G Model BBR-9047; No. of Pages 10

ARTICLE IN PRESS Behavioural Brain Research xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Behavioural Brain Research journal homepage: www.elsevier.com/locate/bbr

Research report

Serotonin in fear conditioning processes Elizabeth P. Bauer ∗ Barnard College, Biology Department, 3009 Broadway, New York, NY 10027, United States

h i g h l i g h t s • • • •

5-HT2A agonists are anxiogenic and enhance fear learning. 5-HT1A agonists are anxiolytic and impair cued and contextual fear learning. Variation in 5-HT genes can influence fear conditioning and extinction. Roles of the amygdala, hippocampus and BNST in mediating acute SSRI effects.

a r t i c l e

i n f o

Article history: Received 9 May 2014 Received in revised form 18 July 2014 Accepted 21 July 2014 Available online xxx Keywords: Serotonin Fear conditioning Amygdala

a b s t r a c t This review describes the latest developments in our understanding of how the serotonergic system modulates Pavlovian fear conditioning, fear expression and fear extinction. These different phases of classical fear conditioning involve coordinated interactions between the extended amygdala, hippocampus and prefrontal cortices. Here, I first define the different stages of learning involved in cued and context fear conditioning and describe the neural circuits underlying these processes. The serotonergic system can be manipulated by administering serotonin receptor agonists and antagonists, as well as selective serotonin reuptake inhibitors (SSRIs), and these can have significant effects on emotional learning and memory. Moreover, variations in serotonergic genes can influence fear conditioning and extinction processes, and can underlie differential responses to pharmacological manipulations. This research has considerable translational significance as imbalances in the serotonergic system have been linked to anxiety and depression, while abnormalities in the mechanisms of conditioned fear contribute to anxiety disorders. © 2014 Published by Elsevier B.V.

1. Introduction One hallmark of several anxiety disorders is an abnormality in acquiring or extinguishing conditioned fear memories [1,2]. Manipulations of the serotonin (5-HT) system are widely used to treat a variety of anxiety disorders such as panic disorder, social phobia, generalized anxiety disorder and obsessive-compulsive disorder [3–7]. Thus, an understanding of how the serotonergic system modulates fear learning processes has been of considerable interest for decades. The advantage of studying classical Pavlovian fear conditioning is that it is a model of emotional learning for which the underlying neural circuitry has been described in detail. The structures involved in fear learning and expression, including the amygdala, hippocampus and prefrontal cortices, contain dense concentrations of 5-HT receptors [8–10]. Moreover, 5-HT levels in the amygdala increase during both cued and context fear conditioning

∗ Corresponding author. Tel.: +1 212 854 2349; fax: +1 212 854 1950. E-mail address: [email protected]

[11–13]. Here, recent advances in our understanding of the neural circuits involved in fear learning, expression and extinction are described. I then review the literature describing how manipulations of the serotonergic system affect each of these behavioral processes, attempt to reconcile seemingly contradictory findings and offer recommendations for future research. 2. The neural circuitry underlying classical fear conditioning Pavlovian fear conditioning is one of the most comprehensively studied behavioral paradigms. In classical fear conditioning, an initially neutral stimulus, such as a tone or light (conditioned stimulus; CS) is paired with an aversive stimulus, such as a brief electrical footshock (unconditioned stimulus; US). As a result of this pairing, the CS acquires aversive properties. Afterwards, when presented alone, the CS elicits responses in the animal that are characteristic of fear, including autonomic changes such as increased heart rate and blood pressure, as well as behavioral reactions such as the cessation of movement (freezing) and/or fear-potentiated startle

http://dx.doi.org/10.1016/j.bbr.2014.07.028 0166-4328/© 2014 Published by Elsevier B.V.

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10 2

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

[14,15]. While cued fear conditioning involves the association of the US with a discrete cued CS, animals are also capable of associating a collection of context cues with the US, producing context fear conditioning [16–18]. If an animal is exposed to repeated presentations of the CS in the absence of the US, a gradual reduction in the ability of the CS to elicit fear responses takes place [19]. This phenomenon, fear extinction, involves learning that the new CS no longer predicts the US. Rather than destroying the original memory, fear extinction is a new learning process that inhibits the original fear memory [20,21]. The gradual reduction in fear during CS presentations is referred to as within-session extinction. This new memory is consolidated and recalled at a later point in time, which is referred to as between-session extinction or extinction retention/recall. To appreciate how the serotonergic system modulates fear learning it is important to keep in mind that fear conditioning involves different stages of memory. Initial learning of the CSUS association during training is referred to as fear acquisition. These memories are thought to consolidate into stable, enduring protein-synthesis dependent memories during the 24 h after fear acquisition [22,23]. When the CS is later presented alone, the fear memory is retrieved, leading to expression of the conditioned fear response. Thus if a manipulation (such as the administration of a drug) occurs immediately before training and impairs both shortterm (STM) and long-term memory (LTM), it is said to disrupt the acquisition of a fear memory. However, a drug administered pretraining which does not affect STM but does impair LTM prevents memory consolidation. Similarly, a drug is given immediately after training which impairs the formation of LTM, also disrupts memory consolidation. Alternatively, if learning has already taken place and a drug is given immediately prior to testing, it can affect expression or recall of the fear memory. In this way, the timing of drug administration determines the stage of learning that is affected. It is important to keep in mind when studying fear conditioning and fear processes that the terms “fear” and “anxiety” can refer to conscious feelings and also to behavioral and physiological responses (see [24–26] for discussions of this idea). All animals, including humans, can detect threats and respond defensively. Studies that measure freezing behavior, or avoidance, or blood pressure are measuring behavioral or autonomic reactions to perceived threats. The conscious perception of fear and anxiety, however, can only be experienced by organisms capable of such conscious experience. In humans, emotions, including panic, can be consciously perceived even when the amygdala or insula is damaged [27,28], even though animals or humans with amygdala damage are not capable of Pavlovian fear conditioning [29,30]. Nevertheless, because the terms “fear conditioning” and “fear expression” are so widely used, this review will continue to use these terms. The robustness of the Pavlovian fear conditioning paradigm has led to a detailed understanding of the key neuronal circuits, neurochemicals and molecular mechanisms underlying fear learning, fear extinction and associative plasticity in general (for recent reviews, see [31–34]). Lesion, inactivation and unit recording studies originally identified the lateral nucleus of the amygdala (LA) as the main input station of the amygdala where CS and US sensory inputs from both cortical and subcortical areas synapse on individual LA neurons [29,35–39]. Activation of individual LA neurons by CS inputs is enhanced by US-mediated depolarization during fear learning [38,40]. As a result of contingent CS-US pairings, subsequent presentations of the CS alone evoke larger responses in LA neurons [37,38,41]. Several recent reviews explore the complex molecular mechanisms underlying the formation of the CS-US association [31,33]. In the original model of fear conditioning, the central nucleus (CE) functioned as the main output of the amygdala. Neurons in

the medial division (CeM) project to several fear effector systems in the hypothalamus and brain stem including the periaqueductal gray (PAG) which controls freezing behavior [15,42,43]. Disinhibition of the CeM allows for the expression of a range of defensive behaviors including freezing [44]. Newer research suggests that neurons in the lateral division of the Ce (CeL) contribute to fear acquisition. Pretraining inactivation of the entire CE, or CeL specifically, interferes with the acquisition of auditory fear conditioning [44,45]. Blockade of NMDA receptors or disruption of protein synthesis within the CE also impairs fear learning [45,46]. Neurons within the CeL receive input from the BLA and project to the CeM [44,47]. Plasticity of these glutamatergic inputs to CeL is induced by fear conditioning such that activation of these CeL neurons is necessary for fear memory recall [48]. Therefore, during CS presentations, subsets of LA and CeL neurons become active, disinhibiting CeM output neurons [44,48]. The bed nucleus of the stria terminalis (BNST), considered part of the extended amygdala, receives a strong glutamatergic input from the basolateral nucleus of the amygdala (BLA; which is composed of the LA and basal nuclei) and is reciprocally connected with the CE [49,50]. The CE and the BNST share a similar pattern of efferent targets including brainstem areas involved in fear and anxiety [49,51]. Unlike the BLA and CE, lesions of the BNST do not disrupt auditory fear conditioning or expression [15]. However, lesions or reversible inactivation of the BNST do attenuate the expression of context fear conditioning [52,53]. The BNST also contributes to an animal’s response to unpredictable stressful events and anxiety [54–57]. These data have led to the hypothesis that there is a distinction between fear (imminent threat or phasic fear) and anxiety (potential threat or sustained fear), which are mediated by the amygdala and BNST, respectively [56,58]. Animals are not only capable of associating discrete cues with an aversive stimulus, but can also associate a collection of contextual cues with a shock. A role for the hippocampus in this process is implicated by studies demonstrating that electrolytic lesions of the dorsal hippocampus before or after training impair context conditioning [16,17,59,60]. More recent experiments using optogenetic techniques reveal that during fear conditioning, subsets of granule cells in the dentate gyrus encode the memory of the context [61,62]. This memory of the context can be recalled when those specific subpopulations of neurons are activated [62,63]. Lesions or functional inactivation of the ventral hippocampus, however, produce inconsistent effects on context fear. Some studies show that the ventral hippocampus is required for the formation of a context representation [64–67], while others fail to replicate these findings [68]. Rather, the dentate gyrus of the ventral hippocampus seems to mediate anxiety [61]. Importantly, only the ventral hippocampus projects directly to the amygdala, thus damage to the ventral portion interferes with the transfer of information from the dorsal hippocampus to the amygdala [69,70]. Moreover, the ventral hippocampus modulates activity of BLA neurons [71]. The ventral hippocampus also sends a strong projection to the prelimbic cortex, which in turn forms a reciprocal connection with the BLA [72,73]. Finally, the BNST also receives input from both the ventral hippocampus and ventral subiculum and BLA [49]. However, lesions of the BNST selectively disrupt the expression of context but not cued fear conditioning [53,74]. After fear learning has taken place, it can be expressed via disinhibition of CeM neurons [44]. The increased responsiveness of CeM output neurons to the CS depends on disinhibition of CeL neurons as well as excitation from glutamatergic neurons in the BLA [44,75,76]. Indeed, the activity of BLA neurons correlates with high or low fear behavior (freezing) [71]. The prelimbic cortex (PL) integrates information from the BLA and ventral hippocampus, with increased activity in this area of the prefrontal cortex also correlating with increased fear expression [77,78].

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

Several lines of evidence implicate the amygdala in extinction learning. Reversible inactivation of the BLA with the GABA agonist muscimol or intra-amygdala infusions of NMDA receptor antagonists block the acquisition of extinction [71,79–82]. At the same time, CeM fear output neurons are inhibited during fear extinction through potentiation of BLA neurons to intercalated cell masses, a process that is dependent on the infralimbic prefrontal cortex (IL) [83]. The IL itself is a critical site of plasticity after extinction is learned and is involved in the retrieval of extinction memories [19]. Indeed, pairing of IL stimulation with a CS is sufficient to simulate extinction memory [84]. Recent data suggests that there exist distinct populations of neurons within the BLA which can be characterized by whether they project to IL or PL. Activation of PL-projecting BLA neurons produces fear-like behavior, while ILprojecting BLA neurons are recruited and exhibit plasticity during fear extinction learning [85].

3. Serotonergic modulation of fear learning and expression Fear learning and extinction are thus mediated by interactions between several nuclei of the amygdala, the hippocampus, the mPFC and the BNST. All of these structures contain dense concentrations of 5-HT receptors [8–10]. Moreover, there is an increase in the concentration of amygdalar 5-HT both during and after behavioral arousal and stress and in response to US presentations during fear conditioning [12,13,86–88]. After context fear conditioning, there are increases in extracellular 5-HT in both the BLA and mPFC [11,12,89]. However, 5-HT modulates activity in these structures in a complex manner. In general, 5-HT can have differential effects on behavior depending on the receptor subtypes present and the behavioral tests used. The 5-HT system is quite complex with multiple 5-HT receptor subtypes which have been classified into seven distinct families (5-HT1–7 ). The 5-HT3 receptor is the only ionotropic 5HT receptor and is permeable to cations [90]. 5-HT2 receptors, G-protein coupled receptors, are positively coupled to phospholipase C; their activation causes an increase in intracellular calcium [90]. Activation of 5-HT1A leads to hyperpolarization via opening of potassium channels [90]. Unlike the 5-HT2 receptors, the 5-HT1A receptor is expressed as both an autoreceptor in raphe nuclei and a heteroreceptor in forebrain regions. As an autoreceptor, it limits the release of 5-HT from raphe neurons, thus affecting overall serotonergic tone [91]. As a heteroreceptor in the amygdala, BNST and hippocampus, it mediates local responses to 5-HT. 5-HT modulation of neural activity is further complicated by the fact that one receptor subtype can exist on both glutamatergic and GABAergic cell types [92] and more than one 5-HT receptor subtype can be present on a neuron [93–95]. Moreover, 5-HT can have differential effects on behavior depending on the brain region that is targeted [96] or the behavioral paradigm used [97–99]. Thus 5-HT modulates the circuits and neurons mediating fear learning in a highly complex manner. 5-HT cell bodies are localized in the mid-brain raphe nuclei, comprised of the dorsal and median raphe nuclei, which send projections throughout the brain. The amygdala receives 5-HT innervation from the dorsal raphe nucleus [100], with the BLA receiving dense projections and the CE receiving weak input. In the BLA, 5-HT produces a primarily inhibitory response by depolarizing GABAergic interneurons [10]. This increase in GABAergic transmission results in feed-forward inhibition of the principal excitatory neurons of the BLA [10,101]. 5-HT2 receptors are found predominantly in the BLA on both the principal excitatory neurons as well as parvalbumin-positive GABAergic interneurons [92]. 5-HT1A receptors are concentrated in the CE, while 5-HT3 receptors are found in both the BLA and CE and are nearly exclusively

3

localized to GABAergic neurons [9,10,102]. The BNST also receives serotonergic input from the dorsal raphe nucleus [103]. Although the response of BNST neurons to 5-HT depends on which specific receptors are activated, the primary response is inhibitory [95]. The prefrontal cortex receives input from both the dorsal and median raphe nuclei [104]. Within the prefrontal cortex the primary response to 5-HT is inhibitory through increased excitability of interneurons and decreased excitability of pyramidal neurons [105]. Similarly, 5-HT in the hippocampus, which receives innervation from the median raphe nucleus [106], depolarizes inhibitory interneurons and reduces pyramidal neuron excitability [107,108]. As a result of the complex nature of the serotonergic system in these structures, an appreciation for how 5-HT affects fear conditioning and fear extinction can be gained by understanding the contribution of specific 5-HT receptor subtypes to these learning processes. Most of these experiments have been done on rodents. 5-HT2A receptors, which are highly enriched in the amygdala, prefrontal cortex and hippocampus [8] modulate learning and memory in a variety of paradigms. Systemic injections of 5-HT2A agonists enhance conditioned avoidance memory [109], enhance eyeblink conditioning [110], and strengthen the consolidation of object memory [111] in animals. Conversely, 5-HT2A antagonists impair conditioned avoidance memory and eyeblink conditioning [110,112]. In humans, reduced binding capacity of 5-HT2A receptors has been observed in normal aging subjects as well as Alzheimer’s patients [113,114]. Moreover, humans who carry a polymorphism in the 5-HT2A receptor gene show impaired consolidation of explicit memory [115]. If the 5-HT2A agonist TCB-2 is administered immediately following fear learning, it enhances both cued and context fear conditioning in mice, suggesting an enhancement of the consolidation of fear memories by this receptor [111]. Similarly, a 5-HT2A agonist injected prior to extinction learning also facilitates extinction learning in both delay and trace fear conditioned mice [111]. Because this study used systemic injections of 5-HT2A agonists, it is unclear which structures mediate the effects of these drugs on fear conditioning and extinction. Considering the dense concentrations of 5-HT2A receptors in the amygdala, hippocampus and cortex, it is possible that agonists of this receptor facilitate their coordinated activity. In the cortex, 5-HT2A receptors influence low-frequency oscillations in the frontal cortex [116] and increase spinogenesis [117]. As described above, the 5-HT2A receptor is positively coupled to phospholipase C and its activation leads to an increase in internal calcium [90]. These receptors are present in the dendrites and dendritic spines of neurons [118] where they directly interact with the PSD-95 and regulate receptor trafficking and signal transduction [119]. Additionally, activation of 5-HT2A receptors enhances cortical presynaptic glutamate release [120,121] and NMDA receptor sensitivity [122]. 5-HT2A receptors facilitate NMDA receptor activity in the cortex [122] and BLA [123]. It should be noted, however, that 5-HT2A knockout mice do not show changes in fear conditioning, although they do exhibit reduced anxiety [124]. Both the 5-HT2A and 5-HT2C receptors in the amygdala are involved in the expression of anxiety-like behaviors. Fear can be characterized by arousal and fight-or-flight responses to imminent threats, while anxiety is characterized more by apprehension, tension, and worry, where the potential threat might occur in the distant future [125]. Infusions of 5-HT2A/C agonists into the BLA, as well as over-expression of 5-HT2C receptors in the BLA produce anxiety-like behaviors in animals as measured by the elevated plus maze and open field test [126–128]. Conversely, both 5-HT2A and 5-HT2C knockout mice show decreases in anxiety-like behavior [124,129]. There have been few direct studies on the contribution of 5-HT2C receptors to fear learning or extinction. However, one study examined the effects of the 5-HT2C receptor antagonist ritanserin on fear conditioning in humans [130]. Skin conductance responses

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10 4

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

to a CS tone were enhanced after the tone was paired with an aversive stimulus, but this effect was blocked by ritanserin, suggesting that activation of 5-HT2C receptors contributes to fear conditioning. Moreover, as described below, co-administration of a 5-HT2C antagonist, but not a 5-HT3 receptor antagonist, blocks the acute effects of SSRI treatment on conditioned fear expression [131]. Together, these data suggest that activation of 5-HT2A/C receptors both increases anxiety and enhances fear learning and extinction. In contrast, 5-HT1A receptor agonists, such as the widely used 8OH-DPAT, produce anxiolytic effects in both humans and animals [132,133], and 5-HT1A knockout mice show increases in anxietylike and depressive-like behavior [134–136]. Several studies have revealed that 5-HT1A receptor agonists decrease context fear when given systemically [137,138]. However, as discussed above, the 5HT1A receptor is expressed as both an autoreceptor in raphe nuclei and a heteroreceptor in forebrain regions. To disambiguate the contribution of these two receptor subtypes to fear conditioning, several studies in animals have used local infusions into the raphe nuclei, the amygdala or the hippocampus. When the 5-HT1A agonist 8-OH-DPAT is infused directly into the median raphe nucleus to stimulate 5-HT1A autoreceptors, it impairs both the acquisition and expression of context fear memories [139]. Testing the contribution of post-synaptic 5-HT1A receptors, pretraining intra-hippocampal infusions of 8-OH-DPAT impair context fear conditioning [140]. Pre-testing infusions of 8-OH-DPAT impair both the expression of context conditioned freezing as well as fear-potentiated startle [141]. In contrast, pre-testing infusions of this agonist into the median raphe nucleus decreases context fear without affecting fear-potentiated startle, which suggests that postsynaptic 5-HT1A receptors play a greater role in the modulation of context fear memories [141]. Similarly, when 5-HT1A agonists are infused into the BLA or the BNST they impair the expression of context fear conditioning [142,143]. Using a conditioned defeat paradigm, it was found that activation of 5-HT1A receptors in the BLA decreases both acquisition and expression [144]. In sum, 5-HT1A agonists, which have been shown to be anxiolytic, decrease both acquisition and expression of fear conditioning. These inhibitory effects may be explained by the fact that activation of 5-HT1A receptors inhibits neuronal activity in both the hippocampus and the BNST [145,146]. A recent study examined the role of 5-HT1A receptors in the BLA, PL and dorsal PAG in low or high-anxious animals on unconditioned and conditioned fear [147]. The examination of individual differences in anxiety and fear conditioning is useful in developing models of psychopathology that affect a relatively small proportion of the population [148,149]. In low-anxious rats, activation of 5-HT1A receptors in either the BLA or dorsal PAG reduced fear potentiated startle. In contrast, in high-anxious rats, 5-HT1A receptor activation in the prelimbic area reduced fear potentiated startle. This type of behavioral study highlights the importance of examining the serotonergic system in a heterogeneous population. The 5-HT3 receptor is the only ionotropic 5-HT receptor and is permeable to cations. Mice in which this receptor is overexpressed demonstrate enhanced context but not cued conditioning, increased exploratory behavior but decreased anxiety-like behavior as measured in the elevated plus maze [150]. Conversely, 5-HT3 antagonists impair the expression of context fear conditioning [89,151]. 5-HT3 antagonists do not block the effects of SSRIs on the expression of cued fear conditioning [131]. Together, these data suggest that 5-HT3 receptors modulate context but not cued fear conditioning. The contribution of 5-HT3 receptors to extinction learning is not clear. 5-HT3 receptor knockout mice show reduced fear extinction learning to both cues and context [152]. However, systemic injections of 5-HT3 receptor antagonists prior to extinction training enhance cued and context fear extinction [153]. This discrepancy is difficult to resolve given the different behavioral

paradigms, species and methods of disrupting 5-HT3 receptor function.

4. Serotonin transporter gene variation Variations in serotonergic genes can influence fear conditioning and extinction processes. The most widely studied is a functional polymorphism in the promotor region of the 5-HT transporter gene (5-HTTPR). The 5-HT transporter 5-HTT regulates 5-HT signaling via reuptake of 5-HT from the extracellular space [154]. In humans, the short (s) allele of the 5-HTT promotor region, which is associated with decreased transcription of the gene compared to the long (l) allele, leads to reduced 5-HTT function and presumably increased concentrations of extracellular 5-HT [155]. At the behavioral level, the s allele is associated with abnormal levels of anxiety [155,156]. Humans with one or two copies of the s allele exhibit greater amygdala activity in response to fearful stimuli [157]. Carriers of the s-allele possess smaller hippocampal and amygdala volumes, as well as a functional uncoupling of the amygdala-cingulate circuit [158,159]. Human carriers of the s-allele show increased fear conditioning as well as stronger fear potentiated startle [160–162]. They also produce enhanced autonomic responses when watching another subject undergo fear conditioning, a process referred to as observational fear learning [163]. Conversely, mice in which 5-HTT is overexpressed throughout the brain show impaired fear learning and reduced theta oscillations in the amygdala [164]. However, 5-HTT knockout mice show intact fear conditioning and initial extinction learning but impaired extinction recall [165,166]. Thus, although variations in 5-HTT affect fear learning and extinction, there are several inconsistent findings. This might be due to different methodologies and subjects, but might also be due to variable correlation of the 5-HTTPR with other genetic alterations. For example, a recent study in humans examined the interaction between two polymorphisms: the 5-HTTPR and variation in the gene encoding the neuropeptide S receptor (NPSR1). Only carriers of both risk alleles exhibited higher fear potentiated startle [167]. Similarly, only human subjects carrying the risk allele of both the 5HTTLPR and that of a polymorphism of the corticotropin releasing hormone receptor 1 (CRHR1–rs878886) show increased fear potentiation of the eyeblink startle reflex [168]. There is a considerable amount of data showing that environment can influence anxiety-like behaviors and anxiety disorders [169–171]. Interestingly, variations in the 5HTTLPR can make individuals more or less susceptible to the effects of environment. For example, humans with one or two copies of the s-allele have an increased risk of developing depression if they experience at least three “stressful” life events during childhood [172]. After repeated social defeat, 5-HTT knockout mice showed higher levels of anxiety-like behavior than normal mice [173]. In addition, homozygous (5-HTT−/− ) and heterozygous (5-HTT+/− ) 5-HTT knockout mice show significantly higher freezing behavior during extinction and extinction recall compared to wild-type mice, and this difference is enhanced after social defeat [165]. Furthermore, socially defeated 5-HTT−/− mice show increased theta correlations between LA and mPFC during extinction learning and recall [165]. A second functional gene variation affecting 5-HTT is the STPP/rs3813034 serotonin transporter polyadenylation polymorphism. Polyadenylation is a posttranscriptional modification; the polymorphism influences the balance of two polyadenylation forms of 5-HTT [174]. The G-allele carriers have reduced 5-HTT levels, which presumably leads to an increase in extracellular 5-HT, and they exhibit increased risk for panic disorder [174]. Carriers of the risk allele also show impaired retention of fear extinction memory, heightened anxiety and depressive symptoms [175]. These

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

data suggest that 5-HTT down-regulation can increase fear learning and impair extinction retention, resulting in persistent fear memories. Future studies should genotype for both 5-HTTPR and STPP in order to clarify their relative phenotypic effects.

5. Acute SSRI administration and fear conditioning SSRIs, such as citalopram, escitalopram, fluoxetine, fluvoxamine and paroxetine are widely used for the treatment of anxiety disorders including panic disorder, generalized anxiety disorder, phobias and obsessive-compulsive disorder [4,6,176,177]. SSRIs act pharmacologically by inhibiting the serotonin transporter, leading to an increase in 5-HT availability. These increased levels of 5-HT produce an increase in anxiety that can be measured in animals using a variety of behavioral tests including the elevated plus maze [178–181], novelty suppressed feeding [182], social interaction [183,184] and the free-exploration test [185]. These behavioral results are consistent with clinical data suggesting that patients experience an initial increase in anxiety and an increased risk of suicide ideation when given SSRIs; these anxiogenic effects eventually subside after chronic treatment [186–192]. Several hypotheses have been suggested for how SSRIs mediate their long-term effects of anxiety including the disinhibition of autoreceptors in the raphe nuclei [193,194], and adaptive changes in glutamatergic neurotransmission [195–197]. Yet despite extensive research, it is still unclear how SSRIs mediate their long-term effects on anxiety. Acute SSRI treatment has been shown to increase extracellular 5-HT within the amygdala, hippocampus and frontal cortex [198–200]. We can therefore gain insight into the role of 5-HT in fear conditioning processes by examining how SSRIs affect fear learning, expression and extinction. Systemic injections of citalopram or fluoxetine prior to training enhance the acquisition of auditory fear conditioning in animals [201–203]. Acute systemic injections of these drugs prior to testing enhance the expression of auditory fear conditioning, an effect which is blocked by coadministration of 5-HT2C antagonists, but not by co-administration of 5-HT3 antagonists [131]. In humans, acute SSRI treatment enhances fear potentiated startle to a discrete cue [204]. Together, these results indicate that overall, acute SSRI treatment increases both fear acquisition and expression when animals or humans are fear conditioned to a discrete cue. Acute SSRI injections increase immediate-early gene expression within the amygdala and BNST [203,205–208]. However, when SSRIs are infused directly into the amygdala, they do not affect fear learning [203]. Rather, direct infusions of SSRIs into the BNST prior to training both enhance immediate early-gene expression in the amygdala and enhance fear conditioning [203]. Although the BNST is not normally required for the acquisition of cued fear memory [53,58], these data suggest that SSRI administration recruits the BNST into the fear circuit during conditioning. The neuronal interactions between the BNST and amygdala during SSRI administration remain to be explored. The effects of SSRIs on the acquisition of context fear conditioning are contradictory, with some finding that SSRI treatment decreases the acquisition of context fear [209,210] while others report an increase in acquisition [202,211]. However, when the SSRI is administered immediately before testing, there is general agreement that it results in a decrease in fear expression [137,202,209,211–214]. SSRIs decrease neuronal activity, immediate-early gene expression and plasticity in the hippocampus [203,215,216]. They may therefore be interfering with the recall of context information [65,217], and thus impairing the expression of context fear conditioning. The contradictory effects of SSRI administration on acquisition of context fear conditioning might be explained by different strategies animals use to form

5

representations of a context. Animals are capable of using the discrete cues that make up a context to successfully fear condition when hippocampal function is impaired [17,217]. If SSRI administration decreases hippocampal activity, animals might use this strategy to successfully fear condition. If this solution is not efficient due to a decrease in salience of the contextual cues used [218], then acute SSRI administration prior to training would impair the animal’s ability to form a representation of the context and would thus impair context fear conditioning. Chronic administration of SSRIs impairs fear extinction learning, while also down-regulating the NR2B subunit of the NMDA receptors in the BLA [219,220]. However no study has observed an effect of acute SSRIs on fear extinction recall, despite an increase in fear expression and within-session fear extinction [131,220]. 6. Serotonin depletion Lastly, reducing central 5-HT levels and examining the effects on fear conditioning can also provide insights into the normal role of 5-HT in this behavioral paradigm. 5-HT can be reduced throughout the brain in several ways. One method is acute tryptophan depletion, which deprives the subject of the precursor of 5-HT [221] and decreases the release of 5-HT [222,223]. Another is to interfere with the synthesis of 5-HT by systemic or local injection of p-chlorophenylalanine (PCPA) which inhibits tryptophan hydroxylase, the enzyme which synthesizes 5-HT from tryptophan [224]. A third method is to destroy serotonergic neurons via a neurotoxin (5,7-DHT) which is taken up into cells by 5-HTT and destroys them [225]. Tryptophan depletion in healthy humans does not affect fearpotentiated startle to short-duration cues, but does decrease fear-potentiated startle to long-duration cues [226]. Previous work has shown that fear potentiated startle to short-duration cues is mediated by the amygdala and can be interpreted as a phasic fear response to an explicit threat, whereas startle to long duration cues is mediated by the BNST and is a more sustained anxietylike response [56,125]. In line with this data, research on rodents found that tryptophan depletion impaired the formation of context fear memory but did not affect cued fear memory [227]. Confusingly, PCPA administration in rodents selectively enhances fear-potentiated startle to short duration cues [228], but does not affect the expression of context fear conditioning [229]. And others have found that depletion of tryptophan in humans leads to attenuated autonomic responses to the CS during fear conditioning [230]. Both tryptophan depletion and PCPA administration reduce 5-HT levels to a modest extent, and they do so globally. Unfortunately, this prevents researchers from drawing conclusions about region-specific effects. Moreover, tryptophan depletion can have differential effects on behavior that can be explained by genetic variation. In a recent study in humans, tryptophan depletion attenuated motivation, as measured by a cued-reinforcement reaction time task. This effect was only seen in subjects with two copies of the s-allele of 5-HTTPR, but not in subjects with the l-allelic variation [231]. 7. Conclusions A considerable body of literature has examined how modulation of the serotonergic system affects Pavlovian fear conditioning and extinction. Strategies for manipulating this system include the administration of agonists or antagonists directed against one of the 17 different 5-HT receptor subtypes, administration of SSRIs and serotonin depletion. Moreover, there is considerable genetic variation among individuals resulting in altered functionality of the 5-HT transporter. Given the complex, and occasionally contradictory

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10 6

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

results reviewed here, it is still possible to draw some general conclusions about the role of 5-HT in fear conditioning processes. Pharmacological manipulations which are anxiogenic generally facilitate cued and context fear conditioning. For example, systemic injections of 5-HT2A receptor agonists, which are anxiogenic when infused directly into the amygdala [126,127] facilitate cued and context fear conditioning as well as fear extinction [111]. Acute administration of SSRIs enhances both acquisition and expression of cued fear conditioning, although it impairs expression of context fear conditioning [131,202,203,212]. And carriers of the s-allele of the 5-HTTPR show both increased anxiety and enhanced fear conditioning [155,156,160–162]. Conversely, administration of 5-HT1A agonists, which are anxiolytic, impair the acquisition and expression of context fear conditioning [132,133,140,141]. Given our knowledge of the neuronal circuitry underlying cued and context fear conditioning, studies which use local infusions rather than systemic manipulations are particularly illuminating. For example, 5-HT1A receptor agonists infused into the hippocampus, BLA or BNST impair the expression of context fear conditioning [140,142,143]. Infusion of an SSRI into the BNST prior to fear conditioning is sufficient to increase fear acquisition and enhance expression of the immediate-early gene Arc in the CE, mirroring the results of acute systemic SSRI administration. While the neuronal interactions between the BNST and amygdala during SSRI administration remain to be explored, these types of experiments underscore the utility of using local drug infusions to identify relevant structures. A second general observation can be made based on the experiments reviewed here. There exist several functional polymorphisms which affect the 5-HT transporter and thus the concentration of 5-HT in the extracellular space. There might be considerable interaction between these genetic alterations, whereby only carriers of several risk alleles exhibit increased fear or anxiety [167,168]. Furthermore, a recent study suggests that human carriers of the s-allele of 5-HTTPR show downregulation of 5-HT1A receptors [232]. This raises the intriguing question of whether pharmacological manipulations targeting the 5-HT1A receptor, for example, show differential efficacy in these individuals. As these functional polymorphisms are discovered and characterized, researchers should perform more extensive genotyping in order to clarify relative phenotypic effects. Moreover, recent experiments suggest that variations in the 5HTTLPR can make both humans and animals more or less susceptible to the effects of environment [165,172]. Finally, in the past few years, new technologies have greatly expanded our ability to study the neural circuitry underlying behavior. Using optogenetic tools, for example, it has been recently discovered that two populations of neurons within the BA which can be characterized by whether they project to PL or IL. BA-PL neurons are activated by fear conditioning; optogenetically inhibiting these neurons promotes extinction. BA-IL neurons, in contrast, are activated by fear extinction and silencing them impairs extinction [85]. Within the CEL , experiments using optogenetic tools have revealed that there are distinct subpopulations which bidirectionally regulate conditioned fear [44,47]. Meanwhile, the effects of optogenetic manipulations can depend on the cell type targeted: stimulating the subpopulation of BLA neurons expressing Thy1 impairs fear learning but strengthens fear extinction [233]. Vectors can selectively target serotonergic neurons [234,235], as well as excitatory pyramidal neurons and astroglia [236]. These techniques have the potential to provide enormous insight into how the serotonergic system modulates the fear memory system. While the relationship between the serotonergic system and fear conditioning processes is not straightforward, there is continued interest in this topic. Modulations of the serotonergic system

have been linked to anxiety and depression, and current anxiolytic and antidepressant treatments target this system. Meanwhile, abnormalities in the mechanisms of conditioned fear contribute to anxiety disorders. There is an extensive body of knowledge concerning the neural circuitry and molecular mechanisms underlying Pavlovian fear conditioning. We should take advantage of this knowledge to ask how 5-HT is acting in specific fear structures and at the cellular and molecular levels at different stages of fear learning and memory formation. References [1] Grillon C. Startle reactivity and anxiety disorders: aversive conditioning, context, and neurobiology. Biol Psychiatry 2002;52(10):958–75. [2] Milad MR, et al. Presence and acquired origin of reduced recall for fear extinction in PTSD: results of a twin study. J Psychiatr Res 2008;42(7):515–20. [3] Bezchlibnyk-Butler K, Aleksic I, Kennedy SH. Citalopram—a review of pharmacological and clinical effects. J Psychiatry Neurosci 2000;25(3):241–54. [4] Gorman JM. Treating generalized anxiety disorder. J Clin Psychiatry 2003;64(Suppl. 2):24–9. [5] Kent JM, Coplan JD, Gorman JM. Clinical utility of the selective serotonin reuptake inhibitors in the spectrum of anxiety. Biol Psychiatry 1998;44(9): 812–24. [6] Sheehan DV, et al. Serotonin in panic disorder and social phobia. Int Clin Psychopharmacol 1993;8(Suppl. 2):63–77. [7] Stein DJ, Stahl S. Serotonin and anxiety: current models. Int Clin Psychopharmacol 2000;15(Suppl. 2):S1–6. [8] Cornea-Hebert V, et al. Cellular and subcellular distribution of the serotonin 5-HT2A receptor in the central nervous system of adult rat. J Comp Neurol 1999;409(2):187–209. [9] Mascagni F, McDonald AJ. A novel subpopulation of 5-HT type 3A receptor subunit immunoreactive interneurons in the rat basolateral amygdala. Neuroscience 2007;144(3):1015–24. [10] Rainnie DG. Serotonergic modulation of neurotransmission in the rat basolateral amygdala. J Neurophysiol 1999;82(1):69–85. [11] Hashimoto S, Inoue T, Koyama T. Effects of conditioned fear stress on serotonin neurotransmission and freezing behavior in rats. Eur J Pharmacol 1999;378(1):23–30. [12] Kawahara H, et al. Psychological stress increases serotonin release in the rat amygdala and prefrontal cortex assessed by in vivo microdialysis. Neurosci Lett 1993;162(1-2):81–4. [13] Yokoyama M, et al. Amygdalic levels of dopamine and serotonin rise upon exposure to conditioned fear stress without elevation of glutamate. Neurosci Lett 2005;379(1):37–41. [14] Campeau S, Davis M. Involvement of subcortical and cortical afferents to the lateral nucleus of the amygdala in fear conditioning measured with fearpotentiated startle in rats trained concurrently with auditory and visual conditioned stimuli. J Neurosci 1995;15(3 Pt 2):2312–27. [15] LeDoux JE, et al. Different projections of the central amygdaloid nucleus mediate autonomic and behavioral correlates of conditioned fear. J Neurosci 1988;8(7):2517–29. [16] Kim JJ, Rison RA, Fanselow MS. Effects of amygdala: hippocampus, and periaqueductal gray lesions on short- and long-term contextual fear. Behav Neurosci 1993;107(6):1093–8. [17] Maren S, Fanselow MS. Electrolytic lesions of the fimbria/fornix: dorsal hippocampus, or entorhinal cortex produce anterograde deficits in contextual fear conditioning in rats. Neurobiol Learn Mem 1997;67(2):142–9. [18] Selden NR, Everitt BJ, Robbins TW. Telencephalic but not diencephalic noradrenaline depletion enhances behavioural but not endocrine measures of fear conditioning to contextual stimuli. Behav Brain Res 1991;43(2):139–54. [19] Quirk GJ, Mueller D. Neural mechanisms of extinction learning and retrieval. Neuropsychopharmacology 2008;33(1):56–72. [20] Bouton ME, et al. Contextual and temporal modulation of extinction: behavioral and biological mechanisms. Biol Psychiatry 2006;60(4):352–60. [21] Ji J, Maren S. Hippocampal involvement in contextual modulation of fear extinction. Hippocampus 2007;17(9):749–58. [22] Dudai Y. The neurobiology of consolidations, or, how stable is the engram? Annu Rev Psychol 2004;55:51–86. [23] McGaugh JL. Memory—a century of consolidation. Science 2000;287(5451): 248–51. [24] LeDoux JE. Evolution of human emotion: a view through fear. Prog Brain Res 2012;195:431–42. [25] LeDoux JE. The slippery slope of fear. Trends Cogn Sci 2013;17(4):155–6. [26] LeDoux JE. Coming to terms with fear. Proc Natl Acad Sci USA 2014; 111(8):2871–8. [27] Damasio A, Damasio H, Tranel D. Persistence of feelings and sentience after bilateral damage of the insula. Cereb Cortex 2013;23(4):833–46. [28] Feinstein JS, et al. Fear and panic in humans with bilateral amygdala damage. Nat Neurosci 2013;16(3):270–2. [29] Amorapanth P, LeDoux JE, Nader K. Different lateral amygdala outputs mediate reactions and actions elicited by a fear-arousing stimulus. Nat Neurosci 2000;3(1):74–9.

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

[30] Bechara A, et al. Double dissociation of conditioning and declarative knowledge relative to the amygdala and hippocampus in humans. Science 1995;269(5227):1115–8. [31] Johansen JP, et al. Molecular mechanisms of fear learning and memory. Cell 2011;147(3):509–24. [32] Orsini CA, Maren S. Neural and cellular mechanisms of fear and extinction memory formation. Neurosci Biobehav Rev 2012;36(7):1773–802. [33] Pape HC, Pare D. Plastic synaptic networks of the amygdala for the acquisition: expression, and extinction of conditioned fear. Physiol Rev 2010;90(2):419–63. [34] Pare D, Duvarci S. Amygdala microcircuits mediating fear expression and extinction. Curr Opin Neurobiol 2012;22(4):717–23. [35] LeDoux JE, et al. The lateral amygdaloid nucleus: sensory interface of the amygdala in fear conditioning. J Neurosci 1990;10(4):1062–9. [36] LeDoux JE, Ruggiero DA, Reis DJ. Projections to the subcortical forebrain from anatomically defined regions of the medial geniculate body in the rat. J Comp Neurol 1985;242(2):182–213. [37] Quirk GJ, Repa C, LeDoux JE. Fear conditioning enhances short-latency auditory responses of lateral amygdala neurons: parallel recordings in the freely behaving rat. Neuron 1995;15(5):1029–39. [38] Repa JC, et al. Two different lateral amygdala cell populations contribute to the initiation and storage of memory. Nat Neurosci 2001;4(7):724–31. [39] Romanski LM, et al. Somatosensory and auditory convergence in the lateral nucleus of the amygdala. Behav Neurosci 1993;107(3):444–50. [40] Blair HT, et al. Synaptic plasticity in the lateral amygdala: a cellular hypothesis of fear conditioning. Learn Mem 2001;8(5):229–42. [41] Collins DR, Pare D. Differential fear conditioning induces reciprocal changes in the sensory responses of lateral amygdala neurons to the CS(+) and CS(−). Learn Mem 2000;7(2):97–103. [42] Hopkins DA, Holstege G. Amygdaloid projections to the mesencephalon: pons and medulla oblongata in the cat. Exp Brain Res 1978;32(4):529–47. [43] Veening JG, Swanson LW, Sawchenko PE. The organization of projections from the central nucleus of the amygdala to brainstem sites involved in central autonomic regulation: a combined retrograde transportimmunohistochemical study. Brain Res 1984;303(2):337–57. [44] Ciocchi S, et al. Encoding of conditioned fear in central amygdala inhibitory circuits. Nature 2010;468(7321):277–82. [45] Wilensky AE, et al. Rethinking the fear circuit: the central nucleus of the amygdala is required for the acquisition, consolidation, and expression of Pavlovian fear conditioning. J Neurosci 2006;26(48):12387–96. [46] Goosens KA, Maren S. Pretraining NMDA receptor blockade in the basolateral complex: but not the central nucleus, of the amygdala prevents savings of conditional fear. Behav Neurosci 2003;117(4):738–50. [47] Haubensak W, et al. Genetic dissection of an amygdala microcircuit that gates conditioned fear. Nature 2010;468(7321):270–6. [48] Li H, et al. Experience-dependent modification of a central amygdala fear circuit. Nat Neurosci 2013;16(3):332–9. [49] Dong HW, Petrovich GD, Swanson LW. Topography of projections from amygdala to bed nuclei of the stria terminalis. Brain Res Brain Res Rev 2001;38(1-2):192–246. [50] Pare D, Smith Y, Pare JF. Intra-amygdaloid projections of the basolateral and basomedial nuclei in the cat: Phaseolus vulgaris-leucoagglutinin anterograde tracing at the light and electron microscopic level. Neuroscience 1995;69(2):567–83. [51] Petrovich GD, Swanson LW. Projections from the lateral part of the central amygdalar nucleus to the postulated fear conditioning circuit. Brain Res 1997;763(2):247–54. [52] Resstel LB, et al. Anxiolytic-like effects induced by acute reversible inactivation of the bed nucleus of stria terminalis. Neuroscience 2008;154(3):869–76. [53] Sullivan GM, et al. Lesions in the bed nucleus of the stria terminalis disrupt corticosterone and freezing responses elicited by a contextual but not by a specific cue-conditioned fear stimulus. Neuroscience 2004;128(1):7–14. [54] Alheid GF. Extended amygdala and basal forebrain. Ann N Y Acad Sci 2003;985:185–205. [55] Dunn JD, Williams TJ. Cardiovascular responses to electrical stimulation of the bed nucleus of the stria terminalis. J Comp Neurol 1995;352(2):227–34. [56] Walker DL, Davis M. Double dissociation between the involvement of the bed nucleus of the stria terminalis and the central nucleus of the amygdala in startle increases produced by conditioned versus unconditioned fear. J Neurosci 1997;17(23):9375–83. [57] Walker DL, Toufexis DJ, Davis M. Role of the bed nucleus of the stria terminalis versus the amygdala in fear, stress, and anxiety. Eur J Pharmacol 2003;463(13):199–216. [58] Walker DL, Davis M. Role of the extended amygdala in short-duration versus sustained fear: a tribute to Dr Lennart Heimer. Brain Struct Funct 2008;213(1–2):29–42. [59] McNish KA, Gewirtz JC, Davis M. Evidence of contextual fear after lesions of the hippocampus: a disruption of freezing but not fear-potentiated startle. J Neurosci 1997;17(23):9353–60. [60] Phillips RG, LeDoux JE. Differential contribution of amygdala and hippocampus to cued and contextual fear conditioning. Behav Neurosci 1992;106(2):274–85. [61] Kheirbek MA, et al. Differential control of learning and anxiety along the dorsoventral axis of the dentate gyrus. Neuron 2013;77(5):955–68. [62] Liu X, et al. Optogenetic stimulation of a hippocampal engram activates fear memory recall. Nature 2012;484(7394):381–5.

7

[63] Ramirez S, et al. Creating a false memory in the hippocampus. Science 2013;341(6144):387–91. [64] Bannerman DM, et al. Ventral hippocampal lesions affect anxiety but not spatial learning. Behav Brain Res 2003;139(1–2):197–213. [65] Matus-Amat P, et al. The role of the dorsal hippocampus in the acquisition and retrieval of context memory representations. J Neurosci 2004;24(10):2431–9. [66] Richmond MA, et al. Dissociating context and space within the hippocampus: effects of complete, dorsal, and ventral excitotoxic hippocampal lesions on conditioned freezing and spatial learning. Behav Neurosci 1999;113(6):1189–203. [67] Rudy JW, Matus-Amat P. The ventral hippocampus supports a memory representation of context and contextual fear conditioning: implications for a unitary function of the hippocampus. Behav Neurosci 2005;119(1):154–63. [68] Kjelstrup KG, et al. Reduced fear expression after lesions of the ventral hippocampus. Proc Natl Acad Sci USA 2002;99(16):10825–30. [69] Fanselow MS, Dong HW. Are the dorsal and ventral hippocampus functionally distinct structures? Neuron 2010;65(1):7–19. [70] Pitkanen A, et al. Reciprocal connections between the amygdala and the hippocampal formation, perirhinal cortex, and postrhinal cortex in rat. A review Ann N Y Acad Sci 2000;911:369–91. [71] Herry C, et al. Switching on and off fear by distinct neuronal circuits. Nature 2008;454(7204):600–6. [72] McDonald AJ, Mascagni F, Guo L. Projections of the medial and lateral prefrontal cortices to the amygdala: a Phaseolus vulgaris leucoagglutinin study in the rat. Neuroscience 1996;71(1):55–75. [73] Vertes RP. Interactions among the medial prefrontal cortex: hippocampus and midline thalamus in emotional and cognitive processing in the rat. Neuroscience 2006;142(1):1–20. [74] Duvarci S, Bauer EP, Pare D. The bed nucleus of the stria terminalis mediates inter-individual variations in anxiety and fear. J Neurosci 2009;29(33):10357–61. [75] Amano T, et al. The fear circuit revisited: contributions of the basal amygdala nuclei to conditioned fear. J Neurosci 2011;31(43):15481–9. [76] Duvarci S, Popa D, Pare D. Central amygdala activity during fear conditioning. J Neurosci 2011;31(1):289–94. [77] Burgos-Robles A, Vidal-Gonzalez I, Quirk GJ. Sustained conditioned responses in prelimbic prefrontal neurons are correlated with fear expression and extinction failure. J Neurosci 2009;29(26):8474–82. [78] Sotres-Bayon F, et al. Gating of fear in prelimbic cortex by hippocampal and amygdala inputs. Neuron 2012;76(4):804–12. [79] Falls WA, Miserendino MJ, Davis M. Extinction of fear-potentiated startle: blockade by infusion of an NMDA antagonist into the amygdala. J Neurosci 1992;12(3):854–63. [80] Lee H, Kim JJ. Amygdalar NMDA receptors are critical for new fear learning in previously fear-conditioned rats. J Neurosci 1998;18(20):8444–54. [81] Lin CH, et al. The similarities and diversities of signal pathways leading to consolidation of conditioning and consolidation of extinction of fear memory. J Neurosci 2003;23(23):8310–7. [82] Parkes SL, Westbrook RF. The basolateral amygdala is critical for the acquisition and extinction of associations between a neutral stimulus and a learned danger signal but not between two neutral stimuli. J Neurosci 2010;30(38):12608–18. [83] Amano T, Unal CT, Pare D. Synaptic correlates of fear extinction in the amygdala. Nat Neurosci 2010;13(4):489–94. [84] Milad MR, Quirk GJ. Neurons in medial prefrontal cortex signal memory for fear extinction. Nature 2002;420(6911):70–4. [85] Senn V, et al. Long-range connectivity defines behavioral specificity of amygdala neurons. Neuron 2014;81(2):428–37. [86] Christianson JP, et al. 5-hydroxytryptamine 2C receptors in the basolateral amygdala are involved in the expression of anxiety after uncontrollable traumatic stress. Biol Psychiatry 2010;67(4):339–45. [87] Mo B, et al. Restraint stress increases serotonin release in the central nucleus of the amygdala via activation of corticotropin-releasing factor receptors. Brain Res Bull 2008;76(5):493–8. [88] Rueter LE, Jacobs BL. A microdialysis examination of serotonin release in the rat forebrain induced by behavioral/environmental manipulations. Brain Res 1996;739(1-2):57–69. [89] Yoshioka M, et al. Effects of conditioned fear stress on 5-HT release in the rat prefrontal cortex. Pharmacol Biochem Behav 1995;51(2-3):515–9. [90] Barnes NM, Sharp T. A review of central 5-HT receptors and their function. Neuropharmacology 1999;38(8):1083–152. [91] Gartside SE, et al. Effects of (-)-tertatolol, (-)-penbutolol and (+/-)-pindolol in combination with paroxetine on presynaptic 5-HT function: an in vivo microdialysis and electrophysiological study. Br J Pharmacol 1999;127(1): 145–52. [92] McDonald AJ, Mascagni F. Neuronal localization of 5-HT type 2A receptor immunoreactivity in the rat basolateral amygdala. Neuroscience 2007;146(1):306–20. [93] Araneda R, Andrade R. 5-Hydroxytryptamine2 and 5-hydroxytryptamine 1A receptors mediate opposing responses on membrane excitability in rat association cortex. Neuroscience 1991;40(2):399–412. [94] Davies MF, et al. Two distinct effects of 5-hydroxytryptamine on single cortical neurons. Brain Res 1987;423(1-2):347–52. [95] Guo JD, et al. Bi-directional modulation of bed nucleus of stria terminalis neurons by 5-HT: molecular expression and functional properties of excitatory 5-HT receptor subtypes. Neuroscience 2009;164(4):1776–93.

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10 8

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

[96] Richardson-Jones JW, et al. Serotonin-1A autoreceptors are necessary and sufficient for the normal formation of circuits underlying innate anxiety. J Neurosci 2011;31(16):6008–18. [97] Griebel G. 5-Hydroxytryptamine-interacting drugs in animal models of anxiety disorders: more than 30 years of research. Pharmacol Ther 1995;65(3): 319–95. [98] Handley SL, McBlane JW. Serotonin mechanisms in animal models of anxiety. Braz J Med Biol Res 1993;26(1):1–13. [99] Sanchez C, Meier E. Behavioral profiles of SSRIs in animal models of depression: anxiety and aggression. Are they all alike? Psychopharmacology (Berl) 1997;129(3):197–205. [100] Vertes RPA. PHA-L analysis of ascending projections of the dorsal raphe nucleus in the rat. J Comp Neurol 1991;313(4):643–68. [101] Stutzmann GE, LeDoux JE. GABAergic antagonists block the inhibitory effects of serotonin in the lateral amygdala: a mechanism for modulation of sensory inputs related to fear conditioning. J Neurosci 1999;19(11):RC8. [102] Li Q, et al. Brain region-specific alterations of 5-HT2A and 5-HT2C receptors in serotonin transporter knockout mice. J Neurochem 2003;84(6): 1256–65. [103] Commons KG, Connolley KR, Valentino RJ. A neurochemically distinct dorsal raphe-limbic circuit with a potential role in affective disorders. Neuropsychopharmacology 2003;28(2):206–15. [104] Azmitia EC, Segal M. An autoradiographic analysis of the differential ascending projections of the dorsal and median raphe nuclei in the rat. J Comp Neurol 1978;179(3):641–67. [105] Zhong P, Yan Z. Differential regulation of the excitability of prefrontal cortical fast-spiking interneurons and pyramidal neurons by serotonin and fluoxetine. PLoS One 2011;6(2):e16970. [106] McQuade R, Sharp T. Functional mapping of dorsal and median raphe 5hydroxytryptamine pathways in forebrain of the rat using microdialysis. J Neurochem 1997;69(2):791–6. [107] Corradetti R, et al. Serotonin blocks the long-term potentiation induced by primed burst stimulation in the CA1 region of rat hippocampal slices. Neuroscience 1992;46(3):511–8. [108] Passani MB, et al. Effects of DAU 6215: a novel 5-hydroxytryptamine3 (5HT3) antagonist on electrophysiological properties of the rat hippocampus. Br J Pharmacol 1994;112(2):695–703. [109] Alhaider AA, Ageel AM, Ginawi OT. The quipazine- and TFMPP-increased conditioned avoidance response in rats: role of 5HT1C/5-HT2 receptors. Neuropharmacology 1993;32(12):1427–32. [110] Harvey JA. Role of the serotonin 5-HT(2A) receptor in learning. Learn Mem 2003;10(5):355–62. [111] Zhang G, et al. Stimulation of serotonin 2A receptors facilitates consolidation and extinction of fear memory in C57BL/6J mice. Neuropharmacology 2013;64:403–13. [112] Ma TC, Yu QH. Effect of 20(S)-ginsenoside-Rg2 and cyproheptadine on twoway active avoidance learning and memory in rats. Arzneimittelforschung 1993;43(10):1049–52. [113] Marner L, et al. Loss of serotonin 2A receptors exceeds loss of serotonergic projections in early Alzheimer’s disease: a combined [11C]DASB and [18F]altanserin-PET study. Neurobiol Aging 2012;33(3):479–87. [114] Meltzer CC, et al. Serotonin in aging, late-life depression, and Alzheimer’s disease: the emerging role of functional imaging. Neuropsychopharmacology 1998;18(6):407–30. [115] Wagner M, et al. The His452Tyr variant of the gene encoding the 5-HT2A receptor is specifically associated with consolidation of episodic memory in humans. Int J Neuropsychopharmacol 2008;11(8):1163–7. [116] Celada P, et al. The hallucinogen DOI reduces low-frequency oscillations in rat prefrontal cortex: reversal by antipsychotic drugs. Biol Psychiatry 2008;64(5):392–400. [117] Yoshida H, et al. Subtype specific roles of serotonin receptors in the spine formation of cortical neurons in vitro. Neurosci Res 2011;71(3): 311–4. [118] Peddie CJ, et al. Colocalisation of serotonin2A receptors with the glutamate receptor subunits NR1 and GluR2 in the dentate gyrus: an ultrastructural study of a modulatory role. Exp Neurol 2008;211(2):561–73. [119] Xia Z, et al. A direct interaction of PSD-95 with 5-HT2A serotonin receptors regulates receptor trafficking and signal transduction. J Biol Chem 2003;278(24):21901–8. [120] Ciranna L. Serotonin as a modulator of glutamate- and GABA-mediated neurotransmission: implications in physiological functions and in pathology. Curr Neuropharmacol 2006;4(2):101–14. [121] Hasuo H, Matsuoka T, Akasu T. Activation of presynaptic 5hydroxytryptamine 2A receptors facilitates excitatory synaptic transmission via protein kinase C in the dorsolateral septal nucleus. J Neurosci 2002;22(17):7509–17. [122] Arvanov VL, et al. A pre- and postsynaptic modulatory action of 5-HT and the 5-HT2A: 2C receptor agonist DOB on NMDA-evoked responses in the rat medial prefrontal cortex. Eur J Neurosci 1999;11(8):2917–34. [123] Chen A, Hough CJ, Li H. Serotonin type II receptor activation facilitates synaptic plasticity via N-methyl-D-aspartate-mediated mechanism in the rat basolateral amygdala. Neuroscience 2003;119(1):53–63. [124] Weisstaub NV, et al. Cortical 5-HT2A receptor signaling modulates anxietylike behaviors in mice. Science 2006;313(5786):536–40. [125] Davis M, et al. Phasic vs sustained fear in rats and humans: role of the extended amygdala in fear vs anxiety. Neuropsychopharmacology 2010;35(1):105–35.

[126] Campbell BM, Merchant KM. Serotonin 2C receptors within the basolateral amygdala induce acute fear-like responses in an open-field environment. Brain Res 2003;993(1-2):1–9. [127] de Mello Cruz AP, et al. Behavioral effects of systemically administered MK-212 are prevented by ritanserin microinfusion into the basolateral amygdala of rats exposed to the elevated plus-maze. Psychopharmacology (Berl) 2005;182(3):345–54. [128] Gibson EL, Barnfield AM, Curzon G. Evidence that mCPP-induced anxiety in the plus-maze is mediated by postsynaptic 5-HT2C receptors but not by sympathomimetic effects. Neuropharmacology 1994;33(3–4):457–65. [129] Heisler LK, et al. Serotonin 5-HT(2C) receptors regulate anxiety-like behavior. Genes Brain Behav 2007;6(5):491–6. [130] Hensman R, et al. Effects of ritanserin on aversive classical conditioning in humans. Psychopharmacology (Berl) 1991;104(2):220–4. [131] Burghardt NS, et al. Acute selective serotonin reuptake inhibitors increase conditioned fear expression: blockade with a 5-HT(2C) receptor antagonist. Biol Psychiatry 2007;62(10):1111–8. [132] De Vry J. 5-HT1A receptor agonists: recent developments and controversial issues. Psychopharmacology (Berl) 1995;121(1):1–26. [133] Lacivita E, et al. 5-HT1A receptor: an old target for new therapeutic agents. Curr Top Med Chem 2008;8(12):1024–34. [134] Gross C, et al. Altered fear circuits in 5-HT(1A) receptor KO mice. Biol Psychiatry 2000;48(12):1157–63. [135] Olivier B, et al. The 5-HT(1A) receptor knockout mouse and anxiety. Behav Pharmacol 2001;12(6-7):439–50. [136] Ramboz S, et al. Serotonin receptor 1A knockout: an animal model of anxietyrelated disorder. Proc Natl Acad Sci USA 1998;95(24):14476–81. [137] Li XB, et al. Effect of chronic administration of flesinoxan and fluvoxamine on freezing behavior induced by conditioned fear. Eur J Pharmacol 2001;425(1):43–50. [138] Nakamura K, Kurasawa M. Anxiolytic effects of aniracetam in three different mouse models of anxiety and the underlying mechanism. Eur J Pharmacol 2001;420(1):33–43. [139] Borelli KG, et al. Effects of inactivation of serotonergic neurons of the median raphe nucleus on learning and performance of contextual fear conditioning. Neurosci Lett 2005;387(2):105–10. [140] Stiedl O, et al. Involvement of the 5-HT1A receptors in classical fear conditioning in C57BL/6J mice. J Neurosci 2000;20(22):8515–27. [141] Almada RC, et al. Serotonergic mechanisms of the median raphe nucleus-dorsal hippocampus in conditioned fear: Output circuit involves the prefrontal cortex and amygdala. Behav Brain Res 2009;203(2): 279–87. [142] Gomes FV, et al. Cannabidiol injected into the bed nucleus of the stria terminalis reduces the expression of contextual fear conditioning via 5-HT1A receptors. J Psychopharmacol 2012;26(1):104–13. [143] Li X, et al. 5-HT1A receptor agonist affects fear conditioning through stimulations of the postsynaptic 5-HT1A receptors in the hippocampus and amygdala. Eur J Pharmacol 2006;532(1-2):74–80. [144] Morrison KE, Cooper MA. A role for 5-HT1A receptors in the basolateral amygdala in the development of conditioned defeat in Syrian hamsters. Pharmacol Biochem Behav 2012;100(3):592–600. [145] Levita L, et al. 5-hydroxytryptamine1A-like receptor activation in the bed nucleus of the stria terminalis: electrophysiological and behavioral studies. Neuroscience 2004;128(3):583–96. [146] Tada K, et al. Endogenous 5-HT inhibits firing activity of hippocampal CA1 pyramidal neurons during conditioned fear stress-induced freezing behavior through stimulating 5-HT1A receptors. Hippocampus 2004;14(2): 143–7. [147] Ferreira R, Nobre MJ. Conditioned fear in low- and high-anxious rats is differentially regulated by cortical subcortical and midbrain 5-HT receptors. Neuroscience 2014;268C:159–68. [148] Bush DE, Sotres-Bayon F, LeDoux JE. Individual differences in fear: isolating fear reactivity and fear recovery phenotypes. J Trauma Stress 2007;20(4):413–22. [149] Goswami S, et al. Impact of predatory threat on fear extinction in Lewis rats. Learn Mem 2010;17(10):494–501. [150] Harrell AV, Allan AM. Improvements in hippocampal-dependent learning and decremental attention in 5-HT(3) receptor overexpressing mice. Learn Mem 2003;10(5):410–9. [151] Nevins ME, Anthony EW. Antagonists at the serotonin-3 receptor can reduce the fear-potentiated startle response in the rat: evidence for different types of anxiolytic activity? J Pharmacol Exp Ther 1994;268(1):248–54. [152] Kondo M, et al. The 5-HT3A receptor is essential for fear extinction. Learn Mem 2014;21(1):1–4. [153] Park SM, Williams CL. Contribution of serotonin type 3 receptors in the successful extinction of cued or contextual fear conditioned responses: interactions with GABAergic signaling. Rev Neurosci 2012;23(5-6):555–69. [154] Blakely RD, et al. Cloning and expression of a functional serotonin transporter from rat brain. Nature 1991;354(6348):66–70. [155] Lesch KP, et al. Association of anxiety-related traits with a polymorphism in the serotonin transporter gene regulatory region. Science 1996;274(5292):1527–31. [156] Katsuragi S, et al. Association between serotonin transporter gene polymorphism and anxiety-related traits. Biol Psychiatry 1999;45(3):368–70. [157] Hariri AR, et al. Serotonin transporter genetic variation and the response of the human amygdala. Science 2002;297(5580):400–3.

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

[158] Eker MC, et al. Smaller hippocampus volume is associated with short variant of 5-HTTLPR polymorphism in medication-free major depressive disorder patients. Neuropsychobiology 2011;63(1):22–8. [159] Pezawas L, et al. 5-HTTLPR polymorphism impacts human cingulateamygdala interactions: a genetic susceptibility mechanism for depression. Nat Neurosci 2005;8(6):828–34. [160] Brocke B, et al. Serotonin transporter gene variation impacts innate fear processing: Acoustic startle response and emotional startle. Mol Psychiatry 2006;11(12):1106–12. [161] Garpenstrand H, et al. Human fear conditioning is related to dopaminergic and serotonergic biological markers. Behav Neurosci 2001;115(2): 358–64. [162] Klumpers F, et al. Genetic variation in serotonin transporter function affects human fear expression indexed by fear-potentiated startle. Biol Psychol 2012;89(2):277–82. [163] Crisan LG, et al. Genetic contributions of the serotonin transporter to social learning of fear and economic decision making. Soc Cogn Affect Neurosci 2009;4(4):399–408. [164] Barkus C, et al. Variation in serotonin transporter expression modulates fearevoked hemodynamic responses and theta-frequency neuronal oscillations in the amygdala. Biol Psychiatry 2013. [165] Narayanan V, et al. Social defeat: impact on fear extinction and amygdalaprefrontal cortical theta synchrony in 5-HTT deficient mice. PLoS One 2011;6(7):e22600. [166] Wellman CL, et al. Impaired stress-coping and fear extinction and abnormal corticolimbic morphology in serotonin transporter knock-out mice. J Neurosci 2007;27(3):684–91. [167] Glotzbach-Schoon E, et al. Contextual fear conditioning in virtual reality is affected by 5HTTLPR and NPSR1 polymorphisms: effects on fear-potentiated startle. Front Behav Neurosci 2013;7:31. [168] Heitland I, et al. Human fear acquisition deficits in relation to genetic variants of the corticotropin releasing hormone receptor 1 and the serotonin transporter. PLoS One 2013;8(5):e63772. [169] Champagne FA, Curley JP. How social experiences influence the brain. Curr Opin Neurobiol 2005;15(6):704–9. [170] Gross C, Hen R. Genetic and environmental factors interact to influence anxiety. Neurotox Res 2004;6(6):493–501. [171] Lesch KP. Gene-environment interaction and the genetics of depression. J Psychiatry Neurosci 2004;29(3):174–84. [172] Caspi A, et al. Influence of life stress on depression: moderation by a polymorphism in the 5-HTT gene. Science 2003;301(5631):386–9. [173] Jansen F, et al. Modulation of behavioural profile and stress response by 5-HTT genotype and social experience in adulthood. Behav Brain Res 2010;207(1):21–9. [174] Gyawali S, et al. Association of a polyadenylation polymorphism in the serotonin transporter and panic disorder. Biol Psychiatry 2010;67(4):331–8. [175] Hartley CA, et al. Serotonin transporter polyadenylation polymorphism modulates the retention of fear extinction memory. Proc Natl Acad Sci USA 2012;109(14):5493–8. [176] Stokes PE, Holtz A. Fluoxetine tenth anniversary update: the progress continues. Clin Ther 1997;19(5):1135–250. [177] van der Kolk BA, et al. Fluoxetine in posttraumatic stress disorder. J Clin Psychiatry 1994;55(12):517–22. [178] Griebel G, et al. Behavioral effects of acute and chronic fluoxetine in WistarKyoto rats. Physiol Behav 1999;67(3):315–20. [179] Kurt M, Arik AC, Celik S. The effects of sertraline and fluoxetine on anxiety in the elevated plus-maze test in mice. J Basic Clin Physiol Pharmacol 2000;11(2):173–80. [180] Ravinder S, Pillai AG, Chattarji S. Cellular correlates of enhanced anxiety caused by acute treatment with the selective serotonin reuptake inhibitor fluoxetine in rats. Front Behav Neurosci 2011;5:88. [181] Silva RC, Brandao ML. Acute and chronic effects of gepirone and fluoxetine in rats tested in the elevated plus-maze: an ethological analysis. Pharmacol Biochem Behav 2000;65(2):209–16. [182] Bodnoff SR, et al. A comparison of the effects of diazepam versus several typical and atypical anti-depressant drugs in an animal model of anxiety. Psychopharmacology (Berl) 1989;97(2):277–9. [183] Bagdy G, et al. Anxiety-like effects induced by acute fluoxetine: sertraline or m-CPP treatment are reversed by pretreatment with the 5-HT2C receptor antagonist SB-242084 but not the 5-HT1A receptor antagonist WAY-100635. Int J Neuropsychopharmacol 2001;4(4):399–408. [184] Dekeyne A, et al. Citalopram reduces social interaction in rats by activation of serotonin (5-HT)(2C) receptors. Neuropharmacology 2000;39(6):1114–7. [185] Belzung C, et al. An investigation of the mechanisms responsible for acute fluoxetine-induced anxiogenic-like effects in mice. Behav Pharmacol 2001;12(3):151–62. [186] Feighner JP, et al. Double-blind comparison of bupropion and fluoxetine in depressed outpatients. J Clin Psychiatry 1991;52(8):329–35. [187] Fergusson D, et al. Association between suicide attempts and selective serotonin reuptake inhibitors: systematic review of randomised controlled trials. BMJ 2005;330(7488):396. [188] Goldstein BJ, Goodnick PJ. Selective serotonin reuptake inhibitors in the treatment of affective disorders—III. Tolerability, safety and pharmacoeconomics. J Psychopharmacol 1998;12(3 Suppl B):S55–87. [189] Gorman JM, et al. An open trial of fluoxetine in the treatment of panic attacks. J Clin Psychopharmacol 1987;7(5):329–32.

9

[190] Masand PS, Gupta S. Selective serotonin-reuptake inhibitors: an update. Harv Rev Psychiatry 1999;7(2):69–84. [191] Spigset O. Adverse reactions of selective serotonin reuptake inhibitors: reports from a spontaneous reporting system. Drug Saf 1999;20(3):277–87. [192] Teicher MH, Glod C, Cole JO. Emergence of intense suicidal preoccupation during fluoxetine treatment. Am J Psychiatry 1990;147(2):207–10. [193] Blier P, Bergeron R. Effectiveness of pindolol with selected antidepressant drugs in the treatment of major depression. J Clin Psychopharmacol 1995;15(3):217–22. [194] Gardier AM, et al. Role of 5-HT1A autoreceptors in the mechanism of action of serotoninergic antidepressant drugs: recent findings from in vivo microdialysis studies. Fundam Clin Pharmacol 1996;10(1):16–27. [195] Riaza Bermudo-Soriano C, et al. New perspectives in glutamate and anxiety. Pharmacol Biochem Behav 2012;100(4):752–74. [196] Sanacora G, Treccani G, Popoli M. Towards a glutamate hypothesis of depression: an emerging frontier of neuropsychopharmacology for mood disorders. Neuropharmacology 2012;62(1):63–77. [197] Skolnick P, et al. Adaptation of N-methyl-D-aspartate (NMDA) receptors following antidepressant treatment: implications for the pharmacotherapy of depression. Pharmacopsychiatry 1996;29(1):23–6. [198] Bosker FJ, et al. Acute and chronic effects of citalopram on postsynaptic 5hydroxytryptamine(1A) receptor-mediated feedback: a microdialysis study in the amygdala. J Neurochem 2001;76(6):1645–53. [199] Hervas I, et al. Role of uptake inhibition and autoreceptor activation in the control of 5-HT release in the frontal cortex and dorsal hippocampus of the rat. Br J Pharmacol 2000;130(1):160–6. [200] Invernizzi R, et al. Effect of 5-HT1A receptor antagonists on citalopraminduced increase in extracellular serotonin in the frontal cortex, striatum and dorsal hippocampus. Neuropharmacology 1997;36(4-5):467–73. [201] Burghardt NS, et al. The selective serotonin reuptake inhibitor citalopram increases fear after acute treatment but reduces fear with chronic treatment: a comparison with tianeptine. Biol Psychiatry 2004;55(12):1171–8. [202] Gravius A, et al. The role of group I metabotropic glutamate receptors in acquisition and expression of contextual and auditory fear conditioning in rats—a comparison. Neuropharmacology 2006;51(7-8):1146–55. [203] Ravinder S, et al. A role for the extended amygdala in the fear-enhancing effects of acute selective serotonin reuptake inhibitor treatment. Transl Psychiatry 2013;3:e209. [204] Grillon C, Levenson J, Pine DS. A single dose of the selective serotonin reuptake inhibitor citalopram exacerbates anxiety in humans: a fear-potentiated startle study. Neuropsychopharmacology 2007;32(1):225–31. [205] Lino-de-Oliveira C, et al. Effects of acute and chronic fluoxetine treatments on restraint stress-induced Fos expression. Brain Res Bull 2001;55(6):747–54. [206] Morelli M, et al. Induction of Fos-like-immunoreactivity in the central extended amygdala by antidepressant drugs. Synapse 1999;31(1):1–4. [207] Singewald N, Salchner P, Sharp T. Induction of c-Fos expression in specific areas of the fear circuitry in rat forebrain by anxiogenic drugs. Biol Psychiatry 2003;53(4):275–83. [208] Veening JG, et al. Patterns of c-fos expression induced by fluvoxamine are different after acute vs chronic oral administration. Eur Neuropsychopharmacol 1998;8(3):213–26. [209] Hashimoto S, et al. Effects of acute citalopram on the expression of conditioned freezing in naive versus chronic citalopram-treated rats. Prog Neuropsychopharmacol Biol Psychiatry 2009;33(1):113–7. [210] Inoue T, et al. Effect of citalopram: a selective serotonin reuptake inhibitor, on the acquisition of conditioned freezing. Eur J Pharmacol 1996;311(1):1–6. [211] Montezinho LP, et al. The effects of acute treatment with escitalopram on the different stages of contextual fear conditioning are reversed by atomoxetine. Psychopharmacology (Berl) 2010;212(2):131–43. [212] Hashimoto S, Inoue T, Koyama T. Serotonin reuptake inhibitors reduce conditioned fear stress-induced freezing behavior in rats. Psychopharmacology (Berl) 1996;123(2):182–6. [213] Muraki I, Inoue T, Koyama T. Effect of co-administration of the selective 5HT1A receptor antagonist WAY 100,635 and selective 5-HT1B/1D receptor antagonist GR 127,935 on anxiolytic effect of citalopram in conditioned fear stress in the rat. Eur J Pharmacol 2008;586(1-3):171–8. [214] Santos JM, Martinez RC, Brandao ML. Effects of acute and subchronic treatments with fluoxetine and desipramine on the memory of fear in moderate and high-intensity contextual conditioning. Eur J Pharmacol 2006;542(13):121–8. [215] Igelstrom KM, Heyward PM. Inhibition of hippocampal excitability by citalopram. Epilepsia 2012;53(11):2034–42. [216] Staubli U, Otaky N. Serotonin controls the magnitude of LTP induced by theta bursts via an action on NMDA-receptor-mediated responses. Brain Res 1994;643(1-2):10–6. [217] Anagnostaras SG, Gale GD, Fanselow MS. Hippocampus and contextual fear conditioning: recent controversies and advances. Hippocampus 2001;11(1):8–17. [218] Frankland PW, et al. The dorsal hippocampus is essential for context discrimination but not for contextual conditioning. Behav Neurosci 1998;112(4):863–74. [219] Burghardt NS, et al. Chronic antidepressant treatment impairs the acquisition of fear extinction. Biol Psychiatry 2012. [220] Lebron-Milad K, et al. Sex differences and estrous cycle in female rats interact with the effects of fluoxetine treatment on fear extinction. Behav Brain Res 2013;253:217–22.

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

G Model BBR-9047; No. of Pages 10 10

ARTICLE IN PRESS E.P. Bauer / Behavioural Brain Research xxx (2014) xxx–xxx

[221] Young SN, et al. Tryptophan depletion causes a rapid lowering of mood in normal males. Psychopharmacology (Berl) 1985;87(2):173–7. [222] Fadda F, et al. A tryptophan-free diet markedly reduces frontocortical 5-HT release: but fails to modify ethanol preference in alcohol-preferring (sP) and non-preferring (sNP) rats. Behav Brain Res 2000;108(2):127–32. [223] Stancampiano R, et al. Acute administration of a tryptophan-free amino acid mixture decreases 5-HT release in rat hippocampus in vivo. Am J Physiol 1997;272(3 Pt 2):R991–4. [224] Jequier E, Lovenberg W, Sjoerdsma A. Tryptophan hydroxylase inhibition: the mechanism by which p-chlorophenylalanine depletes rat brain serotonin. Mol Pharmacol 1967;3(3):274–8. [225] Baumgarten HG, Bjorklund A. Neurotoxic indoleamines and monoamine neurons. Annu Rev Pharmacol Toxicol 1976;16:101–11. [226] Robinson OJ, et al. Acute tryptophan depletion increases translational indices of anxiety but not fear: serotonergic modulation of the bed nucleus of the stria terminalis? Neuropsychopharmacology 2012;37(8):1963–71. [227] Uchida S, et al. Brain neurotransmitter receptor-binding characteristics in rats after oral administration of haloperidol: risperidone and olanzapine. Life Sci 2007;80(17):1635–40. [228] Hughes CR, Keele NB. Phenytoin normalizes exaggerated fear behavior in pchlorophenylalanine (PCPA)-treated rats. Epilepsy Behav 2006;9(4):557–63.

[229] Inoue T. Effects of conditioned fear stress on monoaminergic systems in the rat brain. Hokkaido Igaku Zasshi 1993;68(3):377–90. [230] Hindi Attar C, Finckh B, Buchel C. The influence of serotonin on fear learning. PLoS One 2012;7(8):e42397. [231] Roiser JP, et al. The effects of acute tryptophan depletion and serotonin transporter polymorphism on emotional processing in memory and attention. Int J Neuropsychopharmacol 2007;10(4):449–61. [232] Lothe A, et al. Association between triallelic polymorphism of the serotonin transporter and [18F]MPPF binding potential at 5-HT1A receptors in healthy subjects. Neuroimage 2009;47(2):482–92. [233] Jasnow AM, et al. Thy1-expressing neurons in the basolateral amygdala may mediate fear inhibition. J Neurosci 2013;33(25):10396–404. [234] Benzekhroufa K, et al. Adenoviral vectors for highly selective gene expression in central serotonergic neurons reveal quantal characteristics of serotonin release in the rat brain. BMC Biotechnol 2009;9:23. [235] Benzekhroufa K, et al. Targeting central serotonergic neurons with lentiviral vectors based on a transcriptional amplification strategy. Gene Ther 2009;16(5):681–8. [236] Fenno L, Yizhar O, Deisseroth K. The development and application of optogenetics. Annu Rev Neurosci 2011;34:389–412.

Please cite this article in press as: Bauer EP. Serotonin in fear conditioning processes. Behav Brain Res (2014), http://dx.doi.org/10.1016/j.bbr.2014.07.028

Serotonin in fear conditioning processes.

This review describes the latest developments in our understanding of how the serotonergic system modulates Pavlovian fear conditioning, fear expressi...
2MB Sizes 0 Downloads 9 Views