Skip to main content
Intended for healthcare professionals
Open access
Review article
First published online July 21, 2020

The role of the locus coeruleus in the generation of pathological anxiety

Abstract

This review aims to synthesise a large pre-clinical and clinical literature related to a hypothesised role of the locus coeruleus norepinephrine system in responses to acute and chronic threat, as well as the emergence of pathological anxiety. The locus coeruleus has widespread norepinephrine projections throughout the central nervous system, which act to globally modulate arousal states and adaptive behavior, crucially positioned to play a significant role in modulating both ascending visceral and descending cortical neurocognitive information. In response to threat or a stressor, the locus coeruleus–norepinephrine system globally modulates arousal, alerting and orienting functions and can have a powerful effect on the regulation of multiple memory systems. Chronic stress leads to amplification of locus coeruleus reactivity to subsequent stressors, which is coupled with the emergence of pathological anxiety-like behaviors in rodents. While direct in vivo evidence for locus coeruleus dysfunction in humans with pathological anxiety remains limited, recent advances in high-resolution 7-T magnetic resonance imaging and computational modeling approaches are starting to provide new insights into locus coeruleus characteristics.

Introduction

Pathological anxiety can be defined as a fear-like or defensive physiological and behavioral state that persists in a non-threatening environment (Eysenck, 1992; Rosen and Schulkin, 1998). Pathological anxiety is a core feature of anxiety disorders as defined in the Diagnostic and Statistical Manual of Mental Disorders (5th ed.; DSM-5), including generalised anxiety disorder (GAD), social anxiety disorder (SAD) and panic disorder, and post-traumatic stress disorder (PTSD). These disorders represent the most prevalent class of psychiatric disorders in the United States, with an anxiety disorder showing around an 18% 12-month prevalence rate (Kessler et al., 2005b) and they act as a major risk factor for suicide (Eysenck, 1992; Kessler et al., 2005b; Rosen and Schulkin, 1998). Specific phobia and social phobia have the highest prevalence rates, followed by PTSD, GAD and panic disorder (Kessler et al., 2005b) (Table 1). Panic disorder and GAD, both archetypal examples of disorders of pathological anxiety, represent the sixth leading cause of years lived with disability worldwide (Kessler et al., 2005b)
Table 1. Characterisation of human anxiety and stress-related disorders.
Human anxiety disorder Core symptoms (DSM-5) Research domain criteria (RDoC) domain / construct Average age of onset (years) Prevalence (%, SE) Female:male ratio
        Total Serious Mod Mild  
Panic disorder • Recurrent unexpected panic attacks.
• At least one of:
Persistent concern about having additional attacks
Worry of the implications or consequences of the attack
A significant change in behavior related to the attacks
• Absence of agoraphobia/presence of agoraphobia
• The panic attacks are not caused by the direct physiological effects of a substance or medical condition.
• The panic attacks are not better accounted for by another mental disorder.
Negative valence systems / potential threat (‘anxiety’), acute threat (‘fear’), sustained threat. 30.3; 95% CI = 26.09 to 34.59 2.7 (0.2) 44.8 (3.2) 29.5 (2.7) 25.7 (2.5) 2.1
Specific phobia • Excessive or unreasonable, persistent and intense fear triggered instantaneously by a specific object or situation, out of proportion to the actual danger.
• Avoidance or extreme distress.
• The phobia significantly impacts school, work or personal life.
Negative valence systems / potential threat (‘anxiety’), acute threat (‘fear’), sustained threat. 11.0; 95% CI = 8.25 to 13.65 8.7 (0.4) 21.9 (2.0) 30.0 (2.0) 48.1 (2.1) 1.8
Social anxiety disorder (or social phobia) • Marked and persistent fear of social or performance situations and scrutiny that will be humiliating or embarrassing.
• Exposure to the feared social situation provokes anxiety or panic attack.
• The person recognises that the fear is excessive or unreasonable.
• The feared social or performance situations are avoided or endured with distress.
• The avoidance, anxious anticipation or distress interferes with occupational or academic functioning, social activities or relationships, or there is marked distress.
• The fear or avoidance is not due to the direct physiological effects of a substance or a general medical condition and is not better accounted for by another mental disorder.
Negative valence systems / potential threat (‘anxiety’), acute threat (‘fear’), sustained threat. 14.3; 95% CI = 13.27 to 15.41 6.8 (0.3) 29.9 (2.0) 38.8 (2.5) 31.3 (2.4) 1.6
Generalised anxiety disorder • Excessive anxiety and worry (apprehensive expectation) about a number of events or activities (such as work or school performance).
• The worry is difficult to control.
• The anxiety and worry are associated with three (or more) of the following symptoms:
Restlessness
Being easily fatigued
Difficulty concentrating
Irritability
Muscle tension
Sleep disturbance
• The focus of the anxiety and worry is not confined to features of another mental disorder.
• The anxiety, worry or physical symptoms cause clinically significant distress or impairment in social, occupational or other important areas of functioning.
• The disturbance is not caused by the direct physiological effects of a substance or a general medical condition.
Negative valence systems / potential threat (‘anxiety’), acute threat (‘fear’), sustained threat. Cognitive systems / cognitive control. Arousal and regulatory systems / arousal, sleep-wakefulness. 34.9; 95% CI = 30.88 to 39.01 3.1 (0.2) 32.3 (2.9) 44.6 (4.0) 23.1 (2.9) 1.7
Post-traumatic stress disorder • Exposure to a traumatic event in which the person experienced, witnessed or was confronted with event(s) that involved actual or threatened death or serious injury to self or others.
• The traumatic event is persistently re-experienced as recurrent and intrusive distressing recollections of the event, including images, thoughts or perceptions.
• Persistent avoidance of stimuli associated with the trauma and numbing of general responsiveness.
• Persistent symptoms of increased arousal as indicated by two (or more) of:
Difficulty falling or staying asleep
Irritability or outbursts of anger
Difficulty concentrating
Hypervigilance
Exaggerated startle response
Negative valence systems / potential threat (‘anxiety’), acute threat (‘fear’), sustained threat. Cognitive systems / attention, working memory. Arousal and regulatory systems / arousal. 26.6; 95% CI = 22.13 to 31.06 3.5 (0.3) 36.6 (3.5) 33.1 (2.2) 30.2 (3.4) 1.9
Any anxiety disorder     21.3; 95% CI = 17.46 to 25.07 18.1 (0.7) 22.8 (1.5) 33.7 (1.4) 43.5 (2.1) 1.5
Demographic and feature domain characteristics of disorders characterised by pathological anxiety. Information obtained from Ditlevsen and Elklit (2012); Kessler et al. (2005a, 2012); Lijster et al. (2017).
The Research Domain Criteria (RDoC) initiative (Insel et al., 2010) aims to deconstruct traditional diagnostic categories into constituent domains and constructs that are relevant and testable across species and units of analysis. This approach is aimed to ultimately provide more fundamental measures for diagnostics and treatment determination. Human anxiety disorders are commonly conceptualised as disorders of maladaptive response to acute threat (fear) and potential threat (anxiety), both within the RDoC negative valence systems domain. Threatening environmental stimuli or ‘stressors’ typically induce complex behavioral, neural and endocrine responses, including both fear and anxiety, which are highly conserved across species (Blanchard et al., 2001; Romero, 2004; Ulrich-Lai and Herman, 2009).
Threats or stressors activate brainstem nuclei, particularly the locus coeruleus (LC). The LC has widespread norepinephrine (NE) projections throughout the central nervous system (CNS) thought to primarily function to globally modulate behavior and arousal states. NE has myriad central functions including regulation of CNS cells and circuits (O’Donnell et al., 2012). The LC is the major producer of NE in the CNS and LC activation produces NE release throughout the cortex, acting as a single global regulator (O’Donnell et al., 2012). Tonic, continuous activity of the LC is low during sleep, intermediate during active wake and high in states of distress or anxiety (Atzori et al., 2016). Acute threats also engage the sympathetic nervous system for the behavioral ‘fight or flight’ response (hypothalamus–pituitary–adrenal (HPA) axis), which elevates circulating glucocorticoids for a coordinated physiological and behavioral response (Charmandari et al., 2005) (Figure 1). A ‘normal’ or adaptive response to threat can present as freezing or motor arrest, often used as a primary proxy of fear in rodents or somatomotor agitation or exertion. Both responses are coupled with increased vigilance and arousal, critical for the alerting, orienting and fear learning functions required in a dangerous or uncertain environment (Cardinal et al., 2002; Sara and Bouret, 2012). These partly reflexive responses are collectively referred to as ‘bottom-up’ responses and are critical for an adaptive response to a dynamic environment. Higher order frontal cortical systems provide ‘top-down’ regulation of these responses to blunt or regulate the response to threat if the environment is perceived to be safe (Bishop et al., 2004). While the recruitment of brainstem nuclei responsible for alerting and orienting is critical for a normal, adaptive response to threat, the excessive or overactive engagement of these structures is associated with a maladaptive threat response or a prolonged anxious state (Berridge and Waterhouse, 2003; Ulrich-Lai and Herman, 2009). Alongside this excessive bottom-up response, there can be also deficient cortical top-down regulation of subcortical and midbrain structures, leading to an inability to down-regulate the physiological and behavioral threat response.
Figure 1. Major locus coeruleus (LC) projections throughout the central nervous system play distinct functional roles.
Ascending LC projections innervate the hypothalamus for autonomic and endocrine regulation; the amygdala for salience detection and associative learning; the hippocampus to influence learning, memory and plasticity; and the cortex, for regulation of attention, arousal and the cognitive evaluation of pain. Descending LC projections (gray) reach the periaqueductal gray and other brainstem nuclei, as well as the spinal cord.
In rodents, pathological anxiety can be modeled as a variety of anxiety-like behaviors, each potentially modeling a component of human anxiety disorders. While animal models can capture certain phylogenetically conserved responses to stressors, such as risk aversion and reduced exploratory behaviors, the vast cognitive gap that exists between laboratory animals and humans limits the translation of a wide range of complex psychological characteristics of the human stress response or experience of pathological anxiety. Induction of anxiety-like behaviors in rodents, for example, via genetic manipulation, chronic repeated stress or with predator interactions, is not translatable to humans. Likewise, specific human-experienced stressors, including complex childhood trauma or neglect, financial or work stressors and extreme social judgment cannot be modeled in animals. Furthermore, key clinical features such as worry, rumination, intrusive thoughts, nightmares or catastrophising cannot be modeled in animals. Thus, although animal models can generate multiple translatable features, they do not reflect the full repertoire of symptoms that characterise a human anxiety disorder.
While our ability to measure these constructs differs between humans and rodents, rodent models can provide considerable insight into the pathophysiology of human disorders in some respects. Some primary behavioral anxiety-like examples include excessive fear-like behaviors, such as freezing and reduced social interaction (Rosen and Schulkin, 1998). More translational measures that can be captured by animal models and observed in humans include faster fear learning, reduced extinction learning, reduced exploratory behavior and increased risk aversion (Park and Moghaddam, 2017). Cognitive-behavioral human measures that mirror translational pre-clinical models include fear learning (Lissek, 2012), attentional bias (Shechner et al., 2012), as well as physiological or neural activity during anticipation of negatively valenced stimuli (Bishop et al., 2004). Other biological measures such as pupil dilation and 3-methoxy-4-hydroxyphenylglycol (MHPG) – a major metabolite of NE – can be measured in both humans (Murphy et al., 2014; Raskind et al., 1984; Southwick et al., 1993) and pre-clinical models (Aston-Jones and Cohen, 2005; Gilzenrat et al., 2010; Joshi et al., 2016; Korf et al., 1973a, 1973b). To date, there is predominantly indirect evidence of LC’s role in human pathological anxiety, however more recent work with ultra-high field 7-T magnetic resonance imaging (MRI) is enabling direct examination of the human LC in vivo (Morris et al., 2020; Priovoulos et al., 2018).
This review aims to synthesise the pre-clinical and clinical literature to date related to a hypothesised role of the LC in responses to acute and chronic threat, as well as the emergence of pathological anxiety. By first defining its role in critical cognitive processes, like attention, learning and memory, we aim to lay the groundwork for an understanding of its role in response to threat to inform dimensional (RDoC) mechanistic models across human anxiety disorders with the ultimate goal of improving treatments for these disabling conditions.

Afferents and efferents of the LC-NE system

The LC projects to myriad cortical, subcortical and brainstem nuclei to rapidly and globally modulate neural function (Bremner et al., 1996a, 1996b). It also receives widespread innervation. Extensive work characterising LC afferents and efferents has been largely conducted in animal models. Ascending LC-NE projections diverge into four bundles (Jones et al., 1977) to innervate (1) hypothalamus (Asakura et al., 2000; Jones et al., 1977), particularly the lateral portion including periventricular nucleus and supraoptic nucleus, important for autonomic and endocrine regulation (Jones et al., 1977); (2) ventral and central nucleus of the amygdala (Asakura et al., 2000; Jones et al., 1977), important for salience detection and associative learning (Campese et al., 2017; Chen et al., 1992; Sears et al., 2013); (3) the ‘diagonal band’, medial septum and hippocampus (Haring and Davis, 1985; Jones et al., 1977; Loughlin et al., 1986), that influences learning, memory and plasticity (Ehlers and Todd, 2017; Harley, 1987; Sara, 2009); and (4) the corpus callosum, reaching the cingulum and beyond throughout the cortex (Jones et al., 1977), for regulation of attention, arousal and the cognitive evaluation of pain (David Johnson, 2003; Sara and Bouret, 2012; Scherder et al., 2003; Willis and Westlund, 1997). LC-NE neuron lesions reduce NE in most of these regions, particularly hypothalamus and cerebral cortex (Neophytou et al., 2001). The LC also projects to the bed nucleus of the stria terminalis (BNST) (Asakura et al., 2000), cerebellum (Bremner et al., 1996b), lateral habenula (Purvis et al., 2018) and extensively to the olfactory bulb (Shipley et al., 1985). Descending LC projections pass into the medial forebrain bundle including the periacqueductal gray (PAG), tegmentum and raphe nuclei (Jones et al., 1977). Posterior-ventral LC projects through the length of the spinal cord (Jones and Yang, 1985; Loughlin et al., 1986) and targets parasympathetic neurons of the vagus dorsal motor nucleus (Westlund and Coulter, 1980).
Afferent inputs to the LC are less extensive than its efferents. Midbrain and brainstem projections to LC derive primarily from the ventrolateral (Ennis and Aston-Jones, 1986), rostral (Aston-Jones et al., 1991a) and dorsomedial medulla (Aston-Jones et al., 1991b) – regions that play a major role in the regulation of sympathetic control and behavioral orienting. It receives bidirectional inputs from the ventral tegmental area (VTA) (Deutch et al., 1986; Ornstein et al., 1987), which modulate depressive phenotypes (Isingrini et al., 2016; Weiss et al., 2005; Zhang et al., 2018, 2019), and from suprachiasmatic nucleus (SCN) (Legoratti-Sanchez et al., 1989) for circadian-based regulation of arousal (Aston-Jones et al., 2001). Other brainstem inputs to the LC derive from the hypothalamic paraventricular nucleus (PVN), PAG, raphe nuclei, as well as from the spinal cord, with limited inputs from neighboring nuclei to create a local circuit (Aston-Jones et al., 1991b; Cedarbaum and Aghajanian, 1978). Cortical and subcortical projections to LC originate from insula, central nucleus of the amygdala (Cedarbaum and Aghajanian, 1978), dorsolateral and dorsomedial prefrontal cortex (PFC) (Arnsten and Goldman-Rakic, 1984) and from prelimbic PFC – which may be indirect (Aston-Jones et al., 1991b; Jodo et al., 1998).
The widespread network of the LC system is therefore crucially positioned to play a significant role in modulating both ascending visceral feedback and descending cortical cognitive processing to mediate both psychological and physiological operations (Berntson et al., 2003). There is a growing consensus that distinct projections mediate distinct behaviors, suggesting that the LC is comprised of independent modules (Uematsu et al., 2017). Functionally distinct cell modules have been demonstrated to have specific anatomical projections with distinct functions (Hirschberg et al., 2017; Llorca-Torralba et al., 2019; Uematsu et al., 2015, 2017). For example, discrete projections can have opposite physiological and behavioral effects: amygdala projections can enable aversion learning, whereas PFC projections can enable extinction learning (Uematsu et al., 2017). Similarly, ascending and descending projections can have opposite functions: spinal projections can be analgesic and anti-nociceptive, whereas PFC projections can exacerbate pain responses (Hirschberg et al., 2017). In rodents, there appears to be a developmental genetic basis for some of these functional distinctions (Robertson et al., 2013, 2016). LC projections throughout the cortex are not homogeneous, and they show distinct biochemical and electrophysiological properties, governing varying levels of NE release (Chandler et al., 2014). Further work supports this model of modularity by indicating that LC cell populations are functionally distinct ensembles, since their spiking activity are largely asynchronous (Totah et al., 2018). Interestingly, strong aversive stimuli can cause a robust, unified LC-NE response across most LC cells (Uematsu et al., 2017), suggesting that the LC can provide both specific mediation of discrete behaviors and global mediation of general arousal. This highlights the nexus at which the LC operates and its crucial role as a modulator of highly conserved behavioral responses, discrete higher order cognitive functions and global arousal.

LC modulation of arousal, attention and memory formation

Arousal and attention

Adaptive responses to dynamic environments require intact functioning of attentional and memory-formation systems. Adept attentional direction is required for appropriate memory formation, and a correctly formed memory is required to guide appropriate attention direction. Both of these systems are under tight regulation by the LC, particularly in response to threat-related stimuli or events (Anisman et al., 2000; Ehlers and Todd, 2017; Sara and Bouret, 2012).
For normal attentional function, tonic firing of the LC in the range of 1–3 Hz is needed. Lower than normal activity is associated with hypoarousal and attention deficits, whereas higher tonic firing is associated with hyperarousal and anxious states (Howells et al., 2012). In the normal state and in the absence of threat, the tonic LC-NE system sustains vigilance and orienting functions (David Johnson, 2003). At an early, basic level, LC projections to vestibular nuclei mediate vestibulo-ocular and vestibulospinal reflexes for alerting and vigilance (Balaban, 2002, 2016; Peng et al., 2016). The ability of the LC to direct attention to a given cue seems unrelated to valence (Berridge and Waterhouse, 2003): the LC responds to all novel stimuli and mediates general attentional orienting (Sara and Bouret, 2012; Usher et al., 1999). The central LC-NE system therefore alerts or primes the organism in response to any significant external event (Svensson, 1982). Interestingly, LC lesions reduce exploratory behavior, but only in novel environments (Harro et al., 1995). This, coupled with evidence from electrophysiological studies in primates, suggests that the LC may play a major role in regulating the switch between goal-directed (exploit) and exploratory behaviors in novel environments (Usher et al., 1999).
The LC generally becomes activated in states of heightened vigilance, when a disruptive stimulus requires reorienting behavior (Aston-Jones et al., 1991a). Single unit recordings show that the LC responds to a stimulus predicting a noxious air puff in freely moving cats, but shows no activation during prediction of reward (Rasmussen and Jacobs, 1986). It has been suggested that phasic LC-NE activity acts as a global ‘interrupt’ function to orient attentional and cognitive processing to salient or, specifically, threatening situations (David Johnson, 2003) and its activity is closely related to cortical excitability for wide-scale behavioral and cognitive priming (Sara and Bouret, 2012). However, while the LC responds to all novel stimuli (Berridge and Waterhouse, 2003), and stimuli that require a response (Rajkowski et al., 2004), responses generally habituate over time if the stimulus is not aversive (Sara and Bouret, 2012), implicating a special function for threat. In support of this, LC-NE response to novel light is higher in rodents with greater fear-potentiated startle (Anisman et al., 2000). However, the frontal cortex has the ability to suppress LC activity over time (Sara and Herve-Minvielle, 1995).

Learning and memory

The LC plays a critical role in learning and memory formation, especially for threat-related learning. Multiple neural systems subserve distinct learning and memory processing, broadly including hippocampal context-dependent associative learning, stimulus discrimination, declarative memory and working memory (Olton et al., 1979; Sutherland and McDonald, 1990), amygdala-mediated affective or biologically significant incentive-based associative learning (Phelps, 2004; Sutherland and McDonald, 1990) and dorsal striatal reinforcement-based motor learning (McDonald and White, 1993). The LC-NE system seems to modulate each of these distinct learning and memory systems. First, LC-NE projections to hippocampus regulate long-term potentiation (LTP) and hippocampal plasticity (Harley, 1987), allowing arousal to influence learning (Sara, 2009), to engender subsequent attentional biases (Ehlers and Todd, 2017). Second, an intact direct functional LC-NE projection to the amygdala is necessary for Pavlovian threat learning (Sears et al., 2013) and aversive Pavlovian-to-instrumental transfer (Campese et al., 2017), and NE activity in the amygdala enhances passive-avoidance memory consolidation (Chen et al., 1992). Finally, there are fewer LC-NE terminals in the dorsal striatum, although the dorsal striatum shows high NE turnover rates and interactions between LC-NE neurons and the striatal dopaminergic system seems to mediate the behavioral effects of methamphetamine (Ferrucci et al., 2013; Fornai et al., 1996a, 1996b). Together these findings demonstrate that the LC can have a powerful effect on the regulation of multiple memory systems, including hippocampal plasticity for generation of threat-related attentional biases and amygdala-mediated associative learning for aversive events.

Role of LC in response to acute threat

Threatening environmental cues or events induce a coordinated response (Carrasco and Van de Kar, 2003) that is designed to heighten vigilance and prepare a rapid and flexible behavioral response. Threat or perceived stress increases NE release and HPA axis activation which induces hypothalamic corticotrophin-releasing factor (CRF) release and adrenal production of glucocorticoids including cortisol (Carrasco and Van de Kar, 2003; Makino et al., 2002). CRF and NE work together to promote the response to stress (Gresack and Risbrough, 2011) – inhibiting feeding, increasing blood pressure, stimulating adrenocorticotropic hormone (ACTH) and elevating sympathetic tone (Bailey et al., 2003; Laugero et al., 2001). Altogether these systems prepare the organism for an acute behavioral response. CRF also increases tonic LC firing rate and NE release (Asakura et al., 2000; Fan et al., 2009; Jedema and Grace, 2004), while NE also directly activates the HPA (Calogero et al., 1988), creating a feed-forward system important for anxiety pathogenesis (Owens et al., 1993).
While the role of the HPA axis has been well-defined as a key coordinating system that responds to threat or stress, a significant body of work underscores the critical role of the LC in this response too. It is clear that the LC rapidly responds to threatening stimuli. LC activity increases in monkeys (Grant and Redmond, 1984) and rats within 15 min (Silveira et al., 1993) and 30 min (Day et al., 2004; Sands et al., 2000) after an aversive or threatening stimulus. LC activation occurs following a range of threats or stressors, including the elevated plus maze (Silveira et al., 1993), acute and chronic-restraint stress (Sands et al., 2000), lipopolysaccharide (LPS)-induced sickness (Lacosta et al., 1999), forced swim and a single electric shock (Bruijnzeel et al., 2001), and the impact on increased LC excitability can be long lasting (Borodovitsyna et al., 2018). Threatening predator odor, an ecologically relevant stressor that elicits innate anxiety responses, induces activation of LC (Day et al., 2004; Hayley et al., 2001), BNST, PVN and PAG (Janitzky et al., 2015), leading to anti-predatory responses (Sobrinho and Canteras, 2011). Pain and acute noxious stimuli also activate central NE circuits in rodents (Kowalski et al., 2014) and in humans, as measured by pupilomotry (Chapman et al., 2014).
In humans, subliminal fear activates the LC, alongside higher cortically mediated orienting responses (Liddell et al., 2005). Anticipation of threat engages arousal and increases brainstem auditory evoked potentials (Baas et al., 2006) and pupil dilation (Clewett et al., 2018), both thought to be indirect measures of LC activation, although pupil dilation can be governed by other systems besides NE (Nelson and Mooney, 2016; Reimer et al., 2016). Finally, in humans, even psychological or perceived stress increased LC connectivity with amygdala (van Marle et al., 2010). The LC-NE system is therefore thought to govern a rapid warning response to stress (Lanius et al., 2017).

Other threat response systems

The LC is not, of course, the only neural threat response system. It is worth noting that other cortical and subcortical regions participate in a wider ‘threat circuit’, including amygdala and medial PFC, which are extensively reviewed elsewhere (Simpson et al., 2001; Taylor and Whalen, 2015) (Figure 2). Briefly, the amygdala has been most widely implicated in threat processing (Derntl et al., 2009; Harmer et al., 2006; Isenberg et al., 1999; Johansson and Hansen, 2002; Loughead et al., 2008; Oya et al., 2002), although it seems to serve a higher order, integrative threat learning function compared to the LC’s more rapid alerting function. The amygdala is a site of convergence of exteroceptive information from cortex and thalamus and visceral information from subcortex (Bremner et al., 1996a), where conditioned associations can be formed, activating learned fear responses (Cardinal et al., 2002). Indeed, there is evidence that the amygdala is recruited during early stages of fear learning (Bishop et al., 2007; Davidson, 2002). More generally salient events increase amygdala activation (both appetitive and aversive) (Fitzgerald et al., 2006) and the release of extrahypothalamic CRF, suggesting it drives attention to salient events rather than acting as a specific threat signal (Merali et al., 1998). The medial PFC (in humans, comprised of orbitofrontal cortex, ventomedial PFC, dorsomedial PFC and anterior cingulate cortex (ACC)), which has reciprocal projections with both the amygdala (Aggleton et al., 1980; Carmichael and Price, 1995; Ghashghaei and Barbas, 2002) and the LC, regulates and updates learned negative associations via extinction learning or re-learning new safe/neutral associations (Delgado et al., 2008; Milad et al., 2005; Milad and Quirk, 2002). Optimal emotion regulation requires concomitant activation of medial PFC and deactivation of amygdala (Delgado et al., 2008; Wager et al., 2008), which mitigates anxiogenesis (Bishop et al., 2004; Hare et al., 2008; Hariri et al., 2003; Pezawas et al., 2005; Simpson et al., 2001). In addition, the dorsolateral PFC also plays a regulatory role on medial PFC and amygdala threat or stress-related responses, reducing interference by negative emotion for adaptive cognitive control, important for resilience (Liston et al., 2009; Ochsner et al., 2004; Sinha et al., 2016). While these cortical regions respond to threat or stressors and downregulate the negative emotional and autonomic response to stress, acute threat or stress also has a detrimental effect on PFC function, via increased NE which engages alpha-1 adrenergic receptors to reduce PFC function (Arnsten, 2015), resulting in cortical atrophy, dendritic deterioration and overall reduction in cognitive control of emotion (Arnsten, 2009).
Figure 2. Central actions of the locus coeruleus (LC) in the regulation in threat reactivity.
(a) The adaptive response to acute threat involves a rapid, coordinated response in order to prepare the organism for an acute physiological response and behavioral activation. This involves hypothalamus–pituitary–adrenal (HPA) axis activation, hypothalamic corticotrophin-releasing factor (CRF) release and production of cortisol, which reaches body tissues via peripheral vasculature. Rapid LC activation and norepinephrine (NE) release also occurs, with NE targets throughout the cortex leading to global modulation of arousal and attention. Other NE targets in the amygdala (Amy) and medial prefrontal cortex (MPFC) mediate threat learning and reciprocally regulate the LC (red lines). (b) Chronic threat or stress leads to widespread changes in the central LC-mediated response to subsequent stressors. Chronic stress leads to LC hyper-activity, increased NE in LC, amygdala, hippocampus, MPFC and increased HPA axis activity via reduced HPA regulation. Excessive cortisol and NE relate to maladaptive physiological signs of hyper-arousal and reduced regulation of the LC by the MPFC (dotted red lines), which additionally contribute to reduced regulation of pathological anxiety. Gray lines indicate LC and HPA projections.

Role of LC in the development of pathological anxiety

Thus far, we have highlighted the ‘normal’ or adaptive role of the LC in response to acute threat. The following sections describe the hypothesised role of the LC in responses and processes that lead to maladaptive or pathological states.

Chronic stress

Chronic or repeated stress in rodents can be used as a model for human disorders of pathological anxiety and depression. These rodent studies demonstrate that the LC is involved in several different types of stress response. First, after prolonged restraint stress (30–60 min), LC activity (C-fos) increases (Keshavarzy et al., 2015). Second, chronic or repeated stress (corticosterone administration; Fan et al., 2014) increases tyrosine hydroxylase (TH) in the LC and norepinephrine transporter (NET) in the hippocampus, amygdala and PFC (Fan et al., 2014), leading to increased anxiety and defensive behaviors. After chronic long-term stress, not only does LC activity increases but also its subsequent sensitivity to stress increases. Chronic stress induces amplification of LC reactivity and increased NE release to subsequent stressors in rats (Jedema et al., 2001), possibly related to reduced LC auto-inhibition after stress (Jedema et al., 2008) and blunting of HPA axis regulation in a feedback-facilitation cycle (Makino et al., 2002).
In addition, in rat models of chronic stress, increased NE in the PFC causes further cortical atrophy and dendritic restructuring, resulting in reduced cognitive and attentional control (Liston et al., 2006, 2009). While it is probably too much to assume that the cellular observations made in rats directly translate to the human data, it is nevertheless important to recognise that chronic stress-induced plasticity in the PFC exists across species. These stress-induced alterations could be pathologically exacerbated in patients with PTSD (Bremner et al., 1997).

The emergence of anxiety-like behaviors in rodents

There is considerable evidence indicating a link between increased LC activity and the development of pathological anxiety-like behaviors in animals. For example, anxiety, fear and behavioral inactivation in rats are associated with increased LC activity (Kryzhanovskii et al., 1991). More causal evidence emerges from studies showing that LC activation produces greater anxiety and fear-like behavior (via neurokinin 1 receptors; Hahn and Bannon, 1999) in both rodents (Boulenger and Uhde, 1982) and monkeys (Bunney and Tallman, 1980). Furthermore, increased LC activity and TH expression is associated with the onset of anxiety-like behaviors in a rodent model of chronic pain (Alba-Delgado et al., 2013). This LC-mediated onset of anxiety behaviors is partly mediated via projections to amygdala (McCall et al., 2017) and amygdala CRF inputs to LC, which increases tonic LC activity that promotes anxiety-like behaviors in mice (Curtis et al., 2002; McCall et al., 2015; Van Bockstaele et al., 1998). The increased tonic LC activity appears to be sufficient to induce acute anxiety-like behavior following stress (McCall et al., 2015; Sciolino et al., 2016; Zerbi et al., 2019). Furthermore, transgenic mice with increased LC catecholaminergic neuron density have increased anxiety-like and panic behaviors (Dierssen et al., 2006).
LC lesions seem to reduce anxiety- or fear-like behavior (Boulenger and Uhde, 1982), increase fear extinction (Tsaltas et al., 1984) and do not affect appetitive sucrose conditioning (Tsaltas et al., 1984), indicating its specific role in threat processing. Reduction of alpha2A adrenoreceptor expression in the LC also reduces anxious behavior during the elevated plus maze in rats (Shishkina et al., 2002). Furthermore, blocking the LC increase in TH expression following stress reduces anxiety behaviors (Lee et al., 2012a, 2012b). The specificity of this effect is demonstrated by a studies showing that selective inhibition of LC-NE neurons during stress precludes generation of anxiety-like behaviors in rodents (McCall et al., 2015). Further work has shown that acute stress causes persistent increases in LC firing consistent with long-term expression of anxiety-like behavior in rats (Borodovitsyna et al., 2018). Together these rodent studies strongly indicate the critical role of the LC in anxiety pathogenesis.
There is, however, also contradictory evidence suggesting that reduced LC activity is associated with anxiety. Abolishing LC activity with desipramine increased anxiety behaviors (immobility) (Weiss et al., 1994), and destruction of LC terminals increased anxiety-like behavior in the form of reduced exploration of novel environments (Itoi et al., 2011; Kask et al., 2000) in rodents. This inconsistency in findings of reduced LC activity associated with anxiety may be explained by the specific behavioral measure affected by LC-NE lesions: immobility and reduced exploration. Evidence indicates that LC-NE lesions alter freezing time in rats without affecting the initial locomotor (running and jumping) response to conditioned and unconditioned aversive stimuli, indicating LC lesions may relate to defense rather than behavioral activation for aversive avoidance (Neophytou et al., 2001). However, while the LC has a global function in mediating arousal to strongly aversive stimuli, distinct sub-modules can have distinct and opposite functional roles in mediating learning or responses to aversive stimuli (Hirschberg et al., 2017; Llorca-Torralba et al., 2019; Uematsu et al., 2015, 2017). The distinct roles of the LC-NE system in active versus passive aversive avoidance require further study.

Role of the LC in risk factors for pathological anxiety

Insight into the role of the LC in pathological anxiety can be discerned via examining risk factors for the development of pathological anxiety. For examples, Wistar Kyoto rodents are bred to be susceptible to certain types of stress and exhibit anxiety-like behaviors with excessive responses to stressors. These rats show reduced inhibitory control of the LC (less sensitive alpha-2 adregnergic receptors in LC and reduced inhibitory GABA input), implicating a shift toward enhanced excitatory capacity of the LC as a key mechanism of anxiety (Bruzos-Cidon et al., 2015). Rodents bred to exhibit high anxiety behaviors also show increased activity of the LC, PAG, hypothalamus, as well as reduced activity of the ACC after stress (Salchner et al., 2006). Furthermore, transgenic rats with reduced glial angiotensinogen and enhanced anxiety behaviors show coincident higher LC activity and increased locomotor response to novelty (Ogier et al., 2016). Finally, rat offspring from stressed dams, a risk factor for susceptibility to anxiety, show higher fear responses and corticosterone reactivity, alongside increased LC, amygdala and striatal reactivity and reduced medial PFC reactivity to stress (Sadler et al., 2011). Together, these findings in rodents highlight the role of genetic and prenatal environmental factors that modulate LC and frontal cortical reactivity to stressors. These factors pose clear risk for exaggerated anxiety responses in rodents. Interestingly, in seeming contradiction to these findings, Maudsley non-reactive rats (which are resilient to anxiety) show higher basal LC neuronal activity and a burst-like pattern of firing (Verbanac et al., 1994). More work needs to be done to tease apart the contributions of tonic versus phasic LC firing and specific modular regulation of activity.

Biological sex

Biological sex can also act as a risk factor for pathological anxiety and implicate LC as a differentiating factor in mediating anxiety behaviors. Female rats have larger and more complex LC than males (Bangasser et al., 2011; Valentino et al., 2012) and the constitution of receptor sites on LC structures also seems to differ between sexes. CRF receptors in the cortex (Bangasser et al., 2010) and LC (Bangasser, 2013) are differentially expressed between sexes, making females more sensitive to CRF. Higher female sensitivity to CRF in the LC is linked with more hyperarousal and symptoms of pathological anxiety (Bangasser, 2013). Interestingly, selective LC-NE glucocorticoid receptor ablation in female, but not male rats, results in heightened anxiety-like behaviors (Chmielarz et al., 2013). Chronic alcohol administration also induces anxiogenesis with increased LC activity and in female but not male rats (Retson et al., 2015). Finally, early life stress increases astrocyte function in the LC which leads to anxiety symptoms in female, but not male mice (Nakamoto et al., 2017). Together, these findings indicate that heightened vulnerability to stressors in females may be in part governed by specific differences in LC structural and molecular composition, rendering females more susceptible to pathological anxiety.

Inflammation

Finally, inflammation may mediate the link between LC function and anxiety behaviors. For example, LPS injected into the LC increases astrocyte function and anxiety-like behavior in mice (Nakamoto et al., 2017). Inflammation can also serve as a high-risk state for the development of pathological anxiety. Monoarthritis (a hyper-inflammatory state in rodents) is associated with an anxiety-like phenotype, with increased activation in LC and PFC (Borges et al., 2014) and systemic interleukin-2 injections increases NE in many LC target regions (Lacosta et al., 2000). In a rodent model of anorexia, increased inflammation coincides with increased LC activity and anxiety behaviors (Scharner et al., 2017). Conversely, reducing neuro-inflammation can reduce LC damage and anxiety-like behaviors in a mouse model of Alzheimer’s disease (Braun and Feinstein, 2017). The LC itself seems to have a regulatory, anti-inflammatory effect, acting as a neurotrophic and neuroprotective modulator in a normal state (Feinstein et al., 2016; Lee et al., 2016; Wang et al., 2015). However, the LC’s anti-inflammatory and neuroprotective capacity is reduced following stress (Lee et al., 2016), indicating one mechanism by which the LC mediates the normal and abnormal response to stressors – via neuroimmune regulation.

Role of LC in human patients with pathological anxiety

Patients with pathological anxiety experience clinical symptoms and cognitive disturbances that point toward an underlying disturbance in LC function. Panic disorder is characterised by the recurrent unexpected onset of ‘panic attacks’, a sudden and rapid state of intense fear or sympathetic nervous system arousal, in the absence of any environmental threat, substance or other provoking disorder (American Psychiatric Association, 2013). Individuals with panic disorder also fear future attacks and avoid situations that might trigger an attack, such as crowded public transport. GAD is a condition of excessive, non-specific anxiety or worry, often coupled with restlessness, irritability or fatigue, as well as concentration difficulties (American Psychiatric Association, 2013). SAD is more specific than GAD, characterised by a persistent, excessive fear of social or performance situations in particular, including maladaptive worry about social scrutiny and embarrassment (American Psychiatric Association, 2013). PTSD, previously classified as an anxiety disorder and now classified as a trauma- and stressor-related disorder (Friedman, 2013), can develop in individuals exposed to event(s) that threatened death or serious injury, in which the trauma is re-experienced via nightmares, flashbacks or unwanted memories (American Psychiatric Association, 2013). Patients with PTSD experience negative alterations in mood and cognition, as well as states of heightened physiological arousal and exaggerated startle responses, alongside more general difficulties with concentration and exaggerated avoidance behaviors (American Psychiatric Association, 2013).
Early evidence indicated that patients with PTSD and panic disorder, given yohimbine (an α2 adrenergic receptor antagonist) or CO2 (increases LC-NE), show increased anxiety and panic symptoms (Charney et al., 1984; Gorman et al., 1984; Southwick et al., 1993), implicating a role for LC-NE in anxiety symptoms in humans. PTSD has since been conceptualised as a disorder stemming from conscious and subconscious hyper-response to threat, associated with hyperarousal and a hyper-active ‘alarm system’ neural response including from the LC, amygdala and PFC (Lanius et al., 2017). In support of this model, PTSD patients show more exaggerated heart-rate responses, skin conductance, eye blink responses and LC blood-oxygen-level-dependent (BOLD) activation to loud sounds (Naegeli et al., 2018), as well as higher LC and insula BOLD responses to fearful stimuli (Morey et al., 2015) compared to trauma-exposed controls. There is also evidence linking PTSD with distorted fear learning governed by overactive of LC and insula to fearful stimuli and with increased LC connectivity with amygdala, striatum and insula during direct threatening eye gaze (Steuwe et al., 2015). Contradictory to this model is evidence that patients with PTSD show reduced LC size and reduced LC-NE reuptake availability (Arango et al., 1996; Bracha et al., 2005). However, this may be explained by differences in tonic versus phasic LC responses. For example, there is evidence for increased NE activity in PTSD in response to stressors but not during rest (Bremner et al., 1999). Finally, there is evidence for reduced PFC-mediated cognitive–emotional control during stress in PTSD: traumatic images invoke reduced medial PFC and ACC area 24 blood flow (positron emission tomography (PET)) in patients (Bremner et al., 1999) and patients have reduced ACC response when exposed to emotional conflict (Kim et al., 2008), implicating a mechanism of heightened LC reactivity to threat alongside blunted cortical regulation.
Studies implicating LC in other human disorders of pathological anxiety are less numerous than those in PTSD. LC dysfunction has been implicated in SAD (Marazziti et al., 2015), and worry in GAD has been associated with reduced heart-rate variability (indicating parasympathetic withdrawal) and increased LC–amygdala connectivity (Meeten et al., 2016). The amygdala has been more widely studied in these disorders. For example, phobia is associated with enhanced amygdala activation to threat (Bertolino et al., 2005), suggested to be involved in vigilance-avoidance processing (Larson et al., 2006), which reduces after therapy (Alpers et al., 2009; Goossens et al., 2007). Panic disorder has also been associated with the ‘extended fear network’ including brainstem, cingulate, insula, PFC and amygdala (Sobanski and Wagner, 2017), which cause both physiological and psychological (threat anticipation) symptoms (Windmann, 1998). Other cortical and subcortical regions implicated in human pathological anxiety are more extensively reviewed elsewhere (Simpson et al., 2001; Taylor and Whalen, 2015).

Anxiolytics

Clinical pharmacology on known anxiolytics is another area of research that implicates the function of the LC-NE system in human pathological anxiety. The alpha-1 adrenergic receptor antagonist, prazosin, is effective for treating PTSD (Koola et al., 2014; Raskind et al., 2000) and stress-induced craving in alcohol dependence (Fox et al., 2012). Beta-adrenergic receptors are also necessary for the development of anxiety-like behaviors (Gorman and Dunn, 1993; Wohleb et al., 2011) and regulate the induction of stress-induced gene expression in the brain in mice (Roszkowski et al., 2016). Beta-adrenergic receptor antagonism (‘beta-blockers’) prevents the development of anxiety-like behaviors in mice (Gorman and Dunn, 1993) and humans (Jefferson, 1974), particularly for somatic symptoms (Harris and Aston-Jones, 1993; Hayes and Schulz, 1987; Kelly, 1980; Noyes, 1982). While there is limited evidence for the efficacy of beta-blockers for treating PTSD (Amos et al., 2014) and performance anxiety in healthy individuals (Liebowitz et al., 1985), more recent studies indicate a novel pathway for treating anxiety disorders with beta-blockers. First described by Nader et al. (2000) in 2000, memory reconsolidation is a process whereby memories that are re-activated become labile and vulnerable to manipulation. Several studies have now demonstrated that distribution of traumatic memory reconsolidation with beta-blockers is feasible in humans and effective at relieving PTSD symptoms (Brunet et al., 2008; Evers, 2007; Kindt et al., 2014), although wider replication is currently lacking (Wood et al., 2015). While this novel direction for treatment of pathological anxiety in humans is promising, further work is needed to more directly target LC-NE system dysfunction with pharmacotherapies. One small study of three panic disorder patients showed that inhibition of the LC with the alpha2 agonist, clonidine, resulted in reduced panic and anxiety symptoms (Valenca et al., 2004).
While evidence for targeting LC-NE dysfunction for the treatment of pathological anxiety in humans is largely indirect, more direct evidence from pre-clinical studies implicates modulation of the LC-NE system function as a direct target for a variety of anxiolytics.
First, endogenous neuropeptide Y (NPY) is anxiolytic and reduces the behavioral responses to stress (Desai et al., 2014; Eaton et al., 2007; Kask et al., 2002). A network including LC, amygdala, PAG and hippocampus seems to mediate the anxiolytic effects of NPY (Heilig, 2004; Kask et al., 2002). More specifically, NPY given directly to the LC reduces anxiety and increases exploratory behaviors in rats (Kask et al., 1998) and anxiety-like behavior produced by LC terminal destruction can be attenuated by NPY administration (Kask et al., 2000). NPY given immediately before (Sabban et al., 2015a; Serova et al., 2013) or after (Sabban et al., 2015b) stress (forced swim, elevated plus maze, LC activation) reduces stress-induced physiological and behavioral manifestations of anxiety including ACTH, corticosterone and LC-TH expression (Sayed et al., 2018, #1; Serova et al., 2013}. Interestingly, stress causes a reduction of NPY in LC, nucleus accumbens (Nac) and PVN (Desai et al., 2014), indicating a mechanism by which stress leads to reduced resilience against anxiogenesis.
Benzodiazepines, which are commonly used for anxiolysis, bind to LC neurons (Hellsten et al., 2010) and reduce LC neuronal activity (Soderpalm and Engel, 1989), stress-induced increase in NE (Gray, 1996; Ida et al., 1985; Tanaka et al., 2000), CRF (Skelton et al., 2000) and the neuroendocrine responses to stress (Carrasco and Van de Kar, 2003). Diazepam has been shown to reduce LC responses to negative stimuli (Rasmussen and Jacobs, 1986). In contrast, non-BZ anxiolytics increase LC activity (Sanghera and German, 1983; Trulson and Henderson, 1984).
Chronic selective serotonin reuptake inhibitors (SSRIs), which are antidepressant and anxiolytic, reduce LC-NE neuron firing in control (Szabo and Blier, 2002) and perinatal-protein deprived rats (Sodero et al., 2004). SSRIs seem to reduce LC firing at a rate consistent with their therapeutic effects (Szabo et al., 2000). The SSRI fluoxetine also reduces glucocorticoid receptor expression in LC (Heydendael and Jacobson, 2010). Another non-SSRI antidepressant, imipramine, also reduces stress-induced LC activation (de Medeiros et al., 2005). Selective serotonin-norepinephrine reuptake inhibitors (SNRIs) such as milnacipran, which has also shown to have anxiolytic effects (Dell’Osso et al., 2010), may also activate LC to release 5HT for anxiolysis (Bourin et al., 2005).
Finally, exercise has been shown to reduce anxiety and stress-related markers, and increase galanin in LC (Salim et al., 2010; Sciolino and Holmes, 2012; Sciolino et al., 2015). Exercise or galanin has been shown to reduce stress-induced anxiety behaviors (Sciolino et al., 2015). Chronic galanin antagonism blocked the resilience-inducing influence of exercise (Sciolino et al., 2015).

Future directions in approaches to studying the LC in humans

Emerging advances in neuroimaging technology and computational modeling are starting to highlight pathways by which the LC can be studied in humans in vivo. Advances in MRI acquisition protocols, resolution afforded by higher field MRI and data denoising strategies will allow the investigation of LC structure and function with enhanced spatial and temporal precision.
The past few years have provided in vivo characterisation of the LC in awake humans that reliably correlates to post-mortem analyses of relative LC size, cell distribution, location, age-related size alteration and disease-specific structural changes, notably for Alzheimer’s disease (Chan-Palay and Asan, 1989; Kelly et al., 2017; Theofilas et al., 2017). The T1-weighted turbo spin echo (TSE) technique is the current gold standard of LC structural imaging in humans, offering contrast in the LC due to the presence of neuromelanin (NM), which is MR-visible due to a magnetisation transfer (MT) contrast mechanism. NM contrast has been shown to be a reliable indirect measure of the number of LC cells – providing the basis for in vivo studies of LC microstructure (Betts et al., 2019; Clewett et al., 2016). The metabolic activity of the LC has also been captured with functional MRI, albeit with large voxel sizes. Validation of LC functional MRI (fMRI) has come from reference to simultaneously acquired pupilometry, showing that pupil diameter and LC BOLD activation are tightly correlated (Alnaes et al., 2014; Elman et al., 2017; Murphy et al., 2014). These advances have demonstrated the rapid improvement in fMRI resolution and sensitivity. Nonetheless, while current structural imaging of the LC uses high in-plane resolution (~0.4 mm), it also has large slice thickness (~3 mm). This means that while LC can be localised in the brainstem, the characterisation of its size and shape is still not optimised. Similarly, current functional MRI of LC involves image acquisition at standard voxel sizes of ~3 mm to attain good functional sensitivity. Recent work using high-field MRI and computational segmentation algorithms has been used to measure LC structure and volume with sub-millimeter resolution (Morris et al., 2020). This was successfully performed in human subjects with and without pathological anxiety, demonstrating enlarged LC in patients, associated with poorer attentional control and higher anxious arousal (Morris et al., 2020). Further applications of this work in larger sample sizes and using fMRI will be critical for the translation of the wealth of pre-clinical work described, to clinical settings.
Recent advances in computational modeling approaches are also providing new insights into LC function. Several computational frameworks suggest two underlying modes of LC function. One mode in which tonic LC activity mediates exploration via an increase in gain in sensory representations and general arousal, and the second mode in which phasic LC activity optimises behavior in light of current task performance, recruiting insula-based salience detection and PFC decision-making systems (Aston-Jones and Cohen, 2005; Dowman et al., 2016). Optimal task performance and learning requires an accurate representation of certainty in the environment, which is suggested to be computed by NE and acetylcholine (Yu and Dayan, 2005), both of which modulate general arousal-related pupil dilation (Larsen and Waters, 2018). Indeed, pupil dilation has been linked with the reliability or stability of information in the environment (Nassar et al., 2012), critical for optimal environmental navigation. However, this link between pupil sensitivity and optimal computation of environmental features in an unstable world seems to be disturbed in patients with high trait anxiety (Browning et al., 2015). Together, these computational models provide new and potentially integrative insights into LC and NE function, and their potential disturbance in human disorders of pathological anxiety. Harnessing computational models based on known physiological and biological systems will enable development and understanding of small and large-scale networks that ultimately direct behavior.

Summary

The LC is critically placed to modulate both ascending visceral feedback and descending cortical cognitive processing to mediate both psychological and physiological operations (Berntson et al., 2003). While threat or perceived stress broadly increases sympathetic tone and HPA axis activation (Carrasco and Van de Kar, 2003; Makino et al., 2002), it also recruits LC activity and NE release throughout the CNS. Direction of attention and memory formation, particularly for threat-related stimuli and events, are both under tight regulation by the LC. Indeed the LC responds rapidly to a range of threatening stimuli, including even the mere anticipation of threat (Baas et al., 2006; Clewett et al., 2018), subliminal fear (Liddell et al., 2005) or perceived stress (van Marle et al., 2010) and the broader LC-NE system governs an ‘alarm system’ response to stress (Lanius et al., 2017) across species.
Moreover, chronic or repeated stress (Fan et al., 2014) increases LC, hippocampal, amygdala and PFC activity (Fan et al., 2014) and leads to anxiety and defensive behaviors (Li et al., 2018). After chronic stress, not only does LC activity increase but its sensitivity to subsequent stressors also increases (Jedema et al., 2001), possibly due to reduced LC auto-inhibition after stress (Jedema et al., 2008). Resultant increased tonic LC activity and LC catecholaminergic neuron density is sufficient to lead to anxiety and panic in rodents (Dierssen et al., 2006; McCall et al., 2015). This coincides with human evidence of hyperarousal and exaggerated physiological responses to threat centered on LC (Lanius et al., 2017; Morey et al., 2015; Naegeli et al., 2018). Together this suggests a mechanism by which repeated stress can lead to LC dysregulation and therefore maladaptive exaggerated fear or pathological anxiety responses.
Future innovation in human neuroimaging and neural circuit modeling will advance the investigation of LC structure and function in vivo, allowing greatly enhanced precision for characterising the LC in humans.

Declaration of conflicting interests

The author(s) declared the following potential conflicts of interest with respect to the research, authorship and/or publication of this article: In the past 5 years, Dr. Murrough has provided consultation services and/or served on advisory boards for Allergan, Boehreinger Ingelheim, Clexio Biosciences, Fortress Biotech, FSV7, Global Medical Education (GME), Impel Neuropharma, Janssen Research and Development, Medavante-Prophase, Novartis, Otsuka and Sage Therapeutics. In the past 12 months, Dr. Murrough has provided consultation services and/or served on advisory boards for Boehreinger Ingelheim, Clexio Biosciences, Global Medical Education (GME) and Otsuka. Dr. Murrough is named on a patent pending for neuropeptide Y as a treatment for mood and anxiety disorders and on a patent pending for the use of ezogabine and other KCNQ channel openers to treat depression and related conditions. The Icahn School of Medicine (employer of Dr. Murrough) is named on a patent and has entered into a licensing agreement and will receive payments related to the use of ketamine or esketamine for the treatment of depression. The Icahn School of Medicine is also named on a patent related to the use of ketamine for the treatment of PTSD. Dr. Murrough is not named on these patents and will not receive any payments. Dr. Charney is named as co-inventor on patents filed by the Icahn School of Medicine at Mount Sinai (ISMMS) relating to the treatment for treatment-resistant depression, suicidal ideation and other disorders. ISMMS has entered into a licensing agreement with Janssen Pharmaceuticals, Inc. and it has and will receive payments from Janssen under the license agreement related to these patents for the treatment of treatment-resistant depression and suicidal ideation. Consistent with the ISMMS Faculty Handbook (the medical school policy), Dr. Charney is entitled to a portion of the payments received by the ISMMS. Since SPRAVATO has received regulatory approval for treatment-resistant depression, ISMMS and thus, through the ISMMS, Dr. Charney will be entitled to additional payments, beyond those already received, under the license agreement. Dr. Charney is a named co-inventor on several patents filed by ISMMS for a cognitive training intervention to treat depression and related psychiatric disorders. The ISMMS has entered into a licensing agreement with Click Therapeutics, Inc. and has and will receive payments related to the use of this cognitive training intervention for the treatment of psychiatric disorders. In accordance with the ISMMS Faculty Handbook, Dr. Charney has received a portion of these payments and is entitled to a portion of any additional payments that the medical school might receive from this license with Click Therapeutics. Dr. Charney is a named co-inventor on a patent application filed by the ISMMS for the use of intranasally administered Neuropeptide Y (NPY) for the treatment of mood and anxiety disorders. This intellectual property has not been licensed. Dr. Charney is a named co-inventor on a patent application in the United States, and several issued patents outside the United States filed by the ISMMS related to the use of ketamine for the treatment of PTSD. This intellectual property has not been licensed. The remaining authors disclose no conflicts of interest.

Funding

The author(s) disclosed receipt of the following financial support for the research, authorship and/or publication of this article: Funding for this work was provided by the National Institute of Mental Health (NIMH; grant no. R01MH116953, PI: Murrough) of the National Institutes of Health (NIH). The content is solely the responsibility of the authors and does not necessarily represent the official views of NIH or NIMH. Additional funding was provided by the Ehrenkranz Laboratory for Human Resilience and the Friedman Brain Institute, both components of the Icahn School of Medicine at Mount Sinai.

ORCID iD

References

Aggleton JP, Burton MJ, Passingham RE (1980) Cortical and subcortical afferents to the amygdala of the rhesus monkey (Macaca mulatta). Brain Research 190(2): 347–368.
Alba-Delgado C, Llorca-Torralba M, Horrillo I, et al. (2013) Chronic pain leads to concomitant noradrenergic impairment and mood disorders. Biological Psychiatry 73(1): 54–62.
Alnaes D, Sneve MH, Espeseth T, et al. (2014) Pupil size signals mental effort deployed during multiple object tracking and predicts brain activity in the dorsal attention network and the locus coeruleus. Journal of Vision 14(4): 1.
Alpers GW, Gerdes AB, Lagarie B, et al. (2009) Attention and amygdala activity: An fMRI study with spider pictures in spider phobia. Journal of Neural Transmission (Vienna) 116(6): 747–757.
American Psychiatric Association (2013) Diagnostic and Statistical Manual of Mental Disorders (5th edn). Arlington, VA: American Psychiatric Publishing.
Amos T, Stein DJ, Ipser JC (2014) Pharmacological interventions for preventing post-traumatic stress disorder (PTSD). Cochrane Database of Systematic Reviews 8(7): CD006239.
Anisman H, Kelly O, Hayley S, et al. (2000) Acoustic startle and fear-potentiated startle in rats selectively bred for fast and slow kindling rates: Relation to monoamine activity. European Journal of Neuroscience 12(12): 4405–4416.
Arango V, Underwood MD, Mann JJ (1996) Fewer pigmented locus coeruleus neurons in suicide victims: Preliminary results. Biological Psychiatry 39(2): 112–120.
Arnsten A (2009) Stress signalling pathways that impair prefrontal cortex structure and function. Nature Reviews Neuroscience 10(6): 410–422.
Arnsten AF (2015) Stress weakens prefrontal networks: Molecular insults to higher cognition. Nature Neuroscience 18(10): 1376–1385.
Arnsten AF, Goldman-Rakic PS (1984) Selective prefrontal cortical projections to the region of the locus coeruleus and raphe nuclei in the rhesus monkey. Brain Research 306(1–2): 9–18.
Asakura M, Nagashima H, Fujii S, et al. (2000) [Influences of chronic stress on central nervous systems]. Nihon Shinkei Seishin Yakurigaku Zasshi 20(3): 97–105.
Aston-Jones G, Cohen JD (2005) An integrative theory of locus coeruleus-norepinephrine function: Adaptive gain and optimal performance. Annu Rev Neurosci 28: 403–450.
Aston-Jones G, Chen S, Zhu Y, et al. (2001) A neural circuit for circadian regulation of arousal. Nature Neuroscience 4(7): 732–738.
Aston-Jones G, Chiang C, Alexinsky T (1991a) Discharge of noradrenergic locus coeruleus neurons in behaving rats and monkeys suggests a role in vigilance. Progress in Brain Research 88: 501–520.
Aston-Jones G, Shipley MT, Chouvet G, et al. (1991b) Afferent regulation of locus coeruleus neurons: Anatomy, physiology and pharmacology. Progress in Brain Research 88: 47–75.
Atzori M, Cuevas-Olguin R, Esquivel-Rendon E, et al. (2016) Locus ceruleus norepinephrine release: A central regulator of CNS spatio-temporal activation? Frontiers in Synaptic Neuroscience 8: 25.
Baas JM, Milstein J, Donlevy M, et al. (2006) Brainstem correlates of defensive states in humans. Biological Psychiatry 59(7): 588–593.
Bailey JE, Argyropoulos SV, Lightman SL, et al. (2003) Does the brain noradrenaline network mediate the effects of the CO2 challenge? Journal of Psychopharmacology 17(3): 252–259.
Balaban CD (2002) Neural substrates linking balance control and anxiety. Physiology & Behavior 77(4–5): 469–475.
Balaban CD (2016) Neurotransmitters in the vestibular system. Handbook of Clinical Neurology 137: 41–55.
Bangasser DA (2013) Sex differences in stress-related receptors: ‘Micro’ differences with ‘macro’ implications for mood and anxiety disorders. Biology of Sex Differences 4(1): 2.
Bangasser DA, Curtis A, Reyes BA, et al. (2010) Sex differences in corticotropin-releasing factor receptor signaling and trafficking: Potential role in female vulnerability to stress-related psychopathology. Molecular Psychiatry 15(19): 877896–877904.
Bangasser DA, Zhang X, Garachh V, et al. (2011) Sexual dimorphism in locus coeruleus dendritic morphology: A structural basis for sex differences in emotional arousal. Physiology & Behavior 103(3–4): 342–351.
Berntson GG, Sarter M, Cacioppo JT (2003) Ascending visceral regulation of cortical affective information processing. European Journal of Neuroscience 18(8): 2103–2109.
Berridge CW, Waterhouse BD (2003) The locus coeruleus-noradrenergic system: Modulation of behavioral state and state-dependent cognitive processes. Brain Research Reviews 42(1): 33–84.
Bertolino A, Arciero G, Rubino V, et al. (2005) Variation of human amygdala response during threatening stimuli as a function of 5’HTTLPR genotype and personality style. Biological Psychiatry 57(12): 1517–1525.
Betts MJ, Kirilina E, Otaduy MCG, et al. (2019) Locus coeruleus imaging as a biomarker for noradrenergic dysfunction in neurodegenerative diseases. Brain 142(9): 2558–2571.
Bishop S, Duncan J, Brett M, et al. (2004) Prefrontal cortical function and anxiety: Controlling attention to threat-related stimuli. Nature Neuroscience 7(2): 184–188.
Bishop SJ, Jenkins R, Lawrence AD (2007) Neural processing of fearful faces: Effects of anxiety are gated by perceptual capacity limitations. Cerebral Cortex 17(7): 1595–1603.
Blanchard RJ, McKittrick CR, Blanchard DC (2001) Animal models of social stress: Effects on behavior and brain neurochemical systems. Physiology & Behavior 73(3): 261–271.
Borges G, Neto F, Mico JA, et al. (2014) Reversal of monoarthritis-induced affective disorders by diclofenac in rats. Anesthesiology 120(6): 1476–1490.
Borodovitsyna O, Flamini MD, Chandler DJ (2018) Acute stress persistently alters locus coeruleus function and anxiety-like behavior in adolescent rats. Neuroscience 373: 7–19.
Boulenger JP, Uhde TW (1982) Biological peripheral correlates of anxiety. Encephale 8(2): 119–130.
Bourin M, Masse F, Dailly E, et al. (2005) Anxiolytic-like effect of milnacipran in the four-plate test in mice: Mechanism of action. Pharmacology Biochemistry and Behavior 81(3): 645–656.
Bracha HS, Garcia-Rill E, Mrak RE, et al. (2005) Postmortem locus coeruleus neuron count in three American veterans with probable or possible war-related PTSD. Journal of Neuropsychiatry and Clinical Neurosciences 17(4): 503–509.
Braun D, Feinstein DL (2017) The locus coeruleus neuroprotective drug vindeburnol normalizes behavior in the 5xFAD transgenic mouse model of Alzheimer’s disease. Brain Research 1702: 29–37.
Bremner JD, Innis RB, Ng CK, et al. (1997) Positron emission tomography measurement of cerebral metabolic correlates of yohimbine administration in combat-related posttraumatic stress disorder. Archives of General Psychiatry 54(3): 246–254.
Bremner JD, Krystal JH, Southwick SM, et al. (1996a) Noradrenergic mechanisms in stress and anxiety: I: Preclinical studies. Synapse 23(1): 28–38.
Bremner JD, Krystal JH, Southwick SM, et al. (1996b) Noradrenergic mechanisms in stress and anxiety: II: Clinical studies. Synapse 23(1): 39–51.
Bremner JD, Staib LH, Kaloupek D, et al. (1999) Neural correlates of exposure to traumatic pictures and sound in Vietnam combat veterans with and without posttraumatic stress disorder: A positron emission tomography study. Biological Psychiatry 45(7): 806–816.
Browning M, Behrens TE, Jocham G, et al. (2015) Anxious individuals have difficulty learning the causal statistics of aversive environments. Nature Neuroscience 18(4): 590–596.
Bruijnzeel AW, Stam R, Compaan JC, et al. (2001) Stress-induced sensitization of CRH-ir but not P-CREB-ir responsivity in the rat central nervous system. Brain Research 908(2): 187–196.
Brunet A, Orr SP, Tremblay J, et al. (2008) Effect of post-retrieval propranolol on psychophysiologic responding during subsequent script-driven traumatic imagery in post-traumatic stress disorder. Journal of Psychiatric Research 42(6): 503–506.
Bruzos-Cidon C, Llamosas N, Ugedo L, et al. (2015) Dysfunctional inhibitory mechanisms in locus coeruleus neurons of the Wistar Kyoto rat. International Journal of Neuropsychopharmacology 18(7): pyu122.
Bunney WE Jr, Tallman JF (1980) New biological research relevant to anxiety. Pharmakopsychiatrie, Neuro-Psychopharmakologie 13(5): 273–276.
Calogero AE, Bernardini R, Gold PW, et al. (1988) Regulation of rat hypothalamic corticotropin-releasing hormone secretion in vitro: Potential clinical implications. Advances in Experimental Medicine and Biology 245(3): 167–181.
Campese VD, Soroeta JM, Vazey EM, et al. (2017) Noradrenergic regulation of central amygdala in aversive Pavlovian-to-instrumental transfer. eNeuro 4(5): 0224.
Cardinal RN, Parkinson JA, Hall J, et al. (2002) Emotion and motivation: The role of the amygdala, ventral striatum, and prefrontal cortex. Neuroscience & Biobehavioral Reviews 26(3): 321–352.
Carmichael ST, Price JL (1995) Limbic connections of the orbital and medial prefrontal cortex in macaque monkeys. Journal of Comparative Neurology 363(4): 615–641.
Carrasco GA, Van de Kar LD (2003) Neuroendocrine pharmacology of stress. European Journal of Pharmacology 463(1): 235–272.
Cedarbaum JM, Aghajanian GK (1978) Afferent projections to the rat locus coeruleus as determined by a retrograde tracing technique. Journal of Comparative Neurology 178(1): 1–16.
Chandler DJ, Gao WJ, Waterhouse BD (2014) Heterogeneous organization of the locus coeruleus projections to prefrontal and motor cortices. Proceedings of the National Academy of Sciences of the United States of America 111(18): 6816–6821.
Chan-Palay V, Asan E (1989) Alterations in catecholamine neurons of the locus coeruleus in senile dementia of the Alzheimer type and in Parkinson’s disease with and without dementia and depression. Journal of Comparative Neurology 287(3): 373–392.
Chapman CR, Bradshaw DH, Donaldson GW, et al. (2014) Central noradrenergic mechanisms and the acute stress response during painful stimulation. Journal of Psychopharmacology 28(12): 1135–1142.
Charmandari E, Tsigos C, Chrousos G (2005) Endocrinology of the stress response. Annual Review of Physiology 67(1): 259–284.
Charney DS, Heninger GR, Breier A (1984) Noradrenergic function in panic anxiety: Effects of yohimbine in healthy subjects and patients with agoraphobia and panic disorder. Archives of General Psychiatry 41(8): 751–763.
Chen MF, Chiu TH, Lee EH (1992) Noradrenergic mediation of the memory-enhancing effect of corticotropin-releasing factor in the locus coeruleus of rats. Psychoneuroendocrinology 17(2–3): 113–124.
Chmielarz P, Kusmierczyk J, Parlato R, et al. (2013) Inactivation of glucocorticoid receptor in noradrenergic system influences anxiety- and depressive-like behavior in mice. PLoS ONE 8(8): e72632.
Clewett DV, Huang R, Velasco R, et al. (2018) Locus coeruleus activity strengthens prioritized memories under arousal. Journal of Neuroscience 38(6): 1558–1574.
Clewett DV, Lee TH, Greening S, et al. (2016) Neuromelanin marks the spot: Identifying a locus coeruleus biomarker of cognitive reserve in healthy aging. Neurobiology of Aging 37: 117–126.
Curtis AL, Bello NT, Connolly KR, et al. (2002) Corticotropin-releasing factor neurones of the central nucleus of the amygdala mediate locus coeruleus activation by cardiovascular stress. Journal of Neuroendocrinology 14(8): 667–682.
David Johnson J (2003) Noradrenergic control of cognition: Global attenuation and an interrupt function. Medical Hypotheses 60(5): 689–692.
Davidson RJ (2002) Anxiety and affective style: Role of prefrontal cortex and amygdala. Biological Psychiatry 51(1): 68–80.
Day HE, Masini CV, Campeau S (2004) The pattern of brain c-fos mRNA induced by a component of fox odor, 2,5-dihydro-2,4,5-trimethylthiazoline (TMT), in rats, suggests both systemic and processive stress characteristics. Brain Research 1025(1–2): 139–151.
de Medeiros MA, Carlos Reis L, Eugenio Mello L (2005) Stress-induced c-Fos expression is differentially modulated by dexamethasone, diazepam and imipramine. Neuropsychopharmacology 30(7): 1246–1256.
Delgado MR, Nearing KI, Ledoux JE, et al. (2008) Neural circuitry underlying the regulation of conditioned fear and its relation to extinction. Neuron 59(5): 829–838.
Dell’Osso B, Buoli M, Baldwin DS, et al. (2010) Serotonin norepinephrine reuptake inhibitors (SNRIs) in anxiety disorders: A comprehensive review of their clinical efficacy. Human Psychopharmacology: Clinical and Experimental 25(1): 17–29.
Derntl B, Windischberger C, Robinson S, et al. (2009) Amygdala activity to fear and anger in healthy young males is associated with testosterone. Psychoneuroendocrinology 34(5): 687–693.
Desai SJ, Borkar CD, Nakhate KT, et al. (2014) Neuropeptide Y attenuates anxiety- and depression-like effects of cholecystokinin-4 in mice. Neuroscience 277(2): 818–830.
Deutch AY, Goldstein M, Roth RH (1986) Activation of the locus coeruleus induced by selective stimulation of the ventral tegmental area. Brain Research 363(2): 307–314.
Dierssen M, Gratacos M, Sahun I, et al. (2006) Transgenic mice overexpressing the full-length neurotrophin receptor TrkC exhibit increased catecholaminergic neuron density in specific brain areas and increased anxiety-like behavior and panic reaction. Neurobiology of Disease 24(2): 403–418.
Ditlevsen DN, Elklit A (2012) Gender, trauma type, and PTSD prevalence: A re-analysis of 18 Nordic convenience samples. Annals of General Psychiatry 11: Article 26.
Dowman R, Ritz B, Fowler K (2016) A connectionist modeling study of the neural mechanisms underlying pain’s ability to reorient attention. Cognitive, Affective, & Behavioral Neuroscience 16(2): 689–708.
Eaton K, Sallee FR, Sah R (2007) Relevance of neuropeptide Y (NPY) in psychiatry. Current Topics in Medicinal Chemistry 7(17): 1645–1659.
Ehlers MR, Todd RM (2017) Genesis and maintenance of attentional biases: The role of the locus coeruleus-noradrenaline system. Neural Plasticity 2017(2017): 6817349.
Elman JA, Panizzon MS, Hagler DJ, et al. (2017) Task-evoked pupil dilation and BOLD variance as indicators of locus coeruleus dysfunction. Cortex 97: 60–69.
Ennis M, Aston-Jones G (1986) A potent excitatory input to the nucleus locus coeruleus from the ventrolateral medulla. Neuroscience Letters 71(3): 299–305.
Evers K (2007) Perspectives on memory manipulation: Using beta-blockers to cure post-traumatic stress disorder. Cambridge Quarterly of Healthcare Ethics 16(2): 138–146.
Eysenck MW (1992) Essays in Cognitive Psychology Series. Anxiety: The Cognitive Perspective. Hillsdale, NJ: Lawrence Erlbaum Associates.
Fan JM, Chen XQ, Jin H, et al. (2009) Gestational hypoxia alone or combined with restraint sensitizes the hypothalamic-pituitary-adrenal axis and induces anxiety-like behavior in adult male rat offspring. Neuroscience 159(4): 1363–1373.
Fan Y, Chen P, Li Y, et al. (2014) Corticosterone administration up-regulated expression of norepinephrine transporter and dopamine beta-hydroxylase in rat locus coeruleus and its terminal regions. Journal of Neurochemistry 128(3): 445–458.
Feinstein DL, Kalinin S, Braun D (2016) Causes, consequences, and cures for neuroinflammation mediated via the locus coeruleus: Noradrenergic signaling system. Journal of Neurochemistry 139(Suppl. 2): 154–178.
Ferrucci M, Giorgi FS, Bartalucci A, et al. (2013) The effects of locus coeruleus and norepinephrine in methamphetamine toxicity. Current Neuropharmacology 11(1): 80–94.
Fitzgerald DA, Angstadt M, Jelsone LM, et al. (2006) Beyond threat: Amygdala reactivity across multiple expressions of facial affect. NeuroImage 30(4): 1441–1448.
Fornai F, Bassi L, Torracca MT, et al. (1996a) Region- and neurotransmitter-dependent species and strain differences in DSP-4-induced monoamine depletion in rodents. Neurodegeneration 5(3): 241–249.
Fornai F, Torracca MT, Bassi L, et al. (1996b) Norepinephrine loss selectively enhances chronic nigrostriatal dopamine depletion in mice and rats. Brain Research 735(2): 349–353.
Fox HC, Anderson GM, Tuit K, et al. (2012) Prazosin effects on stress- and cue-induced craving and stress response in alcohol-dependent individuals: Preliminary findings. Alcoholism: Clinical and Experimental Research 36(2): 351–360.
Friedman MJ (2013) Finalizing PTSD in DSM-5: Getting here from there and where to go next. Jouranl of Traumatic Stress 26(5): 548–556.
Ghashghaei HT, Barbas H (2002) Pathways for emotion: Interactions of prefrontal and anterior temporal pathways in the amygdala of the rhesus monkey. Neuroscience 115(4): 1261–1279.
Gilzenrat MS, Nieuwenhuis S, Jepma M, et al. (2010) Pupil diameter tracks changes in control state predicted by the adaptive gain theory of locus coeruleus function. Cognitive, Affective, & Behavioral Neuroscience 10(2): 252–269.
Goossens L, Sunaert S, Peeters R, et al. (2007) Amygdala hyperfunction in phobic fear normalizes after exposure. Biological Psychiatry 62(10): 1119–1125.
Gorman AL, Dunn AJ (1993) Beta-adrenergic receptors are involved in stress-related behavioral changes. Pharmacology Biochemistry and Behavior 45(1): 1–7.
Gorman JM, Askanazi J, Liebowitz MR, et al. (1984) Response to hyperventilation in a group of patients with panic disorder. American Journal of Psychiatry 141(7): 857–861.
Grant SJ, Redmond DE Jr (1984) Neuronal activity of the locus ceruleus in awake Macaca arctoides. Experimental Neurology 84(3): 701–708.
Gray AM (1996) Effect of alprazolam on opiate withdrawal: A combined behavioural and microdialysis study. European Journal of Pharmacology 313(1–2): 73–77.
Gresack JE, Risbrough VB (2011) Corticotropin-releasing factor and noradrenergic signalling exert reciprocal control over startle reactivity. International Journal of Neuropsychopharmacology 14(9): 1179–1194.
Hahn MK, Bannon MJ (1999) Stress-induced C-fos expression in the rat locus coeruleus is dependent on neurokinin 1 receptor activation. Neuroscience 94(4): 1183–1188.
Hare TA, Tottenham N, Galvan A, et al. (2008) Biological substrates of emotional reactivity and regulation in adolescence during an emotional go-nogo task. Biological Psychiatry 63(10): 927–934.
Haring JH, Davis JN (1985) Differential distribution of locus coeruleus projections to the hippocampal formation: Anatomical and biochemical evidence. Brain Research 325(1–2): 366–369.
Hariri AR, Mattay VS, Tessitore A, et al. (2003) Neocortical modulation of the amygdala response to fearful stimuli. Biological Psychiatry 53(6): 494–501.
Harley CW (1987) A role for norepinephrine in arousal, emotion and learning? Limbic modulation by norepinephrine and the Kety hypothesis. Progress in Neuro-Psychopharmacology & Biological Psychiatry 11(4): 419–458.
Harmer CJ, Mackay CE, Reid CB, et al. (2006) Antidepressant drug treatment modifies the neural processing of nonconscious threat cues. Biological Psychiatry 59(9): 816–820.
Harris GC, Aston-Jones G (1993) Beta-adrenergic antagonists attenuate somatic and aversive signs of opiate withdrawal. Neuropsychopharmacology 9(4): 303–311.
Harro J, Oreland L, Vasar E, et al. (1995) Impaired exploratory behaviour after DSP-4 treatment in rats: Implications for the increased anxiety after noradrenergic denervation. European Neuropsychopharmacology 5(4): 447–455.
Hayes PE, Schulz SC (1987) Beta-blockers in anxiety disorders. Journal of Affective Disorders 13(2): 119–130.
Hayley S, Borowski T, Merali Z, et al. (2001) Central monoamine activity in genetically distinct strains of mice following a psychogenic stressor: Effects of predator exposure. Brain Research 892(2): 293–300.
Heilig M (2004) The NPY system in stress, anxiety and depression. Neuropeptides 38(4): 213–224.
Hellsten KS, Sinkkonen ST, Hyde TM, et al. (2010) Human locus coeruleus neurons express the GABA(A) receptor gamma2 subunit gene and produce benzodiazepine binding. Neuroscience Letters 477(2): 77–81.
Heydendael W, Jacobson L (2010) Widespread hypothalamic-pituitary-adrenocortical axis-relevant and mood-relevant effects of chronic fluoxetine treatment on glucocorticoid receptor gene expression in mice. European Journal of Neuroscience 31(5): 892–902.
Hirschberg S, Li Y, Randall A, et al. (2017) Functional dichotomy in spinal- vs prefrontal-projecting locus coeruleus modules splits descending noradrenergic analgesia from ascending aversion and anxiety in rats. Elife 6: e29808.
Howells FM, Stein DJ, Russell VA (2012) Synergistic tonic and phasic activity of the locus coeruleus norepinephrine (LC-NE) arousal system is required for optimal attentional performance. Metabolic Brain Disease 27(3): 267–274.
Ida Y, Tanaka M, Tsuda A, et al. (1985) Attenuating effect of diazepam on stress-induced increases in noradrenaline turnover in specific brain regions of rats: Antagonism by Ro 15-1788. Life Sciences 37(26): 2491–2498.
Insel T, Cuthbert B, Garvey M, et al. (2010) Research domain criteria (RDoC): Toward a new classification framework for research on mental disorders. American Journal of Psychiatry 167(7): 748–751.
Isenberg N, Silbersweig D, Engelien A, et al. (1999) Linguistic threat activates the human amygdala. Proceedings of the National Academy of Sciences of the United States of America 96(18): 10456–10459.
Isingrini E, Perret L, Rainer Q, et al. (2016) Resilience to chronic stress is mediated by noradrenergic regulation of dopamine neurons. Nature Neuroscience 19(4): 560–563.
Itoi K, Sugimoto N, Suzuki S, et al. (2011) Targeting of locus ceruleus noradrenergic neurons expressing human interleukin-2 receptor alpha-subunit in transgenic mice by a recombinant immunotoxin anti-Tac(Fv)-PE38: A study for exploring noradrenergic influence upon anxiety-like and depression-like behaviors. Journal of Neuroscience 31(16): 6132–6139.
Janitzky K, D’Hanis W, Krober A, et al. (2015) TMT predator odor activated neural circuit in C57BL/6J mice indicates TMT-stress as a suitable model for uncontrollable intense stress. Brain Research 1599: 1–8.
Jedema HP, Grace AA (2004) Corticotropin-releasing hormone directly activates noradrenergic neurons of the locus ceruleus recorded in vitro. Journal of Neuroscience 24(43): 9703–9713.
Jedema HP, Finlay JM, Sved AF, et al. (2001) Chronic cold exposure potentiates CRH-evoked increases in electrophysiologic activity of locus coeruleus neurons. Biological Psychiatry 49(4): 351–359.
Jedema HP, Gold SJ, Gonzalez-Burgos G, et al. (2008) Chronic cold exposure increases RGS7 expression and decreases alpha(2)-autoreceptor-mediated inhibition of noradrenergic locus coeruleus neurons. European Journal of Neuroscience 27(9): 2433–2443.
Jefferson JW (1974) Beta-adrenergic receptor blocking drugs in psychiatry. Archives of General Psychiatry 31(5): 681–691.
Jodo E, Chiang C, Aston-Jones G (1998) Potent excitatory influence of prefrontal cortex activity on noradrenergic locus coeruleus neurons. Neuroscience 83(1): 63–79.
Johansson AK, Hansen S (2002) Novelty seeking and harm avoidance in relation to alcohol drinking in intact rats and following axon-sparing lesions to the amygdala and ventral striatum. Alcohol and Alcoholism 37(1): 147–156.
Jones BE, Yang TZ (1985) The efferent projections from the reticular formation and the locus coeruleus studied by anterograde and retrograde axonal transport in the rat. Journal of Comparative Neurology 242(1): 56–92.
Jones EG, Coulter JD, Burton H, et al. (1977) Cells of origin and terminal distribution of corticostriatal fibers arising in the sensory-motor cortex of monkeys. Journal of Comparative Neurology 173(1): 53–80.
Joshi S, Li Y, Kalwani RM, et al. (2016) Relationships between pupil diameter and neuronal activity in the locus coeruleus, colliculi, and cingulate cortex. Neuron 89(1): 221–234.
Kask A, Eller M, Oreland L, et al. (2000) Neuropeptide Y attenuates the effect of locus coeruleus denervation by DSP-4 treatment on social behaviour in the rat. Neuropeptides 34(1): 58–61.
Kask A, Harro J, von Horsten S, et al. (2002) The neurocircuitry and receptor subtypes mediating anxiolytic-like effects of neuropeptide Y. Neuroscience & Biobehavioral Reviews 26(3): 259–283.
Kask A, Rago L, Harro J (1998) Anxiolytic-like effect of neuropeptide Y (NPY) and NPY13-36 microinjected into vicinity of locus coeruleus in rats. Brain Research 788(1–2): 345–348.
Kelly D (1980) Clinical review of beta-blockers in anxiety. Pharmakopsychiatrie, Neuro-Psychopharmakologie 13(5): 259–266.
Kelly SC, He B, Perez SE, et al. (2017) Locus coeruleus cellular and molecular pathology during the progression of Alzheimer’s disease. Acta Neuropathologica Communications 5(1): 8.
Keshavarzy F, Bonnet C, Behzadi G, et al. (2015) Expression patterns of c-Fos early gene and phosphorylated ERK in the rat brain following 1-h immobilization stress: Concomitant changes induced in association with stress-related sleep rebound. Brain Structure and Function 220(3): 1793–1804.
Kessler RC, Berglund P, Demler O, et al. (2005a) Lifetime prevalence and age-of-onset distributions of DSM-IV disorders in the National Comorbidity Survey Replication. Archives of General Psychiatry 62(6): 593–602.
Kessler RC, Chiu WT, Demler O, et al. (2005b) Prevalence, severity, and comorbidity of 12-month DSM-IV disorders in the National Comorbidity Survey Replication. Archives of General Psychiatry 62(6): 617–627.
Kessler RC, Petukhova M, Sampson NA, et al. (2012) Twelve-month and lifetime prevalence and lifetime morbid risk of anxiety and mood disorders in the United States. International Journal of Methods in Psychiatric Research 21(3): 169–184.
Kim MJ, Chey J, Chung A, et al. (2008) Diminished rostral anterior cingulate activity in response to threat-related events in posttraumatic stress disorder. Journal of Psychiatric Research 42(4): 268–277.
Kindt M, Soeter M, Sevenster D (2014) Disrupting reconsolidation of fear memory in humans by a noradrenergic beta-blocker. Journal of Visualized Experiments 94: 52151.
Koola MM, Varghese SP, Fawcett JA (2014) High-dose prazosin for the treatment of post-traumatic stress disorder. Therapeutic Advances in Psychopharmacology 4(1): 43–47.
Korf J, Aghajanian GK, Roth RH (1973a) Increased turnover of norepinephrine in the rat cerebral cortex during stress: Role of the locus coeruleus. Neuropharmacology 12(10): 933–938.
Korf J, Aghajanian GK, Roth RH (1973b) Stimulation and destruction of the locus coeruleus: Opposite effects on 3-methoxy-4-hydroxyphenylglycol sulfate levels in the rat cerebral cortex. European Journal of Pharmacology 21(1): 305–310.
Kowalski PC, Dowben JS, Keltner NL (2014) Biological perspectives: Pain: It’s not all in your head. Perspectives in Psychiatric Care 50(3): 3–6.
Kryzhanovskii GN, Rodina VI, Krupina NA (1991) [Emotional behavioral disturbances in rats administered subconvulsive doses of korazol]. Bulletin of Experimental Biology and Medicine 111(2): 123–126.
Lacosta S, Merali Z, Anisman H (1999) Behavioral and neurochemical consequences of lipopolysaccharide in mice: Anxiogenic-like effects. Brain Research 818(2): 291–303.
Lacosta S, Merali Z, Anisman H (2000) Central monoamine activity following acute and repeated systemic interleukin-2 administration. Neuroimmunomodulation 8(2): 83–90.
Lanius RA, Rabellino D, Boyd JE, et al. (2017) The innate alarm system in PTSD: Conscious and subconscious processing of threat. Current Opinion in Psychology 14(1): 109–115.
Larsen RS, Waters J (2018) Neuromodulatory correlates of pupil dilation. Frontiers in Neural Circuits 12: 21.
Larson CL, Schaefer HS, Siegle GJ, et al. (2006) Fear is fast in phobic individuals: Amygdala activation in response to fear-relevant stimuli. Biological Psychiatry 60(4): 410–417.
Laugero KD, Bell ME, Bhatnagar S, et al. (2001) Sucrose ingestion normalizes central expression of corticotropin-releasing-factor messenger ribonucleic acid and energy balance in adrenalectomized rats: A glucocorticoid-metabolic-brain axis? Endocrinology 142(7): 2796–2804.
Lee B, Shim I, Lee H, et al. (2012a) Effect of ginsenoside Re on depression- and anxiety-like behaviors and cognition memory deficit induced by repeated immobilization in rats. Journal of Microbiolog and Biotechnology 22(5): 708–720.
Lee B, Yun HY, Shim I, et al. (2012b) Bupleurum falcatum prevents depression and anxiety-like behaviors in rats exposed to repeated restraint stress. Journal of Microbiolog and Biotechnology 22(3): 422–430.
Lee HJ, Park HJ, Starkweather A, et al. (2016) Decreased interleukin-4 release from the neurons of the locus coeruleus in response to immobilization stress. Mediators of Inflammation 2016: 3501905.
Legoratti-Sanchez MO, Guevara-Guzman R, Solano-Flores LP (1989) Electrophysiological evidences of a bidirectional communication between the locus coeruleus and the suprachiasmatic nucleus. Brain Research Bulletin 23(4–5): 283–288.
Li L, Feng X, Zhou Z, et al. (2018) Stress accelerates defensive responses to looming in mice and involves a locus coeruleus-superior colliculus projection. Current Biology 28(6): 859–871.e5.
Liddell BJ, Brown KJ, Kemp AH, et al. (2005) A direct brainstem-amygdala-cortical ’alarm’ system for subliminal signals of fear. Neuroimage 24(1): 235–243.
Liebowitz MR, Gorman JM, Fyer AJ, et al. (1985) Social phobia: Review of a neglected anxiety disorder. Archives of General Psychiatry 42(7): 729–736.
Lijster JM, Dierckx B, Utens EM, et al. (2017) The age of onset of anxiety disorders. Canadian Journal of Psychiatry 62(4): 237–246.
Lissek S (2012) Toward an account of clinical anxiety predicated on basic, neurally mapped mechanisms of Pavlovian fear-learning: The case for conditioned overgeneralization. Depress Anxiety 29(4): 257–263.
Liston C, McEwen BS, Casey BJ (2009) Psychosocial stress reversibly disrupts prefrontal processing and attentional control. Proceedings of the National Academy of Sciences of the United States of America 106(3): 912–917.
Liston C, Miller MM, Goldwater DS, et al. (2006) Stress-induced alterations in prefrontal cortical dendritic morphology predict selective impairments in perceptual attentional set-shifting. Journal of Neuroscience 26(30): 7870–7874.
Llorca-Torralba M, Suarez-Pereira I, Bravo L, et al. (2019) Chemogenetic silencing of the locus coeruleus-basolateral amygdala pathway abolishes pain-induced anxiety and enhanced aversive learning in rats. Biological Psychiatry 85(12): 1021–1035.
Loughead J, Gur RC, Elliott M, et al. (2008) Neural circuitry for accurate identification of facial emotions. Brain Research 1194: 37–44.
Loughlin SE, Foote SL, Bloom FE (1986) Efferent projections of nucleus locus coeruleus: Topographic organization of cells of origin demonstrated by three-dimensional reconstruction. Neuroscience 18(2): 291–306.
Morris LS, Tan A, Smith DA, et al. (2020) Sub-millimeter variation in human locus coeruleus is associated with dimensional measures of psychopathology: An in vivo ultra-high field 7-Tesla MRI study. Neuroimage: Clinical 25: 102148.
McCall JG, Al-Hasani R, Siuda ER, et al. (2015) CRH engagement of the locus coeruleus noradrenergic system mediates stress-induced anximorriety. Neuron 87(3): 605–620.
McCall JG, Siuda ER, Bhatti DL, et al. (2017) Locus coeruleus to basolateral amygdala noradrenergic projections promote anxiety-like behavior. Elife 6: e18247.
McDonald RJ, White NM (1993) A triple dissociation of memory systems: Hippocampus, amygdala, and dorsal striatum. Behavioral Neuroscience 107(1): 3–22.
Makino S, Smith MA, Gold PW (2002) Regulatory role of glucocorticoids and glucocorticoid receptor mRNA levels on tyrosine hydroxylase gene expression in the locus coeruleus during repeated immobilization stress. Brain Research 943(2): 216–223.
Marazziti D, Abelli M, Baroni S, et al. (2015) Neurobiological correlates of social anxiety disorder: An update. CNS Spectrums 20(2): 100–111.
Meeten F, Davey GC, Makovac E, et al. (2016) Goal directed worry rules are associated with distinct patterns of amygdala functional connectivity and vagal modulation during perseverative cognition. Frontiers in Human Neuroscience 10: 553.
Merali Z, McIntosh J, Kent P, et al. (1998) Aversive and appetitive events evoke the release of corticotropin-releasing hormone and bombesin-like peptides at the central nucleus of the amygdala. Journal of Neuroscience 18(12): 4758–4766.
Milad MR, Quirk GJ (2002) Neurons in medial prefrontal cortex signal memory for fear extinction. Nature 420(6911): 70–74.
Milad MR, Quinn BT, Pitman RK, et al. (2005) Thickness of ventromedial prefrontal cortex in humans is correlated with extinction memory. Proceedings of the National Academy of Sciences of the United States of America 102(30): 10706–10711.
Morey RA, Dunsmoor JE, Haswell CC, et al. (2015) Fear learning circuitry is biased toward generalization of fear associations in posttraumatic stress disorder. Translational Psychiatry 5(12): e700.
Morris LS, Tan A, Smith DA, et al. (2020) Sub-millimeter variation in human locus coeruleus is associated with dimensional measures of psychopathology: An in vivo ultra-high field 7-Tesla MRI study. NeuroImage: Clinical 25: 102148.
Murphy PR, O’Connell RG, O’Sullivan M, et al. (2014) Pupil diameter covaries with BOLD activity in human locus coeruleus. Human Brain Mapping 35(8): 4140–4154.
Nader K, Schafe GE, Le Doux JE (2000) Fear memories require protein synthesis in the amygdala for reconsolidation after retrieval. Nature 406(6797): 722–726.
Naegeli C, Zeffiro T, Piccirelli M, et al. (2018) Locus coeruleus activity mediates hyperresponsiveness in posttraumatic stress disorder. Biological Psychiatry 83(3): 254–262.
Nakamoto K, Aizawa F, Kinoshita M, et al. (2017) Astrocyte activation in locus coeruleus is involved in neuropathic pain exacerbation mediated by maternal separation and social isolation stress. Frontiers in Pharmacology 8: 401.
Nassar MR, Rumsey KM, Wilson RC, et al. (2012) Rational regulation of learning dynamics by pupil-linked arousal systems. Nature Neuroscience 15(7): 1040–1046.
Nelson A, Mooney R (2016) The basal forebrain and motor cortex provide convergent yet distinct movement-related inputs to the auditory cortex. Neuron 90(3): 635–648.
Neophytou SI, Aspley S, Butler S, et al. (2001) Effects of lesioning noradrenergic neurones in the locus coeruleus on conditioned and unconditioned aversive behaviour in the rat. Progress in Neuropsychopharmacol Biological Psychiatry 25(6): 1307–1321.
Noyes R Jr (1982) Beta-blocking drugs and anxiety. Psychosomatics 23(2): 155–170.
O’Donnell J, Zeppenfeld D, McConnell E, et al. (2012) Norepinephrine: A neuromodulator that boosts the function of multiple cell types to optimize CNS performance. Neurochemical Research 37(11): 2496–2512.
Ochsner KN, Ray RD, Cooper JC, et al. (2004) For better or for worse: Neural systems supporting the cognitive down- and up-regulation of negative emotion. NeuroImage 23(2): 483–499.
Ogier M, Bricca G, Bader M, et al. (2016) Locus coeruleus dysfunction in transgenic rats with low brain angiotensinogen. CNS Neuroscience & Therapeutics 22(3): 230–237.
Olton DS, Becker JT, Handelmann GE (1979) Hippocampus, space, and memory. Behavioral and Brain Sciences 2(3): 313–365.
Ornstein K, Milon H, McRae-Degueurce A, et al. (1987) Biochemical and radioautographic evidence for dopaminergic afferents of the locus coeruleus originating in the ventral tegmental area. Journal of Neural Transmission 70(3–4): 183–191.
Owens MJ, Vargas MA, Nemeroff CB (1993) The effects of alprazolam on corticotropin-releasing factor neurons in the rat brain: Implications for a role for CRF in the pathogenesis of anxiety disorders. Journal of Psychiatric Research 27(Suppl. 1): 209–220.
Oya H, Kawasaki H, Howard MA 3rd, et al. (2002) Electrophysiological responses in the human amygdala discriminate emotion categories of complex visual stimuli. Journal of Neuroscience 22(21): 9502–9512.
Park J, Moghaddam B (2017) Risk of punishment influences discrete and coordinated encoding of reward-guided actions by prefrontal cortex and VTA neurons. Elife 6: e30056.
Peng SY, Zhuang QX, Zhang YX, et al. (2016) Excitatory effect of norepinephrine on neurons in the inferior vestibular nucleus and the underlying receptor mechanism. Journal of Neuroscience Research 94(8): 736–748.
Pezawas L, Meyer-Lindenberg A, Drabant EM, et al. (2005) 5-HTTLPR polymorphism impacts human cingulate-amygdala interactions: A genetic susceptibility mechanism for depression. Nature Neuroscience 8(6): 828–834.
Phelps EA (2004) Human emotion and memory: Interactions of the amygdala and hippocampal complex. Current Opinion in Neurobiology 14(2): 198–202.
Priovoulos N, Jacobs HIL, Ivanov D, et al. (2018) High-resolution in vivo imaging of human locus coeruleus by magnetization transfer MRI at 3T and 7T. NeuroImage 168: 427–436.
Purvis EM, Klein AK, Ettenberg A (2018) Lateral habenular norepinephrine contributes to states of arousal and anxiety in male rats. Behavioural Brain Research 347: 108–115.
Rajkowski J, Majczynski H, Clayton E, et al. (2004) Activation of monkey locus coeruleus neurons varies with difficulty and performance in a target detection task. Journal of Neurophysiology 92(1): 361–371.
Raskind MA, Dobie DJ, Kanter ED, et al. (2000) The alpha1-adrenergic antagonist prazosin ameliorates combat trauma nightmares in veterans with posttraumatic stress disorder: A report of 4 cases. Journal of Clinical Psychiatry 61(2): 129–133.
Raskind MA, Peskind ER, Halter JB, et al. (1984) Norepinephrine and MHPG levels in CSF and plasma in Alzheimer’s disease. Archives of General Psychiatry 41(4): 343–346.
Rasmussen K, Jacobs BL (1986) Single unit activity of locus coeruleus neurons in the freely moving cat. II. Conditioning and pharmacologic studies. Brain Research 371(2): 335–344.
Reimer J, McGinley MJ, Liu Y, et al. (2016) Pupil fluctuations track rapid changes in adrenergic and cholinergic activity in cortex. Nature Communications 7: 13289.
Retson TA, Reyes BA, Van Bockstaele EJ (2015) Chronic alcohol exposure differentially affects activation of female locus coeruleus neurons and the subcellular distribution of corticotropin releasing factor receptors. Progress in Neuro-Psychopharmacology & Biological Psychiatry 56: 66–74.
Robertson SD, Plummer NW, Jensen P (2016) Uncovering diversity in the development of central noradrenergic neurons and their efferents. Brain Research 1641(Pt. B): 234–244.
Robertson SD, Plummer NW, de Marchena J, et al. (2013) Developmental origins of central norepinephrine neuron diversity. Nature Neuroscience 16(8): 1016–1023.
Romero LM (2004) Physiological stress in ecology: Lessons from biomedical research. Trends in Ecology & Evolution 19(5): 249–255.
Rosen JB, Schulkin J (1998) From normal fear to pathological anxiety. Psychological Review 105(2): 325–350.
Roszkowski M, Manuella F, von Ziegler L, et al. (2016) Rapid stress-induced transcriptomic changes in the brain depend on beta-adrenergic signaling. Neuropharmacology 107: 329–338.
Sabban EL, Laukova M, Alaluf LG, et al. (2015a) Locus coeruleus response to single-prolonged stress and early intervention with intranasal neuropeptide Y. Journal of Neurochemistry 135(5): 975–986.
Sabban EL, Serova LI, Alaluf LG, et al. (2015b) Comparative effects of intranasal neuropeptide Y and HS014 in preventing anxiety and depressive-like behavior elicited by single prolonged stress. Behavioural Brain Research 295: 9–16.
Sadler TR, Nguyen PT, Yang J, et al. (2011) Antenatal maternal stress alters functional brain responses in adult offspring during conditioned fear. Brain Research 1385: 163–174.
Salchner P, Sartori SB, Sinner C, et al. (2006) Airjet and FG-7142-induced Fos expression differs in rats selectively bred for high and low anxiety-related behavior. Neuropharmacology 50(8): 1048–1058.
Salim S, Sarraj N, Taneja M, et al. (2010) Moderate treadmill exercise prevents oxidative stress-induced anxiety-like behavior in rats. Behavioural Brain Research 208(2): 545–552.
Sands SA, Strong R, Corbitt J, et al. (2000) Effects of acute restraint stress on tyrosine hydroxylase mRNA expression in locus coeruleus of Wistar and Wistar-Kyoto rats. Brain Research: Molecular Brain Research 75(1): 1–7.
Sanghera MK, German DC (1983) The effects of benzodiazepine and non-benzodiazepine anxiolytics on locus coeruleus unit activity. Journal of Neural Transmission 57(4): 267–279.
Sara SJ (2009) The locus coeruleus and noradrenergic modulation of cognition. Nature Reviews Neuroscience 10(3): 211–223.
Sara SJ, Bouret S (2012) Orienting and reorienting: The locus coeruleus mediates cognition through arousal. Neuron 76(1): 130–141.
Sara SJ, Herve-Minvielle A (1995) Inhibitory influence of frontal cortex on locus coeruleus neurons. Proceedings of the National Academy of Sciences of the United States of America 92(13): 6032–6036.
Sayed S, Van Dam NT, Horn SR, et al. (2018) A randomized dose-ranging study of neuropeptide Y in patients with posttraumatic stress disorder. International Journal of Neuropsychopharmacology 21(1): 3–11.
Scharner S, Prinz P, Goebel-Stengel M, et al. (2017) Activity-based anorexia activates nesfatin-1 immunoreactive neurons in distinct brain nuclei of female rats. Brain Research 1677: 33–46.
Scherder EJ, Sergeant JA, Swaab DF (2003) Pain processing in dementia and its relation to neuropathology. The Lancet Neurology 2(11): 677–686.
Sciolino NR, Holmes PV (2012) Exercise offers anxiolytic potential: A role for stress and brain noradrenergic-galaninergic mechanisms. Neuroscience & Biobehavioral Reviews 36(9): 1965–1984.
Sciolino NR, Plummer NW, Chen YW, et al. (2016) Recombinase-dependent mouse lines for chemogenetic activation of genetically defined cell types. Cell Reports 15(11): 2563–2573.
Sciolino NR, Smith JM, Stranahan AM, et al. (2015) Galanin mediates features of neural and behavioral stress resilience afforded by exercise. Neuropharmacology 89: 255–264.
Sears RM, Fink AE, Wigestrand MB, et al. (2013) Orexin/hypocretin system modulates amygdala-dependent threat learning through the locus coeruleus. Proceedings of the National Academy of Sciences of the United States of America 110(50): 20260–20265.
Serova LI, Tillinger A, Alaluf LG, et al. (2013) Single intranasal neuropeptide Y infusion attenuates development of PTSD-like symptoms to traumatic stress in rats. Neuroscience 236: 298–312.
Shechner T, Britton JC, Perez-Edgar K, et al. (2012) Attention biases, anxiety, and development: Toward or away from threats or rewards? Depress Anxiety 29(4): 282–294.
Shipley MT, Halloran FJ, de la Torre J (1985) Surprisingly rich projection from locus coeruleus to the olfactory bulb in the rat. Brain Research 329(1–2): 294–299.
Shishkina GT, Kalinina TS, Sournina NY, et al. (2002) Effects of antisense oligodeoxynucleotide to the alpha2A-adrenoceptors on the plasma corticosterone level and on elevated plus-maze behavior in rats. Psychoneuroendocrinology 27(5): 593–601.
Silveira MC, Sandner G, Graeff FG (1993) Induction of Fos immunoreactivity in the brain by exposure to the elevated plus-maze. Behavioural Brain Research 56(1): 115–118.
Simpson JR Jr, Drevets WC, Snyder AZ, et al. (2001) Emotion-induced changes in human medial prefrontal cortex: II: During anticipatory anxiety. Proceedings of the National Academy of Sciences of the United States of America 98(2): 688–693.
Sinha R, Lacadie CM, Constable RT, et al. (2016) Dynamic neural activity during stress signals resilient coping. Proceedings of the National Academy of Sciences of the United States of America 113(31): 8837–8842.
Skelton KH, Nemeroff CB, Knight DL, et al. (2000) Chronic administration of the triazolobenzodiazepine alprazolam produces opposite effects on corticotropin-releasing factor and urocortin neuronal systems. Journal of Neuroscience 20(3): 1240–1248.
Sobanski T, Wagner G (2017) Functional neuroanatomy in panic disorder: Status quo of the research. World Journal of Psychiatry 7(1): 12–33.
Sobrinho CR, Canteras NS (2011) A study of the catecholaminergic inputs to the dorsal premammillary nucleus. Neuroscience Letters 501(3): 157–162.
Sodero AO, Valdomero A, Cuadra GR, et al. (2004) Locus coeruleus activity in perinatally protein-deprived rats: Effects of fluoxetine administration. European Journal of Pharmacology 503(1–3): 35–42.
Soderpalm B, Engel JA (1989) Alpha 2-adrenoceptor antagonists potentiate the anticonflict and the rotarod impairing effects of benzodiazepines. Journal of Neural Transmission 76(3): 191–204.
Southwick SM, Krystal JH, Morgan CA, et al. (1993) Abnormal noradrenergic function in posttraumatic stress disorder. Archives of General Psychiatry 50(4): 266–274.
Steuwe C, Daniels JK, Frewen PA, et al. (2015) Effect of direct eye contact in women with PTSD related to interpersonal trauma: Psychophysiological interaction analysis of connectivity of an innate alarm system. Psychiatry Research 232(2): 162–167.
Sutherland RJ, McDonald RJ (1990) Hippocampus, amygdala, and memory deficits in rats. Behavioural Brain Research 37(1): 57–79.
Svensson T (1982) Emerging aspects of the adrenergic nervous systems. Acta Anaesthesiologica Scandinavica 76: 8–11.
Szabo ST, Blier P (2002) Effects of serotonin (5-hydroxytryptamine, 5-HT) reuptake inhibition plus 5-HT(2A) receptor antagonism on the firing activity of norepinephrine neurons. Journal of Pharmacology and Experimental Therapeutics 302(3): 983–991.
Szabo ST, de Montigny C, Blier P (2000) Progressive attenuation of the firing activity of locus coeruleus noradrenergic neurons by sustained administration of selective serotonin reuptake inhibitors. International Journal of Neuropsychopharmacology 3(1): 1–11.
Tanaka M, Yoshida M, Emoto H, et al. (2000) Noradrenaline systems in the hypothalamus, amygdala and locus coeruleus are involved in the provocation of anxiety: Basic studies. European Journal of Pharmacology 405(1–3): 397–406.
Taylor JM, Whalen PJ (2015) Neuroimaging and anxiety: The neural substrates of pathological and non-pathological anxiety. Current Psychiatry Reports 17(6): 49.
Theofilas P, Ehrenberg AJ, Dunlop S, et al. (2017) Locus coeruleus volume and cell population changes during Alzheimer’s disease progression: A stereological study in human postmortem brains with potential implication for early-stage biomarker discovery. Alzheimer’s & Dementia 13(3): 236–246.
Totah NK, Neves RM, Panzeri S, et al. (2018) The locus coeruleus is a complex and differentiated neuromodulatory system. Neuron 99(5): 1055–1068.e6.
Trulson ME, Henderson LJ (1984) Buspirone increases locus coeruleus noradrenergic neuronal activity in vitro. European Journal of Pharmacology 106(1): 195–197.
Tsaltas E, Gray JA, Fillenz M (1984) Alleviation of response suppression to conditioned aversive stimuli by lesions of the dorsal noradrenergic bundle. Behavioural Brain Research 13(2): 115–127.
Uematsu A, Tan BZ, Johansen JP (2015) Projection specificity in heterogeneous locus coeruleus cell populations: Implications for learning and memory. Learning & Memory 22(9): 444–451.
Uematsu A, Tan BZ, Ycu EA, et al. (2017) Modular organization of the brainstem noradrenaline system coordinates opposing learning states. Nature Neuroscience 20(11): 1602–1611.
Ulrich-Lai YM, Herman JP (2009) Neural regulation of endocrine and autonomic stress responses. Nature Reviews Neuroscience 10(6): 397–409.
Usher M, Cohen JD, Servan-Schreiber D, et al. (1999) The role of locus coeruleus in the regulation of cognitive performance. Science 283(5401): 549–554.
Valenca AM, Nardi AE, Mezzasalma MA, et al. (2004) Clonidine in respiratory panic disorder subtype. Arquivos de Neuro-Psiquiatria 62(2B): 396–398.
Valentino RJ, Reyes B, Van Bockstaele E, et al. (2012) Molecular and cellular sex differences at the intersection of stress and arousal. Neuropharmacology 62(1): 13–20.
Van Bockstaele EJ, Colago EE, Valentino RJ (1998) Amygdaloid corticotropin-releasing factor targets locus coeruleus dendrites: Substrate for the co-ordination of emotional and cognitive limbs of the stress response. Journal of Neuroendocrinology 10(10): 743–757.
van Marle HJ, Hermans EJ, Qin S, et al. (2010) Enhanced resting-state connectivity of amygdala in the immediate aftermath of acute psychological stress. NeuroImage 53(1): 348–354.
Verbanac JS, Commissaris RL, Altman HJ, et al. (1994) Electrophysiological characteristics of locus coeruleus neurons in the Maudsley reactive (MR) and non-reactive (MNRA) rat strains. Neuroscience Letters 179(1–2): 137–140.
Wager TD, Davidson ML, Hughes BL, et al. (2008) Prefrontal-subcortical pathways mediating successful emotion regulation. Neuron 59(6): 1037–1050.
Wang X, Zhao B, Li X (2015) Dexmedetomidine attenuates isoflurane-induced cognitive impairment through antioxidant, anti-inflammatory and anti-apoptosis in aging rat. International Journal of Clinical and Experimental Medicine 8(10): 17281–17288.
Weiss JM, Boss-Williams KA, Moore JP, et al. (2005) Testing the hypothesis that locus coeruleus hyperactivity produces depression-related changes via galanin. Neuropeptides 39(3): 281–287.
Weiss JM, Stout JC, Aaron MF, et al. (1994) Depression and anxiety: Role of the locus coeruleus and corticotropin-releasing factor. Brain Research Bulletin 35(5–6): 561–572.
Westlund KN, Coulter JD (1980) Descending projections of the locus coeruleus and subcoeruleus/medial parabrachial nuclei in monkey: Axonal transport studies and dopamine-beta-hydroxylase immunocytochemistry. Brain Research 2(3): 235–264.
Willis WD, Westlund KN (1997) Neuroanatomy of the pain system and of the pathways that modulate pain. Journal of Clinical Neurophysiology 14(1): 2–31.
Windmann S (1998) Panic disorder from a monistic perspective: Integrating neurobiological and psychological approaches. Journal of Anxiety Disorders 12(5): 485–507.
Wohleb ES, Hanke ML, Corona AW, et al. (2011) β-Adrenergic receptor antagonism prevents anxiety-like behavior and microglial reactivity induced by repeated social defeat. Journal of Neuroscience 31(17): 6277–6288.
Wood NE, Rosasco ML, Suris AM, et al. (2015) Pharmacological blockade of memory reconsolidation in posttraumatic stress disorder: Three negative psychophysiological studies. Psychiatry Research 225(1–2): 31–39.
Yu AJ, Dayan P (2005) Uncertainty, neuromodulation, and attention. Neuron 46(4): 681–692.
Zerbi V, Floriou-Servou A, Markicevic M, et al. (2019) Rapid reconfiguration of the functional connectome after chemogenetic locus coeruleus activation. Neuron 103(4): 702–718e705.
Zhang H, Chaudhury D, Nectow AR, et al. (2019) α1- and β3-adrenergic receptor-mediated mesolimbic homeostatic plasticity confers resilience to social stress in susceptible mice. Biological Psychiatry 85(3): 226–236.
Zhang S, Zhang H, Ku SM, et al. (2018) Sex differences in the neuroadaptations of reward-related circuits in response to subchronic variable stress. Neuroscience 376: 108–116.

Cite article

Cite article

Cite article

OR

Download to reference manager

If you have citation software installed, you can download article citation data to the citation manager of your choice

Share options

Share

Share this article

Share with email
EMAIL ARTICLE LINK
Share on social media

Share access to this article

Sharing links are not relevant where the article is open access and not available if you do not have a subscription.

For more information view the Sage Journals article sharing page.

Information, rights and permissions

Information

Published In

Article first published online: July 21, 2020
Issue published: January-December 2020

Keywords

  1. Locus Coeruleus
  2. Pathological Anxiety

Rights and permissions

© The Author(s) 2020.
Creative Commons License (CC BY-NC 4.0)
This article is distributed under the terms of the Creative Commons Attribution-NonCommercial 4.0 License (https://creativecommons.org/licenses/by-nc/4.0/) which permits non-commercial use, reproduction and distribution of the work without further permission provided the original work is attributed as specified on the SAGE and Open Access pages (https://us.sagepub.com/en-us/nam/open-access-at-sage).
Request permissions for this article.
PubMed: 32954002

Authors

Affiliations

Laurel S. Morris
The Depression and Anxiety Center for Discovery and Treatment, Department of Psychiatry, Icahn School of Medicine at Mount Sinai, New York, NY, USA
Jordan G. McCall
Department of Anesthesiology, Washington University in St. Louis, St. Louis, MO, USA
Dennis S. Charney
Dean’s Office, Icahn School of Medicine at Mount Sinai, New York, NY, USA
James W. Murrough
The Depression and Anxiety Center for Discovery and Treatment, Department of Psychiatry, Icahn School of Medicine at Mount Sinai, New York, NY, USA

Notes

Laurel S. Morris, The Depression and Anxiety Center for Discovery and Treatment, Department of Psychiatry, Icahn School of Medicine at Mount Sinai, New York, NY, 10029, USA. Email: [email protected]

Metrics and citations

Metrics

Journals metrics

This article was published in Brain and Neuroscience Advances.

VIEW ALL JOURNAL METRICS

Article usage*

Total views and downloads: 11483

*Article usage tracking started in December 2016


Altmetric

See the impact this article is making through the number of times it’s been read, and the Altmetric Score.
Learn more about the Altmetric Scores



Articles citing this one

Receive email alerts when this article is cited

Web of Science: 0

Crossref: 54

  1. Post-traumatic Stress Disorder: Focus on Neuroinflammation
    Go to citation Crossref Google Scholar
  2. Norepinephrine system at the interface of attention and reward
    Go to citation Crossref Google Scholar
  3. Impact of acute psychosocial stress on attentional control in humans. ...
    Go to citation Crossref Google Scholar
  4. Structural and functional characterization of the locus coeruleus in y...
    Go to citation Crossref Google Scholar
  5. Cannabigerol modulates α2-adrenoceptor and 5-HT1A receptor-mediated el...
    Go to citation Crossref Google Scholar
  6. Locus coeruleus ablation in mice: protocol optimization, stereology an...
    Go to citation Crossref Google Scholar
  7. Prefrontal modulation of anxiety through a lens of noradrenergic signa...
    Go to citation Crossref Google Scholar
  8. Neuropathological correlates of neuropsychiatric symptoms in dementia
    Go to citation Crossref Google Scholar
  9. Whole-brain afferent input mapping to functionally distinct brainstem ...
    Go to citation Crossref Google Scholar
  10. Site-Specific knockdown of microglia in the locus coeruleus regulates ...
    Go to citation Crossref Google Scholar
  11. The Central Noradrenergic System in Neurodevelopmental Disorders: Merg...
    Go to citation Crossref Google Scholar
  12. Mitochondrial dysfunction in animal models of PTSD: Relationships betw...
    Go to citation Crossref Google Scholar
  13. Transcriptome profiles associated with resilience and susceptibility t...
    Go to citation Crossref Google Scholar
  14. Noradrenergic modulation of stress resilience
    Go to citation Crossref Google Scholar
  15. Evaluation of the posterior insular cortex involvement in anxiogenic r...
    Go to citation Crossref Google Scholar
  16. Associations between locus coeruleus MRI contrast and physiological re...
    Go to citation Crossref Google Scholar
  17. First few seconds for flow: A comprehensive proposal of the neurobiolo...
    Go to citation Crossref Google Scholar
  18. In vivo tractography of human locus coeruleus—relation to 7T resting s...
    Go to citation Crossref Google Scholar
  19. From Low-Grade Inflammation in Osteoarthritis to Neuropsychiatric Sequ...
    Go to citation Crossref Google Scholar
  20. Know thy SEFL: Fear sensitization and its relevance to stressor-relate...
    Go to citation Crossref Google Scholar
  21. Sexually dimorphic role of the locus coeruleus PAC1 receptors in regul...
    Go to citation Crossref Google Scholar
  22. Loss of mitochondrial enzyme GPT2 causes early neurodegeneration in lo...
    Go to citation Crossref Google Scholar
  23. The Role of Beta-Adrenergic Receptors in Depression and Resilience
    Go to citation Crossref Google Scholar
  24. Alpha-2 Adrenoreceptor Antagonist Yohimbine Potentiates Consolidation ...
    Go to citation Crossref Google Scholar
  25. Natural locus coeruleus dynamics during feeding
    Go to citation Crossref Google Scholar
  26. Invisible wounds: Suturing the gap between the neurobiology, conventio...
    Go to citation Crossref Google Scholar
  27. The Contribution of Noradrenergic Activity to Anxiety‐Induced Freezing...
    Go to citation Crossref Google Scholar
  28. The interplay of hypoxic and mental stress: Implications for anxiety a...
    Go to citation Crossref Google Scholar
  29. Larval Zebrafish as a Model for Mechanistic Discovery in Mental Health
    Go to citation Crossref Google Scholar
  30. The role of the locus coeruleus in shaping adaptive cortical melodies
    Go to citation Crossref Google Scholar
  31. HIV Tat and cocaine interactively alter genome-wide DNA methylation an...
    Go to citation Crossref Google Scholar
  32. Age differences in diffusivity in the locus coeruleus and its ascendin...
    Go to citation Crossref Google Scholar
  33. Stress induced microglial activation contributes to depression
    Go to citation Crossref Google Scholar
  34. Water and Meadow Views Both Afford Perceived but Not Performance-Based...
    Go to citation Crossref Google Scholar
  35. High trait anxiety blocks olfactory plasticity induced by aversive lea...
    Go to citation Crossref Google Scholar
  36. Pain and psychiatric illness
    Go to citation Crossref Google Scholar
  37. Review of the Midbrain Ascending Arousal Network Nuclei and Implicatio...
    Go to citation Crossref Google Scholar
  38. Clinical Approach to Stress
    Go to citation Crossref Google Scholar
  39. New-Onset Sleepwalking in a Patient Treated With Buspirone
    Go to citation Crossref Google Scholar
  40. Differential Effects of Exercise on fMRI of the Midbrain Ascending Aro...
    Go to citation Crossref Google Scholar
  41. Processing of fMRI-related anxiety and bi-directional information flow...
    Go to citation Crossref Google Scholar
  42. Alterations in reward network functional connectivity are associated w...
    Go to citation Crossref Google Scholar
  43. In search of sex-related mediators of affective illness
    Go to citation Crossref Google Scholar
  44. Modulation of bioelectric cues in the evolution of flying fishes
    Go to citation Crossref Google Scholar
  45. Activation of locus coeruleus to rostromedial tegmental nucleus (RMTg)...
    Go to citation Crossref Google Scholar
  46. Inflammation, Anxiety, and Stress in Attention-Deficit/Hyperactivity D...
    Go to citation Crossref Google Scholar
  47. The neurobiology of human fear generalization: meta-analysis and worki...
    Go to citation Crossref Google Scholar
  48. Fluoroquinolones-Associated Disability: It Is Not All in Your Head
    Go to citation Crossref Google Scholar
  49. Attentional Disengagement and the Locus Coeruleus – Norepinephrine Sys...
    Go to citation Crossref Google Scholar
  50. Mental Resilience and Coping With Stress: A Comprehensive, Multi-level...
    Go to citation Crossref Google Scholar
  51. Is a Mask That Covers the Mouth and Nose Free from Undesirable Side Ef...
    Go to citation Crossref Google Scholar
  52. Locus Coeruleus Malfunction Is Linked to Psychopathology in Prodromal ...
    Go to citation Crossref Google Scholar
  53. Spinally projecting noradrenergic neurons of the locus coeruleus displ...
    Go to citation Crossref Google ScholarPub Med
  54. Chronic Cocaine Exposure Alters Genome-Wide DNA Methylation and Gene E...
    Go to citation Crossref Google Scholar

Figures and tables

Figures & Media

Tables

View Options

View options

PDF/ePub

View PDF/ePub

Get access

Access options

If you have access to journal content via a personal subscription, university, library, employer or society, select from the options below:


Alternatively, view purchase options below:

Access journal content via a DeepDyve subscription or find out more about this option.