Skip to main content
Neuropsychopharmacology logoLink to Neuropsychopharmacology
. 2015 Jul 29;41(1):58–79. doi: 10.1038/npp.2015.180

Stress and Fear Extinction

Stephen Maren 1,*, Andrew Holmes 2
PMCID: PMC4677122  PMID: 26105142

Abstract

Stress has a critical role in the development and expression of many psychiatric disorders, and is a defining feature of posttraumatic stress disorder (PTSD). Stress also limits the efficacy of behavioral therapies aimed at limiting pathological fear, such as exposure therapy. Here we examine emerging evidence that stress impairs recovery from trauma by impairing fear extinction, a form of learning thought to underlie the suppression of trauma-related fear memories. We describe the major structural and functional abnormalities in brain regions that are particularly vulnerable to stress, including the amygdala, prefrontal cortex, and hippocampus, which may underlie stress-induced impairments in extinction. We also discuss some of the stress-induced neurochemical and molecular alterations in these brain regions that are associated with extinction deficits, and the potential for targeting these changes to prevent or reverse impaired extinction. A better understanding of the neurobiological basis of stress effects on extinction promises to yield novel approaches to improving therapeutic outcomes for PTSD and other anxiety and trauma-related disorders.

INTRODUCTION

After all, when a stone is dropped into a pond, the water continues quivering even after the stone has sunk to the bottom

Arthur Golden, Memoirs of a Geisha

Trauma- and stress-related disorders, including posttraumatic stress disorder (PTSD) (DSM-5, 2013), are among the most prevalent and debilitating neuropsychiatric disorders in the world. These disorders not only seriously undermine the mental health of affected individuals but also levy an enormous public health and economic burden on society at large. The numbers are staggering. In the United States, for example, PTSD has an adult lifetime prevalence of 8% women (10%) are diagnosed at twice the rate of men (5%) (Kessler et al, 1995, 2005). In the United States alone, nearly 25 million people (roughly the population of Texas) will develop PTSD at some point in their lives. Shockingly, the rate of PTSD doubles in military veterans. In the last decade, the prevalence of PTSD in soldiers deployed to Iraq and Afghanistan was nearly 14% and the cost of treating these individuals alone is estimated to be over $3 billion USD per year (Ramchand et al, 2008). Children and adolescents also succumb to PTSD after trauma at rates somewhat lower than that experienced by adults (Kessler et al, 2012).

Given the exceptional individual and societal costs of PTSD, there has been a considerable effort aimed at understanding the psychological and neurobiological mechanisms underlying this disorder (Liberzon and Sripada, 2008; Mahan and Ressler, 2012; Pitman et al, 2012; Goswami et al, 2013; Rau et al, 2005). This effort has led to enormous gains in understanding not only the brain circuits mediating stress and fear responses in the face of threat, but also those that are involved in dampening fear and anxiety once threats have passed (Ehrlich et al, 2009; Pape and Pare, 2010; Milad and Quirk, 2012; Duvarci and Paré, 2014). Indeed, harnessing inhibitory brain circuits to dampen fear in the aftermath of trauma may have a central role in therapeutic interventions for PTSD, including cognitive-behavioral therapies. It is widely believed that promoting the function of these inhibitory brain circuits is critical to developing novel therapeutic interventions for PTSD and other trauma- and stress-related disorders (Yehuda and LeDoux, 2007; Milad and Quirk, 2012; Pitman et al, 2012; Singewald et al, 2015).

Yet, despite the promise of new therapeutic avenues for treating PTSD, emerging evidence suggests that the heightened stress that accompanies PTSD may hinder the function of fear-dampening inhibitory circuits that are central to behavioral interventions designed to alleviate the syndrome. In this way, trauma-related stress is not only the ‘stone dropped in the pond' that drives the development of PTSD but is also an enduring emotional state that ‘continues quivering' through the mind and body to maintain PTSD long after trauma.

In the present review, we will consider the neural mechanisms underlying behavioral interventions for trauma- and stress-related disorders, and the modulation of these processes by stress. Our goal is to understand not only how brain circuits involved in the regulation of fear become compromised by stress but also how the impact of stress might be ameliorated to achieve a substantial and lasting therapeutic outcome in patients with PTSD. We rely heavily on work conducted in animal models to elaborate the nature and properties of brain circuits involved in emotional regulation but, where possible, translate this to work in normal human subjects and PTSD patients.

LEARNING AND MEMORY PROCESSES IN TRAUMA-RELATED DISORDERS

Common to all trauma- and stress-related disorders, including PTSD, is the direct or indirect experience (or threat) of death, serious injury, or sexual violence—in other words, experiencing or witnessing a traumatic event precedes the development of PTSD (DSM-5, 2013). Of course, not all individuals that experience or witness trauma develop PTSD—only 10% of individuals that experience trauma will ultimately be diagnosed with PTSD (Kessler et al, 1995; Yehuda and LeDoux, 2007). However, for those individuals who develop PTSD, the memory of the traumatic experience—the sights, sounds, smells, and context of the trauma—has a central role in both the development and expression of the disorder (Holmes and Singewald, 2013; Maren et al, 2013). Critically, trauma-related stimuli serve as haunting, distressing, and intrusive reminders that force PTSD patients to relive the trauma in waking flashbacks and nightmares as they sleep. Moreover, memories of the trauma lead to avoidance behavior wherein those suffering with PTSD avoid contact with individuals, situations, or places that might recall memories of their traumatic experience. In this way, traumatic memories intrude on everyday life and reap immense suffering in PTSD patients.

Therefore, unlike many neuropsychiatric disorders, PTSD is anchored to particular events and experiences in time. As a result, brain systems important for learning and memory have an essential role in the development and expression of the disorder. In the laboratory, the learning and memory processes that contribute to PTSD can be modeled, at least in part, using aversive learning paradigms such as Pavlovian fear conditioning.

TARGETING MEMORIES AS A THERAPEUTIC INTERVENTION

Fear conditioning is a form of associative learning with a long history as a model of emotional learning and memory processes in both animals and humans (Davis, 1992; Fendt and Fanselow, 1999; LeDoux, 2000; Maren, 2001; Sotres-Bayon et al, 2006). For example, in the early part of the 20th century, Watson and Rayner (1920) demonstrated that fear could be learned using classical conditioning procedures in a young boy referred to as ‘Little Albert'. After presenting Little Albert with a white rabbit, Watson hammered a suspended steel bar to produce a frighteningly loud noise that caused Little Albert to tremble and cry. After several pairings of the rabbit and the noise, Albert became visibly upset at the sight of the rabbit alone and generalized his ‘conditioned emotional reaction' of the rabbit to other white, furry objects (Watson and Rayner, 1920; Maren, 2001). Watson and Rayner (1920) concluded ‘it is probable that many of the phobias in psychopathology are true conditioned emotional reactions'.

Twenty years later, Estes and Skinner (1941) developed a quantitative method to measure ‘conditioned anxiety states' in rats. They first trained rats in an instrumental procedure to press a lever for a food reinforcer. Once lever-pressing behavior was established, they performed a Pavlovian fear-conditioning procedure in which an auditory conditioned stimulus (CS, a tone) was paired with an aversive electric footshock unconditioned stimulus (US). After fear conditioning, they assessed whether presentation of the tone alone interrupted lever-pressing for food. Indeed, tone presentation resulted in substantial decreases in lever pressing. Estes and Skinner (1941) inferred that lever pressing was interrupted by a conditioned state of anxiety that competed with the motivation to seek food. Importantly, they also found that continuous presentation of the tone in the absence of shock led to extinction, ie, the tone's ability to decrease lever pressing became weaker with CS-alone exposure. However, the extinction of lever pressing was weak and anxiety to the CS returned over the course of 24 h—a phenomenon termed spontaneous recovery.

These early studies revealed that emotional responses and states, including fear and anxiety, could be learned, and supported the notion that the acquisition of fear through conditioning processes might underlie a variety of disorders including specific phobia and PTSD. Indeed, the ‘conditioning model of anxiety disorders' became central to the development of behavioral therapies for anxiety disorders. For example, Wolpe (1968) contended that the relief of anxiety centered on the ‘deconditioning' of learned fears that underlie states of anxiety in both animals and humans. To this end, Wolpe (1968) developed systematic desensitization, a form of cognitive behavioral therapy that used relaxation methods to inhibit a patient's conditioned anxiety or fear. From a learning theory perspective, systematic desensitization relies on both extinction and counterconditioning, two processes that involve new learning (eg, CS–‘no aversive US' and CS–‘appetitive US', respectively), which interferes with the expression of conditioned fear. Similarly, exposure therapy, which is an effective treatment for many anxiety, trauma- and stress-related disorders, relies on extinction learning to reduce conditioned fear responses to stimuli that provoke anxiety and panic (Bouton et al, 2001; Rothbaum and Davis, 2003; Craske et al, 2008).

THE FRAGILITY OF EXTINCTION

It has long been understood that extinction is less durable than conditioning. Indeed, in his seminal studies of conditioning and extinction learning, Pavlov (1927) described several instances in which conditional responding returned after extinction, including spontaneous recovery and external disinhibition (Figure 1). In the former case, CRs to an extinguished CS returned with the mere passage of time, whereas in the latter case the presentation of novel stimuli reinstated extinguished responding. Consequently, the effectiveness of extinction-based therapies, including exposure therapy, is constrained by phenomena that limit the durability of extinction (Vervliet et al, 2013; Goode and Maren, 2014). In addition to spontaneous recovery and external disinhibition, extinguished CRs have also been found to exhibit renewal when the CS is encountered outside the place or ‘context' in which extinction training was administered, as well as reinstatement following presentation of the US. These recovery phenomena indicate that extinction does not eliminate the conditioning memory. Rather, it generates a new ‘inhibitory' memory (eg, CS–‘no US') that competes with the expression of the conditioning memory (Konorski, 1967; Bouton, 1993). This fragility of extinction memories suggests they might be vulnerable to stress, and as we discuss later there is emerging evidence in both animals and humans indicating just that.

Figure 1.

Figure 1

Relapse of fear after extinction. After the extinction of a conditioned fear response, extinction, the expression of conditioned fear is reduced. However, a number of phenomena are associated with the return or relapse of fear responses after extinction. These include (a) spontaneous recovery with the passage of time, (b) external disinhibition after presentation of a novel stimulus, (c) reinstatement after experiencing a noxious event, including the unconditioned stimulus, and (d) renewal after experiencing the conditioned stimulus outside the extinction context.

BRAIN CIRCUITS FOR FEAR CONDITIONING AND EXTINCTION

Before considering stress effects on extinction learning, it is important to review the neural circuitry underlying this form of learning. Research over several decades has yielded an impressive corpus of work identifying brain regions and circuits involved in fear conditioning and extinction (Fendt and Fanselow, 1999; LeDoux, 2000; Myers and Davis, 2002; Maren and Quirk, 2004; Quirk and Mueller, 2008; Mahan and Ressler, 2012; Duvarci and Paré, 2014; Herry and Johansen, 2014). Importantly, these circuits are not static, but change over time as memories are consolidated, retrieved, reconsolidated, and updated (Maren, 2011).

FEAR, EXTINCTION, AND THE AMYGDALA

At the heart of the emotional learning circuit is the amygdala, a group of heterogeneous nuclei that has an essential role in both fear conditioning and extinction (Maren, 2001; Myers and Davis, 2002; Ehrlich et al, 2009; Pape and Pare, 2010; Duvarci and Paré, 2014). Within the amygdala, the basolateral complex (BLA, including the lateral, basolateral, and basomedial nuclei) has been implicated as a critical site of sensory convergence for CSs (eg, tones, odors, or lights), USs (eg, loud noises, orbital shock, or foot shock), as well as contextual information related to the circumstances surrounding the conditioning experience (Herry and Johansen, 2014). The anatomical pathways conveying this information have been well described and involve both cortical and subcortical routes of transmission. For example, auditory and somatosensory information reach many amygdaloid nuclei, but the shortest-latency single-unit responses are obtained in the lateral nucleus via direct projections from the sensory thalamus (eg, medial geniculate nucleus and posterior intralaminar nuclei) (Bordi and LeDoux, 1994). Cortical areas, including the primary auditory cortex and the perirhinal cortex, also convey sensory information to the amygdala and these areas are sufficient for fear conditioning under some conditions (Medina et al, 2002).

Most circuit models of fear conditioning posit that sensory convergence and associative synaptic plasticity in the BLA is necessary for many forms of aversive learning and memory. In particular, NMDA receptor-dependent long-term potentiation (LTP) in thalamic and cortical afferents on BLA neurons is critical for fear learning and memory (Maren, 1999; Blair et al, 2001; Johansen et al, 2011). For instance, early work showed that intra-BLA infusions of the NMDA receptor antagonist D,L-APV (which antagonizes both GluN2A- and GluN2B-containing receptors) prevent the acquisition and expression of fear conditioning (Miserendino et al, 1990; Maren et al, 1996; Lee and Kim, 1998). In addition, intra-BLA APV also prevents the extinction and reconsolidation of fear (Falls et al, 1992; Ben Mamou et al, 2006; Zimmerman and Maren, 2010), indicating a broad role for BLA NMDA receptors in fear memory processes. The fear expression deficits produced by APV suggest that decreases in BLA excitability (rather than LTP per se) account for deficits in these processes (Maren and Fanselow, 1995; Maren et al, 1996). Interestingly, fear expression deficits are not produced by systemic administration of NMDA receptor antagonists, such as Ro 25-6981 and ifenprodil, with selective actions at NMDA receptors containing the GluN2B subunit, whereas these drugs do reliably impair conditioning, extinction, and reconsolidation (Rodrigues et al, 2001; Ben Mamou et al, 2006; Sotres-Bayon et al, 2007; Mathur et al, 2009). This suggests GluN2B-containing receptors may have a preferential role in the induction of synaptic plasticity critical for the conditioning, extinction, and reconsolidation of fear memories, whereas other types of NMDA receptor heteromer (eg, containing GluN2A not GluN2B) may have a more general role in regulating excitability and therefore fear expression.

Importantly, the NMDA receptor-dependent processes important for emotional learning and memory are anatomically localized to BLA neurons insofar as infusions of NMDA receptor antagonists into the central nucleus of the amygdala (CeA) spare both fear conditioning and extinction (Zimmerman and Maren, 2010). The importance of BLA synaptic plasticity in conditioned fear is further supported by the observation that AMPA receptors, which are involved in LTP expression, are upregulated in LA neurons after fear conditioning (Rumpel, 2005). Moreover, overexpression of CREB, a transcription factor that promotes synaptic plasticity, in LA neurons increases the probability that they will be incorporated into cellular networks required for the conditioning memory (Han et al, 2007). In addition to supporting fear learning, synaptic plasticity in the amygdala is critical for extinction learning.

As noted, extinction is impaired by antagonizing NMDA receptors in the BLA and can be enhanced by NMDA receptor modulators, such as d-cycloserine (Walker et al, 2002). In addition, antagonizing brain-derived neurotrophic factor (BDNF) signaling in the amygdala retards the extinction retention (Chhatwal et al, 2006). Hence, the intracellular cascades associated with extinction appear to recapitulate, at least in part, those involved in fear conditioning (Tronson et al, 2012). For example, extracellular-related kinase (ERK)/mitogen-activated protein kinase (MAPK) activity in the BLA is important for both forms of learning (Herry et al, 2006; Merino and Maren, 2006). However, the two forms of learning are likely to involve different degrees or types of plasticity at excitatory and inhibitory synapses (Maren, 2014a). Indeed, extinction appears to involve plasticity at GABAergic synapses in the amygdala that are critical for fear suppression (Chhatwal et al, 2005; Trouche et al, 2013).

EXTINCTION, CONTEXT, AND THE HIPPOCAMPUS

Although there is ample evidence that synaptic plasticity in the amygdala is critical for encoding conditioning and extinction memories, the amygdala does not act alone. Contexts have multiple roles in the conditioning and extinction of fear (Maren et al, 2013). Animals quickly come to learn that aversive stimuli occur in a particular context: for example, explicitly pairing footshock with a context produces a conditioned fear response to that context. This direct context–US association, similar to a CS–US association, is mediated by associative plasticity in the amygdala among hippocampal efferents to the basolateral and basomedial amygdaloid nuclei. In addition, the BLA modulates contextual encoding by the hippocampus, influencing the strength of memories animals form about the context itself (Huff and Rudy, 2004). Furthermore, projections from the BLA to the ventral hippocampus and entorhinal cortex (the major source of cortical afferents to the hippocampus) modulate both nonassociative (Felix-Ortiz et al, 2013) and associative fear (Sparta et al, 2014). Hence, bidirectional communication between the amygdala and hippocampus is important in establishing contextual representations.

There is considerable evidence indicating that the hippocampus is critical for establishing contextual representations during fear conditioning and extinction (Fanselow, 2000; Maren, 2001; Rudy et al, 2004). Contexts in this regard refer to both exteroceptive and interoceptive information that sets the backdrop for the conditioning and extinction experiences, ie, contexts inform when, where, with whom, and in what internal state an experience occurs. On the one hand, disrupting hippocampal function impairs the formation of contextual representations that are important for contextual fear conditioning. On the other hand, the hippocampus is not essential for contextual fear conditioning in all cases. Animals with pre-training hippocampal lesions acquire conditional freezing to a context normally under some conditions, despite the fact that posttraining lesions are devastating to performance (Maren et al, 1997; Frankland et al, 1998). Moreover, animals that have encoded and consolidated representations of a to-be-conditioned context exhibit normal context conditioning in the absence of the hippocampus (Young et al, 1994). Thus, spared context conditioning after hippocampal damage appears to be mediated by alternate neural systems (Maren et al, 1997; Goshen et al, 2011); however, if detailed contextual memories are acquired by an intact hippocampal system, the retrieval of those representations is reliant on the hippocampus (Rudy, 2009; Sparks et al, 2011). In other words, the hippocampus appears necessary for the encoding and retrieval of detailed contextual representations, whereas other neural systems can mediate the encoding and retrieval of elemental features of context.

Contexts also serve as retrieval cues for specific CS–US relationships learned in those contexts and serve to ‘set the occasion' for these US relationships (Bouton, 1993). This occasion setting function of context is in particular important for extinction, which is often learned in a context different from that in which conditioning is conducted. In a typical experiment, eg, conditioning (CS–US pairings) would be conducted in one context (‘context A') and the following day extinction (CS-alone presentation) would be delivered in a distinct context (‘context B'). This is often done so that fear to the CS can be assessed during extinction training without contamination by fear to the conditioning context, ie, when extinction is performed in the conditioning context, fear is elevated on placement in the context before CS presentation. In this way, conditioning and extinction (represented by CS–US and CS–no US memories, respectively) can be disambiguated by the contexts in which they occurred. Similar to fear memories, extinction memories are also tightly bound to the context in which they are learned, such that when an extinguished CS is encountered outside the extinction context, fear will return—a phenomenon that accounts for fear renewal.

There is compelling evidence indicating that the hippocampus is important for the use of contexts to guide retrieval of fear extinction memories that have been formed in particular contexts, eg, hippocampal inactivation impairs fear renewal (Maren and Holt, 2000; Maren et al, 2013; Preston and Eichenbaum, 2013). Recent work has also revealed that the context-dependent expression of fear memories after extinction involves hippocampal projections to the amygdala and the medial prefrontal cortex (mPFC). Specifically, neuronal activity (as indexed by expression of the immediate early gene product, Fos) is elevated in hippocampal neurons projecting to the mPFC and the amygdala during fear renewal (Knapska and Maren, 2009). This renewal-related elevation in Fos activity is in particular pronounced in a small population of hippocampal neurons that projects to both the mPFC and BLA (Jin and Maren, 2015). Demonstrating the importance of these pathways to the contextual regulation of fear memories, disconnection of the hippocampus from the mPFC and BLA prevents fear renewal (Orsini et al, 2011). Further understanding of these and other neural circuits should help shed light on the basis of the abnormal contextual gating of fear and extinction which, as we discuss later, is evident in patients with PTSD.

mPFC GATING OF FEAR AND EXTINCTION

A growing literature has documented how the mPFC is critical for gating fear expression, in particular after extinction (Quirk and Mueller, 2008; Rozeske et al, 2015). Early work revealed that lesions of the mPFC, in particular those encompassing the infralimbic cortex (IL), produced deficits in the retention of extinction (Morgan and LeDoux, 1995; Quirk et al, 2000). In these studies, there was substantial spontaneous recovery of conditioned fear responses the day after extinction training in rats with mPFC damage. The effects of mPFC lesions on extinction are most robust in paradigms, including conditioned suppression, in which animals are challenged with competing motivational states, whereas the evidence that mPFC lesions affect extinction retrieval in tasks without ongoing appetitive behavior is less clear-cut (Gewirtz et al, 1997; Garcia et al, 2006; Chang and Maren, 2010b). Nonetheless, pharmacological (Hugues et al, 2006; Sotres-Bayon et al, 2007; Laurent and Westbrook, 2008; 2009; Sierra-Mercado et al, 2011), electrical (Milad et al, 2004) or optogenetic (Do-Monte et al, 2015a) manipulations of IL have been found to modulate the acquisition of extinction. In addition, changes in IL spike firing and immediate early gene expression accompany both the acquisition (Milad and Quirk, 2002; Muigg et al, 2008; Fitzgerald et al, 2014b) and expression (Hefner et al, 2008; Knapska and Maren, 2009; Whittle et al, 2010; Chang and Maren, 2010b; Knapska et al, 2012; Orsini et al, 2013) of extinction.

The prevailing view based on these data is that extinction-related plasticity in the IL is involved in establishing extinction memories (Myers and Davis, 2002; Orsini and Maren, 2012). It has been suggested that the IL might gate extinction memories by exciting inhibitory intercalated neurons (ITC) to suppress amygdala output (Quirk et al, 2003; Paré et al, 2004; Likhtik et al, 2008; Ehrlich et al, 2009; Pape and Pare, 2010; Duvarci and Paré, 2014). However, recent findings indicate that the primary functional target of the IL is the BLA rather than the ITCs—consequently, the IL likely modulates amygdala output through the BLA, which in turn recruits the ITCs to suppress fear (Cho et al, 2013; Orsini et al, 2011; Knapska et al, 2012; Strobel et al, 2015). An instructional role for the IL, one that entails guiding plastic changes in the BLA that ultimately underlie extinction memories, would be consistent with recent optogenetics experiments showing the IL and IL inputs to the BLA are necessary for the acquisition of a long-term extinction memory, but is not required for the expression of extinction once learned (Do-Monte et al, 2015a; Bukalo et al, 2015). The IL is not dispensable for extinction expression in all circumstances, however, and may be particularly important when contextual information must be integrated to retrieve an extinction memory (Laurent and Westbrook, 2008; Orsini et al, 2011; Knapska et al, 2012).

In contrast to the IL, the role of the prelimbic cortex (PL) is to drive the expression of conditional fear (Corcoran and Quirk, 2007; Sotres-Bayon et al, 2012). For example, neuronal activity in the PL is positively correlated with conditional freezing behavior (Burgos-Robles et al, 2009) and PL neurons exhibit increases in Fos expression after fear retrieval (Knapska and Maren, 2009; Orsini et al, 2013). PL interneurons have a crucial role in shaping fear responses, such that reduced activity of parvalbumin-positive interneurons in the PL and anterior cingulate cortex (ACC) leads to the disinhibition of principal cell output to the BLA and an increase in conditional fear (Courtin et al, 2014). PL neurons that project to the BLA are also involved in the context-dependent renewal of extinguished fear (Orsini et al, 2011; Knapska et al, 2012). Interestingly, the role of the PL-BLA pathway in conditional fear expression is highly time dependent. Although recently acquired fear can be attenuated by optogenetically silencing PL inputs to the BLA, PL projections to the paraventricular nucleus of the thalamus must be silenced to inhibit older fear memories (Do-Monte et al, 2015b). A final important feature of the PL-BLA and IL-BLA circuits mediating fear and extinction is that they are anatomically and functionally reciprocal. Neuronal projections from the BLA to the PL and IL are recruited to generate fear responses and extinction memories, respectively (Garcia et al, 1999; Senn et al, 2014). The impressive corpus of data dissecting the prefrontal, hippocampal, and amygdalar circuits that subserve extinction in rodents has greatly informed our understanding of vulnerabilities that may contribute to extinction impairments, such as those associated with stress.

STRESS EFFECTS ON EXTINCTION AND UNDERLYING BRAIN CIRCUITS

Stress has far reaching effects on cognition and emotion in both humans and animals. In the context of trauma-related disorders, such as PTSD, stress may hinder interventions, including exposure therapy, aimed at reducing fear. Indeed, ample evidence reveals that patients with PTSD exhibit extinction impairments and stress exposure impairs extinction in both animals and humans.

STRESS EFFECTS ON EXTINCTION: HUMAN STUDIES

A history of exposure to stress increases risk for developing PTSD after traumas such as combat (Karstoft et al, 2015). An abundant literature indicates that stress influences the acquisition and retrieval of extinction in humans (Raio and Phelps, 2015). For example, stress exposure (eg, cold pressor) impairs the extinction (Hartley et al, 2014) of a conditioned galvanic skin response. Moreover, stress exposure before retrieval testing impairs the expression of extinction, resulting in a return of conditional responding (Merz et al, 2014; Raio et al, 2014). Patients with PTSD also exhibit impairments in the acquisition and expression of extinction (Blechert et al, 2007; Guthrie and Bryant, 2006; Lissek et al, 2005; Orr et al, 2000; VanElzakker et al, 2014; Wessa and Flor, 2007). After fear conditioning, PTSD patients show heightened conditional responding during extinction training on a variety of psychophysiological measures (eg, heart rate, skin conductance, and acoustic startle) (Blechert et al, 2007; Orr et al, 2000). The enhanced conditional responding during extinction is often manifest as a loss of differential responding (ie, similarly elevated responses to both the CS+ and CS− relative to controls), which indicates that extinction impairments might reflect overgeneralization of fear (Grillon and Morgan, 1999). In addition to exhibiting larger CRs during extinction training, PTSD patients also have difficulty maintaining extinguished responding over a retention interval, consistent with a deficit in extinction retrieval (Milad et al, 2008, 2009). Whether extinction impairments precede PTSD or are a pre-existing trait is unclear. On the one hand, extinction deficits were found in combat-exposed PTSD patients but not their combat-exposed twins who did not have PTSD, suggesting poor extinction was a consequence rather than a cause of PTSD (Milad et al, 2008). On the other hand, other work has found impairments in extinction learning are not only present before trauma but predict later risk for developing PTSD (Guthrie and Bryant, 2006).

Another important consideration in dissecting the etiology of extinction impairments in PTSD is the fact that women are more than twice as likely as men to develop the disorder (Glover et al, 2015; Shansky, 2015). Women with PTSD also acquire greater levels of conditioned fear than men (Inslicht et al, 2013). Although it is uncertain whether this reflects a premorbid trait in women or a differential sex-related response to the disorder, the latter seems more likely given conditional responding is typically greater in healthy male than in female subjects (Milad et al, 2010, but see Bentz et al, 2013).

The relationship between sex, extinction, and PTSD is also complex. There are recent reports of similar extinction of fear in healthy male and female subjects (Bentz et al, 2013; Lebron-Milad et al, 2012; Merz et al, 2012; Shvil et al, 2014). However, sex differences in extinction retrieval can manifest as a function of hormonal contraceptive usage, estrogen status, and menstrual cycle stage in rats (Graham and Milad, 2013, 2014) and also humans, such that men and early-stage women exhibit better recall than mid-cycle women (Milad et al, 2006). This picture of sex difference may not necessarily hold in PTSD. For instance, Shvil et al (2014) found that male, but not female, patients exhibited poor extinction retrieval in association with elevated activation of the dorsal ACC (dACC), the anatomical analogue of the rodent PL (Milad et al, 2009; Shvil et al, 2014).

This latter study implicates the dACC as a neural correlate of impaired extinction, which is consistent with the widely replicated observation of dACC hyperactivity in PTSD patients responding to a fearful stimulus (Pitman et al, 2012). Other reliable findings in PTSD have been an increased BOLD response to a fear challenge in the amygdala and a corresponding reduction in the activation of the ventromedial PFC (vmPFC, the human analogue of the rodent IL) and the hippocampus (Pitman et al, 2012). Although there have been fewer studies of neural activation specifically during extinction tasks in PTSD patients, Milad et al (2009) found impaired extinction retrieval was associated with decreased BOLD responses in the vmPFC. Thus, current data imply dysregulation of hippocampo–prefrontal–amygdala circuits in PTSD, characterized by over activity of fear-generating brain regions and a difficulty engaging circuits normally involved in the inhibition of conditional fear.

Interestingly, recent work suggests that it may not necessarily be a loss of inhibitory regulation per se that characterizes network changes in PTSD, but rather a loss of context-appropriate engagement of this inhibitory mechanism (Garfinkel et al, 2014; Maren et al, 2013; Rougemont-Bücking et al, 2011). This idea has emerged from studying PTSD patients for neural correlates of the contextual regulation of extinguished fear, a phenomenon that recruits hippocampal and prefrontal cortical circuits to regulate fear expression by the amygdala (Maren et al, 2013; Orsini and Maren, 2012). Recent work showed that PTSD patients exhibited deficits in extinction retrieval, as previously reported, but were also impaired in renewing their fear outside of the extinction context (Garfinkel et al, 2014). This is a surprising outcome insofar as a loss of inhibitory control hypothesized to accompany PTSD would be expected to increase fear to an extinguished CS not only in the extinction context but in any context the CS is encountered. Impaired renewal in PTSD suggests that patients might have hippocampal dysregulation, given the aforementioned work in rodents implicating this region in renewal. Indeed, in addition to the reduced vmPFC activity and heightened amygdala activity reported in prior studies of extinction, PTSD patients exhibited weaker hippocampal activation to the CS− during renewal, relative to combat-exposed controls (Garfinkel et al, 2014). These provocative findings suggest that a major cause of deficient extinction in PTSD patients is the inability to use contextual information to determine when and where it is appropriate to express fear.

STRESS EFFECTS ON EXTINCTION: ANIMAL MODELS

Beginning with evidence that extinction is disrupted in rats and mice subjected to forced swim or restraint (Izquierdo et al, 2006; Miracle et al, 2006), extinction deficits have been reported following exposure to various types of stressor, both acute and chronic. These include maternal separation, predator exposure, social defeat and isolation, and elevated platform exposure (for a cartoon summary, see Figure 2) (Andero et al, 2011; Chauveau et al, 2012; Clay et al, 2011; Deschaux et al, 2013; Dubreucq et al, 2012; Ganon-Elazar and Akirav, 2013; Garcia et al, 2008; Goswami et al, 2010; Green et al, 2011; Ishikawa et al, 2012; Judo et al, 2010; Knox et al, 2012a; Long and Fanselow, 2012; Matsumoto et al, 2008; Matsumoto et al, 2013; Saito et al, 2012; Saito et al, 2013; Segev et al, 2013; Toledo-Rodriguez et al, 2012; Wilber et al, 2011; Wilson et al, 2013; Yamamoto et al, 2008; Yamamoto et al, 2009; Zhang and Rosenkranz, 2013; Zheng et al, 2013; Skelly et al, 2015). Extinction deficits after stress are typically, although not always, manifest as decreases in within-session reductions in conditional responding as well as deficits in the retention of extinction across sessions (ie, extinction retrieval deficits). Echoing the aforementioned sex differences in extinction in human subjects, stress effects on extinction have also been reported to be sex dependent. For example, the same chronic (restraint) stress regimen that impairs extinction in males tends to facilitate extinction in females (Baran et al, 2009, but see (Xiong et al, 2014). Of course, more work is required to further explore sex differences in stress effects on extinction given that the majority of work has been done in males.

Figure 2.

Figure 2

Stressors and stress models affecting fear extinction in rodents. A range of established stressors, applied singly or in combination, either acutely or chronically, has been shown to cause impairments in extinction acquisition or retrieval, sometimes in association with increased fear. Abnormalities in fear and extinction are also found after rodents are subjected to the stress-enhanced fear learning (SEFL) (Rau et al, 2005), single-prolonged stress (SPL) (Yamamoto et al, 2009), and immediate-extinction deficit (IED) (Maren, 2014b) models of stress.

Recent studies have examined whether extinction is affected by exposure to stress regimens that produce other PTSD relevant phenotypes in rodents. In the ‘single prolonged stress' or SPS procedure, animals receive several stressors (restraint, forced swim, and ether anesthesia) in a single session followed by a 1-week period of quiescence. This procedure causes a dysregulation of the hypothalamic–pituitary–adrenal (HPA) axis that mirrors that seen in PTSD patients, and produces several behavioral changes that model aspects of PTSD. SPS is also associated with impairments in extinction retention (but not fear conditioning) to both shock-paired contexts (Yamamoto et al, 2008) and auditory CSs (Knox et al, 2012a). Rats experiencing SPS before fear conditioning and extinction also exhibit greater levels of fear renewal, suggestive of a stress-induced sensitization of fear responding (Knox et al, 2012a). As noted below, SPS-induced deficits in extinction can be prevented by various pharmacological interventions.

Interestingly, in cases where a stress manipulation elevates conditional fear, this fear remains extinction resistant, even when the fear response to the original stressor has itself been successfully extinguished. For example, Long and Fanselow (2012) and Rau et al (2005) have described an enhancement of fear conditioning they term ‘stress-enhanced fear learning' (SEFL) that is resistant to extinction. In this procedure, rats are first submitted to a series of 15 footshocks in a distinct context. One to 7 days later, animals receive a fear-conditioning procedure consisting of either a single unsignaled shock or tone-shock pairing in a novel context. With both procedures, prior shock exposure greatly potentiates or sensitizes the conditional fear response that is acquired during the second learning experience. However, extinguishing fear to the trauma context (the first context) does not eliminate SEFL (Long and Fanselow, 2012). Hence, the capacity of shock exposure to facilitate subsequent fear conditioning does not depend on associations between the trauma context and shock, rather it is mediated by a non-associative sensitization of fear that is resistant to extinction.

Stress-induced impairments in extinction learning are also manifest when extinction occurs soon after fear conditioning (Maren, 2014b). For example, Maren and Chang (2006) submitted rats to an extinction procedure either 15 min or 1 day after a standard, 5-trial auditory fear-conditioning session. Both groups of rats exhibited decrements in conditioned freezing during extinction training, but only rats in the delayed-extinction condition maintained this decrement in freezing over a 24-h retention interval. Rats extinguished 15 min after fear conditioning exhibited a near-complete spontaneous recovery of fear the following day; they showed little evidence of extinction retrieval. This ‘immediate extinction deficit' (IED) has been confirmed in both rats and mice after either context or cued fear conditioning (Chang and Maren, 2009; Kim et al, 2010; Macpherson et al, 2013; Stafford et al, 2013). Interestingly, Maren and Chang (2006) found that recently shocked animals exhibited high levels of sensitized fear in the minutes before extinction training (which was manifest as high baseline, pre-CS freezing behavior), suggesting that prevailing fear at the time of extinction training impairs the acquisition of long-term extinction memories (Chang and Maren, 2009). Indeed, delayed extinction is also impaired if a shock reminder is delivered minutes before extinction training (Maren and Chang, 2006). It has been argued that the IED is a prime example of stress interfering with extinction-mediating processes—in this case, the stress of recently experienced fear conditioning (Maren, 2014b). Consistent with this notion, IEDs are associated with neural abnormalities that are analogous to those produced by other, more ‘explicit' stressors, a set of findings we discuss in more detail later.

GENETIC MODERATION OF RISK FOR PTSD

Risk for stress- and trauma-related conditions is, as with other neuropsychiatric disorders, moderated by genetic factors. The results of twin and family studies estimate a significant heritable contribution to PTSD and there is an ongoing search for the specific genes involved (Hettema et al, 2003; Pitman et al, 2012). Interestingly, a number of the plausible genetic candidates identified to date encode for proteins in the same neurotransmitter and molecular systems that have been shown to regulate the effects of stress on extinction (summarized in Figure 3).

Figure 3.

Figure 3

Candidate gene variants moderating stress-related risk for posttraumatic stress disorder (PTSD). A number of human genetic variants have been found to interact with exposure to stress and trauma, often in childhood, to moderate risk for PTSD. To date, these have included the serotonin transporter (SL6A4) (Caspi et al, 2010; Xie et al, 2012), FK506-binding protein 51 (FKBP5) (Zannas and Binder, 2014), adenylate cyclase-activating polypeptide receptor (ADCYAP1) (Dias et al, 2013), and the β-adrenergic receptor (ADBR2) (Liberzon et al, 2014). Although they have not yet been studied for their potential interaction with stress, variants in the Tolloid-Like 1 Gene (TLL1) (Xie et al, 2013), retinoid-related orphan receptor alpha gene (RORA) (Logue et al, 2013) and fatty acid amide hydrolase (FAAH) (Dincheva, 2015; Gunduz-Cinar et al, 2013b) genes have been recently associated with PTSD or PTSD-related personality traits, including fear extinction and stress-reactivity.

Another notable finding to emerge has been that the influence of certain genes on risk for PTSD is in turn dependent on the amount of stress an individual has experienced in his/her life, especially during childhood. In an exemplar of this kind of interaction, variation in the FK506-binding protein 51 gene (FKBP5) predicts the occurrence of PTSD symptoms in adults as a function of the severity of abuse suffered during childhood (Binder et al, 2008). Further studies have suggested that FKBP5 likely moderates the impact of stress by regulating the function of a key stress-regulating molecule, the glucocorticoid receptor (GR) (Zannas and Binder, 2014). A related recent finding is that epigenetic modification of the GR gene (NR3C1) associates with PTSD prevalence in male Rwandan genocide survivors (Vukojevic et al, 2014). Another interesting link to the GR is the nomination of the β2-adrenergic receptor gene (ADRB2) as a moderator of the lasting effects of stress on risk for PTSD. From an examination of almost 4000 genetic candidates, Liberzon et al (2014) identified a variant in the promoter region of ADRB2 that was associated with rates of PTSD in male and female patients that had experienced childhood adversity. As we discuss at greater length later, there is growing preclinical evidence of interactions between the GR and β2-adrenergic receptor in fear extinction and its mediation by stress.

A number of genes have been linked to stress moderation of PTSD via functional effects on the extinction circuit. For instance, a polymorphism in the SLC6A4 gene, encoding the serotonin transporter, is associated with increased risk for PTSD and depression in individuals with a history of exposure to traumatic life events (Caspi et al, 2010; Xie et al, 2012). A series of elegant studies by Fisher and Hariri (2012), employing functional magnetic resonance imaging, has documented greater amygdala activation and lesser connectivity between the amygdala and vmPFC, to threat in individuals carrying the ‘risk' SLC6A4 variant. In a similar vein, a variant in the gene encoding the adenylate cyclase-activating polypeptide receptor (PAC1R, ADCYAP1) predicts PTSD in women from highly traumatized backgrounds (Dias and Ressler, 2013) in concert with heightened threat activation of the amygdala and hippocampus, and poorer functional connectivity between these two brain regions (Stevens et al, 2014). Effects on these extinction-mediating regions suggest these genes may affect extinction itself. Although additional studies will be needed to firmly establish this link, there is preliminary evidence that these and related variants affect fear and fear extinction in humans and rodents (Dias et al, 2013; Hartley et al, 2012; Lonsdorf et al, 2009).

Among the various functional neural abnormalities detected in traumatized populations, the amygdala response to threat currently represents the most credible neural marker for genetic effects on the extinction circuit. Genes found to predict variant in amygdala reactivity to threat include the aforementioned PTSD-associated FKBP5 as well as the NR3C2 gene encoding the glucocorticoid-activated mineralocorticoid (MR) receptor (Bogdan et al, 2012; White et al, 2012). Another example is the gene for fatty acid amide hydrolase (FAAH), an enzyme that controls the degradation of the endocannabinoid (eCB), anandamide. FAAH gene variation predicts not only amygdala threat reactivity and habituation, but also vmPFC–amygdala coupling, trait anxiety levels, and fear extinction (Dincheva, 2015; Gunduz-Cinar et al, 2013b; Hariri et al, 2009). These findings also have a translational dimension given, as we discuss later, the eCB system, and the amygdala FAAH activity specifically has a key role in regulating fear extinction and the effects of stress in rodents (Gunduz-Cinar et al, 2013a).

In the coming months and years, multiple novel candidate genes moderating PTSD groups will likely be nominated by large-scale genome-wide association studies of PTSD currently underway. The initial findings from these studies are already encouraging, revealing previously unknown associations between PTSD and the retinoid-related orphan receptor-αgene (RORA) (Logue et al, 2013) and genetic loci near the Tolloid-like 1 gene (TLL1) (Xie et al, 2013). This approach will be complemented by preclinical studies of PTSD-relevant phenotypes in rodents, including impaired fear extinction.

Echoing the significant genetic contribution to risk for PTSD in humans, the efficacy of extinction memory formation in rats has a sizeable (greater than one third) heritable component (Shumake et al, 2014). There have been attempts to model individual differences in fear extinction both within and between rodent strains and lines to establish neural and genetic correlates of extinction (for further discussion, see Holmes and Singewald, 2013). Some investigators have studied correlates of extinction differences that are evident within ostensibly homogeneous populations of animals within a single genetic strain (Bush et al, 2007; Shumake et al, 2014). This approach has proven valuable to defining the functional contributions of prefrontal, amygdalar, and hippocampal regions to extinction by showing how neuronal activity in these areas correlates with how well individual animals or subgroups extinguish (Burgos-Robles et al, 2009; Burgos-Robles et al, 2007; Herry and Mons, 2004; Peters et al, 2010). Another useful method has been to exploit marked differences in fear extinction that have been described across different strains of rats and mice (Camp et al, 2009; Chang and Maren, 2010b; Hefner et al, 2008).

How naturally occurring variation in extinction efficacy in rodent populations relates to variation in sensitivity to the effects of stress is not well understood, but a number of findings are consistent with a close relationship between the two domains. For example, strain comparisons have found that certain strains of inbred mice exhibiting impaired extinction also show GR abnormalities and dysregulated neuroendocrine responses to stress challenge (Camp et al, 2012). Another finding of note has been that rodent lines that have been bred for learned helplessness induced by repeated inescapable stress (Shumake et al, 2005) or other stress-related phenotypes, including high contextual fear (Ponder et al, 2007) and unconditioned anxiety-like behavior (Muigg et al, 2008), show impairments in extinction. The implication from these findings is that there may be common genetic factors influencing a collection of stress-related phenotypes that encompass fear extinction.

EXTINCTION AND STRESS IN DEVELOPING ANIMALS

We earlier made the point that genetic influences on PTSD symptoms and circuits are often linked to stress experienced during childhood. In fact, most cases of stress-related disorders emerge during childhood and adolescence (Kessler et al, 2005). Some have posited that this reflects heightened vulnerability stemming from the protracted developmental maturation of corticolimbic systems regulating stress and emotion. The human PFC matures relatively late in development (Giedd and Rapoport, 2010), whereas the amygdala shows exaggerated reactivity to threat in fear-prone adolescent subjects (Blackford and Pine, 2012). How these ontogenetic changes affect extinction and stress effects on extinction could have important implications not only for understanding the pathophysiology of stress-related disorders in children and adolescents, but could also inform decisions about how extinction-based CBT treatments can be best employed in young people. If extinction-based approaches are relatively ineffective in adolescents, then should alternate treatments be considered for this age group? Moreover, if stress shifts the development trajectory of extinction efficacy, does a young patient's history of stressful life events need to be factored into treatment decisions?

In addressing these questions, there is growing evidence that the nature and underlying neural basis of fear and fear extinction also varies profoundly across early development in rodents (Figure 4) (for excellent recent reviews, see Callaghan et al, 2013; Landers and Sullivan, 2012; Pattwell et al, 2013; Shechner et al, 2014). In their third postnatal week, rats and mice show greater reductions in fear after extinction than adults, potentially reflecting an extinction-induced erasure of the original fear memory (Kim and Richardson, 2007). These age-related differences are associated with developmental differences in the extinction circuitry. BLA principal neurons dramatically increase their arborization and spine density during the first postnatal month (Ryan et al, 2014), whereas parvalbumin-positive interneurons in the BLA exhibit more extracellular matrix structures, known as perineuronal nets, at the age when extinction becomes adult-like (Gogolla et al, 2009). Demonstrating the functional importance of these perineuronal nets, their removal via intra-BLA infusion of chondroitinase ABC or chronic fluoxetine treatment recapitulates an infantile form of extinction in adults (Gogolla et al, 2009; Karpova et al, 2012).

Figure 4.

Figure 4

Effects of stress on fear extinction in infant and adolescent rodents. (a) At preweaning ages, infant rodents exhibit an erasure-like form of extinction that is less prone to spontaneous recovery and context-renewal than extinction in post-weaning animals (Kim and Richardson, 2007). (b) Exposing infants to stressors leads to the emergence of the adult form of extinction at preweaning (Callaghan et al, 2013). (c) Adolescent rats exposed to chronic stress typically show impaired extinction when tested as adults (Ishikawa et al, 2012; Judo et al, 2010; Toledo-Rodriguez et al, 2012), although there is preliminary evidence that acute stress in adolescence can facilitate extinction (Schayek and Maroun, 2014).

A related finding is that mutants lacking the extracellular matrix protein, reelin, show a developmental prolongation of infant-like extinction, which extends into the postweaning period (Iafrati et al, 2013). The deficient extinction phenotype in this mutant is associated with impaired dendritic spine density and synaptic plasticity (LTP) in the mPFC (see also Fitzgerald et al, 2015a) and is reversed by a single, systemic administration of either ketamine or a GluN2B NMDA receptor antagonist (Ro 25-6981) (Iafrati et al, 2013). These observations would suggest that functional immaturity of the mPFC might contribute to infant-like extinction. In support of this idea, extinction increases phosphorylated MAPK in the mPFC of adult but not infant rats, and inactivating the infant mPFC does not impair extinction, as it does in adults (Kim et al, 2009).

Rather strikingly, extinction continues to show dynamic changes as development continues. Following a transient period soon after weaning during which extinction demonstrates adult-like properties (erasure resistance), fear memories in adolescent animals show heightened fear or extinction resistance (Hefner and Holmes, 2007; McCallum et al, 2010; Pattwell et al, 2012). Similarly, human adolescents also show poorer fear extinction retrieval as compared with adults and with younger children (Pattwell et al, 2012), indicating the developmental trajectory of extinction is conserved. Deficits in extinction in adolescent rats are associated with the attenuation of mPFC (specifically IL) activation and plasticity, and can be prevented by activating NMDA receptors via d-cycloserine treatment (Kim et al, 2009; Pattwell et al, 2012). This relative hypofunction of the adolescent mPFC echoes the finding, discussed in greater length later, that the mPFC is a major target for the deleterious effects of stress on extinction in adults.

Given the lasting impact of childhood stress on later risk for developing PTSD, understanding how stress might differentially affect extinction as a function of age remains an outstanding question for future work. Pioneering preclinical efforts in this regard, Callaghan and Richardson (2012, 2014) and Cowan et al (2013) have shown that infant (pre-weaning) rats subjected to acute or repeated maternal separation (or corticosterone treatment) show the adult-like, rather than infantile, form of extinction (Figure 4). These observations suggest that stress early in life can catalyze the normal developmental ontogeny of extinction. However, the situation may be more complex in adolescence. One recent study found that acute elevated platform stress in postweaning rats facilitated extinction retrieval (Schayek and Maroun, 2014), which is distinct from the stress-impairing effect of stress in adults. However, in other studies that subject adolescents to repeated stressors (eg, predator odor, elevated platform, social instability, social isolation, or footshock), elevated fear or impaired extinction was evident during extinction training when the animals were later tested in young adulthood (Figure 4) (Ishikawa et al, 2012; Judo et al, 2010; McCormick et al, 2013; Toledo-Rodriguez et al, 2012; Naert 2011; Skelly et al, 2015). Thus, the nature, chronicity, and precise timing of stress exposure in adolescents may be important factors determining how stress alters extinction.

STRESS EFFECTS ON EXTINCTION CIRCUITRY

The effects of stress on fear extinction are mediated, at least in part, by structural and functional alterations within extinction circuits. Here we focus on some of the stress-related abnormalities that have been detected within the prefrontal, hippocampal, and amygdala nodes of the extinction circuitry (for a comprehensive review, see McEwen and Morrison, 2013).

With regards to the amygdala, a series of influential studies by Chattarji and coworkers has shown that chronic immobilization stress leads to dendritic hypertrophy of BLA principal neurons (Mitra et al, 2005; Padival et al, 2013; Vyas et al, 2006; Vyas et al, 2002), mimicking the BLA dendritic hypertrophy in animals with innate extinction deficits (Camp et al, 2012). However, dendritic expansion of BLA neurons does not always parallel deficient extinction. Acute elevated platform stress impairs extinction and causes dendritic retraction (Grillo et al, 2015; Maroun et al, 2013). A more consistent observation across acute and chronic stressors is increased BLA spinogenesis, which can manifest either immediately after stress or following a period of incubation (Maroun et al, 2013; Mitra et al, 2005).

Stress-induced spinogenesis at BLA neurons is accompanied by increased neuronal excitability, NMDAR-mediated synaptic responsivity and plasticity (LTP), as well as elevated levels of BDNF, with a parallel decrease seen in synaptic inhibition (Chauveau et al, 2012; Lakshminarasimhan and Chattarji, 2012; Maroun and Richter-Levin, 2003; Rosenkranz et al, 2010; Suvrathan et al, 2014). Impaired extinction resulting from exposure to chronic stress in adolescence or adulthood has also been associated with increased BLA, CeA, and hippocampal immediate-early gene (c-Fos) activity and metabolism (2-deoxyglucose) (Hoffman et al, 2014; Toledo-Rodriguez et al, 2012). In addition, the extinction-impairing SEFL model alters the BLA expression of glutamate and plasticity-associated genes, including the glutamate transporter, EAAT1, and specific glycine α2 and somatostatin type-2 receptors (Ponomarev et al, 2010). Taken together, these findings are consistent with a stress-induced shift towards greater excitability, at least a subpopulation, of BLA neurons that results in a bias toward sustained fear and weakened extinction (Roozendaal et al, 2009). In agreement with such a scheme, neuropeptide S injected into the LA was recently shown to reverse stress-induced synaptic hyperexcitability and the attendant extinction deficits caused by acute immobilization stress (Chauveau et al, 2012).

The impact of stress is not limited to the amygdala, but encompasses the wider neural circuitry mediating extinction. Of particular note, stress leads to structural and functional changes in mPFC and hippocampal regions known to interact with the amygdala in regulating extinction. However, in contrast to the dendritic hypertrophy seen at BLA neurons after stress, stress leads to the loss of dendritic material at mPFC and hippocampal neurons (Leuner and Gould, 2010). Remarkably, even a brief history of swim stress causes dendritic retraction at principal IL neurons and deficits in extinction (Holmes and Wellman, 2009; Izquierdo et al, 2006). These observations suggest that in agreement with the known importance of the IL for extinction discussed earlier, stress impairment of extinction could be a result of a loss of function in the IL. Indeed, restoring plasticity in the mPFC, via BDNF infusions, can promote extinction and rescue effects of stress (Graybeal et al, 2011; Peters et al, 2010). However, the contribution of the IL may not be so straightforward. Damaging the IL can prevent the impairing effects of stress (Farrell et al, 2010), implying the IL must be intact at the time of stress in order to drive the neural alterations that ultimately manifest as deficits in extinction.

Electrophysiological recordings of the activity of mPFC neurons during extinction reveal a significant blunting of IL (and PL) neuronal firing in chronically restrained rats exhibiting impaired extinction retrieval (Wilber et al, 2011). Similarly, extinction deficits evident during immediate extinction training are accompanied by attenuated mPFC neuronal burst firing as well as IL hypoactivation (indexed by IEG expression) (Chang et al, 2010a; Kim et al, 2010; Stafford et al, 2013). Moreover, pharmacological manipulations that rescue IEDs, including picrotoxin and d-cycloserine, do so in concert with an increase in mPFC neuronal activity (Chang and Maren, 2011). Taken together, these various lines of evidence are consistent with the view that stress influences extinction, at least in part, by interfering with mPFC, although the precise locus of these effects is still not fully clear (Holmes and Wellman, 2009).

One relevant site of neural adaptations induced by stress is at the level of functional connections between the mPFC and BLA. Studies by Maroun and colleagues, and others have demonstrated how extinction deficits produced by acute elevated platform stress are tied to the disruption of synaptic plasticity (LTP) in the mPFC-BLA pathway (Maroun, 2006; Maroun and Richter-Levin, 2003; Richter-Levin and Maroun, 2010; see also Rocher et al, 2004). Interestingly, the same stressor was recently found to enhance mPFC-BLA plasticity in adolescent rats, providing another example of how the neural consequences of stress differ as a function of age (Schayek and Maroun, 2014). In adults, stress-induced plasticity changes in mPFC-BLA pathway are associated with multiple of molecular abnormalities in the mPFC. Among these, stressors including maternal separation, SPS, and chronic corticosterone treatment have been shown to cause extinction deficits that are accompanied by the loss of mPFC levels of glutamate and glutamine, and reductions in the mPFC expression of NMDARs (GluN1 and GluN2B) and l-α-amino-3-hydroxy-5-methyl-isoxazole-4-propionic acid receptors (AMPAR, GluA2/3) (Gourley et al, 2009; Knox et al, 2010; Wilber et al, 2009). This apparent loss of prefrontal glutamate signaling contrasts with the hyperglutamatergic profile of the stressed BLA.

The hippocampus and its connections to the mPFC also undergo signs of functional downregulation in response to stress. For instance, postnatal footshock stress attenuates extinction-related ERK phosphorylation in the CA1 subregion of the hippocampus and the mPFC, and leads to a loss of LTP in the hippocampal-mPFC pathway that can be rescued by systemic treatment with d-cycloserine (Ishikawa et al, 2012; Judo et al, 2010). In addition, exposure to the extinction-disrupting SEFL model leads to decreased synaptic efficacy in the dorsal hippocampus (Deschaux et al, 2014; Spennato et al, 2008). Extinction deficits produced by chronic mild stress (Cerqueira et al, 2007; Garcia et al, 2008) or footshock (Ishikawa et al, 2012; Koseki et al, 2009) are paralleled by a decrease in hippocampal–PL transmission. Moreover, extinction deficits can be mimicked by low-frequency stimulation of the dorsal hippocampus (Garcia et al, 2008) and high-frequency hippocampal stimulation can prevent SEFL-induced extinction deficits (Deschaux et al, 2014) in a manner that requires concurrent activation of the mPFC (Deschaux et al, 2011). These studies suggest a potential causal role for impaired hippocampal recruitment in stress-impaired extinction.

In sum, current findings paint a picture of amygdala hyperexcitability coupled with loss of important sources of input from hippocampal and prefrontal areas in extinction-impaired stressed animals (Figure 5). We next turn to some of the major molecular events that underlie these system-level changes.

Figure 5.

Figure 5

Stress-induced changes in fear extinction-mediating brain regions. Stress produces a range of morphological, electrophysiological and molecular changes in the prefrontal cortex, hippocampus, and amygdala that parallel extinction abnormalities. BLA, basolateral nucleus of the amygdala; CeA, central nucleus of the amygdala; IL, infralimbic cortex; PL, prelimbic cortex.

STRESS, GLUCOCORTICOIDS, AND EXTINCTION

A defining feature of stress is the mobilization of glucocorticoids (cortisol in humans and corticosterone in rodents). Glucocorticoids activate MRs and GRs in extinction-mediating brain regions and there is growing indications of a connection, albeit complex, between glucocorticoids and extinction.

Chronically elevating glucocorticoids in pregnant rats through repeated corticosterone administration (or restraint) impairs extinction and decreases GR expression in the mPFC and hippocampus among other regions, when offspring are subsequently examined as adults (Bingham et al, 2013; Green et al, 2011). Other work has shown that repeated administration of corticosterone in adult rats leads to the loss of NMDARs and AMPARs in the mPFC and again disrupts extinction (Gourley et al, 2009). These stress-related losses in prefrontal and hippocampal glucocorticoids and glutamate signaling may contribute to the impaired capacity for forming extinction memories. This notion concurs with the observation that mouse strains exhibiting poor extinction have a pre-existing loss of hippocampal GRs (Camp et al, 2012) or forebrain MRs (Ter Horst et al, 2012). There are, however, examples in which disruption to extinction following exposure to stressors such as maternal separation and exposure to SPS correlate with increases in mPFC and hippocampal GR expression (Knox et al, 2012a; Knox et al, 2012b; Wilber et al, 2009). Moreover, blocking the SPS-induced upregulation of GRs in these areas by treating rats with the anticonvulsant phenytoin effectively prevents the emergence of extinction deficits (George et al, 2015).

One interpretation of these seemingly contradictory findings is that extinction may be sensitive to the deviations from normal GR signaling, with either increases or decreases being sufficient to disrupt behavior. Indeed, although some of the aforementioned studies (eg, George et al, 2015) suggest that preventing corticohippocampal GR upregulation may be able to prevent stress from impairing extinction, the formation of extinction memories is also known to require glucocorticoids. In a variety of experimental designs, acute administration of exogenous corticosterone or the synthetic glucocorticoid, dexamethasone, facilitates extinction in rats and certain mouse strains (Abrari et al, 2008; Brinks et al, 2009; Cai et al, 2006; Ninomiya et al, 2010; Yang et al, 2006; Yang et al, 2007). Conversely, blocking the synthesis of corticosterone synthesis (with metyrapone) produces deficits in extinction (Barrett and Gonzalez-Lima, 2004; Blundell et al, 2011; Yang et al, 2006; Yang et al, 2007, but see Clay et al, 2011).

The bidirectional modulation of extinction by glucocorticoids likely involves the amygdala given the effects of systemic treatments can be recapitulated by direct delivery into the BLA. Microinfusion of the GR agonist (RU28362) or the GR antagonist (mifepristone) into the BLA before extinction training leads to facilitation and impairment in extinction, respectively (Yang et al, 2006). In addition, the extinction-facilitating effects of systemically administered dexamethasone are prevented by blocking (with AP5 or MK-801) NMDARs in the BLA (Yang et al, 2007), a finding which further implicates the BLA as a key site of glucocorticoid action and suggests that GRs regulate glutamatergic plasticity in this region. Importantly, plasticity-promoting actions of GR signaling in the BLA are not limited to the formation of extinction memory. In the same studies showing that pre-extinction corticosterone treatment improves extinction, the same treatment given before fear conditioning produces stronger, extinction-resistant fear memory (Brinks et al, 2009). This is entirely in agreement with the well-established role GRs have in enhancing emotional memories (Roozendaal and McGaugh, 2011).

From a clinical standpoint, the use of GR-acting drugs as a novel treatment for PTSD has attracted considerable attention (see Figure 6 for a putative model of how GR and NMDA receptor agonists might act). Phobic patients administered the synthetic glucocorticoid (hydrocortisone) before exposure therapy showed improved therapeutic outcome at follow-up (de Quervain et al, 2011; Soravia et al, 2006). Preliminary work with the same drug in PTSD patients has thus far been negative (Suris et al, 2010), but the results of larger clinical trials are expected in the near future (Yehuda et al, 2010) (see clinicaltrials.gov identifier NCT01090518).

Figure 6.

Figure 6

Putative schematic flow of mechanism in the ventromedial prefrontal cortex (vmPFC) and basolateral nucleus of the amygdala (BLA) underlying stress-induced extinction deficits. In the vmPFC, stress has been shown to cause excess glutamate and associated dendritic hypotrophy and downregulation of glutamate receptors. These adaptations could lead to loss of vmPFC modulation of the amygdala and a resulting impairment in extinction that may be prevented by stimulating NMDA receptors using, eg, d-cycloserine. In the BLA, GR downregulation and anandamide depletion is seen after stress, which could in turn lead to reduced engagement of noradrenergic transmission and deficient extinction. There may be multiple points of intervention to reverse this cascade, including stimulating GRs or elevating anandamide with hydrocortisone and FAAH inhibitors, respectively, or blocking α2-adrenoceptors with yohimbine to increase BLA noradrenaline. BLA, basolateral amygdala; NE, norepinephrine; GR, glucocorticoid receptor; vmPFC, ventromedial prefrontal cortex.

STRESS, eCBS, AND EXTINCTION

Recent evidence indicates that glucocorticoids transduce their effects on fear in part through the modulation of eCBs. Stress and glucocorticoids rapidly increase eCBs in extinction-mediating regions such as the BLA and, in turn, eCBs regulate glucocorticoids and the HPA axis to create a functional feedback loop (Hill et al, 2010b). The ability of glucocorticoids (via systemic corticosterone, dexamethasone, or intra-BLA RU28362) to enhance fear memory is occluded by blockade of CB1R (via AM251) in the BLA or hippocampus (Atsak et al, 2014; Atsak et al, 2012; Campolongo et al, 2009; de Oliveira Alvares et al, 2010). Moreover, fear memory enhancement achieved by activating CB1R in the BLA (with WIN55,212-2) is not prevented by blocking GRs (via RU38486), which is consistent with GR being upstream of eCBs (Atsak et al, 2014). However, the finding that the corticosterone synthesis inhibitor metyrapone blocks the enhancement of fear memory by CB1R agonists (Campolongo et al, 2013) suggests that GRs and eCBs likely interact in a more complex, bidirectional manner.

The contribution of eCBs to glucocorticoid-mediated extinction remains to be determined but is a question of significant interest given recent findings. First, improvements in extinction are found after augmenting eCBs levels by inhibiting uptake or reducing degradation of anandamide via FAAH inhibition or point mutation (Chhatwal et al, 2005; de Oliveira Alvares et al, 2008; Dincheva, 2015; Gunduz-Cinar et al, 2013b; Pamplona et al, 2006). Second, the rapid release of corticotropin-releasing hormone (CRH), acting through the CRHR1 receptor, has recently been identified as a mechanism that, by suppressing FAAH activity, causes reductions in BLA anandamide levels that precede, and are independent of, the mobilization of glucocorticoids (Gray et al, 2015). These opposing, temporally dissociated effects of CRH and glucocorticoids on BLA anandamide could impact stress-induced deficits in extinction in important ways that will require further work to parse. A corollary question is whether the pro-extinction efficacy of putative eCB-augmenting treatments in PTSD varies as a function of prior stress history and the functional status of an individual's glucocorticoid system. In normal human subjects, acute pre-extinction administration of dronabinol, a synthetic version of Δ9-tetrahydrocannabinol (Rabinak et al, 2013b), or cannabidiol (Das et al, 2013), a compound with FAAH-inhibiting properties (Leweke et al, 2012), improves extinction retrieval coincident with enhanced recruitment of the vmPFC–amygdala circuit (Rabinak et al, 2013a). How these beneficial effects are impacted by stress remains to be clarified, but we offer an heuristic model for how FAAH inhibitors could combat stress effects on extinction in Figure 6.

STRESS, NOREPINEPHRINE, AND EXTINCTION

The norepinephrine (NE) system provides a further functional link between glucocorticoids and eCBs. Noradrenergic activity in the BLA, acting through the β-adrenoceptor (β-AR), promotes fear memory and the ability of both glucocorticoids and eCBs to enhance fear memory requires the activity of BLA β-ARs (ie, they are prevented by the β-AR blocker propranolol) (Atsak et al, 2014; Roozendaal et al, 2009). On the basis of these observations, some authors suggest that glucocorticoids trigger eCBs to inhibit GABAergic interneurons, which in turn disinhibit NE release onto β-ARs and increases the sensitivity of principal BLA neurons to NE input (Hill and McEwen, 2010a; Morena and Campolongo, 2014). This mechanistic scheme could account for the aforementioned interaction between β-AR gene variation, childhood stress, and risk for PTSD (Liberzon et al, 2014).

The β-AR has also received significant attention clinically as a potential therapeutic for extinction and the β-AR antagonist propranolol has shown some efficacy as an adjunct to exposure therapy, although the results remain preliminary (Brunet et al, 2014). The mechanism by which propranolol would augment extinction remains unclear. Given that extinction is an active learning process requiring the activation of principal BLA ‘extinction' neurons (Herry et al, 2010), blockade of β-ARs would presumably interfere with noradrenergic (as well as glucocorticoid) mediation of extinction. This is borne out by experimental studies in normal human subjects (Bos et al, 2012), which have reported extinction impairments following administration of propranolol (Mueller et al, 2008). However, it may be that under stressful conditions, elevated noradrenergic signaling impedes extinction learning, an effect that might be counteracted by propranolol. Indeed, propranolol has recently been shown to stabilize stress-induced changes in PFC firing and attenuate the immediate extinction deficit (Fitzgerald et al, 2015b).

By the same logic, augmenting noradrenergic activity at the time of extinction memory formation would be expected to promote extinction. In support of this, yohimbine, an α2-adrenoceptor antagonist that increases NE release, promotes extinction in some preclinical studies (Davis et al, 2008; Holmes and Quirk, 2010; Janak and Corbit, 2011). Yohimbine has also been found to improve the efficacy of exposure therapy in patients with claustrophobic and social anxiety disorder (Powers et al, 2009; Smits et al, 2014), with studies in PTSD sufferers currently underway (clinicaltrials.gov identifier NCT01031979) (see Figure 6 for a model of how yohimbine might act to reverse stress-impaired extinction). Again, a caveat is that the efficacy of manipulations that enhance or limit noradrenergic transmission may critically depend on the basal state of NE transmission at the time of the intervention. For example, under stressful conditions and elevated NE transmission it might be expected that a noradrenergic agonist might impair (rather than promote) extinction by exacerbating the deleterious influence of stress on extinction. That is, it is likely that an inverted-U function defines optimal NE signaling for extinction learning and stress may shift this function in a way that alters the efficacy of noradrenergic modulators.

Future directions and clinical implications

The last decade has witnessed remarkable progress that has broadened our understanding of the brain mechanisms for extinction-based therapies and the vulnerabilities of these mechanisms to stress. The challenges, of course, are to develop strategies to dampen stress-induced impairments in extinction, as well as to deepen the endurance and generalization of extinction memories once formed. As we have seen, there are now several pharmacological cognitive enhancers for PTSD (Singewald et al, 2014; Bukalo et al, 2014) and many of these compounds facilitate extinction learning (Fitzgerald et al, 2014a) (Figure 6). Moreover, there are ongoing efforts to improve behavioral therapies by incorporating extinction reminders, eg, to limit relapse of fear after therapy (Vervliet et al, 2013). Yet, in the end, extinction-based therapies will always be limited by the impermanence of this form of learning. Ultimately, a more effective approach might be the targeted erasure of traumatic memories—at least the emotional components of traumatic memories central to PTSD. Interestingly, this approach has met with success in both animals and people.

In the first study of its kind, Monfils et al (2009) examined whether delivering extinction trials soon after reactivating a fear memory would produce a more enduring loss of fear by impairing the reconsolidation of that information. They found that delivering extinction trials from 10 min to 1 h after a single CS reminder (itself an extinction trial) resulted in loss of conditional responding that showed less spontaneous recovery, renewal, and reinstatement than that obtained after a standard extinction procedure. A similar effect has been described in healthy humans (Schiller et al, 2010). In both cases, it has been argued that the retrieval-extinction procedure may produce a more enduring loss of fear by preventing memory reconsolidation, thereby effectively erasing the fear memory. Indeed, there is evidence to suggest, at least in animals, that changes in glutamate receptors at BLA synapses may account for the loss of conditional responding (Clem and Huganir, 2010; Monfils et al, 2009). That said, this retrieval-extinction procedure has not always proved effective at eliminating fear relapse (Chan et al, 2010; Costanzi et al, 2011) and extinction administered soon after conditioning (when fear memories are in an active state) fails to yield long-term extinction (Macpherson et al, 2013; Maren and Chang, 2006; Stafford et al, 2013). Moreover, it is difficult to know whether a manipulation truly erased fear memory or produced a particularly deep extinction (Lattal and Wood, 2013). Clearly, further work is required to understand the conditions under which memories might be erased by extinction procedures (Auber et al, 2013).

In addition to behavioral procedures, pharmacological manipulations have been used to target the reconsolidation of fear memories as previously described (Nader and Hardt, 2009). Of these, the most promising manipulation for use in humans is propranolol (Kindt et al, 2009; Pitman et al, 2006). For example, Kindt et al (2009) found that fear memories (indexed by startle) in healthy humans could be erased (ie, the CR failed to reinstate) by reactivating the fear memory under propranolol. Much of this work has been interpreted in terms of propranolol's properties as a protein synthesis inhibitor to disrupt the reconsolidation of reactivated fear memories (Otis et al, 2015; Soeter and Kindt, 2011). Despite the early promise in healthy humans, results have been underwhelming in clinical populations studied to date (Wood et al, 2015) (additional data are pending, see clinicaltrials.gov identifier NCT01127568). One possibility is that propranolol or retrieval-extinction procedures, which have typically been tested in isolation, might be particularly effective if administered together.

Of course, it has been argued that extinction memories are not necessarily impermanent (Lattal and Wood, 2013). By this view, either enhanced extinction or reconsolidation failure could account for the enduring loss of fear after a variety of behavioral or pharmacological manipulations. From this perspective, molecular strategies to strengthen extinction memory hold promise for achieving lasting suppression of conditional fear. In this regard, there is considerable interest in the possibility that epigenetic modifications can sustain long-term extinction memories. For example, some very recent work indicates that specific histone deacetylase (HDAC) inhibitors can facilitate the consolidation of fear extinction (Bowers et al, 2015; Stafford et al, 2012; Whittle et al, 2013; Pizzimenti and Lattal, 2015). Moreover, HDAC2-targeted inhibitors coupled with memory reactivation and extinction can lead to persistent loss of even remote fear memories that are normally resistant to updating (Gräff et al, 2014). This suggests that memory reactivation coupled with extinction procedures and memory-promoting adjuncts may hold the key for establishing deep extinction memories that are resistant to fear relapse.

There is clearly much yet to learn about the conditions that dampen the deleterious effects of stress on extinction, promote and deepen extinction memories, and erase unwanted fears. Ultimately, it can be said that the field has made considerable progress on all fronts, producing important mechanistic information on extinction learning that is being translated to human clinical trials and novel therapies. The clinical implications of this progress are enormous; there is no doubt that this work will herald a new era of neurobiologically informed pharmacotherapies to revolutionize the treatment of stress and trauma-related disorders.

FUNDING AND DISCLOSURE

The authors declare no conflict of interest.

Acknowledgments

AH is supported by the NIAAA Intramural Research Program, Henry Jackson Foundation for the Advancement of Military Medicine, and the Department of Defense in the Center for Neuroscience and Regenerative Medicine. SM is supported by grants from the NIH (R01MH065961) and a McKnight Foundation Memory and Cognitive Disorders Award.

References

  1. Abrari K, Rashidy-Pour A, Semnanian S, Fathollahi Y (2008). Administration of corticosterone after memory reactivation disrupts subsequent retrieval of a contextual conditioned fear memory: dependence upon training intensity. Neurobiol Learn Mem 89: 178–184. [DOI] [PubMed] [Google Scholar]
  2. Andero R, Heldt SA, Ye K, Liu X, Armario A, Ressler KJ (2011). Effect of 7,8-dihydroxyflavone, a small-molecule TrkB agonist, on emotional learning. Am J Psychiatry 168: 163–172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Atsak P, Hauer D, Campolongo P, Schelling G, Fornari RV, Roozendaal B (2014). Endocannabinoid signaling within the basolateral amygdala integrates multiple stress hormone effects on memory consolidation. Neuropsychopharmacology 40: 1485–1494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Atsak P, Hauer D, Campolongo P, Schelling G, McGaugh JL, Roozendaal B (2012). Glucocorticoids interact with the hippocampal endocannabinoid system in impairing retrieval of contextual fear memory. Proc Natl Acad Sci USA 109: 3504–3509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Auber A, Tedesco V, Jones CE, Monfils M-H, Chiamulera C (2013). Post-retrieval extinction as reconsolidation interference: methodological issues or boundary conditions? Psychopharmacology (Berl) 226: 631–647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Baran SE, Armstrong CE, Niren DC, Hanna JJ, Conrad CD (2009). Chronic stress and sex differences on the recall of fear conditioning and extinction. Neurobiol Learn Mem 91: 323–332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Barrett D, Gonzalez-Lima F (2004). Behavioral effects of metyrapone on Pavlovian extinction. Neurosci Lett 371: 91–96. [DOI] [PubMed] [Google Scholar]
  8. Ben Mamou C, Gamache K, Nader K (2006). NMDA receptors are critical for unleashing consolidated auditory fear memories. Nat Neurosci 9: 1237–1239. [DOI] [PubMed] [Google Scholar]
  9. Bentz D, Michael T, Wilhelm FH, Hartmann FR, Kunz S, Rohr von IRR et al (2013). Influence of stress on fear memory processes in an aversive differential conditioning paradigm in humans. Psychoneuroendocrinology 38: 1186–1197. [DOI] [PubMed] [Google Scholar]
  10. Binder EB, Bradley RG, Liu W, Epstein MP, Deveau TC, Mercer KB et al (2008). Association of FKBP5 polymorphisms and childhood abuse with risk of posttraumatic stress disorder symptoms in adults. JAMA 299: 1291–1305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bingham BC, Sheela Rani CS, Frazer A, Strong R, Morilak DA (2013). Exogenous prenatal corticosterone exposure mimics the effects of prenatal stress on adult brain stress response systems and fear extinction behavior. Psychoneuroendocrinology 38: 2746–2757. [DOI] [PubMed] [Google Scholar]
  12. Blackford JU, Pine DS (2012). Neural substrates of childhood anxiety disorders: a review of neuroimaging findings. Child Adolesc Psychiatr Clin N Am 21: 501–525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Blair HT, Schafe GE, Bauer EP, Rodrigues SM, LeDoux JE (2001). Synaptic plasticity in the lateral amygdala: a cellular hypothesis of fear conditioning. Learn Mem 8: 229–242. [DOI] [PubMed] [Google Scholar]
  14. Blechert J, Michael T, Vriends N, Margraf J, Wilhelm FH (2007). Fear conditioning in posttraumatic stress disorder: evidence for delayed extinction of autonomic, experiential, and behavioural responses. Behav Res Ther 45: 2019–2033. [DOI] [PubMed] [Google Scholar]
  15. Blundell J, Blaiss CA, Lagace DC, Eisch AJ, Powell CM (2011). Block of glucocorticoid synthesis during re-activation inhibits extinction of an established fear memory. Neurobiol Learn Mem 95: 453–460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Bogdan R, Williamson DE, Hariri AR (2012). Mineralocorticoid receptor Iso/Val (rs5522) genotype moderates the association between previous childhood emotional neglect and amygdala reactivity. Am J Psychiatry 169: 515–522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Bordi F, LeDoux J (1994). Response properties of single units in areas of rat auditory thalamus that project to the amygdala. I. Acoustic discharge patterns and frequency receptive fields. Exp Brain Res 98: 261–274. [DOI] [PubMed] [Google Scholar]
  18. Bos MG, Beckers T, Kindt M (2012). The effects of noradrenergic blockade on extinction in humans. Biol Psychol 89: 598–605. [DOI] [PubMed] [Google Scholar]
  19. Bouton M (1993). Context, time, and memory retrieval in the interference paradigms of Pavlovian learning. Psychol Bull 114: 80–99. [DOI] [PubMed] [Google Scholar]
  20. Bouton M, Mineka S, Barlow D (2001). A modern learning theory perspective on the etiology of panic disorder. Psychol Rev 108: 4–32. [DOI] [PubMed] [Google Scholar]
  21. Bowers ME, Xia B, Carreiro S, Ressler KJ (2015). The class I HDAC inhibitor RGFP963 enhances consolidation of cued fear extinction. Learn Mem 22: 225–231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Brinks V, de Kloet ER, Oitzl MS (2009). Corticosterone facilitates extinction of fear memory in BALB/c mice but strengthens cue related fear in C57BL/6 mice. Exp Neurol 216: 375–382. [DOI] [PubMed] [Google Scholar]
  23. Brunet A, Thomas E, Saumier D, Ashbaugh AR, Azzoug A, Pitman RK et al (2014). Trauma reactivation plus propranolol is associated with durably low physiological responding during subsequent script-driven traumatic imagery. Can J Psychiatry 59: 228–232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Bukalo O, Pinard C, Holmes A (2014). Mechanisms to medicines: modeling the neural circuitry of fear to reveal new anxiolytic drug targets. Brit J Pharmacol 171: 4690–4718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Bukalo O, Pinard CR, Silverstein S, Brehm C, Whittle N, Colacicco G et al (2015). Prefrontal inputs to the amygdala instruct fear extinction memory formation. Science (in press). [DOI] [PMC free article] [PubMed]
  26. Burgos-Robles A, Vidal-Gonzalez I, Quirk GJ (2009). Sustained conditioned responses in prelimbic prefrontal neurons are correlated with fear expression and extinction failure. J Neurosci 29: 8474–8482. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Burgos-Robles A, Vidal-Gonzalez I, Santini E, Quirk GJ (2007). Consolidation of fear extinction requires NMDA receptor-dependent bursting in the ventromedial prefrontal cortex. Neuron 53: 871–880. [DOI] [PubMed] [Google Scholar]
  28. Bush DE, Sotres-Bayon F, LeDoux JE (2007). Individual differences in fear: isolating fear reactivity and fear recovery phenotypes. J Trauma Stress 20: 413–422. [DOI] [PubMed] [Google Scholar]
  29. Cai WH, Blundell J, Han J, Greene RW, Powell CM (2006). Postreactivation glucocorticoids impair recall of established fear memory. J Neurosci 26: 9560–9566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Callaghan BL, Graham BM, Li S, Richardson R (2013). From resilience to vulnerability: mechanistic insights into the effects of stress on transitions in critical period plasticity. Front Psychiatry 4: 90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Callaghan BL, Richardson R (2012). The effect of adverse rearing environments on persistent memories in young rats: removing the brakes on infant fear memories. Transl Psychiatry 2: e138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Callaghan BL, Richardson R (2014). Early emergence of adult-like fear renewal in the developing rat after chronic corticosterone treatment of the dam or the pups. Behav Neurosci 128: 594–602. [DOI] [PubMed] [Google Scholar]
  33. Camp M, Norcross M, Whittle N, Feyder M, D'Hanis W, Yilmazer-Hanke D et al (2009). Impaired Pavlovian fear extinction is a common phenotype across genetic lineages of the 129 inbred mouse strain. Genes Brain Behav 8: 744–752. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Camp MC, Macpherson KP, Lederle L, Graybeal C, Gaburro S, Debrouse LM et al (2012). Genetic strain differences in learned fear inhibition associated with variation in neuroendocrine, autonomic, and amygdala dendritic phenotypes. Neuropsychopharmacology 37: 1534–1547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Campolongo P, Morena M, Scaccianoce S, Trezza V, Chiarotti F, Schelling G et al (2013). Novelty-induced emotional arousal modulates cannabinoid effects on recognition memory and adrenocortical activity. Neuropsychopharmacology 38: 1276–1286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Campolongo P, Roozendaal B, Trezza V, Hauer D, Schelling G, McGaugh JL et al (2009). Endocannabinoids in the rat basolateral amygdala enhance memory consolidation and enable glucocorticoid modulation of memory. Proc Natl Acad Sci USA 106: 4888–4893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Caspi A, Hariri AR, Holmes A, Uher R, Moffitt TE (2010). Genetic sensitivity to the environment: the case of the serotonin transporter gene and its implications for studying complex diseases and traits. Am J Psychiatry 167: 509–527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Cerqueira JJ, Mailliet F, Almeida OF, Jay TM, Sousa N (2007). The prefrontal cortex as a key target of the maladaptive response to stress. J Neurosci 27: 2781–2787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Chan WYM, Leung HT, Westbrook RF, Mcnally GP (2010). Effects of recent exposure to a conditioned stimulus on extinction of Pavlovian fear conditioning. Learn Mem 17: 512–521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Chang C-H, Berke JD, Maren S (2010. a). Single-unit activity in the medial prefrontal cortex during immediate and delayed extinction of fear in rats. PLoS One 5: e11971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Chang C-H, Maren S (2009). Early extinction after fear conditioning yields a context-independent and short-term suppression of conditional freezing in rats. Learn Mem 16: 62–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Chang C-H, Maren S (2010. b). Strain difference in the effect of infralimbic cortex lesions on fear extinction in rats. Behav Neurosci 124: 391–397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Chang C-H, Maren S (2011). Medial prefrontal cortex activation facilitates re-extinction of fear in rats. Learn Mem 18: 221–225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Chauveau F, Lange MD, Jungling K, Lesting J, Seidenbecher T, Pape HC (2012). Prevention of stress-impaired fear extinction through neuropeptide S action in the lateral amygdala. Neuropsychopharmacology 37: 1588–1599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Chhatwal JP, Davis M, Maguschak KA, Ressler KJ (2005). Enhancing cannabinoid neurotransmission augments the extinction of conditioned fear. Neuropsychopharmacology 30: 516–524. [DOI] [PubMed] [Google Scholar]
  46. Chhatwal J, Myers K, Ressler K, Davis M (2005). Regulation of gephyrin and GABAA receptor binding within the amygdala after fear acquisition and extinction. J Neurosci 25: 502–506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Chhatwal J, Stanek-Rattiner L, Davis M, Ressler K (2006). Amygdala BDNF signaling is required for consolidation but not encoding of extinction. Nat Neurosci 9: 870–872. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Cho J-H, Deisseroth K, Bolshakov VY (2013). Synaptic encoding of fear extinction in mPFC-amygdala circuits. Neuron 80: 1491–1507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Clay R, Hebert M, Gill G, Stapleton LA, Pridham A, Coady M et al (2011). Glucocorticoids are required for extinction of predator stress-induced hyperarousal. Neurobiol Learn Mem 96: 367–377. [DOI] [PubMed] [Google Scholar]
  50. Clem RL, Huganir RL (2010). Calcium-permeable AMPA receptor dynamics mediate fear memory erasure. Science 330: 1108–1112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Corcoran K, Quirk G (2007). Activity in prelimbic cortex is necessary for the expression of learned, but not innate, fears. J Neurosci 27: 840–844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Costanzi M, Cannas S, Saraulli D, Rossi-Arnaud C, Cestari V (2011). Extinction after retrieval: Effects on the associative and nonassociative components of remote contextual fear memory. Learn Mem 18: 508–518. [DOI] [PubMed] [Google Scholar]
  53. Courtin J, Chaudun F, Rozeske RR, Karalis N, Gonzalez-Campo C, Wurtz H et al (2014). Prefrontal parvalbumin interneurons shape neuronal activity to drive fear expression. Nature 505: 92–96. [DOI] [PubMed] [Google Scholar]
  54. Cowan CS, Callaghan BL, Richardson R (2013). Acute early-life stress results in premature emergence of adult-like fear retention and extinction relapse in infant rats. Behav Neurosci 127: 703–711. [DOI] [PubMed] [Google Scholar]
  55. Craske MG, Kircanski K, Zelikowsky M, Mystkowski J, Chowdhury N, Baker A (2008). Optimizing inhibitory learning during exposure therapy. Behav Res Ther 46: 5–27. [DOI] [PubMed] [Google Scholar]
  56. Das RK, Kamboj SK, Ramadas M, Yogan K, Gupta V, Redman E et al (2013). Cannabidiol enhances consolidation of explicit fear extinction in humans. Psychopharmacology (Berl) 226: 781–792. [DOI] [PubMed] [Google Scholar]
  57. Davis AR, Shields AD, Brigman JL, Norcross M, McElligott ZA, Holmes A et al (2008). Yohimbine impairs extinction of cocaine-conditioned place preference in an alpha2-adrenergic receptor independent process. Learn Mem 15: 667–676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Davis M (1992). The role of the amygdala in fear and anxiety. Annu Rev Neurosci 15: 353–375. [DOI] [PubMed] [Google Scholar]
  59. de Oliveira Alvares L, Engelke DS, Diehl F, Scheffer-Teixeira R, Haubrich J, de Freitas Cassini L et al (2010). Stress response recruits the hippocampal endocannabinoid system for the modulation of fear memory. Learn Mem 17: 202–209. [DOI] [PubMed] [Google Scholar]
  60. de Oliveira Alvares L, Pasqualini Genro B, Diehl F, Molina VA, Quillfeldt JA (2008). Opposite action of hippocampal CB1 receptors in memory reconsolidation and extinction. Neuroscience 154: 1648–1655. [DOI] [PubMed] [Google Scholar]
  61. de Quervain DJ, Bentz D, Michael T, Bolt OC, Wiederhold BK, Margraf J et al (2011). Glucocorticoids enhance extinction-based psychotherapy. Proc Natl Acad Sci USA 108: 6621–6625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Deschaux O, Koumar O, Canini F, Moreau J, Garcia R (2014). High-frequency stimulation of the hippocampus blocks fear learning sensitization and return of extinguished fear. Neuroscience 286C: 423–429. [DOI] [PubMed] [Google Scholar]
  63. Deschaux O, Motanis H, Spennato G, Moreau JL, Garcia R (2011). Re-emergence of extinguished auditory-cued conditioned fear following a sub-conditioning procedure: effects of hippocampal and prefrontal tetanic stimulations. Neurobiol Learn Mem 95: 510–518. [DOI] [PubMed] [Google Scholar]
  64. Deschaux O, Zheng X, Lavigne J, Nachon O, Cleren C, Moreau JL et al (2013). Post-extinction fluoxetine treatment prevents stress-induced reemergence of extinguished fear. Psychopharmacology (Berl) 225: 209–216. [DOI] [PubMed] [Google Scholar]
  65. Dias BG, Ressler KJ (2013). PACAP and the PAC1 receptor in post-traumatic stress disorder. Neuropsychopharmacology 38: 245–246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Dincheva (2015). FAAH genetic variation enhances fronto-amygdala function in mouse and human Nat Commun 6: 6395. [DOI] [PMC free article] [PubMed]
  67. Do-Monte FH, Manzano-Nieves G, Quiñones-Laracuente K, Ramos-Medina L, Quirk GJ (2015. a). Revisiting the role of infralimbic cortex in fear extinction with optogenetics. J Neurosci 35: 3607–3615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Do-Monte FH, Quiñones-Laracuente K, Quirk GJ (2015. b). A temporal shift in the circuits mediating retrieval of fear memory. Nature 519: 460–463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. DSM-5 (2013) Diagnostic and Statistical Manual of Mental Disorders. Fourth Edition. APA Press: Washington, D.C. [Google Scholar]
  70. Dubreucq S, Matias I, Cardinal P, Haring M, Lutz B, Marsicano G et al (2012). Genetic dissection of the role of cannabinoid type-1 receptors in the emotional consequences of repeated social stress in mice. Neuropsychopharmacology 37: 1885–1900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Duvarci S, Paré D (2014). Amygdala microcircuits controlling learned fear. Neuron 82: 966–980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Ehrlich I, Humeau Y, Grenier F, Ciocchi S, Herry C et al (2009). Amygdala inhibitory circuits and the control of fear memory. Neuron 62: 757–771. [DOI] [PubMed] [Google Scholar]
  73. Estes W, Skinner B (1941). Some quantitative properties of anxiety. J Exp Psychol 29: 390. [Google Scholar]
  74. Falls W, Miserendino M, Davis M (1992). Extinction of fear-potentiated startle: blockade by infusion of an NMDA antagonist into the amygdala. J Neurosci 12: 854–863. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Fanselow MS (2000). Contextual fear, gestalt memories, and the hippocampus. Behav Brain Res 110: 73–81. [DOI] [PubMed] [Google Scholar]
  76. Farrell MR, Sayed JA, Underwood AR, Wellman CL (2010). Lesion of infralimbic cortex occludes stress effects on retrieval of extinction but not fear conditioning. Neurobiol Learn Mem 94: 240–246. [DOI] [PubMed] [Google Scholar]
  77. Felix-Ortiz AC, Beyeler A, Seo C, Leppla CA, Wildes CP, Tye KM (2013). BLA to vHPC inputs modulate anxiety-related behaviors. Neuron 79: 658–664. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Fendt M, Fanselow M (1999). The neuroanatomical and neurochemical basis of conditioned fear. Neurosci Biobehav Rev 23: 743–760. [DOI] [PubMed] [Google Scholar]
  79. Fisher PM, Hariri AR (2012). Linking variability in brain chemistry and circuit function through multimodal human neuroimaging. Genes Brain Behav 11: 633–642. [DOI] [PubMed] [Google Scholar]
  80. Fitzgerald PJ, Seemann JR, Maren S (2014. a). Can fear extinction be enhanced? A review of pharmacological and behavioral findings. Brain Res Bull 105: 46–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Fitzgerald PJ, Whittle N, Flynn SM, Graybeal C, Pinard CR, Gunduz-Cinar O et al (2014. b). Prefrontal single-unit firing associated with deficient extinction in mice. Neurobiol Learn Mem 113: 69–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Fitzgerald PJ, Pinard CR, Camp MC, Feyder M, Sah A, Bergstrom HC et al (2015. a). Durable fear memories require PSD-95 Mol Psychiatry 20: 901–912. [DOI] [PMC free article] [PubMed]
  83. Fitzgerald PJ, Giustino TF, Seemann JR, Maren S (2015. b). Noradrenergic blockade stabilizes fear extinction under stress. Proc Natl Acad Sci USA (doi:doi:10.1073/pnas.1500682112. [DOI] [PMC free article] [PubMed]
  84. Frankland P, Cestari V, Filipkowski R, McDonald R, Silva A (1998). The dorsal hippocampus is essential for context discrimination but not for contextual conditioning. Behav Neurosci 112: 863–874. [DOI] [PubMed] [Google Scholar]
  85. Ganon-Elazar E, Akirav I (2013). Cannabinoids and traumatic stress modulation of contextual fear extinction and GR expression in the amygdala-hippocampal-prefrontal circuit. Psychoneuroendocrinology 38: 1675–1687. [DOI] [PubMed] [Google Scholar]
  86. Garcia R, Chang C-H, Maren S (2006). Electrolytic lesions of the medial prefrontal cortex do not interfere with long-term memory of extinction of conditioned fear. Learn Mem 13: 14–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Garcia R, Vouimba R, Baudry M, Thompson R (1999). The amygdala modulates prefrontal cortex activity relative to conditioned fear. Nature 402: 294–296. [DOI] [PubMed] [Google Scholar]
  88. Garcia R, Spennato G, Nilsson-Todd L, Moreau JL, Deschaux O (2008). Hippocampal low-frequency stimulation and chronic mild stress similarly disrupt fear extinction memory in rats. Neurobiol Learn Mem 89: 560–566. [DOI] [PubMed] [Google Scholar]
  89. Garfinkel SN, Abelson JL, King AP, Sripada RK, Wang X, Gaines LM et al (2014). Impaired contextual modulation of memories in PTSD: an fMRI and psychophysiological study of extinction retention and fear renewal. J Neurosci 34: 13435–13443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. George SA, Rodriguez-Santiago M, Riley J, Rodriguez E, Liberzon I (2015). The effect of chronic phenytoin administration on single prolonged stress induced extinction retention deficits and glucocorticoid upregulation in the rat medial prefrontal cortex. Psychopharmacology (Berl) 232: 47–56. [DOI] [PubMed] [Google Scholar]
  91. Gewirtz J, Falls W, Davis M (1997). Normal conditioned inhibition and extinction of freezing and fear-potentiated startle following electrolytic lesions of medical prefrontal cortex in rats. Behav Neurosci 111: 712–726. [DOI] [PubMed] [Google Scholar]
  92. Giedd JN, Rapoport JL (2010). Structural MRI of pediatric brain development: what have we learned and where are we going? Neuron 67: 728–734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Glover EM, Jovanovic T, Norrholm SD (2015). Estrogen and extinction of fear memories: implications for posttraumatic stress disorder treatment. Biol Psychiatry (doi:10.1016/j.biopsych.2015.02.007. [DOI] [PMC free article] [PubMed]
  94. Gogolla N, Caroni P, Luthi A, Herry C (2009). Perineuronal nets protect fear memories from erasure. Science 325: 1258–1261. [DOI] [PubMed] [Google Scholar]
  95. Goode TD, Maren S (2014). Animal models of fear relapse. ILAR J 55: 246–258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Goshen I, Brodsky M, Prakash R, Wallace J, Gradinaru V, Ramakrishnan C et al (2011). Dynamics of retrieval strategies for remote memories. Cell 147: 678–689. [DOI] [PubMed] [Google Scholar]
  97. Goswami S, Cascardi M, Rodriguez-Sierra OE, Duvarci S, Pare D (2010). Impact of predatory threat on fear extinction in Lewis rats. Learn Mem 17: 494–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Goswami S, Rodríguez-Sierra O, Cascardi M, Paré D (2013). Animal models of post-traumatic stress disorder: face validity. Front Neurosci 7: 89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Gourley SL, Kedves AT, Olausson P, Taylor JR (2009). A history of corticosterone exposure regulates fear extinction and cortical NR2B, GluR2/3, and BDNF. Neuropsychopharmacology 34: 707–716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Gräff J, Joseph NF, Horn ME, Samiei A, Meng J, Seo J et al (2014). Epigenetic priming of memory updating during reconsolidation to attenuate remote fear memories. Cell 156: 261–276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Graham BM, Milad MR (2013). Blockade of estrogen by hormonal contraceptives impairs fear extinction in female rats and women. Biol Psychiatry 73: 371–378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Graham BM, Milad MR (2014). Inhibition of estradiol synthesis impairs fear extinction in male rats. Learn Mem 21: 347–350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Gray JM, Vecchiarelli HA, Morena M, Lee TT, Hermanson DJ, Kim AB et al (2015). Corticotropin-releasing hormone drives anandamide hydrolysis in the amygdala to promote anxiety. J Neurosci 35: 3879–3892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Graybeal C, Feyder M, Schulman E, Saksida LM, Bussey TJ, Brigman JL et al (2011). Paradoxical reversal learning enhancement by stress or prefrontal cortical damage: rescue with BDNF. Nat Neurosci 14: 1507–1509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Green MK, Rani CS, Joshi A, Soto-Pina AE, Martinez PA, Frazer A et al (2011). Prenatal stress induces long term stress vulnerability, compromising stress response systems in the brain and impairing extinction of conditioned fear after adult stress. Neuroscience 192: 438–451. [DOI] [PubMed] [Google Scholar]
  106. Grillo CA, Risher M, Macht VA, Bumgardner AL, Hang A, Gabriel C et al (2015). Repeated restraint stress-induced atrophy of glutamatergic pyramidal neurons and decreases in glutamatergic efflux in the rat amygdala are prevented by the antidepressant agomelatine. Neuroscience 284: 430–443. [DOI] [PubMed] [Google Scholar]
  107. Grillon C, Morgan CA (1999). Fear-potentiated startle conditioning to explicit and contextual cues in Gulf War veterans with posttraumatic stress disorder. J Abnorm Psychol 108: 134–142. [DOI] [PubMed] [Google Scholar]
  108. Gunduz-Cinar O, Hill MN, McEwen BS, Holmes A (2013. a). Amygdala FAAH and anandamide: mediating protection and recovery from stress. Trends Pharmacol Sci 34: 637–644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Gunduz-Cinar O, Macpherson KP, Cinar R, Gamble-George J, Sugden K, Williams B et al (2013. b). Convergent translational evidence of a role for anandamide in amygdala-mediated fear extinction, threat processing and stress-reactivity. Mol Psychiatry 18: 813–823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Guthrie RM, Bryant RA (2006). Extinction learning before trauma and subsequent posttraumatic stress. Psychosom Med 68: 307–311. [DOI] [PubMed] [Google Scholar]
  111. Han J, Kushner S, Yiu A, Cole C, Matynia A, Brown R et al (2007). Neuronal competition and selection during memory formation. Science 316: 457–460. [DOI] [PubMed] [Google Scholar]
  112. Hariri AR, Gorka A, Hyde LW, Kimak M, Halder I, Ducci F et al (2009). Divergent effects of genetic variation in endocannabinoid signaling on human threat- and reward-related brain function. Biol Psychiatry 66: 9–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Hartley CA, Gorun A, Reddan MC, Ramirez F, Phelps EA (2014). Stressor controlability modulates fear extincion in humans. Neurbiol Learn Mem 113: 149–156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Hartley CA, McKenna MC, Salman R, Holmes A, Casey BJ, Phelps EA et al (2012). Serotonin transporter polyadenylation polymorphism modulates the retention of fear extinction memory. Proc Natl Acad Sci USA 109: 5493–5498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Hefner K, Holmes A (2007). Ontogeny of fear-, anxiety- and depression-related behavior across adolescence in C57BL/6 J mice. Behav Brain Res 176: 210–215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Hefner K, Whittle N, Juhasz J, Norcross M, Karlsson RM, Saksida LM et al (2008). Impaired fear extinction learning and cortico-amygdala circuit abnormalities in a common genetic mouse strain. J Neurosci 28: 8074–8085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Herry C, Ferraguti F, Singewald N, Letzkus JJ, Ehrlich I, Luthi A (2010). Neuronal circuits of fear extinction. Eur J Neurosci 31: 599–612. [DOI] [PubMed] [Google Scholar]
  118. Herry C, Johansen JP (2014). Encoding of fear learning and memory in distributed neuronal circuits. Nat Neurosci 17: 1644–1654. [DOI] [PubMed] [Google Scholar]
  119. Herry C, Mons N (2004). Resistance to extinction is associated with impaired immediate early gene induction in medial prefrontal cortex and amygdala. Eur J Neurosci 20: 781–790. [DOI] [PubMed] [Google Scholar]
  120. Herry C, Trifilieff P, Micheau J, Lüthi A, Mons N (2006). Extinction of auditory fear conditioning requires MAPK/ERK activation in the basolateral amygdala. Eur J Neurosci 24: 261–269. [DOI] [PubMed] [Google Scholar]
  121. Hettema JM, Annas P, Neale MC, Kendler KS, Fredrikson M (2003). A twin study of the genetics of fear conditioning. Arch Gen Psychiatry 60: 702–708. [DOI] [PubMed] [Google Scholar]
  122. Hill MN, McEwen BS (2010. a). Involvement of the endocannabinoid system in the neurobehavioural effects of stress and glucocorticoids. Prog Neuropsychopharmacol Biol Psychiatry 34: 791–797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Hill MN, Patel S, Campolongo P, Tasker JG, Wotjak CT, Bains JS (2010. b). Functional interactions between stress and the endocannabinoid system: from synaptic signaling to behavioral output. J Neurosci 30: 14980–14986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Hoffman AN, Lorson NG, Sanabria F, Foster Olive M, Conrad CD (2014). Chronic stress disrupts fear extinction and enhances amygdala and hippocampal Fos expression in an animal model of post-traumatic stress disorder. Neurobiol Learn Mem 112: 139–147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Holmes A, Quirk GJ (2010). Pharmacological facilitation of fear extinction and the search for adjunct treatments for anxiety disorders–the case of yohimbine. Trends Pharmacol Sci 31: 2–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Holmes A, Singewald N (2013). Individual differences in recovery from traumatic fear. Trends Neurosci 36: 23–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Holmes A, Wellman CL (2009). Stress-induced prefrontal reorganization and executive dysfunction in rodents. Neurosci Biobehav Rev 33: 773–783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Huff N, Rudy J (2004). The amygdala modulates hippocampus-dependent context memory formation and stores cue-shock associations. Behav Neurosci 118: 53–62. [DOI] [PubMed] [Google Scholar]
  129. Hugues S, Chessel A, Lena I, Marsault R, Garcia R (2006). Prefrontal infusion of PD098059 immediately after fear extinction training blocks extinction-associated prefrontal synaptic plasticity and decreases prefrontal ERK2 phosphorylation. Synapse 60: 280–287. [DOI] [PubMed] [Google Scholar]
  130. Iafrati J, Orejarena MJ, Lassalle O, Bouamrane L, Chavis P (2013). Reelin, an extracellular matrix protein linked to early onset psychiatric diseases, drives postnatal development of the prefrontal cortex via GluN2B-NMDARs and the mTOR pathway. Mol Psychiatry 19: 417–426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Inslicht SS, Metzler TJ, Garcia NM, Pineles SL, Milad MR, Orr SP et al (2013). Sex differences in fear conditioning in posttraumatic stress disorder. J Psychiatr Res 47: 64–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Ishikawa S, Saito Y, Yanagawa Y, Otani S, Hiraide S, Shimamura K et al (2012). Early postnatal stress alters extracellular signal-regulated kinase signaling in the corticolimbic system modulating emotional circuitry in adult rats. Eur J Neurosci 35: 135–145. [DOI] [PubMed] [Google Scholar]
  133. Izquierdo A, Wellman CL, Holmes A (2006). Brief uncontrollable stress causes dendritic retraction in infralimbic cortex and resistance to fear extinction in mice. J Neurosci 26: 5733–5738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Janak PH, Corbit LH (2011). Deepened extinction following compound stimulus presentation: noradrenergic modulation. Learn Mem 18: 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Jin J, Maren S (2015). Fear renewal preferentially activates ventral hippocampal neurons projecting to both amygdala and prefrontal cortex in rats. Sci Rep 5: 8388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Johansen JP, Cain CK, Ostroff LE, LeDoux JE (2011). Molecular mechanisms of fear learning and memory. Cell 147: 509–524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Judo C, Matsumoto M, Yamazaki D, Hiraide S, Yanagawa Y, Kimura S et al (2010). Early stress exposure impairs synaptic potentiation in the rat medial prefrontal cortex underlying contextual fear extinction. Neuroscience 169: 1705–1714. [DOI] [PubMed] [Google Scholar]
  138. Karpova NN, Pickenhagen A, Lindholm J, Tiraboschi E, Kulesskaya N, Agustsdottir A et al (2012). Fear erasure in mice requires synergy between antidepressant drugs and extinction training. Science 334: 1731–1734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Karstoft KI, Statnikov A, Andersen SB, Madsen T, Galatzer-Levy IR (2015). Early identification of posttraumatic stress following military deployment: application of machine learning methods to a prospective study of Danish soldiers. J Affect Disord 184: 170–175. [DOI] [PubMed] [Google Scholar]
  140. Kessler RC, Demler O, Frank RG, Olfson M, Pincus HA, Walters EE et al (2005). Prevalence and treatment of mental disorders, 1990 to 2003. N Engl J Med 352: 2515–2523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Kessler RC, Avenevoli S, Costello EJ, Georgiades K, Green JG, Gruber MJ et al (2012). Prevalence, persistence, and sociodemographic correlates of DSM-IV disorders in the National Comorbidity Survey Replication Adolescent Supplement. Arch Gen Psychiatry 69: 372–380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Kessler RC, Berglund P, Demler O, Jin R, Merikangas KR, Walters EE (2005). Lifetime prevalence and age-of-onset distributions of DSM-IV disorders in the National Comorbidity Survey Replication. Arch Gen Psychiatry 62: 593–602. [DOI] [PubMed] [Google Scholar]
  143. Kessler RC, Sonnega A, Bromet E, Hughes M, Nelson CB (1995). Posttraumatic stress disorder in the National Comorbidity Survey. Arch Gen Psychiatry 52: 1048–1060. [DOI] [PubMed] [Google Scholar]
  144. Kim JH, Hamlin AS, Richardson R (2009). Fear extinction across development: the involvement of the medial prefrontal cortex as assessed by temporary inactivation and immunohistochemistry. J Neurosci 29: 10802–10808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Kim JH, Richardson R (2007). A developmental dissociation in reinstatement of an extinguished fear response in rats. Neurobiol Learn Mem 88: 48–57. [DOI] [PubMed] [Google Scholar]
  146. Kim SC, Jo YS, Kim IH, Kim H, Choi JS (2010). Lack of medial prefrontal cortex activation underlies the immediate extinction deficit. J Neurosci 30: 832–837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Kindt M, Soeter M, Vervliet B (2009). Beyond extinction: erasing human fear responses and preventing the return of fear. Nat Neurosci 12: 256–258. [DOI] [PubMed] [Google Scholar]
  148. Knapska E, Macias M, Mikosz M, Nowak A, Owczarek D, Wawrzyniak M et al (2012). Functional anatomy of neural circuits regulating fear and extinction. Proc Natl Acad Sci USA 109: 17093–17098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Knapska E, Maren S (2009). Reciprocal patterns of c-Fos expression in the medial prefrontal cortex and amygdala after extinction and renewal of conditioned fear. Learn Mem 16: 486–493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Knox D, George SA, Fitzpatrick CJ, Rabinak CA, Maren S, Liberzon I (2012. a). Single prolonged stress disrupts retention of extinguished fear in rats. Learn Mem 19: 43–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Knox D, Nault T, Henderson C, Liberzon I (2012. b). Glucocorticoid receptors and extinction retention deficits in the single prolonged stress model. Neuroscience 223: 163–173. [DOI] [PubMed] [Google Scholar]
  152. Knox D, Perrine SA, George SA, Galloway MP, Liberzon I (2010). Single prolonged stress decreases glutamate, glutamine, and creatine concentrations in the rat medial prefrontal cortex. Neurosci Lett 480: 16–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Konorski J (1967) Integrative Activity of the Brain: An Interdisciplinary Approach. University of Chicago Press: Chicago. [Google Scholar]
  154. Koseki H, Matsumoto M, Togashi H, Miura Y, Fukushima K, Yoshioka M (2009). Alteration of synaptic transmission in the hippocampal-mPFC pathway during extinction trials of context-dependent fear memory in juvenile rat stress models. Synapse 63: 805–813. [DOI] [PubMed] [Google Scholar]
  155. Lakshminarasimhan H, Chattarji S (2012). Stress leads to contrasting effects on the levels of brain derived neurotrophic factor in the hippocampus and amygdala. PLoS One 7: e30481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Landers MS, Sullivan RM (2012). The development and neurobiology of infant attachment and fear. Dev Neurosci 34: 101–114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Lattal KM, Wood MA (2013). Epigenetics and persistent memory: implications for reconsolidation and silent extinction beyond the zero. Nat Neurosci 16: 124–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Laurent V, Westbrook R (2008). Distinct contributions of the basolateral amygdala and the medial prefrontal cortex to learning and relearning extinction of context conditioned fear. Learn Mem 15: 657–666. [DOI] [PubMed] [Google Scholar]
  159. Laurent V, Westbrook R (2009). Inactivation of the infralimbic but not the prelimbic cortex impairs consolidation and retrieval of fear extinction. Learn Mem 16: 520–529. [DOI] [PubMed] [Google Scholar]
  160. Lebron-Milad K, Abbs B, Milad MR, Linnman C, Rougemount-Bücking A, Zeidan MA et al (2012). Sex differences in the neurobiology of fear conditioning and extinction: a preliminary fMRI study of shared sex differences with stress-arousal circuitry. Biol Mood Anxiety Disord 2: 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. LeDoux JE (2000). Emotion circuits in the brain. Annu Rev Neurosci 23: 155–184. [DOI] [PubMed] [Google Scholar]
  162. Lee H, Kim J (1998). Amygdalar NMDA receptors are critical for new fear learning in previously fear-conditioned rats. J Neurosci 18: 8444–8454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Leuner B, Gould E (2010). Structural plasticity and hippocampal function. Annu Rev Psychol 61: C111–C113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Leweke FM, Piomelli D, Pahlisch F, Muhl D, Gerth CW, Hoyer C et al (2012). Cannabidiol enhances anandamide signaling and alleviates psychotic symptoms of schizophrenia. Transl Psychiatry 2: e94. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Liberzon I, King AP, Ressler KJ, Almli LM, Zhang P, Ma ST et al (2014). Interaction of the ADRB2 gene polymorphism with childhood trauma in predicting adult symptoms of posttraumatic stress disorder. JAMA Psychiatry 71: 1174–1182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Liberzon I, Sripada C (2008). The functional neuroanatomy of PTSD: a critical review. Prog Brain Res 167: 151–169. [DOI] [PubMed] [Google Scholar]
  167. Likhtik E, Popa D, Apergis-Schoute J, Fidacaro GA, Paré D (2008). Amygdala intercalated neurons are required for expression of fear extinction. Nature 454: 642–645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Lissek S, Powers AS, McClure EB, Phelps EA, Woldehawariat G, Grillon C et al (2005). Classical fear conditioning in the anxiety disorders: a meta-analysis. Behav Res Ther 43: 1391–1424. [DOI] [PubMed] [Google Scholar]
  169. Logue MW, Baldwin C, Guffanti G, Melista E, Wolf EJ, Reardon AF et al (2013). A genome-wide association study of post-traumatic stress disorder identifies the retinoid-related orphan receptor alpha (RORA) gene as a significant risk locus. Mol Psychiatry 18: 937–942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Long VA, Fanselow MS (2012). Stress-enhanced fear learning in rats is resistant to the effects of immediate massed extinction. Stress 15: 627–636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Lonsdorf TB, Weike AI, Nikamo P, Schalling M, Hamm AO, Ohman A (2009). Genetic gating of human fear learning and extinction: possible implications for gene-environment interaction in anxiety disorder. Psychol Sci 20: 198–206. [DOI] [PubMed] [Google Scholar]
  172. Macpherson K, Whittle N, Camp M, Gunduz-Cinar O, Singewald N, Holmes A (2013). Temporal factors in the extinction of fear in inbred mouse strains differing in extinction efficacy. Biol Mood Anxiety Disord 3: 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Mahan AL, Ressler KJ (2012). Fear conditioning, synaptic plasticity and the amygdala: implications for posttraumatic stress disorder. Trends Neurosci 35: 24–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Maren S (1999). Long-term potentiation in the amygdala: a mechanism for emotional learning and memory. Trends Neurosci 22: 561–567. [DOI] [PubMed] [Google Scholar]
  175. Maren S (2001). Neurobiology of Pavlovian fear conditioning. Annu Rev Neurosci 24: 897–931. [DOI] [PubMed] [Google Scholar]
  176. Maren S (2011). Seeking a spotless mind: extinction, deconsolidation, and erasure of fear memory. Neuron 70: 830–845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Maren S (2014. a). Out with the old and in with the new: synaptic mechanisms of extinction in the amygdala. Brain Res (doi:10.1016/j.brainres.2014.10.010. [DOI] [PMC free article] [PubMed]
  178. Maren S (2014. b). Nature and causes of the immediate extinction deficit: A brief review. Neurobiol Learn Mem 113: 19–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Maren S, Aharonov G, Fanselow MS (1997). Neurotoxic lesions of the dorsal hippocampus and Pavlovian fear conditioning in rats. Behav Brain Res 88: 261–274. [DOI] [PubMed] [Google Scholar]
  180. Maren S, Aharonov G, Stote DL, Fanselow MS (1996). N-methyl-D-aspartate receptors in the basolateral amygdala are required for both acquisition and expression of conditional fear in rats. Behav Neurosci 110: 1365–1374. [DOI] [PubMed] [Google Scholar]
  181. Maren S, Chang C-H (2006). Recent fear is resistant to extinction. Proc Natl Acad Sci USA 103: 18020–18025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Maren S, Fanselow MS (1995). Synaptic plasticity in the basolateral amygdala induced by hippocampal formation stimulation in vivo. J Neurosci 15: 7548–7564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Maren S, Holt W (2000). The hippocampus and contextual memory retrieval in Pavlovian conditioning. Behav Brain Res 110: 97–108. [DOI] [PubMed] [Google Scholar]
  184. Maren S, Phan KL, Liberzon I (2013). The contextual brain: implications for fear conditioning, extinction and psychopathology. Nat Rev Neurosci 14: 417–428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Maren S, Quirk GJ (2004). Neuronal signalling of fear memory. Nat Rev Neurosci 5: 844–852. [DOI] [PubMed] [Google Scholar]
  186. Maroun M (2006). Stress reverses plasticity in the pathway projecting from the ventromedial prefrontal cortex to the basolateral amygdala. Eur J Neurosci 24: 2917–2922. [DOI] [PubMed] [Google Scholar]
  187. Maroun M, Ioannides PJ, Bergman KL, Kavushansky A, Holmes A, Wellman CL (2013). Fear extinction deficits following acute stress associate with increased spine density and dendritic retraction in basolateral amygdala neurons. Eur J Neurosci 38: 2611–2620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Maroun M, Richter-Levin G (2003). Exposure to acute stress blocks the induction of long-term potentiation of the amygdala-prefrontal cortex pathway in vivo. J Neurosci 23: 4406–4409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Mathur P, Graybeal C, Feyder M, Davis MI, Holmes A (2009). Fear memory impairing effects of systemic treatment with the NMDA NR2B subunit antagonist, Ro 25-6981, in mice: attenuation with ageing. Pharmacol Biochem Behav 91: 453–460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Matsumoto M, Togashi H, Konno K, Koseki H, Hirata R, Izumi T et al (2008). Early postnatal stress alters the extinction of context-dependent conditioned fear in adult rats. Pharmacol Biochem Behav 89: 247–252. [DOI] [PubMed] [Google Scholar]
  191. Matsumoto Y, Morinobu S, Yamamoto S, Matsumoto T, Takei S, Fujita Y et al (2013). Vorinostat ameliorates impaired fear extinction possibly via the hippocampal NMDA-CaMKII pathway in an animal model of posttraumatic stress disorder. Psychopharmacology (Berl) 229: 51–62. [DOI] [PubMed] [Google Scholar]
  192. McCallum J, Kim JH, Richardson R (2010). Impaired extinction retention in adolescent rats: effects of D-cycloserine. Neuropsychopharmacology 35: 2134–2142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. McCormick CM, Mongillo DL, Simone JJ (2013). Age and adolescent social stress effects on fear extinction in female rats. Stress 16: 678–688. [DOI] [PubMed] [Google Scholar]
  194. McEwen BS, Morrison JH (2013). The brain on stress: vulnerability and plasticity of the prefrontal cortex over the life course. Neuron 79: 16–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Medina JF, Christopher Repa J, Mauk MD, LeDoux JE (2002). Parallels between cerebellum- and amygdala-dependent conditioning. Nat Rev Neurosci 3: 122–131. [DOI] [PubMed] [Google Scholar]
  196. Merino SM, Maren S (2006). Hitting Ras where it counts: Ras antagonism in the basolateral amygdala inhibits long-term fear memory. Eur J Neurosci 23: 196–204. [DOI] [PubMed] [Google Scholar]
  197. Merz CJ, Hamacher-Dang TC, Wolf OT (2014). Exposure to stress attenuates fear retrieval in healthy men. Psychoneuroendocrinology 41: 89–96. [DOI] [PubMed] [Google Scholar]
  198. Merz CJ, Tabbert K, Schweckendiek J, Klucken T, Vaitl D, Stark R et al (2012). Neuronal correlates of extinction learning are modulated by sex hormones. Soc Cogn Affect Neurosci 7: 819–830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Milad M, Quirk G (2012). Fear extinction as a model for translational neuroscience: ten years of progress. Annu Rev Psychol 63: 129–151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Milad M, Vidal-Gonzalez I, Quirk G (2004). Electrical stimulation of medial prefrontal cortex reduces conditioned fear in a temporally specific manner. Behav Neurosci 118: 389–394. [DOI] [PubMed] [Google Scholar]
  201. Milad MR, Goldstein JM, Orr SP, Wedig MM, Klibanski A, Pitman RK et al (2006). Fear conditioning and extinction: influence of sex and menstrual cycle in healthy humans. Behav Neurosci 120: 1196–1203. [DOI] [PubMed] [Google Scholar]
  202. Milad MR, Orr SP, Lasko NB, Chang Y, Rauch SL, Pitman RK (2008). Presence and acquired origin of reduced recall for fear extinction in PTSD: results of a twin study. J Psychiatr Res 42: 515–520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Milad MR, Pitman RK, Ellis CB, Gold AL, Shin LM, Lasko NB et al (2009). Neurobiological basis of failure to recall extinction memory in posttraumatic stress disorder. Biol Psychiatry 66: 1075–1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Milad MR, Quirk GJ (2002). Neurons in medial prefrontal cortex signal memory for fear extinction. Nature 420: 70–74. [DOI] [PubMed] [Google Scholar]
  205. Milad MR, Zeidan MA, Contero A, Pitman RK, Klibanski A, Rauch SL et al (2010). The influence of gonadal hormones on conditioned fear extinction in healthy humans. Neuroscience 168: 652–658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Miracle AD, Brace MF, Huyck KD, Singler SA, Wellman CL (2006). Chronic stress impairs recall of extinction of conditioned fear. Neurobiol Learn Mem 85: 213–218. [DOI] [PubMed] [Google Scholar]
  207. Miserendino M, Sananes C, Melia K, Davis M (1990). Blocking of acquisition but not expression of conditioned fear-potentiated startle by NMDA antagonists in the amygdala. Nature 345: 716–718. [DOI] [PubMed] [Google Scholar]
  208. Mitra R, Jadhav S, McEwen BS, Vyas A, Chattarji S (2005). Stress duration modulates the spatiotemporal patterns of spine formation in the basolateral amygdala. Proc Natl Acad Sci USA 102: 9371–9376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Monfils M-H, Cowansage KK, Klann E, LeDoux JE (2009). Extinction-reconsolidation boundaries: key to persistent attenuation of fear memories. Science 324: 951–955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Morena M, Campolongo P (2014). The endocannabinoid system: an emotional buffer in the modulation of memory function. Neurobiol Learn Mem 112: 30–43. [DOI] [PubMed] [Google Scholar]
  211. Morgan M, LeDoux J (1995). Differential contribution of dorsal and ventral medial prefrontal cortex to the acquisition and extinction of conditioned fear in rats. Behav Neurosci 109: 681–688. [DOI] [PubMed] [Google Scholar]
  212. Mueller D, Porter JT, Quirk GJ (2008). Noradrenergic signaling in infralimbic cortex increases cell excitability and strengthens memory for fear extinction. J Neurosci 28: 369–375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Muigg P, Hetzenauer A, Hauer G, Hauschild M, Gaburro S, Frank E et al (2008). Impaired extinction of learned fear in rats selectively bred for high anxiety–evidence of altered neuronal processing in prefrontal-amygdala pathways. Eur J Neurosci 28: 2299–2309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Myers K, Davis M (2002). Behavioral and neural analysis of extinction. Neuron 36: 567–584. [DOI] [PubMed] [Google Scholar]
  215. Nader K, Hardt O (2009). A single standard for memory: the case for reconsolidation. Nat Rev Neurosci 10: 224–234. [DOI] [PubMed] [Google Scholar]
  216. Naert A, Callaerts-Vegh Z, D'Hooge R (2011). Nocturnal hyperactivity, increased social novelty preference and delayed extinction of fear responses in postweaning socially isolated mice. Brain Res Bull 85: 354e362. [DOI] [PubMed] [Google Scholar]
  217. Ninomiya EM, Martynhak BJ, Zanoveli JM, Correia D, da Cunha C, Andreatini R (2010). Spironolactone and low-dose dexamethasone enhance extinction of contextual fear conditioning. Prog Neuropsychopharmacol Biol Psychiatry 34: 1229–1235. [DOI] [PubMed] [Google Scholar]
  218. Orr SP, Metzger LJ, Lasko NB, Macklin ML, Peri T, Pitman RK (2000). De novo conditioning in trauma-exposed individuals with and without posttraumatic stress disorder. J Abnorm Psychol 109: 290–298. [PubMed] [Google Scholar]
  219. Orsini CA, Kim JH, Knapska E, Maren S (2011). Hippocampal and prefrontal projections to the basal amygdala mediate contextual regulation of fear after extinction. J Neurosci 31: 17269–17277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Orsini CA, Maren S (2012). Neural and cellular mechanisms of fear and extinction memory formation. Neurosci Biobehav Rev 36: 1773–1802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Orsini CA, Yan C, Maren S (2013). Ensemble coding of context-dependent fear memory in the amygdala. Front Behav Neurosci 7: 199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Otis JM, Werner CT, Mueller D (2015). Noradrenergic regulation of fear and drug-associated memory reconsolidation. Neuropsychopharmacology 40: 793–803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Padival MA, Blume SR, Rosenkranz JA (2013). Repeated restraint stress exerts different impact on structure of neurons in the lateral and basal nuclei of the amygdala. Neuroscience 246: 230–242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Pamplona FA, Prediger RD, Pandolfo P, Takahashi RN (2006). The cannabinoid receptor agonist WIN 55,212-2 facilitates the extinction of contextual fear memory and spatial memory in rats. Psychopharmacology (Berl) 188: 641–649. [DOI] [PubMed] [Google Scholar]
  225. Pape H-C, Pare D (2010). Plastic synaptic networks of the amygdala for the acquisition, expression, and extinction of conditioned fear. Physiol Rev 90: 419–463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Paré D, Quirk GJ, LeDoux JE (2004). New vistas on amygdala networks in conditioned fear. J Neurophysiol 92: 1–9. [DOI] [PubMed] [Google Scholar]
  227. Pattwell SS, Duhoux S, Hartley CA, Johnson DC, Jing D, Elliott MD et al (2012). Altered fear learning across development in both mouse and human. Proc Natl Acad Sci USA 109: 16318–16323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Pattwell SS, Lee FS, Casey BJ (2013). Fear learning and memory across adolescent development: hormones and behavior special issue: puberty and adolescence. Horm Behav 64: 380–389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. Pavlov IPConditioned reflexes: an investigation of the physiological activity of the cerebral cortex. Dover Publications: New York; (1927). [Google Scholar]
  230. Peters J, Dieppa-Perea LM, Melendez LM, Quirk GJ (2010). Induction of fear extinction with hippocampal-infralimbic BDNF. Science 328: 1288–1290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Pitman R, Brunet A, Orr S, Tremblay J, Nader K (2006). A novel treatment for post-traumatic stress disorder by reconsolidation blockade with propranolol. Neuropsychopharmacology 31: S8–S9. [Google Scholar]
  232. Pitman RK, Rasmusson AM, Koenen KC, Shin LM, Orr SP, Gilbertson MW et al (2012). Biological studies of post-traumatic stress disorder. Nat Rev Neurosci 13: 769–787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. Pizzimenti CL, Lattal KM (2015). Epigenetics and memory: causes, consequences and treatments for post-traumatic stress disorder and addiction. Genes Brain Behav 14: 73–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Ponder CA, Kliethermes CL, Drew MR, Muller J, Das K, Risbrough VB et al (2007). Selection for contextual fear conditioning affects anxiety-like behaviors and gene expression. Genes Brain Behav 6: 736–749. [DOI] [PubMed] [Google Scholar]
  235. Ponomarev I, Rau V, Eger EI, Harris RA, Fanselow MS (2010). Amygdala transcriptome and cellular mechanisms underlying stress-enhanced fear learning in a rat model of posttraumatic stress disorder. Neuropsychopharmacology 35: 1402–1411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Powers MB, Smits JA, Otto MW, Sanders C, Emmelkamp PM (2009). Facilitation of fear extinction in phobic participants with a novel cognitive enhancer: a randomized placebo controlled trial of yohimbine augmentation. J Anxiety Disord 23: 350–356. [DOI] [PubMed] [Google Scholar]
  237. Preston AR, Eichenbaum H (2013). Interplay of hippocampus and prefrontal cortex in memory. Curr Biol 23: R764–R773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Quirk G, Mueller D (2008). Neural mechanisms of extinction learning and retrieval. Neuropsychopharmacology 33: 56–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  239. Quirk GJ, Likhtik E, Pelletier JG, Paré D (2003). Stimulation of medial prefrontal cortex decreases the responsiveness of central amygdala output neurons. J Neurosci 23: 8800–8807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  240. Quirk GJ, Russo GK, Barron JL, Lebron K (2000). The role of ventromedial prefrontal cortex in the recovery of extinguished fear. J Neurosci 20: 6225–6231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Rabinak CA, Angstadt M, Lyons M, Mori S, Milad MR, Liberzon I et al (2013. a). Cannabinoid modulation of prefrontal-limbic activation during fear extinction learning and recall in humans. Neurobiol Learn Mem. 113: 125–134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Rabinak CA, Angstadt M, Sripada CS, Abelson JL, Liberzon I, Milad MR et al (2013. b). Cannabinoid facilitation of fear extinction memory recall in humans. Neuropharmacology 64: 396–402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Ramchand R, Karney BR, Osilla KC, Burns RM (2008) Prevalence of PTSD, depression, and TBI among returning servicemembers Tanielian T, Jaycox LH (eds). Invisible Wounds of War: Psychological and Cognitive Injuries, Their Consequences, and Services to Assist Recovery. RAND Corporation: Santa Monica, CA, USA, pp 35–85. [Google Scholar]
  244. Rau V, DeCola JP, Fanselow MS (2005). Stress-induced enhancement of fear learning: an animal model of posttraumatic stress disorder. Neurosci Biobehav Rev 29: 1207–1223. [DOI] [PubMed] [Google Scholar]
  245. Riao CM, Brignoni-Perez E, Goldman R, Phelps EA (2014). Acute stress impairs the retrieval of extinction memory in humans. Neurobiol Learn Mem 112: 212–221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  246. Riao CM, Phelps EA (2015). The influence of acute stress on the regulation of conditioned fear. Neurobiol Stress 1: 134–146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  247. Richter-Levin G, Maroun M (2010). Stress and amygdala suppression of metaplasticity in the medial prefrontal cortex. Cereb Cortex 20: 2433–2441. [DOI] [PubMed] [Google Scholar]
  248. Rocher C, Spedding M, Munoz C, Jay TM (2004). Acute stress-induced changes in hippocampal/prefrontal circuits in rats: effects of antidepressants. Cereb Cortex 14: 224–229. [DOI] [PubMed] [Google Scholar]
  249. Rodrigues SM, Schafe GE, LeDoux JE (2001). Intra-amygdala blockade of the NR2B subunit of the NMDA receptor disrupts the acquisition but not the expression of fear conditioning. J Neurosci 21: 6889–6896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  250. Roozendaal B, McEwen BS, Chattarji S (2009). Stress, memory and the amygdala. Nat Rev Neurosci 10: 423–433. [DOI] [PubMed] [Google Scholar]
  251. Roozendaal B, McGaugh JL (2011). Memory modulation. Behav Neurosci 125: 797–824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  252. Rosenkranz JA, Venheim ER, Padival M (2010). Chronic stress causes amygdala hyperexcitability in rodents. Biol Psychiatry 67: 1128–1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  253. Rothbaum B, Davis M (2003). Applying learning principles to the treatment of post-trauma reactions. Ann NY Acad Sci 1008: 112–121. [DOI] [PubMed] [Google Scholar]
  254. Rougemont-Bücking A, Linnman C, Zeffiro TA, Zeidan MA, Lebron-Milad K, Rodriguez-Romaguera J et al (2011). Altered processing of contextual information during fear extinction in PTSD: an fMRI study. CNS Neurosci Ther 17: 227–236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Rozeske RR, Valerio S, Chaudun F, Herry C (2015). Prefrontal neuronal circuits of contextual fear conditioning. Genes Brain Behav 14: 22–36. [DOI] [PubMed] [Google Scholar]
  256. Rudy J, Huff N, Matus-Amat P (2004). Understanding contextual fear conditioning: insights from a two-process model. Neurosci Biobehav Rev 28: 675–685. [DOI] [PubMed] [Google Scholar]
  257. Rudy JW (2009). Context representations, context functions, and the parahippocampal-hippocampal system. Learn Mem 16: 573–585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  258. Rumpel S (2005). Postsynaptic receptor trafficking underlying a form of associative learning. Science 308: 83–88. [DOI] [PubMed] [Google Scholar]
  259. Ryan SJ, Ehrlich DE, Rainnie DG (2014). Morphology and dendritic maturation of developing principal neurons in the rat basolateral amygdala. Brain Struct Funct. [DOI] [PMC free article] [PubMed]
  260. Saito Y, Matsumoto M, Otani S, Yanagawa Y, Hiraide S, Ishikawa S et al (2012). Phase-dependent synaptic changes in the hippocampal CA1 field underlying extinction processes in freely moving rats. Neurobiol Learn Mem 97: 361–369. [DOI] [PubMed] [Google Scholar]
  261. Saito Y, Matsumoto M, Yanagawa Y, Hiraide S, Inoue S, Kubo Y et al (2013). Facilitation of fear extinction by the 5-HT(1 A) receptor agonist tandospirone: possible involvement of dopaminergic modulation. Synapse 67: 161–170. [DOI] [PubMed] [Google Scholar]
  262. Schayek R, Maroun M (2014). Differences in stress-induced changes in extinction and prefrontal plasticity in postweanling and adult animals. Biol Psychiatry (doi:10.1016/j.biopsych.2014.10.004). [DOI] [PubMed]
  263. Schiller D, Monfils M-H, Raio CM, Johnson DC, LeDoux JE, Phelps EA (2010). Preventing the return of fear in humans using reconsolidation update mechanisms. Nature 463: 49–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Segev A, Rubin AS, Abush H, Richter-Levin G, Akirav I (2013). Cannabinoid receptor activation prevents the effects of chronic mild stress on emotional learning and LTP in a rat model of depression. Neuropsychopharmacology 39: 919–933. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Senn V, Wolff SBE, Herry C, Grenier F, Ehrlich I, Gründemann J et al (2014). Long-range connectivity defines behavioral specificity of amygdala neurons. Neuron 81: 428–437. [DOI] [PubMed] [Google Scholar]
  266. Shansky RM (2015). Sex differences in PTSD resilience and susceptibility: challenges for animal models of fear learning. Neurobiol Stress 1: 60–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  267. Shechner T, Hong M, Britton JC, Pine DS, Fox NA (2014). Fear conditioning and extinction across development: evidence from human studies and animal models. Biol Psychol 100: 1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  268. Shumake J, Barrett D, Gonzalez-Lima F (2005). Behavioral characteristics of rats predisposed to learned helplessness: reduced reward sensitivity, increased novelty seeking, and persistent fear memories. Behav Brain Res 164: 222–230. [DOI] [PubMed] [Google Scholar]
  269. Shumake J, Furgeson-Moreira S, Monfils MH (2014). Predictability and heritability of individual differences in fear learning. Anim Cogn 17: 1207–1221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Shvil E, Sullivan GM, Schafer S, Markowitz JC, Campeas M, Wager TD et al (2014). Sex differences in extinction recall in posttraumatic stress disorder: a pilot fMRI study. Neurobiol Learn Mem 113: 101–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Sierra-Mercado D, Padilla-Coreano N, Quirk G (2011). Dissociable roles of prelimbic and infralimbic cortices, ventral hippocampus, and basolateral amygdala in the expression and extinction of conditioned fear. Neuropsychopharmacology 36: 529–538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Singewald N, Schmuckermair C, Whittle N, Holmes A, Ressler KJ (2015). Pharmacology of cognitive enhancers for exposure-based therapy of fear, anxiety and trauma-related disorders. Pharmacol Ther 149: 150–190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Skelly MJ, Chappell AE, Carter E, Weiner JL (2015). Adolescent social isolation increases anxiety-like behavior and ethanol intake and impairs fear extinction in adulthood: possible role of disrupted noradrenergic signaling. Neuropharmacology 97: 149–159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  274. Smits JA, Rosenfield D, Davis ML, Julian K, Handelsman PR, Otto MW et al (2014). Yohimbine enhancement of exposure therapy for social anxiety disorder: a randomized controlled trial. Biol Psychiatry 75: 840–846. [DOI] [PubMed] [Google Scholar]
  275. Soeter M, Kindt M (2011). Disrupting reconsolidation: pharmacological and behavioral manipulations. Learn Mem 18: 357–366. [DOI] [PubMed] [Google Scholar]
  276. Soravia LM, Heinrichs M, Aerni A, Maroni C, Schelling G, Ehlert U et al (2006). Glucocorticoids reduce phobic fear in humans. Proc Natl Acad Sci USA 103: 5585–5590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  277. Sotres-Bayon F, Bush DEA, LeDoux JE (2007). Acquisition of fear extinction requires activation of NR2B-containing NMDA receptors in the lateral amygdala. Neuropsychopharmacology 32: 1929–1940. [DOI] [PubMed] [Google Scholar]
  278. Sotres-Bayon F, Cain C, LeDoux J (2006). Brain mechanisms of fear extinction: historical perspectives on the contribution of prefrontal cortex. Biol Psychiatry 60: 329–336. [DOI] [PubMed] [Google Scholar]
  279. Sotres-Bayon F, Sierra-Mercado D, Pardilla-Delgado E, Quirk GJ (2012). Gating of fear in prelimbic cortex by hippocampal and amygdala inputs. Neuron 76: 804–812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Sparks F, Lehmann H, Sutherland R (2011). Between-systems memory interference during retrieval. Eur J Neurosci 34: 780–786. [DOI] [PubMed] [Google Scholar]
  281. Sparta DR, Smithuis J, Stamatakis AM, Jennings JH, Kantak PA, Ung RL et al (2014). Inhibition of projections from the basolateral amygdala to the entorhinal cortex disrupts the acquisition of contextual fear. Front Behav Neurosci 8: 129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  282. Spennato G, Zerbib C, Mondadori C, Garcia R (2008). Fluoxetine protects hippocampal plasticity during conditioned fear stress and prevents fear learning potentiation. Psychopharmacology (Berl) 196: 583–589. [DOI] [PubMed] [Google Scholar]
  283. Stafford JM, Maughan DK, Ilioi EC, Lattal KM (2013). Exposure to a fearful context during periods of memory plasticity impairs extinction via hyperactivation of frontal-amygdalar circuits. Learn Mem 20: 156–163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Stafford JM, Raybuck JD, Ryabinin AE, Lattal KM (2012). Increasing histone acetylation in the hippocampus-infralimbic network enhances fear extinction. Biol Psychiatry 72: 25–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  285. Stevens JS, Almli LM, Fani N, Gutman DA, Bradley B, Norrholm SD et al (2014). PACAP receptor gene polymorphism impacts fear responses in the amygdala and hippocampus. Proc Natl Acad Sci USA 111: 3158–3163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Strobel C, Marek R, Gooch HM, Sullivan RKP, Sah P (2015). Prefrontal and auditory input to intercalated neurons of the amygdala. Cell Rep (doi:10.1016/j.celrep.2015.02.008. [DOI] [PubMed]
  287. Suris A, North C, Adinoff B, Powell CM, Greene R (2010). Effects of exogenous glucocorticoid on combat-related PTSD symptoms. Ann Clin Psychiatry 22: 274–279. [PMC free article] [PubMed] [Google Scholar]
  288. Suvrathan A, Bennur S, Ghosh S, Tomar A, Anilkumar S, Chattarji S (2014). Stress enhances fear by forming new synapses with greater capacity for long-term potentiation in the amygdala. Philos Trans R Soc Lond B Biol Sci 369: 20130151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  289. Ter Horst JP, Carobrez AP, van der Mark MH, de Kloet ER, Oitzl MS (2012). Sex differences in fear memory and extinction of mice with forebrain-specific disruption of the mineralocorticoid receptor. Eur J Neurosci 36: 3096–3102. [DOI] [PubMed] [Google Scholar]
  290. Toledo-Rodriguez M, Pitiot A, Paus T, Sandi C (2012). Stress during puberty boosts metabolic activation associated with fear-extinction learning in hippocampus, basal amygdala and cingulate cortex. Neurobiol Learn Mem 98: 93–101. [DOI] [PubMed] [Google Scholar]
  291. Tronson NC, Corcoran KA, Jovasevic V, Radulovic J (2012). Fear conditioning and extinction: emotional states encoded by distinct signaling pathways. Trends Neurosci 35: 145–155.22118930 [Google Scholar]
  292. Trouche S, Sasaki JM, Tu T, Reijmers LG (2013). Fear extinction causes target-specific remodeling of perisomatic inhibitory synapses. Neuron 80: 1054–1065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  293. VanElzakker MB, Dahlgren MK, Davis FC, Dubois S, Shin LM (2014). From Pavlov to PTSD: the extinction of conditioned fear in rodents, humans, and anxiety disorders. Neurobiol Learn Mem 113: 3–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Vervliet B, Craske MG, Hermans D (2013). Fear extinction and relapse: state of the art. Annu Rev Clin Psychol 9: 215–248. [DOI] [PubMed] [Google Scholar]
  295. Vukojevic V, Kolassa IT, Fastenrath M, Gschwind L, Spalek K, Milnik A et al (2014). Epigenetic modification of the glucocorticoid receptor gene is linked to traumatic memory and post-traumatic stress disorder risk in genocide survivors. J Neurosci 34: 10274–10284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  296. Vyas A, Jadhav S, Chattarji S (2006). Prolonged behavioral stress enhances synaptic connectivity in the basolateral amygdala. Neuroscience 143: 387–393. [DOI] [PubMed] [Google Scholar]
  297. Vyas A, Mitra R, Shankaranarayana Rao BS, Chattarji S (2002). Chronic stress induces contrasting patterns of dendritic remodeling in hippocampal and amygdaloid neurons. J Neurosci 22: 6810–6818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  298. White MG, Bogdan R, Fisher PM, Munoz KE, Williamson DE, Hariri AR (2012). FKBP5 and emotional neglect interact to predict individual differences in amygdala reactivity. Genes Brain Behav 11: 869–878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  299. Walker DL, Ressler KJ, Lu K-T, Davis M (2002). Facilitation of conditioned fear extinction by systemic administration or intra-amygdala infusions of D-cycloserine as assessed with fear-potentiated startle in rats. J Neurosci 22: 2343–2351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  300. Watson J, Rayner R (1920). Conditioned emotional reactions. J Exp Psychol 3: 1–14. [DOI] [PubMed] [Google Scholar]
  301. Wessa M, Flor H (2007). Failure of extinction of fear responses in posttraumatic stress disorder: evidence from second-order conditioning. Am J Psychiatry 164: 1684–1692. [DOI] [PubMed] [Google Scholar]
  302. Whittle N, Hauschild M, Lubec G, Holmes A, Singewald N (2010). Rescue of impaired fear extinction and normalization of cortico-amygdala circuit dysfunction in a genetic mouse model by dietary zinc restriction. J Neurosci 30: 13586–13596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  303. Whittle N, Schmuckermair C, Gunduz Cinar O, Hauschild M, Ferraguti F, Homes A et al (2013). Deep brain stimulation, histone deacetylase inhibitors and glutamatergic drugs rescue resistance to fear extinction in a genetic mouse model. Neuropsychopharmacology 64: 414–423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  304. Wilber AA, Southwood CJ, Wellman CL (2009). Brief neonatal maternal separation alters extinction of conditioned fear and corticolimbic glucocorticoid and NMDA receptor expression in adult rats. Dev Neurobiol 69: 73–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. Wilber AA, Walker AG, Southwood CJ, Farrell MR, Lin GL, Rebec GV et al (2011). Chronic stress alters neural activity in medial prefrontal cortex during retrieval of extinction. Neuroscience 174: 115–131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Wilson CA, Vazdarjanova A, Terry AV Jr. (2013). Exposure to variable prenatal stress in rats: effects on anxiety-related behaviors, innate and contextual fear, and fear extinction. Behav Brain Res 238: 279–288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Wolpe J (1968). Psychotherapy by reciprocal inhibition. Conditional Reflex: A Pavlovian Journal of Research & Therapy 3: 234–240. [DOI] [PubMed] [Google Scholar]
  308. Wood NE, Rosasco ML, Suris AM, Spring JD, Marin MF, Lasko NB et al (2015). Pharmacological blockade of memory reconsolidation in posttraumatic stress disorder: three negative psychophysiological studies. Psychiatry Res 225: 31–39. [DOI] [PubMed] [Google Scholar]
  309. Xie P, Kranzler HR, Farrer L, Gelernter J (2012). Serotonin transporter 5-HTTLPR genotype moderates the effects of childhood adversity on posttraumatic stress disorder risk: a replication study. Am J Med Genet B Neuropsychiatr Genet 159B: 644–652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  310. Xie P, Kranzler HR, Yang C, Zhao H, Farrer LA, Gelernter J (2013). Genome-wide association study identifies new susceptibility loci for posttraumatic stress disorder. Biol Psychiatry 74: 656–663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  311. Xiong GJ, Yang Y, Wang LP, Xu L, Mao RR (2014). Maternal separation exaggerates spontaneous recovery of extinguished contextual fear in adult female rats. Behav Brain Res 269: 75–80. [DOI] [PubMed] [Google Scholar]
  312. Yamamoto S, Morinobu S, Fuchikami M, Kurata A, Kozuru T, Yamawaki S (2008). Effects of single prolonged stress and D-cycloserine on contextual fear extinction and hippocampal NMDA receptor expression in a rat model of PTSD. Neuropsychopharmacology 33: 2108–2116. [DOI] [PubMed] [Google Scholar]
  313. Yamamoto S, Morinobu S, Takei S, Fuchikami M, Matsuki A, Yamawaki S et al (2009). Single prolonged stress: toward an animal model of posttraumatic stress disorder. Depress Anxiety 26: 1110–1117. [DOI] [PubMed] [Google Scholar]
  314. Yang YL, Chao PK, Lu KT (2006). Systemic and intra-amygdala administration of glucocorticoid agonist and antagonist modulate extinction of conditioned fear. Neuropsychopharmacology 31: 912–924. [DOI] [PubMed] [Google Scholar]
  315. Yang YL, Chao PK, Ro LS, Wo YY, Lu KT (2007). Glutamate NMDA receptors within the amygdala participate in the modulatory effect of glucocorticoids on extinction of conditioned fear in rats. Neuropsychopharmacology 32: 1042–1051. [DOI] [PubMed] [Google Scholar]
  316. Yehuda R, Bierer LM, Pratchett L, Malowney M (2010). Glucocorticoid augmentation of prolonged exposure therapy: rationale and case report. Eur J Psychotraumatol 1: 5643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  317. Yehuda R, LeDoux J (2007). Response variation following trauma: a translational neuroscience approach to understanding PTSD. Neuron 56: 19–32. [DOI] [PubMed] [Google Scholar]
  318. Young S, Bohenek D, Fanselow M (1994). NMDA processes mediate anterograde amnesia of contextual fear conditioning induced by hippocampal damage: immunization against amnesia by context preexposure. Behav Neurosci 108: 19–29. [DOI] [PubMed] [Google Scholar]
  319. Zannas AS, Binder EB (2014). Gene-environment interactions at the FKBP5 locus: sensitive periods, mechanisms and pleiotropism. Genes Brain Behav 13: 25–37. [DOI] [PubMed] [Google Scholar]
  320. Zhang W, Rosenkranz JA (2013). Repeated restraint stress enhances cue-elicited conditioned freezing and impairs acquisition of extinction in an age-dependent manner. Behav Brain Res 248: 12–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  321. Zheng X, Deschaux O, Lavigne J, Nachon O, Cleren C, Moreau JL et al (2013). Prefrontal high-frequency stimulation prevents sub-conditioning procedure-provoked, but not acute stress-provoked, reemergence of extinguished fear. Neurobiol Learn Mem 101: 33–38. [DOI] [PubMed] [Google Scholar]
  322. Zimmerman JM, Maren S (2010). NMDA receptor antagonism in the basolateral but not central amygdala blocks the extinction of Pavlovian fear conditioning in rats. Eur J Neurosci 31: 1664–1670. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Neuropsychopharmacology are provided here courtesy of Nature Publishing Group

RESOURCES