015.00298 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Etkin A, Egner T, Kalisch R. Emotional processing in anterior cingulate and medial prefrontal cortex. Trends Cogn Sci. (2011) 15:85–93. 10.1016/j.tics.2010.11.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Helpman L, Marin M-F, Papini S, Zhu X, Sullivan GM, Schneier F, et al. Neural changes in extinction recall following prolonged exposure treatment for PTSD: a longitudinal fMRI study. NeuroImage (2016) 12:715–23. 10.1016/j.nicl.2016.10.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Duvarci S, Pare D. Amygdala microcircuits controlling learned fear. Neuron (2014) 82:966–80. 10.1016/j.neuron.2014.04.042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Asede D, Bosch D, Luthi A, Ferraguti F, Ehrlich I. Sensory inputs to intercalated cells provide fear-learning modulated inhibition to the basolateral amygdala. Neuron (2015) 86:541–54. 10.1016/j.neuron.2015.03.008 [DOI] [PubMed] [Google Scholar]
  • 67.Rauch SL, Whalen PJ, Shin LM, McInerney SC, Macklin ML, Lasko NB, et al. Exaggerated amygdala response to masked facial stimuli in posttraumatic stress disorder: a functional MRI study. Biol Psychiatry (2000) 47:769–76. 10.1016/S0006-3223(00)00828-3 [DOI] [PubMed] [Google Scholar]
  • 68.Bryant R, Felmingham K, Kemp A, Das P, Hughes G, Peduto A, et al. Amygdala and ventral anterior cingulate activation predicts treatment response to cognitive behaviour therapy for post-traumatic stress disorder. Psychol Med. (2008) 38:555. 10.1017/S0033291707002231 [DOI] [PubMed] [Google Scholar]
  • 69.Stark EA, Parsons C, Van Hartevelt T, Charquero-Ballester M, McManners H, Ehlers A, et al. Post-traumatic stress influences the brain even in the absence of symptoms: a systematic, quantitative meta-analysis of neuroimaging studies. Neurosci Biobehav Rev. (2015) 56:207–21. 10.1016/j.neubiorev.2015.07.007 [DOI] [PubMed] [Google Scholar]
  • 70.Uddin LQ, Kinnison J, Pessoa L, Anderson ML. Beyond the tripartite cognition–emotion–interoception model of the human insular cortex. J Cogn Neurosci. (2014) 26:16–27. 10.1162/jocn_a_00462 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Butler RW, Braff DL, Rausch JL, Jenkins MA, Sprock J, Geyer MA. Physiological evidence of exaggerated startle response in a subgroup of Vietnam veterans with combat-related PTSD. Am J Psychiatry (1990) 147:1308–12. 10.1176/ajp.147.10.1308 [DOI] [PubMed] [Google Scholar]
  • 72.Metzger LJ, Orr SP, Lasko NB, Berry NJ, Pitman RK. Evidence for diminished P3 amplitudes in PTSDa. Ann N Y Acad Sci. (1997) 821:499–503. 10.1111/j.1749-6632.1997.tb48315.x [DOI] [PubMed] [Google Scholar]
  • 73.Nieuwenhuis S, Aston-Jones G, Cohen JD. Decision making, the P3, and the locus coeruleus–norepinephrine system. Psychol Bull. (2005) 131:510–32. 10.1037/0033-2909.131.4.510 [DOI] [PubMed] [Google Scholar]
  • 74.Nawijn L, van Zuiden M, Frijling JL, Koch SBJ, Veltman DJ, Olff M. Reward functioning in PTSD: a systematic review exploring the mechanisms underlying anhedonia. Neurosci Biobehav Rev. (2015) 51:189–204. 10.1016/j.neubiorev.2015.01.019 [DOI] [PubMed] [Google Scholar]
  • 75.Elman I, Lowen S, Frederick BB, Chi W, Becerra L, Pitman RK. Functional neuroimaging of reward circuitry responsivity to monetary gains and losses in posttraumatic stress disorder. Biol Psychiatry (2009) 66:1083–90. 10.1016/j.biopsych.2009.06.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Felmingham KL, Falconer EM, Williams L, Kemp AH, Allen A, Peduto A, et al. Reduced amygdala and ventral striatal activity to happy faces in PTSD is associated with emotional numbing. PLOS ONE (2014) 9:e103653. 10.1371/journal.pone.0103653 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Brady KT, Killeen TK, Brewerton T, Lucerini S. Comorbidity of psychiatric disorders and posttraumatic stress disorder. J Clin Psychiatry (2000) 61(Suppl. 7): 22–32. Available online at: http://www.psychiatrist.com/jcp/article/pages/2000/v61s07/v61s0704.aspx [PubMed] [Google Scholar]
  • 78.Mitchell KS, Wolf EJ. PTSD, food addiction, and disordered eating in a sample of primarily older veterans: the mediating role of emotion regulation. Psychiatry Res. (2016) 243:23–9. 10.1016/j.psychres.2016.06.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Hsieh Y-P, Shen AC-T, Wei H-S, Feng J-Y, Huang SC-Y, Hwa H-L. Associations between child maltreatment, PTSD, and internet addiction among Taiwanese students. Comput Hum Behav. (2016) 56:209–14. 10.1016/j.chb.2015.11.048 [DOI] [Google Scholar]
  • 80.Campbell DG, Felker BL, Liu C-F, Yano EM, Kirchner JE, Chan D, et al. Prevalence of depression–PTSD comorbidity: implications for clinical practice guidelines and primary care-based interventions. J Gen Intern Med. (2007) 22:711–8. 10.1007/s11606-006-0101-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.James LM, Strom TQ, Leskela J. Risk-taking behaviors and impulsivity among veterans with and without PTSD and mild TBI. Mil Med. (2014) 179:357–63. 10.7205/MILMED-D-13-00241 [DOI] [PubMed] [Google Scholar]
  • 82.Ruth Pat-Horenczyk PD, Osnat Peled MA, Tomer Miron MA, Daniel Brom PD, Yael Villa PD, Claude M, et al. Risk-taking behaviors among israeli adolescents exposed to recurrent terrorism: provoking danger under continuous threat? Am J Psychiatry (2007) 164:66–72. 10.1176/ajp.2007.164.1.66 [DOI] [PubMed] [Google Scholar]
  • 83.Tsigos C, Chrousos GP. Hypothalamic–pituitary–adrenal axis, neuroendocrine factors and stress. J Psychos Res. (2002) 53:865–71. 10.1016/S0022-3999(02)00429-4 [DOI] [PubMed] [Google Scholar]
  • 84.Penzo MA, Robert V, Tucciarone J, De Bundel D, Wang M, Van Aelst L, et al. The paraventricular thalamus controls a central amygdala fear circuit. Nature (2015) 519:455–9. 10.1038/nature13978 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Gray TS, Carney ME, Magnuson DJ. Direct projections from the central amygdaloid nucleus to the hypothalamic paraventricular nucleus: possible role in stress-induced adrenocorticotropin release. Neuroendocrinology (1989) 50:433–46. 10.1159/000125260 [DOI] [PubMed] [Google Scholar]
  • 86.Herman J, McKlveen J, Solomon M, Carvalho-Netto E, Myers B. Neural regulation of the stress response: glucocorticoid feedback mechanisms. Brazil J Med Biol Res. (2012) 45:292–8. 10.1590/S0100-879X2012007500041 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Harris A, Seckl J. Glucocorticoids, prenatal stress and the programming of disease. Horm Behav. (2011) 59:279–89. 10.1016/j.yhbeh.2010.06.007 [DOI] [PubMed] [Google Scholar]
  • 88.Harris AP, Holmes MC, de Kloet ER, Chapman KE, Seckl JR. Mineralocorticoid and glucocorticoid receptor balance in control of HPA axis and behaviour. Psychoneuroendocrinology (2013) 38:648–58. 10.1016/j.psyneuen.2012.08.007 [DOI] [PubMed] [Google Scholar]
  • 89.Liberzon I, Young EA. Effects of stress and glucocorticoids on CNS oxytocin receptor binding. Psychoneuroendocrinology (1997) 22:411–22. 10.1016/S0306-4530(97)00045-0 [DOI] [PubMed] [Google Scholar]
  • 90.Perrine SA, Eagle AL, George SA, Mulo K, Kohler RJ, Gerard J, et al. Severe, multimodal stress exposure induces PTSD-like characteristics in a mouse model of single prolonged stress. Behav Brain Res. (2016) 303:228–37. 10.1016/j.bbr.2016.01.056 [DOI] [PubMed] [Google Scholar]
  • 91.Yehuda R, Golier JA, Halligan SL, Meaney M, Bierer LM. The ACTH response to dexamethasone in PTSD. Am J Psychiatry (2004) 161:1397–403. 10.1176/appi.ajp.161.8.1397 [DOI] [PubMed] [Google Scholar]
  • 92.Yehuda R, Halligan SL, Golier JA, Grossman R, Bierer LM. Effects of trauma exposure on the cortisol response to dexamethasone administration in PTSD and major depressive disorder. Psychoneuroendocrinology (2004) 29:389–404. 10.1016/S0306-4530(03)00052-0 [DOI] [PubMed] [Google Scholar]
  • 93.Yehuda R, Yang R-K, Buchsbaum MS, Golier JA. Alterations in cortisol negative feedback inhibition as examined using the ACTH response to cortisol administration in PTSD. Psychoneuroendocrinology (2006) 31:447–51. 10.1016/j.psyneuen.2005.10.007 [DOI] [PubMed] [Google Scholar]
  • 94.Koren D, Arnon I, Klein E. Acute stress response and posttraumatic stress disorder in traffic accident victims: a one-year prospective, follow-up study. Am J Psychiatry (1999) 156:367–73. 10.1176/ajp.156.3.367 [DOI] [PubMed] [Google Scholar]
  • 95.O'Donnell ML, Elliott P, Lau W, Creamer M. PTSD symptom trajectories: from early to chronic response. Behav Res Ther. (2007) 45:601–6. 10.1016/j.brat.2006.03.015 [DOI] [PubMed] [Google Scholar]
  • 96.Knox D, Nault T, Henderson C, Liberzon I. Glucocorticoid receptors and extinction retention deficits in the single prolonged stress model. Neuroscience (2012) 223:163–73. 10.1016/j.neuroscience.2012.07.047 [DOI] [PubMed] [Google Scholar]
  • 97.Yamamoto S, Morinobu S, Takei S, Fuchikami M, Matsuki A, Yamawaki S, et al. Single prolonged stress: toward an animal model of posttraumatic stress disorder. Depress Anxiety (2009) 26:1110–7. 10.1002/da.20629 [DOI] [PubMed] [Google Scholar]
  • 98.Foa EB, Steketee G, Rothbaum BO. Behavioral/cognitive conceptualizations of post-traumatic stress disorder. Behav Ther. (1989) 20:155–76. 10.1016/S0005-7894(89)80067-X [DOI] [Google Scholar]
  • 99.Parsons RG, Ressler KJ. Implications of memory modulation for post-traumatic stress and fear disorders. Nat Neurosci. (2013) 16:146–53. 10.1038/nn.3296 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Souza RR, Noble LJ, McIntyre CK. Using the single prolonged stress model to examine the pathophysiology of PTSD. Front Pharmacol. (2017) 8:615. 10.3389/fphar.2017.00615 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Graham BM, Milad MR. Fear conditioning and extinction. In: Neurophenotypes, ed Stein DJ. Berlin: Springer; (2016). p. 139–55. [Google Scholar]
  • 102.Shvil E, Rusch HL, Sullivan GM, Neria Y. Neural, psychophysiological, and behavioral markers of fear processing in PTSD: a review of the literature. Curr Psychiatry Rep. (2013) 15:358. 10.1007/s11920-013-0358-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Norrholm SD, Jovanovic T, Gerardi M, Breazeale KG, Price M, Davis M, et al. Baseline psychophysiological and cortisol reactivity as a predictor of PTSD treatment outcome in virtual reality exposure therapy. Behav Res Ther. (2016) 82:28–37. 10.1016/j.brat.2016.05.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Blanchard EB, Hickling EJ, Veazey CH, Buckley TC, Freidenberg BM, Walsh JD, et al. Treatment-related changes in cardiovascular reactivity to trauma cues in motor vehicle accident-related PTSD. Behav Ther. (2002) 33:417–26. 10.1016/S0005-7894(02)80036-3 [DOI] [Google Scholar]
  • 105.Toledano D, Gisquet-Verrier P. Only susceptible rats exposed to a model of PTSD exhibit reactivity to trauma-related cues and other symptoms: an effect abolished by a single amphetamine injection. Behav Brain Res. (2014) 272:165–74. 10.1016/j.bbr.2014.06.039 [DOI] [PubMed] [Google Scholar]
  • 106.Le Dorze C, Gisquet-Verrier P. Sensitivity to trauma-associated cues is restricted to vulnerable traumatized rats and reinstated after extinction by yohimbine. Behav Brain Res. (2016) 313:120–34. 10.1016/j.bbr.2016.07.006 [DOI] [PubMed] [Google Scholar]
  • 107.Blechert J, Michael T, Vriends N, Margraf J, Wilhelm FH. Fear conditioning in posttraumatic stress disorder: evidence for delayed extinction of autonomic, experiential, and behavioural responses. Behav Res Ther. (2007) 45:2019–33. 10.1016/j.brat.2007.02.012 [DOI] [PubMed] [Google Scholar]
  • 108.Guthrie RM, Bryant RA. Extinction learning before trauma and subsequent posttraumatic stress. Psychosomatic Med. (2006) 68:307–11. 10.1097/01.psy.0000208629.67653.cc [DOI] [PubMed] [Google Scholar]
  • 109.Keller SM, Schreiber WB, Staib JM, Knox D. Sex differences in the single prolonged stress model. Behav Brain Res. (2015) 286:29–32. 10.1016/j.bbr.2015.02.034 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Keller SM, Schreiber WB, Stanfield BR, Knox D. Inhibiting corticosterone synthesis during fear memory formation exacerbates cued fear extinction memory deficits within the single prolonged stress model. Behav Brain Res. (2015) 287:182–6. 10.1016/j.bbr.2015.03.043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Knox D, George SA, Fitzpatrick CJ, Rabinak CA, Maren S, Liberzon I. Single prolonged stress disrupts retention of extinguished fear in rats. Learn Memory (2012) 19:43–9. 10.1101/lm.024356.111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.George SA, Rodriguez-Santiago M, Riley J, Rodriguez E, Liberzon I. The effect of chronic phenytoin administration on single prolonged stress induced extinction retention deficits and glucocorticoid upregulation in the rat medial prefrontal cortex. Psychopharmacology (2015) 232:47–56. 10.1007/s00213-014-3635-x [DOI] [PubMed] [Google Scholar]
  • 113.Vanderheyden WM, George SA, Urpa L, Kehoe M, Liberzon I, Poe GR. Sleep alterations following exposure to stress predict fear-associated memory impairments in a rodent model of PTSD. Exp Brain Res. (2015) 233:2335–46. 10.1007/s00221-015-4302-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Lin C-C, Tung C-S, Liu Y-P. Escitalopram reversed the traumatic stress-induced depressed and anxiety-like symptoms but not the deficits of fear memory. Psychopharmacology (2016) 233:1135–46. 10.1007/s00213-015-4194-5 [DOI] [PubMed] [Google Scholar]
  • 115.Eskandarian S, Vafaei A, Vaezi GH, Taherian F, Kashefi A, Rashidy-Pour A. Effects of systemic administration of oxytocin on contextual fear extinction in a rat model of post-traumatic stress disorder. Basic Clin Neurosci. (2013) 4:315–22. Available online at: http://bcn.iums.ac.ir/article-1-435-en.html [PMC free article] [PubMed] [Google Scholar]
  • 116.Ganon-Elazar E, Akirav I. Cannabinoids and traumatic stress modulation of contextual fear extinction and GR expression in the amygdala-hippocampal-prefrontal circuit. Psychoneuroendocrinology (2013) 38:1675–87. 10.1016/j.psyneuen.2013.01.014 [DOI] [PubMed] [Google Scholar]
  • 117.Yamamoto S, Morinobu S, Fuchikami M, Kurata A, Kozuru T, Yamawaki S. Effects of single prolonged stress and D-cycloserine on contextual fear extinction and hippocampal NMDA receptor expression in a rat model of PTSD. Neuropsychopharmacology (2008) 33:2108–16. 10.1038/sj.npp.1301605 [DOI] [PubMed] [Google Scholar]
  • 118.Knox D, Stanfield BR, Staib JM, David NP, Keller SM, DePietro T. Neural circuits via which single prolonged stress exposure leads to fear extinction retention deficits. Learn Memory (2016) 23:689–98. 10.1101/lm.043141.116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Harada K, Yamaji T, Matsuoka N. Activation of the serotonin 5-HT 2C receptor is involved in the enhanced anxiety in rats after single-prolonged stress. Pharmacol Biochem Behav. (2008) 89:11–6. 10.1016/j.pbb.2007.10.016 [DOI] [PubMed] [Google Scholar]
  • 120.Imanaka A, Morinobu S, Toki S, Yamawaki S. Importance of early environment in the development of post-traumatic stress disorder-like behaviors. Behav Brain Res. (2006) 173:129–37. 10.1016/j.bbr.2006.06.012 [DOI] [PubMed] [Google Scholar]
  • 121.Iwamoto Y, Morinobu S, Takahashi T, Yamawaki S. Single prolonged stress increases contextual freezing and the expression of glycine transporter 1 and vesicle-associated membrane protein 2 mRNA in the hippocampus of rats. Prog Neuropsychopharmacol Biol Psychiatry (2007) 31:642–51. 10.1016/j.pnpbp.2006.12.010 [DOI] [PubMed] [Google Scholar]
  • 122.Kohda K, Harada K, Kato K, Hoshino A, Motohashi J, Yamaji T, et al. Glucocorticoid receptor activation is involved in producing abnormal phenotypes of single-prolonged stress rats: a putative post-traumatic stress disorder model. Neuroscience (2007) 148:22–33. 10.1016/j.neuroscience.2007.05.041 [DOI] [PubMed] [Google Scholar]
  • 123.Liu F-F, Yang L-D, Sun X-R, Zhang H, Pan W, Wang X-M, et al. NOX2 mediated-parvalbumin interneuron loss might contribute to anxiety-like and enhanced fear learning behavior in a rat model of post-traumatic stress disorder. Mol Neurobiol. (2015) 53:6680–9. 10.1007/s12035-015-9571-x [DOI] [PubMed] [Google Scholar]
  • 124.Miao Y-L, Guo W-Z, Shi W-Z, Fang W-W, Liu Y, Liu J, et al. Midazolam ameliorates the behavior deficits of a rat posttraumatic stress disorder model through dual 18 kDa translocator protein and central benzodiazepine receptor and neurosteroidogenesis. PLoS ONE (2014) 9:e101450. 10.1371/journal.pone.0101450 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Mirshekar M, Abrari K, Goudarzi I, Rashidy-Pour A. Effects of β-estradiol on enhanced conditioned fear induced by single prolonged stress and shock in rats. Basic Clin Neurosci. (2012) 3:5–11. 10.1007/s00213-006-0545-6 [DOI] [Google Scholar]
  • 126.Saffari S, Abrari K, Rezaei A, Rashidy-Pour A, Goudarzi I, Salmani ME. Correlation of fear memory in a PTSD animal model and hippocampal BDNF in response to β-estradiol treatment. J Paramed Sci. (2015) 6:22–34. 10.22037/jps.v6i3.9766 [DOI] [Google Scholar]
  • 127.Takahashi T, Morinobu S, Iwamoto Y, Yamawaki S. Effect of paroxetine on enhanced contextual fear induced by single prolonged stress in rats. Psychopharmacology (2006) 189:165–73. 10.1007/s00213-006-0545-6 [DOI] [PubMed] [Google Scholar]
  • 128.Takei S, Morinobu S, Yamamoto S, Fuchikami M, Matsumoto T, Yamawaki S. Enhanced hippocampal BDNF/TrkB signaling in response to fear conditioning in an animal model of posttraumatic stress disorder. J Psychiatr Res. (2011) 45:460–8. 10.1016/j.jpsychires.2010.08.009 [DOI] [PubMed] [Google Scholar]
  • 129.Wang H, Zuo D, He B, Qiao F, Zhao M, Wu Y. Conditioned fear stress combined with single-prolonged stress: a new PTSD mouse model. Neurosci Res. (2012) 73:142–52. 10.1016/j.neures.2012.03.003 [DOI] [PubMed] [Google Scholar]
  • 130.Feng D, Guo B, Liu G, Wang B, Wang W, Gao G, et al. FGF2 alleviates PTSD symptoms in rats by restoring GLAST function in astrocytes via the JAK/STAT pathway. Eur Neuropsychopharmacol. (2015) 25:1287–99. 10.1016/j.euroneuro.2015.04.020 [DOI] [PubMed] [Google Scholar]
  • 131.Santiago PN, Ursano RJ, Gray CL, Pynoos RS, Spiegel D, Lewis-Fernandez R, et al. A systematic review of PTSD prevalence and trajectories in DSM-5 defined trauma exposed populations: intentional and non-intentional traumatic events. PLoS ONE (2013) 8:e59236. 10.1371/journal.pone.0059236 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Brand L, Groenewald I, Stein DJ, Wegener G, Harvey BH. Stress and re-stress increases conditioned taste aversion learning in rats: possible frontal cortical and hippocampal muscarinic receptor involvement. Eur J Pharmacol. (2008) 586:205–11. 10.1016/j.ejphar.2008.03.004 [DOI] [PubMed] [Google Scholar]
  • 133.Jovanovic T, Kazama A, Bachevalier J, Davis M. Impaired safety signal learning may be a biomarker of PTSD. Neuropharmacology (2012) 62:695–704. 10.1016/j.neuropharm.2011.02.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Grillon C. Startle reactivity and anxiety disorders: aversive conditioning, context, and neurobiology. Biol Psychiatry (2002) 52:958–75. 10.1016/S0006-3223(02)01665-7 [DOI] [PubMed] [Google Scholar]
  • 135.Ornitz EM, Pynoos RS. Startle modulation in children with posttraumatic stress disorder. Am J Psychiatry (1989) 146:866–70. 10.1176/ajp.146.7.866 [DOI] [PubMed] [Google Scholar]
  • 136.Vrana SR, Calhoun PS, McClernon FJ, Dennis MF, Lee ST, Beckham JC. Effects of smoking on the acoustic startle response and prepulse inhibition in smokers with and without posttraumatic stress disorder. Psychopharmacology (2013) 230:477–85. 10.1007/s00213-013-3181-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Schmidt U, Kaltwasser SF, Wotjak CT. Biomarkers in posttraumatic stress disorder: Overview and implications for future research. Dis Mark. (2013) 35:43–54. 10.1155/2013/835876 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Ganon-Elazar E, Akirav I. Cannabinoids prevent the development of behavioral and endocrine alterations in a rat model of intense stress. Neuropsychopharmacology (2012) 37:456–66. 10.1038/npp.2011.204 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Serova L, Tillinger A, Alaluf L, Laukova M, Keegan K, Sabban E. Single intranasal neuropeptide Y infusion attenuates development of PTSD-like symptoms to traumatic stress in rats. Neuroscience (2013) 236:298–312. 10.1016/j.neuroscience.2013.01.040 [DOI] [PubMed] [Google Scholar]
  • 140.Khan S, Liberzon I. Topiramate attenuates exaggerated acoustic startle in an animal model of PTSD. Psychopharmacology (2004) 172:225–9. 10.1007/s00213-003-1634-4 [DOI] [PubMed] [Google Scholar]
  • 141.Berlant J. Topiramate as a therapy for chronic posttraumatic stress disorder. Psychiatry (Edgmont) (2006) 3:40–5. [PMC free article] [PubMed] [Google Scholar]
  • 142.Twamley EW, Allard CB, Thorp SR, Norman SB, Cissell SH, Berardi KH, et al. Cognitive impairment and functioning in PTSD related to intimate partner violence. J Int Neuropsychol Soc. (2009) 15:879–87. 10.1017/S135561770999049X [DOI] [PubMed] [Google Scholar]
  • 143.Qureshi SU, Long ME, Bradshaw MR, Pyne JM, Magruder KM, Kimbrell T, et al. Does PTSD impair cognition beyond the effect of trauma? J Neuropsychiatry Clin Neurosci. (2011) 23:16–28. 10.1176/appi.neuropsych.23.1.16 [DOI] [PubMed] [Google Scholar]
  • 144.Aupperle RL, Melrose AJ, Stein MB, Paulus MP. Executive function and PTSD: disengaging from trauma. Neuropharmacology (2012) 62:686–94. 10.1016/j.neuropharm.2011.02.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Danese A, Moffitt TE, Arseneault L, Bleiberg BA, Dinardo PB, Gandelman SB, et al. The origins of cognitive deficits in victimized children: implications for neuroscientists and clinicians. Am J Psychiatry (2017) 174:349–61. 10.1176/appi.ajp.2016.16030333 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.DiGangi JA, Gomez D, Mendoza L, Jason LA, Keys CB, Koenen KC. Pretrauma risk factors for posttraumatic stress disorder: a systematic review of the literature. Clin Psychol Rev. (2013) 33:728–44. 10.1016/j.cpr.2013.05.002 [DOI] [PubMed] [Google Scholar]
  • 147.Ji L-L, Peng J-B, Fu C-H, Cao D, Li D, Tong L, et al. Activation of Sigma-1 receptor ameliorates anxiety-like behavior and cognitive impairments in a rat model of post-traumatic stress disorder. Behav Brain Res. (2016) 311:408–16. 10.1016/j.bbr.2016.05.056 [DOI] [PubMed] [Google Scholar]
  • 148.Eagle AL, Fitzpatrick CJ, Perrine SA. Single prolonged stress impairs social and object novelty recognition in rats. Behav Brain Res. (2013) 256:591–7. 10.1016/j.bbr.2013.09.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Huang Z-l, Liu R, Bai X-y, Zhao G, Song J-k, Wu S, et al. Protective effects of the novel adenosine derivative WS0701 in a mouse model of posttraumatic stress disorder. Acta Pharmacol Sin. (2014) 35:24–32. 10.1038/aps.2013.143 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.George SA, Rodriguez-Santiago M, Riley J, Abelson JL, Floresco SB, Liberzon I. Alterations in cognitive flexibility in a rat model of post-traumatic stress disorder. Behav Brain Res. (2015) 286:256–64. 10.1016/j.bbr.2015.02.051 [DOI] [PubMed] [Google Scholar]
  • 151.Piao C, Deng X, Wang X, Yuan Y, Liu Z, Liang J. Altered function in medial prefrontal cortex and nucleus accumbens links to stress-induced behavioral inflexibility. Behav Brain Res. (2017) 317:16–26. 10.1016/j.bbr.2016.09.017 [DOI] [PubMed] [Google Scholar]
  • 152.Solan EA. Effects of Single Prolonged Stress on Set Shifting Errors. Kalamazoo, MI: Kalamazoo College; (2011). [Google Scholar]
  • 153.George SA, Knox D, Curtis AL, Aldridge JW, Valentino RJ, Liberzon I. Altered locus coeruleus–norepinephrine function following single prolonged stress. Eur J Neurosci. (2013) 37:901–9. 10.1111/ejn.12095 [DOI] [PubMed] [Google Scholar]
  • 154.Knox D, Perrine SA, George SA, Galloway MP, Liberzon I. Single prolonged stress decreases glutamate, glutamine, and creatine concentrations in the rat medial prefrontal cortex. Neurosci Lett. (2010) 480:16–20. 10.1016/j.neulet.2010.05.052 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Aston-Jones G, Cohen JD. An integrative theory of locus coeruleus-norepinephrine function: adaptive gain and optimal performance. Annu Rev Neurosci. (2005) 28:403–50. 10.1146/annurev.neuro.28.061604.135709 [DOI] [PubMed] [Google Scholar]
  • 156.Logue SF, Gould TJ. The neural and genetic basis of executive function: attention, cognitive flexibility, and response inhibition. Pharmacol Biochem Behav. (2014) 123:45–54. 10.1016/j.pbb.2013.08.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Germain A. Sleep disturbances as the hallmark of PTSD: where are we now? Am J Psychiatry (2013) 170:372–82. 10.1176/appi.ajp.2012.12040432 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Pace-Schott EF, Germain A, Milad MR. Sleep and REM sleep disturbance in the pathophysiology of PTSD: the role of extinction memory. Biol Mood Anxiety Disord. (2015) 5:1. 10.1186/s13587-015-0018-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Rusch HL, Guardado P, Baxter T, Mysliwiec V, Gill JM. Improved sleep quality is associated with reductions in depression and PTSD arousal symptoms and increases in IGF-1 concentrations. J Clin Sleep Med. (2015) 11:615–23. 10.5664/jcsm.4770 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Bonn-Miller MO, Babson KA, Vandrey R. Using cannabis to help you sleep: heightened frequency of medical cannabis use among those with PTSD. Drug Alcohol Depend. (2014) 136:162–5. 10.1016/j.drugalcdep.2013.12.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Chakravorty S, Grandner MA, Mavandadi S, Perlis ML, Sturgis EB, Oslin DW. Suicidal ideation in veterans misusing alcohol: relationships with insomnia symptoms and sleep duration. Addict Behav. (2014) 39:399–405. 10.1016/j.addbeh.2013.09.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Cox RC, Olatunji BO. A systematic review of sleep disturbance in anxiety and related disorders. J Anxiety Disord. (2016) 37:104–29. 10.1016/j.janxdis.2015.12.001 [DOI] [PubMed] [Google Scholar]
  • 163.Vanderheyden WM, Poe GR, Liberzon I. Trauma exposure and sleep: using a rodent model to understand sleep function in PTSD. Exp Brain Res. (2014) 232:1575–84. 10.1007/s00221-014-3890-4 [DOI] [PubMed] [Google Scholar]
  • 164.Vanderheyden W, Urpa L, Poe G. Increase in rem sleep following trauma exposure. Sleep Med. (2013) 14:e293. 10.1016/j.sleep.2013.11.718 [DOI] [Google Scholar]
  • 165.Nedelcovych MT, Gould RW, Zhan X, Bubser M, Gong X, Grannan M, et al. A rodent model of traumatic stress induces lasting sleep and quantitative electroencephalographic disturbances. ACS Chem Neurosci. (2015) 6:485–93. 10.1021/cn500342u [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Yehuda R, Halligan SL, Bierer LM. Relationship of parental trauma exposure and PTSD to PTSD, depressive and anxiety disorders in offspring. J Psychiatric Res. (2001) 35:261–70. 10.1016/S0022-3956(01)00032-2 [DOI] [PubMed] [Google Scholar]
  • 167.Lee B, Sur B, Cho S-G, Yeom M, Shim I, Lee H, et al. Ginsenoside Rb1 rescues anxiety-like responses in a rat model of post-traumatic stress disorder. J Nat Med. (2016) 70:133–44. 10.1007/s11418-015-0943-3 [DOI] [PubMed] [Google Scholar]
  • 168.Lee B, Sur B, Yeom M, Shim I, Lee H, Hahm D-H. Effects of systemic administration of ibuprofen on stress response in a rat model of post-traumatic stress disorder. Korean J Physiol Pharmacol. (2016) 20:357–66. 10.4196/kjpp.2016.20.4.357 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Patki G, Li L, Allam F, Solanki N, Dao AT, Alkadhi K, et al. Moderate treadmill exercise rescues anxiety and depression-like behavior as well as memory impairment in a rat model of posttraumatic stress disorder. Physiol Behav. (2014) 130:47–53. 10.1016/j.physbeh.2014.03.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Peng Y, Feng S-F, Wang Q, Wang H-N, Hou W-G, Xiong L, et al. Hyperbaric oxygen preconditioning ameliorates anxiety-like behavior and cognitive impairments via upregulation of thioredoxin reductases in stressed rats. Progr Neuro Psychopharmacol Biol Psychiatry (2010) 34:1018–25. 10.1016/j.pnpbp.2010.05.016 [DOI] [PubMed] [Google Scholar]
  • 171.Qi J, Chen C, Lu Y-C, Zhang T, Xu H, Cui Y-Y, et al. Activation of extracellular signal-regulated kinase1/2 in the medial prefrontal cortex contributes to stress-induced hyperalgesia. Mol Neurobiol. (2014) 50:1013–23. 10.1007/s12035-014-8707-8 [DOI] [PubMed] [Google Scholar]
  • 172.Qiu Z-K, Zhang G-H, He J-L, Ma J-C, Zeng J, Shen D, et al. Free and Easy Wanderer Plus (FEWP) improves behavioral deficits in an animal model of post-traumatic stress disorder by stimulating allopregnanolone biosynthesis. Neurosci Lett. (2015) 602:162–6. 10.1016/j.neulet.2015.06.055 [DOI] [PubMed] [Google Scholar]
  • 173.Solanki N, Alkadhi I, Atrooz F, Patki G, Salim S. Grape powder prevents cognitive, behavioral, and biochemical impairments in a rat model of posttraumatic stress disorder. Nutr Res. (2015) 35:65–75. 10.1016/j.nutres.2014.11.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Nie H, Peng Z, Lao N, Wang H, Chen Y, Fang Z, et al. Rosmarinic acid ameliorates PTSD-like symptoms in a rat model and promotes cell proliferation in the hippocampus. Progr Neuro Psychopharmacol Biol Psychiatry (2014) 51:16–22. 10.1016/j.pnpbp.2014.01.002 [DOI] [PubMed] [Google Scholar]
  • 175.Peng Z, Zhang R, Wang H, Chen Y, Xue F, Wang L, et al. Ziprasidone ameliorates anxiety-like behaviors in a rat model of PTSD and up-regulates neurogenesis in the hippocampus and hippocampus-derived neural stem cells. Behav Brain Res. (2013) 244:1–8. 10.1016/j.bbr.2013.01.032 [DOI] [PubMed] [Google Scholar]
  • 176.Yu L, Wang L, Zhao X, Song M, Wang X. Role of single prolonged stress in acquisition of alcohol conditioned place preference in rats. Life Sci. (2016) 151:259–63. 10.1016/j.lfs.2016.03.004 [DOI] [PubMed] [Google Scholar]
  • 177.Serova L, Laukova M, Alaluf L, Pucillo L, Sabban E. Intranasal neuropeptide Y reverses anxiety and depressive-like behavior impaired by single prolonged stress PTSD model. Eur Neuropsychopharmacol. (2014) 24:142–7. 10.1016/j.euroneuro.2013.11.007 [DOI] [PubMed] [Google Scholar]
  • 178.Sabban EL, Serova LI, Alaluf LG, Laukova M, Peddu C. Comparative effects of intranasal neuropeptide Y and HS014 in preventing anxiety and depressive-like behavior elicited by single prolonged stress. Behav Brain Res. (2015) 295:9–16. 10.1016/j.bbr.2014.12.038 [DOI] [PubMed] [Google Scholar]
  • 179.Sabban EL, Alaluf LG, Serova LI. Potential of neuropeptide Y for preventing or treating post-traumatic stress disorder. Neuropeptides (2016) 56:19–24. 10.1016/j.npep.2015.11.004 [DOI] [PubMed] [Google Scholar]
  • 180.Lisieski MJ, Perrine SA. Binge-pattern cocaine administration causes long-lasting behavioral hyperarousal but does not enhance vulnerability to single prolonged stress in rats. Psychiatry Res. (2017) 257(Suppl. C):95–101. 10.1016/j.psychres.2017.07.02 [DOI] [PubMed] [Google Scholar]
  • 181.Wu Z-M, Yang L-H, Cui R, Ni G-L, Wu F-T, Liang Y. Contribution of hippocampal 5-HT3 receptors in hippocampal autophagy and extinction of conditioned fear responses after a single prolonged stress exposure in rats. Cell Mol Neurobiol. (2016) 1–12. 10.1007/s10571-016-0395-7 [DOI] [PubMed] [Google Scholar]
  • 182.Wu Z-M, Zheng C-H, Zhu Z-H, Wu F-T, Ni G-L, Liang Y. SiRNA-mediated serotonin transporter knockdown in the dorsal raphe nucleus rescues single prolonged stress-induced hippocampal autophagy in rats. J Neurolog Sci. (2016) 360:133–40. 10.1016/j.jns.2015.11.056 [DOI] [PubMed] [Google Scholar]
  • 183.Harvey BH, Brand L, Jeeva Z, Stein DJ. Cortical/hippocampal monoamines, HPA-axis changes and aversive behavior following stress and restress in an animal model of post-traumatic stress disorder. Physiol Behav. (2006) 87:881–90. 10.1016/j.physbeh.2006.01.033 [DOI] [PubMed] [Google Scholar]
  • 184.Eagle AL, Knox D, Roberts MM, Mulo K, Liberzon I, Galloway MP, et al. Single prolonged stress enhances hippocampal glucocorticoid receptor and phosphorylated protein kinase B levels. Neurosci Res. (2013) 75:130–7. 10.1016/j.neures.2012.11.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Rosso IM, Weiner MR, Crowley DJ, Silveri MM, Rauch SL, Jensen JE. Insula and anterior cingulate GABA levels in posttraumatic stress disorder: preliminary findings using magnetic resonance spectroscopy. Depress Anxiety (2014) 31:115–23. 10.1002/da.22155 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Gewirtz JC, McNish KA, Davis M. Lesions of the bed nucleus of the stria terminalis block sensitization of the acoustic startle reflex produced by repeated stress, but not fear-potentiated startle. Progr Neuro Psychopharmacol Biol Psychiatry (1998) 22:625–48. 10.1016/S0278-5846(98)00028-1 [DOI] [PubMed] [Google Scholar]
  • 187.Lebow M, Neufeld-Cohen A, Kuperman Y, Tsoory M, Gil S, Chen A. Susceptibility to PTSD-like behavior is mediated by corticotropin-releasing factor receptor type 2 levels in the bed nucleus of the stria terminalis. J Neurosci. (2012) 32:6906–16. 10.1523/JNEUROSCI.4012-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Guina J, Welton RS, Broderick PJ, Correll TL, Peirson RP. DSM-5 Criteria and its implications for diagnosing PTSD in military service members and veterans. Curr Psychiatry Rep. (2016) 18:43. 10.1007/s11920-016-0686-1 [DOI] [PubMed] [Google Scholar]
  • 189.Kashdan TB, Elhai JD, Frueh BC. Anhedonia and emotional numbing in combat veterans with PTSD. Behav Res Ther. (2006) 44:457–67. 10.1016/j.brat.2005.03.001 [DOI] [PubMed] [Google Scholar]
  • 190.Enman NM, Arthur K, Ward SJ, Perrine SA, Unterwald EM. Anhedonia, reduced cocaine reward, and dopamine dysfunction in a rat model of posttraumatic stress disorder. Biol Psychiatry. (2015) 78:871–9. 10.1016/j.biopsych.2015.04.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Berridge KC. Measuring hedonic impact in animals and infants: microstructure of affective taste reactivity patterns. Neurosci Biobehav Rev. (2000) 24:173–98. 10.1016/S0149-7634(99)00072-X [DOI] [PubMed] [Google Scholar]
  • 192.Knutson B, Burgdorf J, Panksepp J. Ultrasonic vocalizations as indices of affective states in rats. Psychol Bull. (2002) 128:961. 10.1037/0033-2909.128.6.961 [DOI] [PubMed] [Google Scholar]
  • 193.Maier SF, Seligman ME. Learned helplessness at fifty: insights from neuroscience. Psychol Rev. (2016) 123:349. 10.1037/rev0000033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Yu B, Wen L, Xiao B, Han F, Shi Y. Single prolonged stress induces ATF6 alpha-dependent endoplasmic reticulum stress and the apoptotic process in medial frontal cortex neurons. BMC Neurosci. (2014) 15:115. 10.1186/s12868-014-0115-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Serova LI, Laukova M, Alaluf LG, Sabban EL. Intranasal infusion of melanocortin receptor four (MC4R) antagonist to rats ameliorates development of depression and anxiety related symptoms induced by single prolonged stress. Behav Brain Res. (2013) 250:139–47. 10.1016/j.bbr.2013.05.006 [DOI] [PubMed] [Google Scholar]
  • 196.Porsolt RD, Le Pichon M, Jalfre M. Depression: a new animal model sensitive to antidepressant treatments. Nature (1977) 266:730–2. 10.1038/266730a0 [DOI] [PubMed] [Google Scholar]
  • 197.Detke MJ, Rickels M, Lucki I. Active behaviors in the rat forced swimming test differentially produced by serotonergic and noradrenergic antidepressants. Psychopharmacology (1995) 121:66–72. 10.1007/BF02245592 [DOI] [PubMed] [Google Scholar]
  • 198.Harvey BH, Oosthuizen F, Brand L, Wegener G, Stein DJ. Stress–restress evokes sustained iNOS activity and altered GABA levels and NMDA receptors in rat hippocampus. Psychopharmacology (2004) 175:494–502. 10.1007/s00213-004-1836-4 [DOI] [PubMed] [Google Scholar]
  • 199.Mezadri TJ, Batista GM, Portes AC, Marino-Neto J, Lino-de-Oliveira C. Repeated rat-forced swim test: reducing the number of animals to evaluate gradual effects of antidepressants. J Neurosci Methods (2011) 195:200–5. 10.1016/j.jneumeth.2010.12.015 [DOI] [PubMed] [Google Scholar]
  • 200.Molendijk ML, de Kloet ER. Immobility in the forced swim test is adaptive and does not reflect depression. Psychoneuroendocrinology (2015) 62:389–91. 10.1016/j.psyneuen.2015.08.028 [DOI] [PubMed] [Google Scholar]
  • 201.De Kloet E, Molendijk M. Coping with the forced swim stressor: towards understanding an adaptive mechanism. Neural Plasticity (2016) 2016:6503162. 10.1155/2016/6503162 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Landgraf D, Long J, Der-Avakian A, Streets M, Welsh DK. Dissociation of learned helplessness and fear conditioning in mice: a mouse model of depression. PLoS ONE (2015) 10:e0125892. 10.1371/journal.pone.0125892 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Chourbaji S, Zacher C, Sanchis-Segura C, Dormann C, Vollmayr B, Gass P. Learned helplessness: validity and reliability of depressive-like states in mice. Brain Res Brain Res Protoc. (2005) 16:70–8. 10.1016/j.brainresprot.2005.09.002 [DOI] [PubMed] [Google Scholar]
  • 204.Vollmayr B, Henn FA. Learned helplessness in the rat: improvements in validity and reliability. Brain Res Brain Res Protoc. (2001) 8:1–7. 10.1016/S1385-299X(01)00067-8 [DOI] [PubMed] [Google Scholar]
  • 205.McCauley JL, Killeen T, Gros DF, Brady KT, Back SE. Posttraumatic stress disorder and co-occurring substance use disorders: advances in assessment and treatment. Clin Psychol. (2012) 19:283–304. 10.1111/cpsp.12006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Breslau N, Davis GC, Andreski P, Peterson EL, Schultz LR. Sex differences in posttraumatic stress disorder. Arch Gen Psychiatry (1997) 54:1044–8. 10.1001/archpsyc.1997.01830230082012 [DOI] [PubMed] [Google Scholar]
  • 207.Fishbain DA, Pulikal A, Lewis JE, Gao J. Chronic pain types differ in their reported prevalence of Post-Traumatic Stress Disorder (PTSD) and there is consistent evidence that chronic pain is associated with PTSD: an evidence-based structured systematic review. Pain Med. (2016) 18:711–735. 10.1093/pm/pnw065 [DOI] [PubMed] [Google Scholar]
  • 208.Sareen J, Cox BJ, Stein MB, Afifi TO, Fleet C, Asmundson GJ. Physical and mental comorbidity, disability, and suicidal behavior associated with posttraumatic stress disorder in a large community sample. Psychos Med. (2007) 69:242–8. 10.1097/PSY.0b013e31803146d8 [DOI] [PubMed] [Google Scholar]
  • 209.Pace TW, Heim CM. A short review on the psychoneuroimmunology of posttraumatic stress disorder: from risk factors to medical comorbidities. Brain Behav Immunity (2011) 25:6–13. 10.1016/j.bbi.2010.10.003 [DOI] [PubMed] [Google Scholar]
  • 210.Leslie K, Jacobsen Steven M, Southwick Thomas R., Kosten Substance use disorders in patients with posttraumatic stress disorder: a review of the literature. Am J Psychiatry (2001) 158:1184–90. 10.1176/appi.ajp.158.8.1184 [DOI] [PubMed] [Google Scholar]
  • 211.Najavits LM, Hien D. Helping vulnerable populations: a comprehensive review of the treatment outcome literature on substance use disorder and PTSD. J Clin Psychol. (2013) 69:433–79. 10.1002/jclp.21980 [DOI] [PubMed] [Google Scholar]
  • 212.Brady KT, Back SE, Coffey SF. Substance abuse and posttraumatic stress disorder. Curr Dir Psychol Sci. (2004) 13:206–9. 10.1111/j.0963-7214.2004.00309.x [DOI] [Google Scholar]
  • 213.Hruska B, Delahanty DL. PTSD-SUD biological mechanisms: self-medication and beyond. In: Ouimette P, Read JP, editors. Trauma and Substance Abuse: Causes, Consequences, and Treatment of Comorbid Disorders. Washington, DC: American Psychological Association; (2014). p. 35–52. [Google Scholar]
  • 214.Steketee JD, Kalivas PW. Drug wanting: behavioral sensitization and relapse to drug-seeking behavior. Pharmacol Rev. (2011) 63:348–65. 10.1124/pr.109.001933 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Matchynski-Franks JJ, Susick LL, Schneider BL, Perrine SA, Conti AC. Impaired ethanol-induced sensitization and decreased cannabinoid receptor-1 in a model of posttraumatic stress disorder. PLoS ONE (2016) 11:e0155759. 10.1371/journal.pone.0155759 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Eagle AL, Singh R, Kohler RJ, Friedman AL, Liebowitz CP, Galloway MP, et al. Single prolonged stress effects on sensitization to cocaine and cocaine self-administration in rats. Behav Brain Res. (2015) 284:218–24. 10.1016/j.bbr.2015.02.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Eagle AL, Perrine SA. Methamphetamine-induced behavioral sensitization in a rodent model of posttraumatic stress disorder. Drug Alcohol Depend. (2013) 131:36–43. 10.1016/j.drugalcdep.2013.04.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Toledano D, Tassin J-P, Gisquet-Verrier P. Traumatic stress in rats induces noradrenergic-dependent long-term behavioral sensitization: role of individual differences and similarities with dependence on drugs of abuse. Psychopharmacology (2013) 230:465–76. 10.1007/s00213-013-3179-5 [DOI] [PubMed] [Google Scholar]
  • 219.Holly EN, Miczek KA. Capturing individual differences: challenges in animal models of posttraumatic stress disorder and drug abuse. Biol Psychiatry (2015) 78:816–8. 10.1016/j.biopsych.2015.09.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Saladin ME, Brady KT, Dansky BS, Kilpatrick DG. Understanding comorbidity between ptsd and substance use disorders: two preliminary investigations. Addict Behav. (1995) 20:643–55. 10.1016/0306-4603(95)00024-7 [DOI] [PubMed] [Google Scholar]
  • 221.Brady KT, Dansky BS, Sonne SC, Saladin ME. Posttraumatic stress disorder and cocaine dependence: order of onset. Am J Addict. (1998) 7:128–35. 10.1111/j.1521-0391.1998.tb00327.x [DOI] [PubMed] [Google Scholar]
  • 222.Carri-Ann Gibson MD D. Review of posttraumatic stress disorder and chronic pain: the path to integrated care. J Rehabil Res Dev. (2012) 49:753 10.1682/JRRD.2011.09.0158 [DOI] [PubMed] [Google Scholar]
  • 223.Moeller-Bertram T, Keltner J, Strigo IA. Pain and post traumatic stress disorder – review of clinical and experimental evidence. Neuropharmacology (2012) 62:586–97. 10.1016/j.neuropharm.2011.04.028 [DOI] [PubMed] [Google Scholar]
  • 224.Zhang Y, Gandhi PR, Standifer KM. Increased nociceptive sensitivity and nociceptin/orphanin FQ levels in a rat model of PTSD. Mol Pain (2012) 8:1. 10.1186/1744-8069-8-76 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.He Y-Q, Chen Q, Ji L, Wang Z-G, Bai Z-H, Stephens RL, et al. PKCγ receptor mediates visceral nociception and hyperalgesia following exposure to PTSD-like stress in the spinal cord of rats. Mol Pain (2013) 9:35. 10.1186/1744-8069-9-35 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Le Dorze C, Gisquet-Verrier P. Effects of multiple brief exposures to trauma-associated cues on traumatized resilient and vulnerable rats. Brain Res. (2016) 1652:71–80. 10.1016/j.brainres.2016.10.002 [DOI] [PubMed] [Google Scholar]
  • 227.Tolin DF, Foa EB. Sex differences in trauma and posttraumatic stress disorder: a quantitative review of 25 years of research. Psychol Bull. (2006) 132:959. 10.1037/0033-2909.132.6.959 [DOI] [PubMed] [Google Scholar]
  • 228.Hourani L, Williams J, Bray R, Kandel D. Gender differences in the expression of PTSD symptoms among active duty military personnel. J Anxiety Disord. (2015) 29:101–8. 10.1016/j.janxdis.2014.11.007 [DOI] [PubMed] [Google Scholar]
  • 229.Carragher N, Sunderland M, Batterham PJ, Calear AL, Elhai JD, Chapman C, et al. Discriminant validity and gender differences in DSM-5 posttraumatic stress disorder symptoms. J Affect Disord. (2016) 190:56–67. 10.1016/j.jad.2015.09.071 [DOI] [PubMed] [Google Scholar]
  • 230.King MW, Street AE, Gradus JL, Vogt DS, Resick PA. gender differences in posttraumatic stress symptoms among OEF/OIF veterans: an item response theory analysis. J Traum Stress (2013) 26:175–83. 10.1002/jts.21802 [DOI] [PubMed] [Google Scholar]
  • 231.Carmassi C, Stratta P, Massimetti G, Bertelloni CA, Conversano C, Cremone IM, et al. New DSM-5 maladaptive symptoms in PTSD: gender differences and correlations with mood spectrum symptoms in a sample of high school students following survival of an earthquake. Ann Gen Psychiatry (2014) 13:28. 10.1186/s12991-014-0028-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Cover KK, Maeng LY, Lebron-Milad K, Milad MR. Mechanisms of estradiol in fear circuitry: implications for sex differences in psychopathology. Transl Psychiatry (2014) 4:e422. 10.1038/tp.2014.67 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Brewin CR, Andrews B, Valentine JD. Meta-analysis of risk factors for posttraumatic stress disorder in trauma-exposed adults. J Consult Clin Psychol. (2000) 68:748. 10.1037/0022-006X.68.5.748 [DOI] [PubMed] [Google Scholar]
  • 234.Koenen KC, Moffitt TE, Poulton R, Martin J, Caspi A. Early childhood factors associated with the development of post-traumatic stress disorder: results from a longitudinal birth cohort. Psychol Med. (2007) 37:181–92. 10.1017/S0033291706009019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Yehuda R, Teicher MH, Trestman RL, Levengood RA, Siever LJ. Cortisol regulation in posttraumatic stress disorder and major depression: a chronobiological analysis. Biol Psychiatry (1996) 40:79–88. 10.1016/0006-3223(95)00451-3 [DOI] [PubMed] [Google Scholar]
  • 236.Yehuda R, Southwick SM, Nussbaum G, Wahby V, Giller EL, Jr, Mason JW. Low urinary cortisol excretion in patients with posttraumatic stress disorder. J Nerv Ment Dis. (1990) 178:366–9. 10.1097/00005053-199006000-00004 [DOI] [PubMed] [Google Scholar]
  • 237.Wheler GT, Brandon D, Clemons A, Riley C, Kendall J, Loriaux DL, et al. Cortisol production rate in posttraumatic stress disorder. J Clin Endocrinol Metab. (2006) 91:3486–9. 10.1210/jc.2006-0061 [DOI] [PubMed] [Google Scholar]
  • 238.Young EA, Breslau N. Saliva cortisol in posttraumatic stress disorder: a community epidemiologic study. Biol Psychiatry (2004) 56:205–9. 10.1016/j.biopsych.2004.05.011 [DOI] [PubMed] [Google Scholar]
  • 239.Meewisse M-L, Reitsma JB, De Vries G-J, Gersons BP, Olff M. Cortisol and post-traumatic stress disorder in adults. Brit J Psychiatry (2007) 191:387–92. 10.1192/bjp.bp.106.024877 [DOI] [PubMed] [Google Scholar]
  • 240.Yehuda R. Status of glucocorticoid alterations in post-traumatic stress disorder. Ann N Y Acad Sci. (2009) 1179:56–69. 10.1111/j.1749-6632.2009.04979.x [DOI] [PubMed] [Google Scholar]
  • 241.Liberzon I, Krstov M, Young EA. Stress-restress: effects on ACTH and fast feedback. Psychoneuroendocrinology (1997) 22:443–53. 10.1016/S0306-4530(97)00044-9 [DOI] [PubMed] [Google Scholar]
  • 242.Oitzl MS, van Haarst AD, Sutanto W, de Kloet ER. Corticosterone, brain mineralocorticoid receptors (MRs) and the activity of the hypothalamic-pituitary-adrenal (HPA) axis: the Lewis rat as an example of increased central MR capacity and a hyporesponsive HPA axis. Psychoneuroendocrinology (1995) 20:655–75. 10.1016/0306-4530(95)00003-7 [DOI] [PubMed] [Google Scholar]
  • 243.Du Zhe HF, Yuxiu S. Changes of glucocorticoid receptor in hypothalamus of rat with posttraumatic stress disorder [J]. Chin J Anat. (2008) 5:022 Available online at: http://caod.oriprobe.com/articles/15033138/Changes_of_glucocorticoid_receptor_in_hypothalamus_of_rat_with_posttra.htm [Google Scholar]
  • 244.Wang H-T, Han F, Shi Y-X. Activity of the 5-HT1A receptor is involved in the alteration of glucocorticoid receptor in hippocampus and corticotropin-releasing factor in hypothalamus in SPS rats. Int J Mol Med. (2009) 24:227. 10.3892/ijmm_00000225 [DOI] [PubMed] [Google Scholar]
  • 245.Wen L, Han F, Shi Y. Changes in the glucocorticoid receptor and Ca2+/calreticulin-dependent signalling pathway in the medial prefrontal cortex of rats with post-traumatic stress disorder. J Mol Neurosci. (2015) 56:24–34. 10.1007/s12031-014-0464-7 [DOI] [PubMed] [Google Scholar]
  • 246.Li M, Han F, Shi Y. Expression of locus coeruleus mineralocorticoid receptor and glucocorticoid receptor in rats under single-prolonged stress. Neurol Sci. (2011) 32:625–31. 10.1007/s10072-011-0597-1 [DOI] [PubMed] [Google Scholar]
  • 247.Cui H, Sakamoto H, Higashi S, Kawata M. Effects of single-prolonged stress on neurons and their afferent inputs in the amygdala. Neuroscience (2008) 152:703–12. 10.1016/j.neuroscience.2007.12.028 [DOI] [PubMed] [Google Scholar]
  • 248.Liberzon I, Lopez J, Flagel S, Vazquez D, Young E. Differential regulation of hippocampal glucocorticoid receptors mRNA and fast feedback: relevance to post-traumatic stress disorder. J Neuroendocrinol. (1999) 11:11–7. 10.1046/j.1365-2826.1999.00288.x [DOI] [PubMed] [Google Scholar]
  • 249.Zhang J-H, Han F, Shi Y-X. Single prolonged stress induces changes in the expression of mineralocorticoid receptor in the medial prefrontal cortex in a rat model of post-traumatic stress disorder. Mol Med Rep. (2012) 6:330–4. 10.3892/mmr.2012.937 [DOI] [PubMed] [Google Scholar]
  • 250.de Quervain D, Schwabe L, Roozendaal B. Stress, glucocorticoids and memory: implications for treating fear-related disorders. Nat Rev Neurosci. (2017) 18:7–19. 10.1038/nrn.2016.155 [DOI] [PubMed] [Google Scholar]
  • 251.Raglan GB, Schmidt LA, Schulkin J. The role of glucocorticoids and corticotropin-releasing hormone regulation on anxiety symptoms and response to treatment. Endocr Connect. (2017) 6:R1–R7. 10.1530/EC-16-0100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Chambers R, Bremner J, Moghaddam B, Southwick S, Charney D, Krystal J. editors. Glutamate and post-traumatic stress disorder: toward a psychobiology of dissociation. Semin Clin Neuropsychiatry (1999) 4:274–81. 10.153/SCNP00400274 [DOI] [PubMed] [Google Scholar]
  • 253.Meyerhoff DJ, Mon A, Metzler T, Neylan TC. Cortical gamma-aminobutyric acid and glutamate in posttraumatic stress disorder and their relationships to self-reported sleep quality. Sleep (2014) 37:893–900. 10.5665/sleep.3654 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Steckler T, Risbrough V. Pharmacological treatment of PTSD–established and new approaches. Neuropharmacology (2012) 62:617–27. 10.1016/j.neuropharm.2011.06.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Mahan AL, Ressler KJ. Fear conditioning, synaptic plasticity and the amygdala: implications for posttraumatic stress disorder. Trends Neurosci. (2012) 35:24–35. 10.1016/j.tins.2011.06.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Michels L, Schulte-Vels T, Schick M, O'Gorman RL, Zeffiro T, Hasler G, et al. Prefrontal GABA and glutathione imbalance in posttraumatic stress disorder: preliminary findings. Psychiatry Res. (2014) 224:288–95. 10.1016/j.pscychresns.2014.09.007 [DOI] [PubMed] [Google Scholar]
  • 257.Yang ZY, Quan H, Peng ZL, Zhong Y, Tan ZJ, Gong QY. Proton magnetic resonance spectroscopy revealed differences in the glutamate+ glutamine/creatine ratio of the anterior cingulate cortex between healthy and pediatric post-traumatic stress disorder patients diagnosed after 2008 Wenchuan earthquake. Psychiatry Clin Neurosci. (2015) 69:782–90. 10.1111/pcn.12332 [DOI] [PubMed] [Google Scholar]
  • 258.Averill LA, Purohit P, Averill CL, Boesl MA, Krystal JH, Abdallah CG. Glutamate dysregulation and glutamatergic therapeutics for PTSD: evidence from human studies. Neurosci Lett. (2017) 649:147–55. 10.1016/j.neulet.2016.11.064 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Geuze E, Van Berckel B, Lammertsma A, Boellaard R, De Kloet C, Vermetten E, et al. Reduced GABA A benzodiazepine receptor binding in veterans with post-traumatic stress disorder. Mol Psychiatry. (2008) 13:74 10.1038/sj.mp.4002054 [DOI] [PubMed] [Google Scholar]
  • 260.Lim S-I, Song K-H, Yoo C-H, Woo D-C, Choe B-Y. Decreased glutamatergic activity in the frontal cortex of single prolonged stress model: in vivo and ex vivo proton MR spectroscopy. Neurochem Res. (2017) 1–12. 10.1007/s11064-017-2232-x [DOI] [PubMed] [Google Scholar]
  • 261.Matsumoto Y, Morinobu S, Yamamoto S, Matsumoto T, Takei S, Fujita Y, et al. Vorinostat ameliorates impaired fear extinction possibly via the hippocampal NMDA-CaMKII pathway in an animal model of posttraumatic stress disorder. Psychopharmacology (2013) 229:51–62. 10.1007/s00213-013-3078-9 [DOI] [PubMed] [Google Scholar]
  • 262.Yamamoto S, Morinobu S, Iwamoto Y, Ueda Y, Takei S, Fujita Y, et al. Alterations in the hippocampal glycinergic system in an animal model of posttraumatic stress disorder. J Psychiatr Res. (2010) 44:1069–74. 10.1016/j.jpsychires.2010.03.013 [DOI] [PubMed] [Google Scholar]
  • 263.Hofmann SG. D-cycloserine for treating anxiety disorders: making good exposures better and bad exposures worse. Depress Anxiety (2014) 31:175. 10.1002/da.22257 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Reul J, Nutt D. Glutamate and cortisol—a critical confluence in PTSD? J Psychopharmacol. (2008) 22:469–72. 10.1177/0269881108094617 [DOI] [PubMed] [Google Scholar]
  • 265.Davis L, Hamner M, Bremner JD. Pharmacotherapy for PTSD: effects on PTSD symptoms and the brain. In: Posttraumatic Stress Disorder: From Neurobiology to Treatment. Douglas Bremner J, editor. (2016). p. 385–412. [Google Scholar]
  • 266.Puglisi-Allegra S, Andolina D. Serotonin and stress coping. Behav Brain Res. (2015) 277:58–67. 10.1016/j.bbr.2014.07.052 [DOI] [PubMed] [Google Scholar]
  • 267.Lanfumey L, Mongeau R, Cohen-Salmon C, Hamon M. Corticosteroid–serotonin interactions in the neurobiological mechanisms of stress-related disorders. Neurosci Biobehav Rev. (2008) 32:1174–84. 10.1016/j.neubiorev.2008.04.006 [DOI] [PubMed] [Google Scholar]
  • 268.Han F, Xiao B, Wen L, Shi Y. Effects of fluoxetine on the amygdala and the hippocampus after administration of a single prolonged stress to male Wistar rates: In vivo proton magnetic resonance spectroscopy findings. Psychiatry Res. (2015) 232:154–61. 10.1016/j.pscychresns.2015.02.011 [DOI] [PubMed] [Google Scholar]
  • 269.Garabadu D, Ahmad A, Krishnamurthy S. Risperidone attenuates modified stress–re-stress paradigm-induced mitochondrial dysfunction and apoptosis in rats exhibiting post-traumatic stress disorder-like symptoms. J Mol Neurosci. (2015) 56:299–312. 10.1007/s12031-015-0532-7 [DOI] [PubMed] [Google Scholar]
  • 270.Heisler L, Zhou L, Bajwa P, Hsu J, Tecott L. Serotonin 5-HT2C receptors regulate anxiety-like behavior. Genes Brain Behav. (2007) 6:491–6. 10.1111/j.1601-183X.2007.00316.x [DOI] [PubMed] [Google Scholar]
  • 271.Liu D, Xiao B, Han F, Luo F, Wang E, Shi Y. Changes in 5-HT1A receptor expression in the oculomotor nucleus in a rat model of post-traumatic stress disorder. J Mol Neurosci. (2013) 49:360–8. 10.1007/s12031-012-9874-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Kondo M, Nakamura Y, Ishida Y, Shimada S. The 5-HT3 receptor is essential for exercise-induced hippocampal neurogenesis and antidepressant effects. Mol Psychiatry (2015) 20:1428–37. 10.1038/mp.2014.153 [DOI] [PubMed] [Google Scholar]
  • 273.Southwick SM, Bremner JD, Rasmusson A, Morgan CA, Arnsten A, Charney DS. Role of norepinephrine in the pathophysiology and treatment of posttraumatic stress disorder. Biol Psychiatry (1999) 46:1192–204. 10.1016/S0006-3223(99)00219-X [DOI] [PubMed] [Google Scholar]
  • 274.Geracioti TD, Jr, Baker DG, Ekhator NN, West SA, Hill KK, Bruce AB, et al. CSF norepinephrine concentrations in posttraumatic stress disorder. Am J Psychiatry (2001) 158:1227–30. 10.1176/appi.ajp.158.8.1227 [DOI] [PubMed] [Google Scholar]
  • 275.Mellman TA, Kumar A, Kulick-Bell R, Kumar M, Nolan B. Nocturnal/daytime urine noradrenergic measures and sleep in combat-related PTSD. Biol Psychiatry (1995) 38:174–9. 10.1016/0006-3223(94)00238-X [DOI] [PubMed] [Google Scholar]
  • 276.Morgan C, III, Grillon C, Southwick SM, Nagy LM, Davis M, Krystal JH, et al. Yohimbine facilitated acoustic startle in combat veterans with post-traumatic stress disorder. Psychopharmacology (1995) 117:466–71. 10.1007/BF02246220 [DOI] [PubMed] [Google Scholar]
  • 277.Sabban EL, Laukova M, Alaluf LG, Olsson E, Serova LI. Locus coeruleus response to single-prolonged stress and early intervention with intranasal neuropeptide Y. J Neurochem. (2015) 135:975–86. 10.1111/jnc.13347 [DOI] [PubMed] [Google Scholar]
  • 278.Wise RA. Dopamine, learning and motivation. Nat Rev Neurosci. (2004) 5:483–94. 10.1038/nrn1406 [DOI] [PubMed] [Google Scholar]
  • 279.Grace AA. Dysregulation of the dopamine system in the pathophysiology of schizophrenia and depression. Nat Rev Neurosci. (2016) 17:524–32. 10.1038/nrn.2016.57 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Camardese G, Di Giuda D, Di Nicola M, Cocciolillo F, Giordano A, Janiri L, et al. Imaging studies on dopamine transporter and depression: a review of literature and suggestions for future research. J Psychiatr Res. (2014) 51:7–18. 10.1016/j.jpsychires.2013.12.006 [DOI] [PubMed] [Google Scholar]
  • 281.Nutt DJ, Lingford-Hughes A, Erritzoe D, Stokes PR. The dopamine theory of addiction: 40 years of highs and lows. Nat Rev Neurosci. (2015) 16:305–12. 10.1038/nrn3939 [DOI] [PubMed] [Google Scholar]
  • 282.Abraham AD, Neve KA, Lattal KM. Dopamine and extinction: a convergence of theory with fear and reward circuitry. Neurobiol Learn Mem. (2014) 108:65–77. 10.1016/j.nlm.2013.11.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Norman SB, Myers US, Wilkins KC, Goldsmith AA, Hristova V, Huang Z, et al. Review of biological mechanisms and pharmacological treatments of comorbid PTSD and substance use disorder. Neuropharmacology (2012) 62:542–51. 10.1016/j.neuropharm.2011.04.032 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Geracioti TD, Jr, Jefferson-Wilson L, Strawn JR, Baker DG, Dashevsky BA, Horn PS, et al. Effect of traumatic imagery on cerebrospinal fluid dopamine and serotonin metabolites in posttraumatic stress disorder. J Psychiatric Res. (2013) 47:995–8. 10.1016/j.jpsychires.2013.01.023 [DOI] [PubMed] [Google Scholar]
  • 285.Hoexter MQ, Fadel G, Felício AC, Calzavara MB, Batista IR, Reis MA, et al. Higher striatal dopamine transporter density in PTSD: an in vivo SPECT study with [99mTc] TRODAT-1. Psychopharmacology (2012) 224:337–45. 10.1007/s00213-012-2755-4 [DOI] [PubMed] [Google Scholar]
  • 286.Banerjee SB, Morrison FG, Ressler KJ. Genetic approaches for the study of PTSD: advances and challenges. Neurosci Lett. (2017) 649:139–46. 10.1016/j.neulet.2017.02.058 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Kessler RC, Berglund P, Demler O, Jin R, Merikangas KR, Walters EE. Lifetime prevalence and age-of-onset distributions of DSM-IV disorders in the National Comorbidity Survey Replication. Arch Gen Psychiatry (2005) 62:593–602. 10.1001/archpsyc.62.6.593 [DOI] [PubMed] [Google Scholar]
  • 288.Sun R, Zhang W, Bo J, Zhang Z, Lei Y, Huo W, et al. Spinal activation of alpha7-nicotinic acetylcholine receptor attenuates posttraumatic stress disorder-related chronic pain via suppression of glial activation. Neuroscience (2017) 344:243–54. 10.1016/j.neuroscience.2016.12.029 [DOI] [PubMed] [Google Scholar]
  • 289.Tanev KS, Orr SP, Pace-Schott EF, Griffin M, Pitman RK, Resick PA. Positive association between nightmares and heart rate response to loud tones: relationship to parasympathetic dysfunction in PTSD nightmares. J Nerv Ment Dis. (2017) 205:308–12. 10.1097/NMD.0000000000000641 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Di Marzo V, Melck D, Bisogno T, De Petrocellis L. Endocannabinoids: endogenous cannabinoid receptor ligands with neuromodulatory action. Trends Neurosci. (1998) 21:521–8. 10.1016/S0166-2236(98)01283-1 [DOI] [PubMed] [Google Scholar]
  • 291.Neumeister A. The endocannabinoid system provides an avenue for evidence-based treatment development for PTSD. Depress Anxiety (2013) 30:93–6. 10.1002/da.22031 [DOI] [PubMed] [Google Scholar]
  • 292.Roitman P, Mechoulam R, Cooper-Kazaz R, Shalev A. Preliminary, open-label, pilot study of add-on oral Δ9-tetrahydrocannabinol in chronic post-traumatic stress disorder. Clin Drug Investig. (2014) 34:587–91. 10.1007/s40261-014-0212-3 [DOI] [PubMed] [Google Scholar]
  • 293.Fraser GA. The use of a synthetic cannabinoid in the management of Treatment-Resistant nightmares in posttraumatic stress disorder (PTSD). CNS Neurosci Therapeut. (2009) 15:84–8. 10.1111/j.1755-5949.2008.00071.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Jetly R, Heber A, Fraser G, Boisvert D. The efficacy of nabilone, a synthetic cannabinoid, in the treatment of PTSD-associated nightmares: a preliminary randomized, double-blind, placebo-controlled cross-over design study. Psychoneuroendocrinology (2015) 51:585–8. 10.1016/j.psyneuen.2014.11.002 [DOI] [PubMed] [Google Scholar]
  • 295.Rabinak CA, Angstadt M, Sripada CS, Abelson JL, Liberzon I, Milad MR, et al. Cannabinoid facilitation of fear extinction memory recall in humans. Neuropharmacology (2013) 64:396–402. 10.1016/j.neuropharm.2012.06.063 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Cougle JR, Bonn-Miller MO, Vujanovic AA, Zvolensky MJ, Hawkins KA. Posttraumatic stress disorder and cannabis use in a nationally representative sample. Psychol Addict Behav. (2011) 25:554. 10.1037/a0023076 [DOI] [PubMed] [Google Scholar]
  • 297.Ramot A, Akirav I. Cannabinoid receptors activation and glucocorticoid receptors deactivation in the amygdala prevent the stress-induced enhancement of a negative learning experience. Neurobiol Learn Mem. (2012) 97:393–401. 10.1016/j.nlm.2012.03.003 [DOI] [PubMed] [Google Scholar]
  • 298.Carter CS. Oxytocin pathways and the evolution of human behavior. Annu Rev Psychol. (2014) 65:17–39. 10.1146/annurev-psych-010213-115110 [DOI] [PubMed] [Google Scholar]
  • 299.Feldman R, Vengrober A, Ebstein R. Affiliation buffers stress: cumulative genetic risk in oxytocin–vasopressin genes combines with early caregiving to predict PTSD in war-exposed young children. Transl Psychiatry (2014) 4:e370. 10.1038/tp.2014.6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Olff M, Koch SB, Nawijn L, Frijling JL, Van Zuiden M, Veltman DJ. Social support, oxytocin and PTSD. Eur J Psychotraumatol. (2014) 5:26513. 10.3402/ejpt.v5.26513 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Koch SB, van Zuiden M, Nawijn L, Frijling JL, Veltman DJ, Olff M. Intranasal oxytocin as strategy for medication-enhanced psychotherapy of PTSD: Salience processing and fear inhibition processes. Psychoneuroendocrinology (2014) 40:242–56. 10.1016/j.psyneuen.2013.11.018 [DOI] [PubMed] [Google Scholar]
  • 302.Frijling JL. Preventing PTSD with oxytocin: effects of oxytocin administration on fear neurocircuitry and PTSD symptom development in recently trauma-exposed individuals. Eur J Psychotraumatol. (2017) 8:1302652. 10.1080/20008198.2017.1302652 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.van Zuiden M, Frijling JL, Nawijn L, Koch SBJ, Goslings JC, Luitse JS, et al. Intranasal oxytocin to prevent posttraumatic stress disorder symptoms: a randomized controlled trial in emergency department patients. Biol Psychiatry (2017) 81:1030–40. 10.1016/j.biopsych.2016.11.012 [DOI] [PubMed] [Google Scholar]
  • 304.Verma D, Tasan RO, Herzog H, Sperk G. editors. The role of NPY in expression and extinction of conditioned fear. In 15th Scientific Symposium of the Austrian Pharmacological Society. London: Springer; (2009). [Google Scholar]
  • 305.Sutcliffe JG, de Lecea L. The hypocretins: setting the arousal threshold. Nat Rev Neurosci. (2002) 3:339–49. 10.1038/nrn808 [DOI] [PubMed] [Google Scholar]
  • 306.Enman NM, Sabban EL, McGonigle P, Van Bockstaele EJ. Targeting the neuropeptide Y system in stress-related psychiatric disorders. Neurobiol Stress. (2015) 1:33–43. 10.1016/j.ynstr.2014.09.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Schmeltzer SN, Herman JP, Sah R. Neuropeptide Y (NPY) and posttraumatic stress disorder (PTSD): a translational update. Exp Neurol. (2016) 284:196–210. 10.1016/j.expneurol.2016.06.020 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Laukova M, Alaluf LG, Serova LI, Arango V, Sabban EL. Early intervention with intranasal NPY prevents single prolonged stress-triggered impairments in hypothalamus and ventral hippocampus in male rats. Endocrinology (2014) 155:3920–33. 10.1210/en.2014-1192 [DOI] [PubMed] [Google Scholar]
  • 309.Maras PM, Baram TZ. Sculpting the hippocampus from within: stress, spines, and CRH. Trends Neurosci. (2012) 35:315–24. 10.1016/j.tins.2012.01.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Nimchinsky EA, Sabatini BL, Svoboda K. Structure and function of dendritic spines. Annu Rev Physiol. (2002) 64:313–53. 10.1146/annurev.physiol.64.081501.160008 [DOI] [PubMed] [Google Scholar]
  • 311.Yuste R, Bonhoeffer T. Morphological changes in dendritic spines associated with long-term synaptic plasticity. Annu Rev Neurosci. (2001) 24:1071–89. 10.1146/annurev.neuro.24.1.1071 [DOI] [PubMed] [Google Scholar]
  • 312.Han F, Xiao B, Wen L. Loss of glial cells of the hippocampus in a rat model of post-traumatic stress disorder. Neurochem Res. (2015) 40:942–51. 10.1007/s11064-015-1549-6 [DOI] [PubMed] [Google Scholar]
  • 313.Li XM, Han F, Liu DJ, Shi YX. Single-prolonged stress induced mitochondrial-dependent apoptosis in hippocampus in the rat model of post-traumatic stress disorder. J Chem Neuroanat. (2010) 40:248–55. 10.1016/j.jchemneu.2010.07.001 [DOI] [PubMed] [Google Scholar]
  • 314.Cohen S, Kozlovsky N, Matar MA, Kaplan Z, Zohar J, Cohen H. Post-exposure sleep deprivation facilitates correctly timed interactions between glucocorticoid and adrenergic systems, which attenuate traumatic stress responses. Neuropsychopharmacology (2012) 37:2388–404. 10.1038/npp.2012.94 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Golub Y, Kaltwasser SF, Mauch CP, Herrmann L, Schmidt U, Holsboer F, et al. Reduced hippocampus volume in the mouse model of Posttraumatic Stress Disorder. J Psychiatr Res. (2011) 45:650–9. 10.1016/j.jpsychires.2010.10.014 [DOI] [PubMed] [Google Scholar]
  • 316.Campbell JS, Seger R, Graves JD, Graves L, Jensen A, Krebs EG. The MAP kinase cascade. Recent Prog Horm Res. (2013) 50:131–59. 10.3892/ijmm_00000225 [DOI] [PubMed] [Google Scholar]
  • 317.Bockaert J, Marin P. mTOR in brain physiology and pathologies. Physiol Rev. (2015) 95:1157–87. 10.1152/physrev.00038.2014 [DOI] [PubMed] [Google Scholar]
  • 318.Cargnello M, Roux PP. Activation and function of the MAPKs and their substrates, the MAPK-activated protein kinases. Microbiol Mol Biol Rev. (2011) 75:50–83. 10.1128/MMBR.00031-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Thomas GM, Huganir RL. MAPK cascade signalling and synaptic plasticity. Nat Rev Neurosci. (2004) 5:173. 10.1038/nrn1346 [DOI] [PubMed] [Google Scholar]
  • 320.Wang H-T, Han F, Gao J-L, Shi Y-X. Increased phosphorylation of extracellular signal-regulated kinase in the medial prefrontal cortex of the single-prolonged stress rats. Cell Mol Neurobiol. (2010) 30:437–44. 10.1007/s10571-009-9468-1 [DOI] [PubMed] [Google Scholar]
  • 321.Wen L, Han F, Shi Y, Li X. Role of the endoplasmic reticulum pathway in the medial prefrontal cortex in post-traumatic stress disorder model rats. J Mol Neurosci. (2016) 59:471–82. 10.1007/s12031-016-0755-2 [DOI] [PubMed] [Google Scholar]
  • 322.Xiao B, Han F, Wang H-T, Shi Y-X. Single-prolonged stress induces increased phosphorylation of extracellular signal-regulated kinase in a rat model of post-traumatic stress disorder. Mol Med Rep. (2011) 4:445–9. 10.3892/mmr.2011.459 [DOI] [PubMed] [Google Scholar]
  • 323.Liu H, Li H, Xu A, Kan Q, Liu B. Role of phosphorylated ERK in amygdala neuronal apoptosis in single-prolonged stress rats. Mol Med Rep. (2010) 3:1059–63. 10.3892/mmr.2010.362 [DOI] [PubMed] [Google Scholar]
  • 324.Hafezi-Moghadam A, Simoncini T, Yang Z, Limbourg FP, Plumier JC, Rebsamen MC, et al. Acute cardiovascular protective effects of corticosteroids are mediated by non-transcriptional activation of endothelial nitric oxide synthase. Nat Med. (2002) 8:473–9. 10.1038/nm0502-473 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Moosavi M, Maghsoudi N, Zahedi-Asl S, Naghdi N, Yousefpour M, Trounce IA. The role of PI3/Akt pathway in the protective effect of insulin against corticosterone cell death induction in hippocampal cell culture. Neuroendocrinology (2008) 88:293–8. 10.1159/000150441 [DOI] [PubMed] [Google Scholar]
  • 326.Fifield K, Hebert M, Angel R, Adamec R, Blundell J. Inhibition of mTOR kinase via rapamycin blocks persistent predator stress-induced hyperarousal. Behav Brain Res. (2013) 256:457–63. 10.1016/j.bbr.2013.08.047 [DOI] [PubMed] [Google Scholar]
  • 327.Blundell J, Kouser M, Powell CM. Systemic inhibition of mammalian target of rapamycin inhibits fear memory reconsolidation. Neurobiol Learn Mem. (2008) 90:28–35. 10.1016/j.nlm.2007.12.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Chen C, Ji M, Xu Q, Zhang Y, Sun Q, Liu J, et al. Sevoflurane attenuates stress-enhanced fear learning by regulating hippocampal BDNF expression and Akt/GSK-3beta signaling pathway in a rat model of post-traumatic stress disorder. J Anesthesia (2015) 29:600–8. 10.1007/s00540-014-1964-x [DOI] [PubMed] [Google Scholar]
  • 329.Sui ZX, Liu H, Wang HT, Yang ZL, Liu JG, Xu AJ, et al. [Alteration of apoptosis and Akt/mTOR signal pathway in hippocampal neurons of rat with post-traumatic stress]. Sichuan Da Xue Xue Bao Yi Xue Ban (2014) 45:221–4. Available online at: http://scdx.cnjournals.com/scdxxbyxben/ch/reader/view_abstract.aspx?file_no=20140209&flag=1 [PubMed] [Google Scholar]
  • 330.Mitra R, Sapolsky RM. Acute corticosterone treatment is sufficient to induce anxiety and amygdaloid dendritic hypertrophy. Proc Natl Acad Sci U.S.A. (2008) 105:5573–8. 10.1073/pnas.0705615105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Yoshii T, Sakamoto H, Kawasaki M, Ozawa H, Ueta Y, Onaka T, et al. The single-prolonged stress paradigm alters both the morphology and stress response of magnocellular vasopressin neurons. Neuroscience (2008) 156:466–74. 10.1016/j.neuroscience.2008.07.049 [DOI] [PubMed] [Google Scholar]
  • 332.Urbanska M, Blazejczyk M, Jaworski J. Molecular basis of dendritic arborization. Acta Neurobiol Exp. (2008) 68:264. Available online at: https://ane.pl/pdf/6830.pdf [DOI] [PubMed] [Google Scholar]
  • 333.Ding J, Han F, Shi Y. Single-prolonged stress induces apoptosis in the amygdala in a rat model of post-traumatic stress disorder. J Psychiatric Res. (2010) 44:48–55. 10.1016/j.jpsychires.2009.06.001 [DOI] [PubMed] [Google Scholar]
  • 334.Zhao W, Han F, Shi Y. IRE1α pathway of endoplasmic reticulum stress induces neuronal apoptosis in the locus coeruleus of rats under single prolonged stress. Progr NeuroPsychopharmacol Biol Psychiatry (2016) 69:11–8. 10.1016/j.pnpbp.2016.03.008 [DOI] [PubMed] [Google Scholar]
  • 335.Han F, Yan S, Shi Y. Single-prolonged stress induces endoplasmic reticulum-dependent apoptosis in the hippocampus in a rat model of post-traumatic stress disorder. PLoS ONE (2013) 8:e69340. 10.1371/journal.pone.0069340 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Wen L, Xiao B, Shi Y, Han F. PERK signalling pathway mediates single prolonged stress-induced dysfunction of medial prefrontal cortex neurons. Apoptosis : an international journal on programmed cell death. (2017) 22:753–68. 10.1007/s10495-017-1371-5 [DOI] [PubMed] [Google Scholar]
  • 337.Xiao B, Yu B, Wang H-t, Han F, Shi Y-x. Single-prolonged stress induces apoptosis by activating cytochrome C/caspase-9 pathway in a rat model of post-traumatic stress disorder. Cell Mol Neurobiol. (2011) 31:37–43. 10.1007/s10571-010-9550-8 [DOI] [PubMed] [Google Scholar]
  • 338.Su YA, Wu J, Zhang L, Zhang Q, Su DM, He P, et al. Dysregulated mitochondrial genes and networks with drug targets in postmortem brain of patients with Posttraumatic Stress Disorder (PTSD) revealed by human mitochondria-focused cDNA microarrays. Int J Biol Sci. (2008) 4:223–35. 10.7150/ijbs.4.223 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 339.Gill JM, Saligan L, Woods S, Page G. PTSD is associated with an excess of inflammatory immune activities. Perspect Psychiatr Care (2009) 45:262–77. 10.1111/j.1744-6163.2009.00229.x [DOI] [PubMed] [Google Scholar]
  • 340.Najjar S, Pearlman DM, Alper K, Najjar A, Devinsky O. Neuroinflammation and psychiatric illness. J Neuroinflamm. (2013) 10:1. 10.1186/1742-2094-10-43 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Pace TWW, Wingenfeld K, Schmidt I, Meinlschmidt G, Hellhammer DH, Heim CM. Increased peripheral NF-κB pathway activity in women with childhood abuse-related posttraumatic stress disorder. Brain Behav Immunity (2012) 26:13–7. 10.1016/j.bbi.2011.07.232 [DOI] [PubMed] [Google Scholar]
  • 342.von Känel R, Hepp U, Kraemer B, Traber R, Keel M, Mica L, et al. Evidence for low-grade systemic proinflammatory activity in patients with posttraumatic stress disorder. J Psychiatr Res. (2007) 41:744–52. 10.1016/j.jpsychires.2006.06.009 [DOI] [PubMed] [Google Scholar]
  • 343.Spitzer C, Barnow S, Völzke H, Wallaschofski H, John U, Freyberger HJ, et al. Association of posttraumatic stress disorder with low-grade elevation of C-reactive protein: evidence from the general population. J Psychiatr Res. (2010) 44:15–21. 10.1016/j.jpsychires.2009.06.002 [DOI] [PubMed] [Google Scholar]
  • 344.Gill JM, Saligan L, Lee H, Rotolo S, Szanton S. Women in recovery from PTSD have similar inflammation and quality of life as non-traumatized controls. J Psychos Res. (2013) 74:301–6. 10.1016/j.jpsychores.2012.10.013 [DOI] [PubMed] [Google Scholar]
  • 345.Silverman MN, Sternberg EM. Glucocorticoid regulation of inflammation and its functional correlates: from HPA axis to glucocorticoid receptor dysfunction. Ann N Y Acad Sci. (2012) 1261:55–63. 10.1111/j.1749-6632.2012.06633.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 346.Peng Z, Wang H, Zhang R, Chen Y, Xue F, Nie H, et al. Gastrodin ameliorates anxiety-like behaviors and inhibits IL-1 [beta] level and p38 MAPK Phosphorylation of hippocampus in the rat model of posttraumatic stress disorder. Physiol Res. (2013) 62:537. Available online at: http://www.biomed.cas.cz/physiolres/pdf/62/62_537.pdf [DOI] [PubMed] [Google Scholar]
  • 347.Sun R, Zhao Z, Feng J, Bo J, Rong H, Lei Y, et al. Glucocorticoid-potentiated spinal microglia activation contributes to preoperative anxiety-induced postoperative hyperalgesia. Mol Neurobiol. (2016) 54:4316–28. 10.1007/s12035-016-9976-1 [DOI] [PubMed] [Google Scholar]
  • 348.Sun R, Zhang Z, Lei Y, Liu Y, Lu C, Rong H, et al. Hippocampal activation of microglia may underlie the shared neurobiology of comorbid posttraumatic stress disorder and chronic pain. Mol Pain. (2016) 12. 10.1177/1744806916679166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 349.Haim LB, Rowitch DH. Functional diversity of astrocytes in neural circuit regulation. Nat Rev Neurosci. (2017) 18:31. 10.1038/nrn.2016.159 [DOI] [PubMed] [Google Scholar]
  • 350.Albrecht A, Ivens S, Papageorgiou IE, Caliskan G, Saiepour N, Bruck W, et al. Shifts in excitatory/inhibitory balance by juvenile stress: a role for neuron-astrocyte interaction in the dentate gyrus. Glia (2016) 64:911–22. 10.1002/glia.22970 [DOI] [PubMed] [Google Scholar]
  • 351.Xia L, Zhai M, Wang L, Miao D, Zhu X, Wang W. FGF2 blocks PTSD symptoms via an astrocyte-based mechanism. Behav Brain Res. (2013) 256:472–80. 10.1016/j.bbr.2013.08.048 [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials


Articles from Frontiers in Psychiatry are provided here courtesy of Frontiers Media SA

RESOURCES