<<

Accepted Manuscript

Cannabinoid interventions for PTSD: Where to next?

Luke Ney, Allison Matthews, Raimondo Bruno, Kim Felmingham

PII: S0278-5846(19)30034-X DOI: https://doi.org/10.1016/j.pnpbp.2019.03.017 Reference: PNP 9622 To appear in: Progress in Neuropsychopharmacology & Biological Psychiatry Received date: 14 January 2019 Revised date: 20 March 2019 Accepted date: 29 March 2019

Please cite this article as: L. Ney, A. Matthews, R. Bruno, et al., interventions for PTSD: Where to next?, Progress in Neuropsychopharmacology & Biological Psychiatry, https://doi.org/10.1016/j.pnpbp.2019.03.017

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. ACCEPTED MANUSCRIPT 1

Cannabinoid interventions for PTSD: Where to next?

Luke Ney1, Allison Matthews1, Raimondo Bruno1 and Kim Felmingham2

1School of Psychology, University of Tasmania, Australia

2School of Psychological Sciences, University of Melbourne, Australia

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 2

Abstract

Cannabinoids are a promising method for pharmacological treatment of post- traumatic stress disorder (PTSD). Despite considerable research devoted to the effect of cannabinoid modulation on PTSD symptomology, there is not a currently agreed way by which the cannabinoid system should be targeted in humans. In this review, we present an overview of recent research identifying neurological pathways by which different cannabinoid-based treatments may exert their effects on PTSD symptomology. We evaluate the strengths and weaknesses of each of these different approaches, including recent challenges presented to favourable options such as fatty acid amide hydrolase (FAAH) inhibitors. This article makes the strengths and challenges of different potential cannabinoid treatments accessible to psychological researchers interested in cannabinoid therapeutics and aims to aid selection of appropriate tools for future clinical trials.

ACCEPTED MANUSCRIPT

Keywords: endocannabinoids; posttraumatic stress disorder; translational medical research; ; psychiatric disorders; traumatic stress; .

ACCEPTED MANUSCRIPT 3

1. Introduction

1.1 Post-traumatic Stress Disorder

Post-traumatic Stress Disorder (PTSD) is a serious mental health condition that may develop following a traumatic experience. Symptoms include re-experiencing distressing aspects of the trauma, hyper-arousal, avoidance of trauma reminders, sleep disturbances and negative cognitions and mood (American Psychiatric Association, 2013). PTSD can be a chronic condition and is associated with significant distress and functional impairments. The gold-standard psychological treatment for PTSD is exposure therapy, where patients are encouraged to gradually confront reminders of their trauma, enabling them to learn to regulate their fear of the no-longer threatening situation (Graham, Callaghan, & Richardson,

2014). Using this technique, many PTSD patients experience a significant reduction in their symptoms. Unfortunately, approximately 40% of patients experience only minimal or partial response to exposure treatment (Bisson, Roberts, Andrew, Cooper, & Lewis, 2013), hence the focus of PTSD research is often to enhance the outcomes of exposure therapy. The use of pharmacological adjuncts to optimize exposure therapy and enhance treatment response has been a recent area of active research, though has yet to emerge with a promising pharmaceutical for this purpose (Singewald, Schmuckermair, Whittle, Holmes, & Ressler,

2015; Stein, Ipser, & Seedat, 2006).

1.1.1 PTSD Symptomology and Mechanisms ACCEPTED MANUSCRIPT PTSD symptomology is believed to be largely maintained by a combination of dysregulated biological stress responding and maladaptive memory processes attributable to the extreme stress caused by a traumatic event (Pitman et al., 2012; Yehuda et al., 2015; Zuj,

Palmer, Hsu, et al., 2016). At the time of trauma, sympathetic and central stress hormones

(for example noradrenaline and cortisol) are released in excessive levels as part of the stress

ACCEPTED MANUSCRIPT 4 response. Excessive release of these hormones assigns emotional salience to the events, which in turn leads to over-consolidation of the trauma and associated environment or context in which the trauma occurs (Pitman & Delahanty, 2005). This is achieved by a well- known mechanism by which emotional experiences are preferentially consolidated into long- term memory (McGaugh & Roozendaal, 2002; Roozendaal & McGaugh, 2011); however, in

PTSD this occurs at an extreme and maladaptive level due in part to the excessive release of stress hormones (Pitman & Delahanty, 2005; Pitman et al., 2012). Over-consolidation of trauma memories leads to fractured and poorly contextualised memory traces, which may spontaneously arise post-trauma and are known as intrusive memories (Brewin, Gregory,

Lipton, & Burgess, 2010; Pitman et al., 2012). Studies in humans have shown that interactions between noradrenaline and cortisol can predict the number of intrusive memories following exposure to emotional or traumatic images (Bryant, McGrath, & Felmingham,

2013; Chou, La Marca, Steptoe, & Brewin, 2014; Nicholson, Bryant, & Felmingham, 2014); however, it is clear that not all people exposed to a traumatic event will develop these symptoms. Therefore, PTSD researchers are often interested in understanding predisposing factors that determine whether someone will develop symptoms such as emotional or intrusive memories following trauma.

Following trauma, a significant proportion of persons who eventually develop PTSD show chronically decreased cortisol levels, which is due to increased negative feedback of the hypothalamic-pituitary-adrenal (HPA) axis (Galatzer-Levy, Ma, Statnikov, Yehuda, & Shalev, ACCEPTED MANUSCRIPT 2017; Yehuda, 2009). This means that PTSD subjects show decreased cortisol reactivity to stress (Griffin, Resick, & Yehuda, 2005; Stein, Yehuda, Koverola, & Hanna, 1997).

Conversely, chronic noradrenergic hyperactivity has been reported in PTSD subjects and is believed, along with disruption to enhanced reduction in natural HPA reactivity, to contribute to hyperarousal features of the disorder (Pitman et al., 2012; Southwick et al., 1999).

ACCEPTED MANUSCRIPT 5

Following trauma, there is also believed to be a deficient in what is termed “fear extinction” in the aetiology and maintenance of PTSD (Milad & Quirk, 2012; Zuj, Palmer,

Lommen, & Felmingham, 2016). Fear extinction is part of a larger framework based on classical conditioning, whereby “fear” – whether it be purely physiological and unconscious or entirely conscious – is acquired during a stressful or traumatic experience. In the most common model of fear learning, fear of a stimulus or a situation is maintained via a process called associative learning, where an aversive experience in the past may provoke fearful responses to an otherwise benign stimulus or situation well into the future (LeDoux, 2014).

In PTSD, fear learning occurs at the time of trauma and sustained fear of situations, stimuli and other reminders associated with the trauma contribute to the bulk of PTSD symptoms, including hyperarousal, situational avoidance and intrusive memories (LeDoux, 2014; Milad

& Quirk, 2012; Shechner, Hong, Britton, Pine, & Fox, 2014). In contrast to fear learning,

“fear extinction learning” describes the process by which a previously threatening or aversive situation or stimulus is learnt to be no longer threatening or aversive; this occurs through multiple experiences where the situation or stimulus is not paired with any aversive outcome.

Fear extinction learning, then, describes the extinction of the fear memory and is the result of safety learning superseding the fear memory (Bouton, 2004; Milad & Quirk, 2012).

Similarly, “fear extinction recall” refers to the ability to retain fear extinction learning over a prolonged period of time; often the fear memory will resurface (known as return of fear) and the extinction memory must be learnt again. ACCEPTED MANUSCRIPT Critically, it is proposed that both fear extinction learning and fear extinction recall are inversely compromised in PTSD (Parsons & Hurd, 2015; Pitman et al., 2012; Zuj, Palmer,

Lommen, et al., 2016). Firstly, standardised paradigms designed to test the efficacy of fear extinction processes have shown that PTSD subjects display poorer fear extinction learning and recall than controls (Milad et al., 2009; Norrholm et al., 2011; Shvil et al., 2014; Zuj,

ACCEPTED MANUSCRIPT 6

Palmer, Hsu, et al., 2016). Secondly, efficacy of fear extinction prior to trauma exposure is correlated with the severity of PTSD symptoms post-trauma (Guthrie & Bryant, 2006;

Lommen, Engelhard, Sijbrandij, van den Hout, & Hermans, 2013). Finally, in the classical model of fear extinction learning, extinction memories are formed via a neural circuit involving the ventral medial prefrontal cortex (vmPFC), hippocampus and amygdala

(LeDoux, 2014; LeDoux, Iwata, Cicchetti, & Reis, 1988; Phelps, Delgado, Nearing, &

LeDoux, 2004). In this model, fear extinction is achieved by successful inhibitory signals from the higher-order vmPFC region attenuating fear signals from the amygdala. Similarly, contextual information of the fear memory stored in the hippocampus must also be over- ridden by extinction learning. Notably, it has been demonstrated using fMRI in humans that fear extinction is most successful with increased vmPFC volume and activation (Milad et al.,

2009; Milad et al., 2005; Milad et al., 2007). Similarly, vmPFC and hippocampus activity is higher in healthy controls compared to PTSD patients during extinction recall (Milad, Pitman, et al., 2009). Moreover, cortisol reactivity is essential for fear extinction learning (Rodrigues,

LeDoux, & Sapolsky, 2009), which is important given the lowered levels of cortisol production in PTSD subjects (Pitman et al., 2012; Yehuda, 2009).

1.2 The

Recently, the endogenous cannabinoid (endocannabinoid) system has been recognised as a promising target for PTSD treatment (Hill, Campolongo, Yehuda, & Patel, 2018; ACCEPTED MANUSCRIPT Neumeister, Seidel, Ragen, & Pietrzak, 2015; Steenkamp, Blessing, Galatzer-Levy, Hollahan,

& Anderson, 2017; Trezza & Campolongo, 2013; Zer-Aviv, Segev, & Akirav, 2016). The endocannabinoid system is comprised of distinct cannabinoid receptors (CB1 and CB2), as well as endogenous cannabinoid ligands, namely arachidonoyl ethanolamine (AEA; Devane, et al. 1992) and 2-arachidonoyl glycerol (2-AG; Mechoulam et al. 1995; Suiguira et al., 1995).

ACCEPTED MANUSCRIPT 7

AEA and 2-AG are derivatives of , and are metabolised by fatty acid amide hydrolase (FAAH) and (MAGL), respectively (Howlett et al., 2002).

Exogenous phytocannabinoids, such as delta9- (THC) and

(CBD) found in cannabis, are also effectors of the endocannabinoid system; though CBD is thought to exert its effects primarily through non-endocannabinoid pathways such as serotonin and adenosine receptors (McPartland, Duncan, Di Marzo, & Pertwee, 2015).

Cannabinoid receptors are profusely located across both the central and peripheral nervous systems and are heavily involved in many neuro-modulatory functions through retrograde signalling (Howlett et al., 2002; Kano, Ohno-Shosaku, Hashimotodani, Uchigashima, &

Watanabe, 2009). In addition to the cannabinoid receptors, cannabinoid ligands also show affinity for transient potential vanilloid (TPRV) channels, G-protein coupled receptors 55, 18 and 119 (GPR55, GPR18, GPR119), glycine receptors and peroxisome proliferator-activated receptors, as well as other targets such as the opioid and serotonin systems (De Petrocellis & Di Marzo, 2010; Ligresti, De Petrocellis, & Di Marzo, 2016;

Morales, Hurst, & Reggio, 2017). The endocannabinoid system is involved in a plethora of functions, including sleeping, eating, immune function, pleasure and emotional homeostasis

(Kano et al., 2009; Mechoulam & Parker, 2013).

1.2.1 Endocannabinoids and PTSD

Recently, the endocannabinoid system has also been recognised as central to stress ACCEPTED MANUSCRIPT responses, emotional memories and fear extinction (Hill et al., 2018; Morena, Patel, Bains, &

Hill, 2016; Ney, Matthews, Bruno, & Felmingham, 2018), which represent important therapeutic targets. AEA levels tonically gate the stress response through the HPA axis, and rapidly decrease following acute stress via FAAH activation (Gray et al., 2015; Gunduz-

Cinar et al., 2013; Hill et al., 2009; Morena et al., 2016; Natividad et al., 2017), allowing

ACCEPTED MANUSCRIPT 8 activation of the HPA stress response. The resulting increase in glucocorticoids mobilizes 2-

AG, which mediates glucocorticoid-driven negative feedback and termination of the stress response (Bedse et al., 2017; Di et al., 2016; Evanson, Tasker, Hill, Hillard, & Herman, 2010;

Hill et al., 2011; Morena et al., 2016). Hence, the endocannabinoid system is a crucial mediator of the HPA response under stress. Consequently, symptoms in PTSD that are a product of maladaptive stress responding such as intrusive memories, hyperarousal and poor sleep quality also appear to be facilitated by decreased endocannabinoid signalling (Hill et al.,

2018). Most importantly, preclinical models show that fear extinction – central to exposure therapy itself – is facilitated by enhanced endocannabinoid signalling and reduced by endocannabinoid inhibition (Aisenberg, Serova, Sabban, & Akirav, 2017; Bitencourt,

Pamplona, & Takahashi, 2008; Do Monte, Souza, Bitencourt, Kroon, & Takahashi, 2013).

Accordingly, PTSD has been associated with decreased endocannabinoid concentrations in war veterans (Wilker et al., 2016) and following the World Trade Centre attacks (Hill et al.,

2013), implying that endocannabinoids are involved in trauma responses in humans.

Although the science concerning the contribution of endocannabinoids to PTSD is robust and well-accepted (reviewed in Hill, et al. 2017; Morena et al. 2016; Ney, et al. 2018), utilisation of this knowledge is still in its beginning stages. For instance, there is to date only one published double-blind, cross-over randomised controlled trial where ten PTSD patients were given to improve nightmare symptomology (Jetly, Heber, Fraser, & Boisvert,

2015). Similarly, only a few other studies have administered cannabinoids in open-label ACCEPTED MANUSCRIPT prospective (Roitman, Mechoulam, Cooper-Kazaz, & Shalev, 2014) or retrospective studies to PTSD patients (Cameron, Watson, & Robinson, 2014; Elms, Shannon, Hughes, & Lewis,

2018; Fraser, 2009), or in experimental studies to healthy participants (Das et al., 2013;

Klumpers et al., 2012; Rabinak et al., 2014; Rabinak et al., 2013). Generally, the results of studies involving standardized extinction learning and recall tasks in healthy humans have

ACCEPTED MANUSCRIPT 9 been promising. However, there is a current need for randomised controlled trials where

PTSD patients’ symptomology is measured during and after cannabinoid therapy.

1.2.2 Challenges to Cannabinoid-Based PTSD Treatment

Aside from the tentative legality of cannabis-based products for medical purposes in most of the Western world, there are several barriers to testing therapeutic outcomes of cannabinoids for PTSD patients. Firstly, there are demonstrable adverse health risks associated with whole-plant cannabis use, with chronic recreational use associated with dependence and decreased mental health across a number of domains including cognition and risk of psychosis (Hall & Degenhardt, 2009; Moore et al., 2007). Similarly, chronic use of marijuana is associated with global cortical downregulation of CB1 receptor availability, as well as across regions such as the temporal lobe, nucleus accumbens and cingulate cortex

(Ceccarini et al., 2015; Hirvonen et al., 2012). Downregulation of cannabinoid receptors results in reduced endogenous cannabinoid functioning, reflecting the tolerance and dependence effect chronic use on the endogenous system. Increased understanding of appropriate dosing levels must be further established before definitive clinical trials can begin, and this may increasingly reflect a move away from whole plant products due to difficulty in regulating levels of the active compounds in this form of the product. Further, the endocannabinoid system acts in different ways even in the same disease states (Di Marzo,

2008), meaning that for many conditions the direction in which the system should be ACCEPTED MANUSCRIPT modulated is unclear and upstream options for treatment should also be researched. There is potential that this problem might ultimately reflect endocannabinoids being a downstream focus for PTSD, as in schizophrenia where the traditional dopamine hypothesis as a treatment point is more recently being challenged by upstream targets such as glutamate (Howes &

Kapur, 2009; Moghaddam & Javitt, 2012).

ACCEPTED MANUSCRIPT 10

1.3 Focus of this Review

There is also indecision amongst researchers as to the best pharmacological avenue to pursue. Modulation of the endocannabinoid system can theoretically occur through multiple means, and each of these methods has been reviewed previously (Di Marzo, 2008; Micale, Di

Marzo, Sulcova, Wotjak, & Drago, 2013). For the remainder of this review, we will consider and compare each of the main ways in which pharmaceuticals may modulate the endocannabinoid system in a PTSD context, with reference to the strengths and weaknesses of each approach. These approaches will be limited to the primary methods used in human and animal research to achieve direct or indirect agonism of cannabinoid receptors and inhibition of degradation or reuptake of active cannabinoid ligands. Previous reviews have collated studies in the broader spectrum of anxiety research (Di Marzo, 2008; Micale et al.,

2013) as well as in PTSD (Berardi, Schelling, & Campolongo, 2016), though there have been several important developments in recent years that may have changed the direction in which this field is going. This review is therefore not intended to be an exhaustive review of the studies conducted using each methodology, or by any means of the chemistry or pharmacological studies underlying each technique, but will provide essential information and updates on types of cannabinoid treatments under development specifically for PTSD.

Interested readers should note that previous reviews (Berardi, et al. 2016; Micale, et al. 2013) outline a wider range of neurochemical interactions between the endocannabinoid and ACCEPTED MANUSCRIPT various other systems in PTSD mechanisms and symptoms; here we restrict our review of the literature to the most recent findings. Finally, there are many specialized and specific processes involved in memory in PTSD which have been comprehensively reviewed previously (eg. Brewin, 2011), but given that endocannabinoid research has largely focused

ACCEPTED MANUSCRIPT 11 on fear extinction recall and emotional memory consolidation most of our review relates to these processes.

1.4 Method

Studies were selected for this review during several systematic searches of electronic databases Web of Science, PubMed and PsycINFO. Keywords searched were “cannabinoid”,

“fear”, “FAAH”, “extinction”, “endocannabinoid”. “MAGL”, “cannabidiol”, “THC”,

“cannabis”, tetrahydrocannabinol”, “trauma” or “PTSD” and searches were restricted to

January 2016 until the present date. We only included papers that were printed in English.

The reference lists of eligible papers were checked for additional relevant studies. Studies that did not address mechanisms directly relevant to PTSD were excluded.

2. Multiple Avenues for Endocannabinoid Modulation

2.1 THC and its recently identified effects on PTSD symptomology

The most obvious way to affect the endocannabinoid system is through direct agonism of cannabinoid receptors. THC is a with high affinity to both CB1 and CB2 receptors, and, despite having multiple other molecular targets, is thought to exert most of its effects through these receptors (Ligresti et al., 2016). Similar to endogenous cannabinoid agonists, THC exerts its effect through CB1 receptors via presynaptic inhibition

(Kano et al., 2009), though it is only one of the two phytocannabinoids that directly activates ACCEPTED MANUSCRIPT the cannabinoid receptors, with the other being (Ligresti et al., 2016; Showalter,

Compton, Martin, & Abood, 1996). As a result, carriers of the polymorphism C385_A, implicated in lower levels of the FAAH enzyme and therefore higher levels of native cannabinoid ligands, are four times less likely to be dependent on THC due to the increased presence of endocannabinoids in these people (Tyndale, Payne, Gerber, & Sipe, 2007).

ACCEPTED MANUSCRIPT 12

Several synthetic analogues of THC have been developed for therapeutic purposes, with (Marinol, Solvay Pharmaceuticals) and nabilone (Cesamet, Valeant

Pharmaceuticals) licensed for the treatment of nausea and vomiting caused by chemotherapy, and Sativex (GW Pharmaceuticals) for the treatment of late-stage cancer pain and spasticity in multiple sclerosis patients (Ligresti et al., 2016). While dronabinol and nabilone are synthetic THC analogues, Sativex is 1:1 THC:CBD.

THC has previously been shown to improve performance on fear extinction and analogue emotional memory tasks in animals, as well as reducing the stress response and promoting anti-anxiety behaviours at mild to moderate doses (Berardi et al., 2016; Micale et al., 2013). The most recent animal studies have continued to show this, with THC reducing anxiety produced by 2-AG synthesis inhibition (Bedse et al., 2017), by decreasing fearful behaviours to conditioned stimuli through medial prefrontal cortex (mPFC) CB1-dependent disruption of memory reconsolidation (Stern et al., 2015) and by reducing acquisition of fear learning (Ruhl, Zeymer, & von der Emde, 2017) (Table 1).

In humans, THC has been shown to improve fear extinction recall in healthy participants compared to placebo (Rabinak et al., 2013), possibly by bolstering the vmPFC and hippocampal responses during extinction recall as shown by fMRI (Rabinak et al., 2014).

Marketed versions or synthetic analogues of THC have also improved PTSD symptoms, with nabilone and THC improving nightmares, sleep quality, hyperarousal and overall symptomology in PTSD patients (Cameron et al., 2014; Fraser, 2009; Jetly et al., 2015; ACCEPTED MANUSCRIPT Roitman et al., 2014). However, THC administered prior to memory retrieval was recently found to increase false recollection of emotional and non-emotional stimuli in healthy participants (Doss, Weafer, Gallo, & de Wit, 2018), suggesting that distortion, rather than reconsolidation, of memories may be an unwanted side-effect of THC treatment. Further,

THC administration to rat nucleus accumbens increased salience of fear memories, which

ACCEPTED MANUSCRIPT 13 further suggests that THC may in some situations contribute to unwanted memory traces

(Fitoussi, Zunder, Tan, & Laviolette, 2018). Interestingly, the effects of THC were dopamine-dependent, such that antagonism on nucleus accumbens dopamine receptors completely blocked this effect (Fitoussi et al., 2018). This suggests that future clinical studies should focus on discerning which aspects of symptomology with be improved or worsened given cannabinoid treatment. Randomised clinical trials using THC, dronabinol or nabilone for PTSD patients are currently being conducted (NCT03008005, NCT02759185,

NCT02069366, NCT02517424 and NCT03251326).

2.1.1 Synthetic agonists and recently identified effects on PTSD symptomology

THC is a partial, non-selective agonist of CB1 and CB2 receptors, and has over time become less stigmatised as a tool for research purposes (Ligresti et al., 2016). Consequently, endocannabinoid research could not have possibly evolved without the development of potent agonists selective to CB1 and/or CB2 receptors (Pertwee, 2006). Many such agonists have been developed, with some of the most notable being WIN55,212-2 (Compton, Gold, Ward,

Balster, & Martin, 1992), HU-210 (Mechoulam et al., 1987) and CP-55,940 (Howlett,

Champion, Wilken, & Mechoulam, 1990), which are salient due to their continued and frequent use in preclinical research (Figure 1). The favourable effects of these potent agonists in animal models of traumatic stress have been previously reviewed (Berardi et al.,

2016; Micale et al., 2013). ACCEPTED MANUSCRIPT

Insert Figure 1 about here

The most recent studies have continued to show that WIN55,212-2 and CP-55,940 produce anxiolytic effects in rodents (Bedse et al., 2017; Burstein, Shoshan, Doron, & Akirav,

ACCEPTED MANUSCRIPT 14

2018; Faraji, Komaki, & Salehi, 2017; Lisboa et al., 2018; Zubedat & Akirav, 2017). These recent studies have contributed to our understanding of how cannabinoids might influence

PTSD mechanisms, beyond what has been previously reported. Firstly, whereas mice susceptible to acute stress showed increased metabotropic glutamate receptor 5 levels in the hippocampus when acutely administered vehicle, acute dosing of WIN55,212-2 regulated these receptor densities, presumably reversing the stress-susceptible phenotype (Sun et al.,

2017). Both WIN55,212-2 and CP-55,940 have also recently been shown to improve fear extinction learning (Burstein et al., 2018; Ghasemi, Abrari, Goudarzi, & Rashidy-Pour, 2017;

Korem, Lange, Hillard, & Akirav, 2017; Lisboa et al., 2018; Sachser, Crestani, Quillfeldt,

Souza, & de Oliveira Alvares, 2015; Shoshan & Akirav, 2017; Zubedat & Akirav, 2017), and to reduce fear memories more generally (Nasehi, Shahbazzadeh, Ebrahimi-Ghiri, &

Zarrindast, 2018; Santana et al., 2016; Shoshan, Segev, Abush, Mizrachi Zer-Aviv, & Akirav,

2017). In accordance with past literature (Berardi et al., 2016), these effects were enhanced by glucocorticoid receptor antagonism and mediated by hippocampal cornu ammonis 1 (CA1) and basolateral amygdala (BLA) CB1 receptors (Santana et al., 2016; Shoshan & Akirav,

2017; Zubedat & Akirav, 2017), and may also involve BLA β1-adrenoceptors (Nasehi,

Shahbazzadeh, et al., 2018). These memory effects may be caused by improved neuroplasticity, shown when WIN55,212-2 reversed the effects of acute stress on long-term potentiation pathways in the nucleus accumbens and CA1 (Segev & Akirav, 2016; Shoshan et al., 2017). ACCEPTED MANUSCRIPT Interestingly, antagonism of 5-HT4 receptors reduced or enhanced the ameliorative effects of CB1 agonism on fear conditioning at higher or lower doses of 5-HT4 antagonism, respectively (Nasehi, Farrahizadeh, Ebrahimi-Ghiri, & Zarrindast, 2016). Further,

WIN55,212-2 decreased hippocampal and PFC brain-derived neurotrophic factor (BDNF) concentrations, which were associated with PTSD symptomology in a rodent model of the

ACCEPTED MANUSCRIPT 15 disorder (Burstein et al., 2018; Korem et al., 2017). Most recently, CB2 CA1 agonism using

GP1a also had an impairing effect on passive avoidance memory consolidation in male mice, implicating a potential role for the peripheral cannabinoid system as well (Nasehi, Gerami-

Majd, Khakpai, & Zarrindast, 2018). Finally, the beneficial effect of WIN55,212-2 on fear extinction and anxiety behaviour was associated with reduced interleukin 1β mRNA following repeated social defeat stress and contextual fear conditioning, suggesting that cannabinoids act on fear and stress partly by reducing neuroinflammation (Lisboa et al.,

2018). Whilst these findings agree with the growing view of an expanded endocannabinoid system (Berardi et al., 2016; Hill et al., 2018; Morena et al., 2016; Ney et al., 2018), they also contribute valuable new insight into the molecular interactions between BDNF, endocannabinoid and serotonin systems during fear extinction. Combined with recent research identifying interactions between the dopamine and endocannabinoid (Fitoussi et al.,

2018; Wenzel et al., 2018) systems during fear memory consolidation and extinction, understanding of endocannabinoid interactions highlights the importance of examining dopaminergic and similar classical memory and cognition circuits in the context of PTSD

(Abraham, Neve, & Lattal, 2014; Lee, Wang, & Tsien, 2016).

2.1.2 Disadvantages and status of direct agonists of the endocannabinoid system

There are several disadvantages to directly stimulating cannabinoid receptors pharmacologically. Firstly, direct agonism of cannabinoid receptors displays marked ACCEPTED MANUSCRIPT biphasic effects. For instance, high doses of THC and other agonists result in anxiogenic, rather than anxiolytic, effects in rodents (Fitoussi et al., 2018; Micale et al., 2013; Patel & Hillard, 2006; Rey, Purrio, Viveros, & Lutz, 2012; Sbarski & Akirav, 2018;

Todd & Arnold, 2016), and in humans (7.5mg compared to 12.5mg oral THC; Childs, Lutz,

& de Wit, 2017). Heavy use of whole plant marijuana in humans can also result in

ACCEPTED MANUSCRIPT 16 psychiatric symptoms such as depression, cognitive impairment and psychosis (Hall &

Degenhardt, 2009), as well as decreases in memory capacity compared to lower doses

(Calabrese & Rubio-Casillas, 2018). Since synthetic cannabinoid agonists are considerably more potent than THC, these effects may potentially occur even at low doses in humans

(Castaneto et al., 2014). Further, effective treatment of PTSD is likely to require ongoing medication until symptoms are fully alleviated, meaning that chronic dosing will likely be necessary. that are recreationally available, such as “Spice”, have high efficiency as well as potency and are reported in numerous case studies and anecdotal reports to have far stronger physiological and psychological effects than THC (Spaderna,

Addy, & D'Souza, 2013). There is also evidence suggesting that these enhanced effects are accompanied by higher rates of adverse events, such as increased rates of physical insults such as tachycardia, as well as increased risk of mental health problems such as psychotic features (Bush & Woodwell, 2013; Castaneto et al., 2014; Fattore, 2016; Forrester,

Kleinschmidt, Schwarz, & Young, 2011; Hoyte et al., 2012; Spaderna et al., 2013).

Heavy cannabis use may also result in downregulation of endogenous cannabinoid signalling and decreased performance on fear extinction tasks in humans, which presumably occurs through sustained reliance on exogenous signalling (Hirvonen et al., 2012; McPartland,

Guy, & Di Marzo, 2014; Papini et al., 2017). Importantly, recent naturalistic studies of cannabis use in PTSD have reported null (Johnson et al., 2016) or negative (Wilkinson,

Stefanovics, & Rosenheck, 2015) associations between cannabis use and symptom severity, ACCEPTED MANUSCRIPT despite patients using the drug for temporary relief. Conversely, at therapeutic doses nabilone and dronabinol appear to be generally well tolerated with few severe side effects

(U.S. Food and Drug Administration, 2004, 2006) though have not been thoroughly tested in humans with PTSD despite promising early trials in patient populations (Cameron et al., 2014;

Jetly et al., 2015). These reports therefore highlight the importance of conducting safety

ACCEPTED MANUSCRIPT 17 trials for synthetic cannabinoid agonists and trialling cannabinoids in PTSD at controlled doses despite a discouraging literature surrounding the effects of whole-plant marijuana.

Perhaps the greatest problem with direct stimulation of cannabinoid receptors is the involvement of these receptors, particularly CB1, in producing unwanted intoxicant effects in areas that are not a therapeutic target. This problem is evident in animal studies where administration to one region, such as the mPFC with THC, can produce therapeutic effects such as reduced fear responding and anxiety, whereas administration to another region, such as the nucleus accumbens with THC, can produce negative effects such as potentiating fear reactivity (Fitoussi et al., 2018; Stern et al., 2015). Activation of cannabinoid receptors in the brain can produce intoxicant effects, which alter emotions, thought processes and perception

(Pertwee, 2006). Intoxicant effects are sought by many people for recreational purposes, meaning that potent cannabinoid agonists have high potential for misuse (Ware, Martel,

Jovey, Lynch, & Singer, 2018). For these reasons, there is a long-held hesitancy in advocating for chronic use of cannabinoid agonists for medical conditions, particularly psychiatric conditions and conditions comorbid with substance use disorders where patients are likely to seek intoxicant effects (Table 1). In the pharmacological community, attention towards treating PTSD – a psychiatric disorder marked by high rates of comorbid substance misuse – has instead turned to the development of indirect agonists that are unlikely to produce intoxicant effects, which will be reviewed next. ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 18

Table 1. Direct Agonists of the endocannabinoid system

Recognised clinical Newly recognised mechanisms of Advantages Disadvantages Status

effects in PTSD action in PTSD processes

Delta9- Reduces nightmares Role of NAc dopamine in Easy to manufacture Unwanted effects (eg. Being tested in humans, tetrahydro Improves sleep modulating susceptibility to acute Effective in PTSD subjects intoxicant effects) some synthetic forms cannabinol quality stress Some understanding of health Tolerance and dependence due approved for some

Improved fear effects to downregulation of indications in some

extinction recall# Partial agonist at CB1 and endocannabinoid system countries

CB2 Potential for misuse

Synthetic Not tested in Decreasing HC and PFC BDNF More potent agonist compared Central side effects (off target) Untested in humans cannabinoid humans due to lack Biphasic interaction with 5-HT4 to THC Tolerance and dependence due agonists of safety studies NAc β-catenin-dependent Reliable PTSD-related health to downregulation of

WIN55,212-2; Via CB2 in CA1 outcomes in animals endocannabinoid system

CP-55,940; Interactions with BLA β1- Potential for misuse

HU-210; adrenoceptors Health hazard for recreational ACCEPTEDVia reduced neuroinflammation MANUSCRIPTusers (eg. HU-210 in Spice) Note: NAc = nucleus accumbens, HC = hippocampus, PFC = prefrontal cortex, BDNF = brain-derived neurotrophic factor, CB2 = cannabinoid receptor type 2, CA1 = cornu ammonis-1, BLA = basolateral amygdala #Tested only in healthy humans in fear extinction studies at time of publication

ACCEPTED MANUSCRIPT 19

2.2 Inhibition of Degradation – Enzyme and Transport Inhibition

Another way of enhancing endocannabinoid signalling is through inhibition of the catabolic enzymes FAAH and MAGL (Micale et al., 2013). MAGL is the primary metabolic enzyme for 2-AG and FAAH is primarily responsible for metabolism of AEA, though also plays a small part in 2-AG hydrolysis (Howlett et al., 2002). FAAH belongs to the amidase signature family, which are part of the serine hydrolase superfamily (McKinney & Cravatt,

2005). Similarly, MAGL is part of the serine hydrolase superfamily, though 2-AG is hydrolysed through multiple enzymes, with MAGL being the primary one (Murataeva,

Straiker, & Mackie, 2014). Theoretically, inhibition of active ligand degradation will occur locally, which is in contrast to direct agonists of cannabinoid receptors which will tend to produce global effects (Di Marzo, 2009; Kano et al., 2009). This is because endogenous cannabinoids such as AEA and 2-AG are produced on demand and will only be synthesised at locations local to insult (Kano et al., 2009). FAAH inhibition can therefore theoretically target specific pathways and regions, and should not produce unwanted intoxicant and other like side-effects (Bisogno & Maccarrone, 2013).

Multiple FAAH inhibitors have been developed and these vary in FAAH enzyme selectivity and efficacy. URB597 is a potent and selective inhibitor of FAAH (Fegley et al.,

2005; Kathuria et al., 2003) and has been the most popular choice for anxiety and PTSD research though has not progressed to research in humans. Three FAAH inhibitors (SSR-

411298, JNJ-42165279 and PF-04457845) have progressed to Phase I human trials, and two ACCEPTED MANUSCRIPT (PF-04457845 and SSR-411298) to Phase II trials (Ahn et al., 2011; Johnson et al., 2011;

Keith et al., 2015; Postnov et al., 2018). In the latter trials (PF-04457845 for knee pain and

SSR-411298 for depression) there was no benefit in humans despite promising animal data

(Huggins, Smart, Langman, Taylor, & Young, 2012; Sanofi, 2013). New inhibitors are being

ACCEPTED MANUSCRIPT 20 continuously developed, including a dual MAGL/FAAH inhibitor (JZL195) (Long et al.,

2009) which may act to increase local levels of both AEA and 2-AG.

Insert Figure 2 about here

2.2.1 Recently identified effects of catabolic inhibition of endocannabinoids on PTSD symptomology

Generally speaking, past reviews have found that FAAH inhibition favourably enhances fear extinction and other processes relevant to PTSD (Berardi et al., 2016; Micale et al., 2013). FAAH inhibitors, but not the MAGL inhibitor JZL184, have been found to decrease corticotrophin-releasing hormone mRNA in the hypothalamus and pituitary following stress (Bedse et al., 2014) as well as reducing stress and anxiety behaviours in rodents (Bedse et al., 2018; Danandeh et al., 2018; Griebel et al., 2018; Heinz, Genewsky, &

Wotjak, 2017; Sartim, Moreira, & Joca, 2017), though it did not reduce corticosterone levels

(Bedse et al., 2014; Griebel et al., 2018; Roberts, Stuhr, Hutz, Raff, & Hillard, 2014; Shoshan

& Akirav, 2017). One recent study showed, however, that the FAAH inhibitor URB597 did reduce corticosterone responses to immobilisation stress in young rats, but only those who had been previously exposed to stress during the neonatal stage of life. This may suggest that

FAAH inhibition may preferentially reduce stress responses in those who had been exposed to early life trauma and have a consequent developmentally-founded stress phenotype ACCEPTED MANUSCRIPT (McLaughlin, Verlezza, Gray, Hill, & Walker, 2016). Interestingly, the MAGL/FAAH inhibitor JZL195 did not decrease anxiety behaviour, whereas MAGL and FAAH inhibitors

JZL184 and URB597 alone did, suggesting that the potential of dual FAAH/MAGL inhibitors for PTSD may be limited (Table 2) (Bedse et al., 2018).

ACCEPTED MANUSCRIPT 21

Recent studies have also continued to show that URB597 promotes favourable outcomes in aversive conditioning paradigms (Aisenberg et al., 2017; Burstein et al., 2018;

Fidelman, Mizrachi Zer-Aviv, Lange, Hillard, & Akirav, 2018; Morena et al., 2017; Segev et al., 2018; Zer-Aviv & Akirav, 2016), as do FAAH inhibitors AM3506 (Gunduz-Cinar, et al.,

2013) and OL-135 (Bowers & Ressler, 2015; Burman et al., 2016). Recent studies have also identified that the effects of FAAH inhibition on extinction memory occur in the BLA, mPFC and CA1 through normalisation of CB1 receptors, as well as enhanced long-term potentiation and normalised BDNF levels, which is a similar mechanism to that observed in direct CB1 agonists (Aisenberg et al., 2017; Burstein et al., 2018; Fidelman et al., 2018; Sartim et al.,

2017; Segev et al., 2018; Zer-Aviv & Akirav, 2016). The effect of FAAH inhibition on memory consolidation of stressful events more generally is inconsistent, with some studies finding FAAH inhibition enhances memory traces (Morena et al., 2014; Ratano, Palmery,

Trezza, & Campolongo, 2017) and others finding it reduces memory (Aisenberg et al., 2017).

One recent study found that URB597 improved impaired hippocampal long-term potentiation and performance on short-term memory tasks, and improved performance on overly expressive BLA-dependent memory tasks (Shoshan et al., 2017). At this stage, the role in memory consolidation of FAAH inhibition is ambiguous, as directionality appears highly sensitive to context, dosage, brain location and other factors (Morena & Campolongo, 2014).

2.2.2 Alternatives to FAAH inhibition ACCEPTED MANUSCRIPT As FAAH degrades not just AEA but also other ethanolomides that have discrete molecular targets apart from the endocannabinoid system, FAAH inhibition has the potential to have off-target effects. For example, TRPV1 channels are activated by – part of the endocannabinoid ligand family – and are recognised to produce anxiety when activated in rodents (Faraji et al., 2017). For this reason, TRPV1

ACCEPTED MANUSCRIPT 22 channels are recognised to be one of the unwanted targets incidentally activated during

FAAH inhibition using some inhibitors, such as URB597 (Bisogno et al., 1998; Micale et al.,

2009). An approach to targeting the endocannabinoid system that counteracts this problem is by using N-arachidonoyl-serotonin (AA-5-HT), which is a dual inhibitor of FAAH and

TRPV1 channels. AA-5-HT administered intraperitoneally, to the mPFC or to the BLA of rodents reduced anxiety (John & Currie, 2012; Micale et al., 2009) and stress responses

(Sartim et al., 2017), and was more recently found to reduce fear memories when administered to the dorsal hippocampus (Gobira et al., 2017). Unfortunately, studies using this promising compound have been sparse, despite recent literature suggesting that low

CB1:TRPV1 ratio potentiates fear learning (Back & Carobrez, 2018).

Recent studies have also begun examining the effect of MAGL inhibitors such as

JZL184, and have generally found anxiolytic effects (Aliczki et al., 2013; Almeida-Santos,

Moreira, Guimaraes, & Aguiar, 2017; Bedse et al., 2018; Bedse et al., 2017; Bluett et al.,

2017; Bosch-Bouju, Larrieu, Linders, Manzoni, & Laye, 2016; Morena et al., 2017), but see

(Heinz et al., 2017). Limited PTSD-oriented studies have found decreased stress responses following MAGL inhibition (Bluett et al., 2017; Roberts et al., 2014) along with decreased memory consolidation following stress (Morena et al., 2015), though possibly in anti- extinction learning effects as well (Hartley et al., 2016; Llorente-Berzal et al., 2015). MAGL inhibition is a rapidly growing area of research (Aghazadeh Tabrizi et al., 2018; Granchi,

Caligiuri, Minutolo, Rizzolio, & Tuccinardi, 2017) and it will become increasingly clear over ACCEPTED MANUSCRIPT the next few years how 2-AG augmentation may benefit therapeutic outcomes, with the first

Phase I clinical trials currently being conducted (NCT03447756).

2.2.3 Recently identified effects of reuptake inhibition of endocannabinoids on PTSD symptomology

ACCEPTED MANUSCRIPT 23

Similar to catabolic enzyme inhibitors, previous research on the effect of cellular reuptake inhibitors with respect to PTSD and other psychiatric disorders have been reported

(Berardi et al., 2016; Micale et al., 2013). As the mechanisms for cellular uptake of the endocannabinoids is poorly understood (Nicolussi & Gertsch, 2015), relatively little research has yet been conducted using reuptake inhibitors. Reuptake inhibitors of endocannabinoids share a similar mechanism to other drugs, such as selective serotonin reuptake inhibitors, whereby cellular transport of the active ligands after exerting their effects is inhibited, which delays reuptake. AM404 (Beltramo et al., 1997) is the most commonly used endocannabinoid reuptake inhibitor, though it primarily deactivates AEA transporters through poorly understood mechanisms (Nicolussi & Gertsch, 2015), and has off-target effects at

TRPV1 as well as cyclooxygenase 1 and cyclooxygenase 2 for which activation can result in increased anxiety and reduced recovery from PTSD symptomology in rodents (Faraji et al.,

2017; Gamble-George et al., 2016; Hogestatt et al., 2005; Wang et al., 2018). Other reuptake inhibitors, such as VDM11, AM1172 or the recently synthesized and highly selective AEA and 2-AG transport inhibitor WOBE437 (Reynoso-Moreno, Chicca, Flores-Soto, Viveros-

Paredes, & Gertsch, 2018), are yet to garner attention in PTSD research. Regardless, the most recent studies using AM404 found that administration in rodent paradigms reduced the expression of fear (Llorente-Berzal et al., 2015) and potentiated learning of safety cues, an effect that was dependent on CB1 receptor integrity in the CA1 (Micale et al., 2017).

Therefore, AM404 exerts anxiolytic effects through the cannabinoid system, though the off- ACCEPTED MANUSCRIPT target effects are also well documented. Importantly, recent developments have seen the synthesis of increasingly selective and potent endocannabinoid reuptake inhibitors, which may be a promising target for therapeutic modulation in the near future (Chicca et al., 2017).

2.2.4 Disadvantages and status of catabolic/reuptake inhibitors of endocannabinoids

ACCEPTED MANUSCRIPT 24

Despite the fact that these studies agree with past reports on the potential efficacy of endocannabinoid hydrolysis inhibition for PTSD treatment (Berardi et al., 2016; Micale et al.,

2013), recent developments have sparked concern in the safety of this approach (Table 2).

Full assessment of the safety and efficacy of preclinical MAGL and FAAH inhibitors is difficult to achieve for a number of reasons. Firstly, the potential for off-target effects, due to the formulation of most of these compounds from existing serine hydrolase inhibitors, is problematic to establish due to the size of the serine hydrolase superfamily (Bisogno &

Maccarrone, 2013; Di Marzo, 2008; Zhang et al., 2007). For instance, URB597, which is most commonly utilised in animal research, displayed off-target effects for triacylglycerol hydrolase of at least the same potency as FAAH (Lichtman et al., 2004), making it unsuitable for use in humans. Further, FAAH in humans and animals displays structural differences (Di

Venere et al., 2012), which may explain why some highly efficacious inhibitors developed for animals are ineffective in humans (Bisogno & Maccarrone, 2013). In a sparse and growing literature, one recent study found the MAGL inhibition resulted in negative consequences for memory and learning in rodents, suggesting that the safety profile of such compounds may be poor (Griebel et al., 2015).

Unfortunately, a recent Phase I clinical trial using the FAAH inhibitor BIA 10-2474 reported that five out of their six participants in a fifth cohort of the study had experienced serious adverse events, including one death (Kerbrat et al., 2016). Indeed, a subsequent study found that BIA 10-2474 targeted multiple serine hydrolases in addition to FAAH; many of ACCEPTED MANUSCRIPT which were lipophilic and therefore possible contributors to the neurological damage induced in participants (van Esbroeck et al., 2017). Participants of this particular trial tolerated BIA

10-2474 well at 20mg or 100mg acutely, but not 50mg over six days (Kerbrat et al., 2016).

This event represents a setback in the development of FAAH inhibitors for clinical purposes, as most current clinical trials using any related compounds have been halted, despite the US

ACCEPTED MANUSCRIPT 25

Food and Drug Administration ruling that FAAH inhibitors aside from BIA 10-2474 are likely to be safe (U.S. Food and Drug Administration, 2016). Such a serious adverse event advocates for more rigorous safety profiling prior to bringing new drugs to clinical trial, which will need to be increasingly rigorous for FAAH inhibitors given how easy it is to miss off-target effects (Bisogno & Maccarrone, 2013; Piscitelli & Di Marzo, 2012).

Further to this, a recent human study examined the frequency of the low-FAAH expressing allele of the single nucleotide polymorphism rs324420 in relation to adult depression and anxiety following childhood adversity (Lazary, Eszlari, Juhasz, & Bagdy,

2016). It was found in this study that higher adult anxiety and depression was associated with genetically reduced FAAH expression only if participants reported childhood adversity.

Therefore, despite other reports suggesting stress-resilient phenotypes in low-FAAH expressing rs324420 carriers (Dincheva et al., 2015; Spagnolo et al., 2016), more research is needed to understand under which conditions and for what populations FAAH inhibition will be efficacious.

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 26

Table 2. Degradation/reuptake inhibitors of endocannabinoid ligands

Effects in Mechanisms of Action (in Advantages Disadvantages Status

humans rodents)

FAAH inhibitors Not tested Decrease CRH mRNA Potentially more localised Binding with off-targets likely Due to SAEs in humans,

URB597, SSR- in PTSD Normalised BDNF, LTP and treatment FAAH does not only metabolise trials have been slowed.

411298, JNJ- research CB1 expression in BLA, Effective for PTSD AEA/common eCBs

42165279^, PF- mPFC and CA1 processes Available research suggests

04457845^, May avoid central side altered effectiveness in humans

AM3506, OL-135, effects by interacting with

BIA 10-2474 ‘on-demand’ nature of eCBs

Dual FAAH/TRPV1 Not tested Inhibition of FAAH and Potentially more localised Untested in humans Untested in humans, slow inhibitors in humans TRPV1 channels in mPFC, treatment Mixed effectiveness findings progress

AA-5-HT, BLA and dorsal HC Mostly effective for PTSD

OMDM198 Increased AEA and 2-AG processes

levels Avoids off-target ACCEPTEDenhancement MANUSCRIPT of TRPV1 MAGL inhibitors Not tested Increased HC 2-AG levels Reduces stress responses Possible aversive central health Due to possible negative

JZL-184 in humans Increased BLA 2-AG levels Potentially more localised effects health effects clinical trials

Involves GABA neurons treatment May impair fear extinction are not yet being widely

ACCEPTED MANUSCRIPT 27

Limited testing in humans conducted

Dual FAAH/MAGL Not tested Increased AEA and 2-AG Potentially more localised Untested in humans Promising in other areas, inhibitors in humans levels treatment Failed to reduce anxiety though the first PTSD

JZL-195 Enhancement of both AEA Limited testing in animals animal findings were

and 2-AG signalling negative

Reuptake inhibitors Not tested Increased AEA Potentially more localised Untested in humans AM404 is metabolite of

AM404, VDM11, in humans Anxiolytic effect was TRPV1 treatment Limited testing in animals (Sharma et al.,

AM1172, WOBE437 and CB1 dependent Preliminary results imply Poorly understood mechanisms 2017), implying possible

CB1 dependent in CA1 effective for PTSD Off-target at TRPV1 safety for humans

processes

Note: ^indicates FAAH inhibitors that have progressed to clinical trial but have not been used in PTSD animal research.

AEA = arachidonoyl ethanolamine, 2-AG = 2-arachidonoyl glycerol, NAc = nucleus accumbens, HC = hippocampus, PFC = prefrontal cortex, BDNF = brain- derived neurotrophic factor, CB1 = cannabinoid receptor type 1, CA1 = cornu ammonis-1, BLA = basolateral amygdala, LTP = long-term potentiation, CRH = corticotropin-releasing hormone, eCB = endocannabinoids ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 28

2.3 Cannabidiol

The endocannabinoid system may also be stimulated indirectly. Cannabidiol (CBD) is one way of doing this (Figure 3). CBD is the primary non-intoxicant phytocannabinoid native to the plant. CBD has very low affinity for both CB1 and CB2 receptors, and exerts effects at moderate levels at multiple targets in both the central and peripheral nervous systems (Ligresti et al., 2016; McPartland et al., 2015). Many of CBD’s effects rely on the integrity of cannabinoid receptors (Bitencourt et al., 2008; Campos, Ortega, et al., 2013; Capasso et al., 2008), and CBD may possibly exert some of its effects through inhibition of the endocannabinoid catabolic enzymes (Bisogno et al., 2001; McPartland et al.,

2015). Therefore, CBD is believed to primarily activate the cannabinoid receptors indirectly and work through the endocannabinoid system, though there is evidence of affinity for multiple other molecular targets where this may be achieved (Ligresti et al., 2016;

Mechoulam, Peters, Murillo-Rodriguez, & Hanus, 2007). Firstly, it was established relatively early that CBD is an inverse agonist of both CB1 and CB2 receptors (Mechoulam et al., 2007), as it was shown to antagonise cannabinoid receptor agonists in vivo (Pertwee,

Ross, Craib, & Thomas, 2002). However, a recent meta-analysis recognised that CBD also enhances the effects of cannabinoid receptor agonists in some situations (McPartland et al.,

2015; Todd & Arnold, 2016). CBD also binds to and desensitises TRPV1 channels (Iannotti et al., 2014), which explains its anxiolytic action to some extent (Campos & Guimarães, 2009;

Ligresti et al., 2016). More likely, since CBD has some small affinity for the 5-HT ACCEPTED MANUSCRIPT 1A receptor (Campos & Guimaraes, 2008; McPartland et al., 2015; Russo, Burnett, Hall, &

Parker, 2005), activity through serotonergic pathways may explain part of its anxiolytic

(Campos, Moreira, Gomes, Del Bel, & Guimaraes, 2012; Lee, Bertoglio, Guimaraes, &

Stevenson, 2017) and antidepressant effects (Sales, Crestani, Guimaraes, & Joca, 2018).

CBD also exhibits anti-cancer, anti-epileptic and immune boosting properties, which are

ACCEPTED MANUSCRIPT 29

probably mediated by activity at GPR18, GPR55, α3 glycine and peroxisome proliferator- activated receptor gamma receptors, as well as inhibition of adenosine reuptake (Carrier,

Auchampach, & Hillard, 2006; Izzo, Borrelli, Capasso, Di Marzo, & Mechoulam, 2009;

Ligresti et al., 2016; McPartland et al., 2015). Therefore, as an inverse agonist with low affinity for cannabinoid receptors CBD has some direct influence on the endocannabinoid system, though is reported to primarily work as indirect agent on endocannabinoid functioning via independent paths.

Notably, CBD is recognised as having a relatively safe profile in doses up to 1500mg

(Zuardi, Morais, Guimaraes, & Mechoulam, 1995), and only minor adverse effects are noted in studies of chronic dosing of between 400-800mg/day in humans (Bergamaschi, Queiroz,

Crippa, & Zuardi, 2011; Iffland & Grotenhermen, 2017; World Health Organisation, 2017).

A recent World Health Organisation report concluded that there is no evidence of potential for abuse or dependence in humans, and no reported recreational profile or public health concerns relating to CBD use (World Health Organisation, 2017). CBD appears to have low toxicity and is well tolerated in most recipients (World Health Organisation, 2017). CBD is reported to exhibit some metabolic drug interactions through the CYP450 enzyme family that may affect its use in combination with some common medications, particularly anti-epileptics.

Further, in vitro studies have demonstrated that CBD may show some cytotoxicity, though moderate or major adverse effects have only been observed at extreme doses in a large range of clinical trials (Bergamaschi et al., 2011). Importantly, CBD is known to counteract many ACCEPTED MANUSCRIPT of the central intoxicant effects of THC when administered simultaneously as a potential inverse agonist (World Health Organisation, 2017). This means that adverse effects associated with direct cannabinoid receptor activation in the central nervous system are avoided, though CBD may also potentiate these effects in some cases in the peripheral system

(McPartland et al., 2015). In a minority of studies CBD has displayed biphasic effects, with

ACCEPTED MANUSCRIPT 30 small doses (.5mg/kg in zebrafish or 100mg in humans) being ineffective and large doses

(10mg/kg in zebrafish or 900mg in humans) being anxiogenic rather than anxiolytic (Nazario et al., 2015; Zuardi et al., 2017), however most animal and human studies report anxiolytic effects exclusively (Blessing, Steenkamp, Manzanares, & Marmar, 2015). Overall, CBD has been shown to have potent antioxidant, antipsychotic, anxiolytic, antiemetic and anti- inflammatory properties, among others (Izzo et al., 2009; Ligresti et al., 2016; Zuardi, 2008).

Insert Figure 3 about here

2.3.1 Recently identified effects of cannabidiol on PTSD symptomology

Preclinical research also shows great promise for CBD as an enhancer of fear extinction and therapeutic consolidation of emotional memories (Table 3, previously reviewed in Blessing, Steenkamp, Manzanares, & Marmar, 2015; Campos et al., 2012; Lee et al., 2017). CBD exerts anti-stress effects following both acute and chronic stress (Campos,

Ferreira, & Guimaraes, 2012; Campos, Ortega, et al., 2013; Granjeiro, Gomes, Guimaraes,

Correa, & Resstel, 2011), which, similar to its anxiolytic effects, appears to recruit a serotonergic mechanism through the 5-HT1A receptors in addition to the endocannabinoid system (Campos, Ferreira, et al., 2012; Lee et al., 2017; Resstel et al., 2009). With respect to the existing endocannabinoid-mediated stress model (Balsevich, Petrie, & Hill, 2017; Morena et al., 2016), however, the exact mechanism of CBD on stress responses is incompletely ACCEPTED MANUSCRIPT understood, as few studies have provided evidence of endocannabinoid mechanisms following CBD and during stress (Campos, Ortega, et al., 2013; Stern, de Carvalho, Bertoglio,

& Takahashi, 2018). CBD has in most studies enhanced fear extinction learning and retention (Blessing et al., 2015; Campos, Moreira, et al., 2012; Lee et al., 2017). Recent studies have supported this effect by showing impaired fear memory consolidation through

ACCEPTED MANUSCRIPT 31 the mPFC which was associated with decreased dopamine release (Rossignoli et al., 2017), the mesolimbic serotonergic system (Gomes et al., 2012; Norris et al., 2016) and hippocampal CB1 and CB2 receptors (Stern et al., 2017). Fear extinction is enhanced by

CBD only if cannabinoid receptors are not antagonised (Bitencourt et al., 2008; Do Monte et al., 2013), though may interact with the strength of fear conditioning with weak and strong conditioning resulting in heightened and decreased fear expression, respectively (Song,

Stevenson, Guimaraes, & Lee, 2016). Further, recent research has shown that CBD administered prior to fear conditioning can increase generalised fear and impair subsequent extinction through increased hippocampal spine density growth, implying that human trials involving experimental fear conditioning and extinction will require careful consideration of timing of administration of CBD doses (Uhernik, Montoya, Balkissoon, & Smith, 2018).

There is very little published research on the effect of CBD on memory consolidation and retrieval relevant to intrusive memories in PTSD, which may be a topic of future investigation due to CBD’s strong anxiolytic and anti-stress effects. This is likely to be due to the multi-target nature of CBD; whereas the memory literature is well defined at specific molecular targets. Research on memory processes has found that CBD exerts biphasic effects on memory. Larger doses have been shown to impair memory (Nazario et al., 2015) and lower doses have been shown to rescue memory impaired due to inflammation (Cassol-Jr et al., 2010; Fagherazzi et al., 2012) and THC use (Englund et al., 2013; Wright, Vandewater, &

Taffe, 2013). More specifically to PTSD, limited research suggests that CBD administration ACCEPTED MANUSCRIPT may improve performance on inhibitory avoidance tasks (Campos, de Paula Soares, et al.,

2013; Soares et al., 2010), which is likely to simulate memory and learning in rodents (Gold,

1986). However, these studies were not specifically targeted towards stress-related emotional memory performance, so whether CBD may provide a therapeutic avenue in that respect is uncertain.

ACCEPTED MANUSCRIPT 32

Despite its excellent safety profile and promising preclinical background, there have been very few human PTSD or stress studies utilising CBD. Elms et al. (2018) recently reported significant reductions in symptomology in ten out of eleven PTSD patients prescribed 25-100mg of CBD per day over an eight week trial. Experimental research has shown that CBD inhaled compared to placebo improved performance on a standard fear conditioning and extinction task (Das et al., 2013). Further, in a case study of a young girl with PTSD, CBD oil led to reduced anxiety and improved sleep (Shannon & Opila-Lehman,

2016). Improving impoverished sleep is considered to be critical to improving waking PTSD symptoms, especially intrusive memories (Germain, 2013; Pace-Schott, Germain, & Milad,

2015). In a recent series of 72 case control studies involving patients with anxiety or sleep problems, four-fifths of participants experienced a reduction in anxiety symptoms and two- thirds showed improvements in sleep quality following sustained CBD treatment (Shannon,

Lewis, Lee, & Hughes, 2019). Importantly, CBD (Chagas et al., 2014; Shannon et al., 2019), as well as CBD compared to placebo (Carlini & Cunha, 1981), has been shown to improve sleep quality and duration in participants with insomnia and other sleep disorders, though this is probably only likely to occur in higher doses with 15mg of CBD promoting alertness (see

Nicholson, Turner, Stone, & Robson, 2004). Similarly, biphasic effects of CBD dosage were recently observed during a stressful public speaking test, with 300mg, but not 100mg or

900mg, reducing subjective anxiety post-speech (Zuardi et al., 2017). There are currently three clinical trials being conducted for CBD in PTSD (NCT03518801, NCT02517424 and ACCEPTED MANUSCRIPT NCT03248167), however initial trials have been promising (Elms et al., 2018).

2.3.3 Disadvantages and status of Cannabidiol

Whereas there are previous incidences of serious adverse effects of synthetic FAAH inhibition (Kerbrat et al., 2016), CBD appears to impair endocannabinoid ligand metabolism

ACCEPTED MANUSCRIPT 33 to some extent without the side effects seen in other FAAH inhibitors. However, there are several disadvantages to using CBD. Firstly, the mechanisms for how CBD exerts its effects are largely unknown. CBD is thought to act through multiple different pathways, including, but not limited to, 5-HT1A receptors, adenosine reuptake, GPR receptors and TRPV channels

(Ligresti et al., 2016; McPartland et al., 2015; Morales, Hurst, et al., 2017). Another current challenge is how much CBD needed to be an effective dose in humans. As an oral dose, large amounts of cannabidiol in either capsules or in oil is required for beneficial effects, due to low bioavailability of CBD on ingestion, with preliminary clinical figures ranging from 6% to a more optimistic 19% oral bioavailability following first pass metabolism (Hawksworth &

McArdle, 2004; Scuderi et al., 2009). Due to studies generally showing that large amounts of

CBD (between 150-600mg per day) are required for effective treatment of anxiety and other similar disorders (World Health Organisation, 2017; Zhornitsky & Potvin, 2012), there are some concerns that CBD may not be currently cost effective for consumers (Breuer et al.,

2016). It is reported that inhaled CBD has higher bioavailability, with an average of 31% in cannabis users (Consroe, Kennedy, & Schram, 1991). However, due to the relatively brief duration of action of smoked CBD (Huestis, 2007) and the general discommendation of inhalation of medicinal products, slow-release oral medicinal products will most likely be preferable for the treatment of chronic ailments.

Enhancing the efficacy of CBD for medical purposes is a current topic of pharmacological research. There are seven known CBD-type molecules aside from CBD ACCEPTED MANUSCRIPT found in cannabis sativa, some of which have begun to show therapeutic potential for epilepsy and cancer (Morales, Reggio, & Jagerovic, 2017). Recent synthetic CBD derivatives have also shown promise for enhancing the potency and overall efficacy of the drug, though have not necessarily been targeted towards psychiatric illness (Morales, Reggio, et al., 2017). For instance, recent alterations on the alkyl chain resulted in KLS-13019, which

ACCEPTED MANUSCRIPT 34 has a far higher potency and safety rating than pure CBD, with capacity for improved bioavailability (Kinney et al., 2016). Other approaches include hydroxyl and other alkyl chain (dimethylheptyl) substitutions (Hanus et al., 2005; Mechoulam, Kogan, Gallily, &

Breuer, 2008; Takeda, Usami, Yamamoto, & Watanabe, 2009), though these have had mixed success in terms of improved potency or safety. Another promising recent approach has been the addition of fluorine atoms to CBD (Breuer et al., 2016). Many pharmaceutical drugs are fluorinated to enhance their stability, potency and bioavailability, which is thought to occur through improved receptor binding and selectivity with the addition of fluorine atoms to drug compounds (Muller, Faeh, & Diederich, 2007). Recently, a Israeli-Brazilian initiative synthesized several fluorinated CBD compounds, of which HUF-101 exhibited enhanced efficacy for decreasing anxiety behaviours (Breuer et al., 2016) and increased antinociception in rats compared to CBD (Silva et al., 2017). In both of these studies, HUF-101 administration produced the desired effect at lower doses than CBD, suggesting that it may be a potential therapeutic option (Breuer et al., 2016; Silva et al., 2017). However, fluorination can also affect the pharmacological profile and targets of compounds, and the toxicity profile of a fluorinated compound must be firmly established before human use

(Liang, Neumann, & Ritter, 2013). To date, trials using several brands of fluorinated synthetic cannabinoid drugs have reported kidney toxicity in humans (Banister et al., 2015).

Therefore, further refinement of CBD and CBD derivatives may be necessary before these products are ready for the consumer market. ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 35

Table 3. Cannabidiol modulation of the endocannabinoid system

Recognised clinical Newly recognised mechanisms of Advantages Disadvantages Status

effects action in PTSD processes

Cannabidiol Improves sleep HC CB1 and CB2 dependent Efficacious in healthy and Difficult and expensive to Being tested in humans,

quality Interacts with dopamine and clinical subjects manufacture emerging medical

Improved fear serotonin systems Excellent safety profile Poorly understood recognition for some

extinction recall Avoids central side effects mechanisms of action indications in some

Reduced anxiety, Efficacious in animal models Possible biphasic dosing and countries

PTSD symptoms of PTSD timing effects

Halogenated Not tested in CB2 dependent More potent Needs further investigation Untested in humans, derivatives humans Cheaper due to potency needs further

HUF-101 investigation

C40-Alkyl Not tested in Not tested More potent No follow-up studies, needs Needs further

Chain humans Cheaper due to potency further investigation investigation derivatives Improved safety profile KLS-13019 ACCEPTED MANUSCRIPT Note: Other cannabidiol derivatives are described in Morales, Reggio, et al. (2017), though have not been described in PTSD animal research

HC = hippocampus, CB1 = cannabinoid receptor type 1, CB2 = cannabinoid receptor type 2

ACCEPTED MANUSCRIPT 36

3. Summary/Conclusions

Mounting preclinical evidence shows that downregulation of the endocannabinoid system is pivotal to the development and maintenance of PTSD, as well as other anxiety disorders (Hill et al., 2018; Morena et al., 2016; Ney et al., 2018; Patel, Hill, Cheer, Wotjak,

& Holmes, 2017). The most recent research has continued to show that cannabinoids act through CB1 receptors in the classical memory systems most relevant to fear circuitry, such as the CA1 of the hippocampus, the BLA and the mPFC. Studies have also begun to show that the endocannabinoid system interacts with multiple signalling systems across different brain regions to influence PTSD aetiology and maintenance. For instance, recent research has shown that cannabinoid modulation of fear and extinction memory is dependent on the dopamine and serotonin systems (Campos, Ferreira, et al., 2012; Fitoussi et al., 2018; Nasehi et al., 2016; Rossignoli et al., 2017; Wenzel et al., 2018), which is congruent with the increasing recognition in the PTSD literature of importance of mesolimbic reward system in

PTSD maintenance (Abraham et al., 2014; Kalebasi et al., 2015; Lee et al., 2016), as well as the expanded endocannabinoid system (De Petrocellis & Di Marzo, 2010; Ligresti et al., 2016;

Ney et al., 2018). Further, although classically assumed that CB1 was the primary receptor involved in modulating these processes, some research has begun to identify a role for CB2 receptors in the hippocampus (Nasehi, Gerami-Majd, et al., 2018; Stern et al., 2017). Finally, cannabinoid modulation may exert effects on memory processes through alteration of BDNF concentrations in the hippocampus and the BLA, as well as altering long-term potentiation in ACCEPTED MANUSCRIPT hippocampal neurons (Burstein et al., 2018; Korem et al., 2017; Segev & Akirav, 2016;

Segev et al., 2018; Shoshan et al., 2017). Development of new chemicals that combine aspects of FAAH inhibition with MAGL and TRPV1 inhibition have been mildly successful

(Bedse et al., 2018; Llorente-Berzal et al., 2015), but are still in their infancy.

ACCEPTED MANUSCRIPT 37

Despite early clinical evidence showing improvements in PTSD symptomology with regular THC and CBD treatment (Cameron et al., 2014; Elms et al., 2018; Fraser, 2009; Jetly et al., 2015; Passie, Emrich, Karst, Brandt, & Halpern, 2012; Shannon & Opila-Lehman,

2016), chronic THC administration is problematic for multiple reasons including off-target effects, endocannabinoid system downregulation, difficulties regulating dosing and mental health complications (Ceccarini et al., 2015; Hall & Degenhardt, 2009; Hirvonen et al., 2012;

Johnson et al., 2016; Moreira, Grieb, & Lutz, 2009; Ware et al., 2018; Wilkinson et al., 2015).

Therefore, the future of cannabinoid-based treatments for PTSD and related disorders is through inhibition of catabolic enzymes or through cannabidiol, which is the non-intoxicant compound found in cannabis. Unfortunately, there are currently significant setbacks in the advancements of these treatments as well. For instance, a clinical trial using FAAH inhibitor

BIA 10-2474 resulted in serious adverse events including death (Kerbrat et al., 2016), and other FAAH inhibitors have not shown the same efficacy in humans as compared to animals

(Huggins et al., 2012; Sanofi, 2013). Further, cannabidiol is currently expensive to manufacture and requires large amounts when orally ingested to be a feasible treatment option.

3.1 Future Directions

Recent manipulation of endocannabinoids in animal models of anxiety and PTSD have not only informed us about the strengths and differences of alternative pharmacological ACCEPTED MANUSCRIPT techniques, but have also increased our understanding about how the endocannabinoid system interactions with neighbouring systems to modulate mood, memory and symptomology.

Future research will benefit from focusing on the role of the endocannabinoid system in other systems relevant to PTSD, such as the dopaminergic and serotonergic systems (Abraham et al., 2014; Lee et al., 2016). Given the recent move towards understanding shared

ACCEPTED MANUSCRIPT 38 mechanisms between psychiatric fields, such as between PTSD and the psychotic illnesses

(Duncan et al., 2018), there is increasing need for molecular mediators between these illnesses that explains the similarities. The endocannabinoid system is a likely system where shared mechanisms may be identified; thus exploration of its varied molecular interactions in anxiety and PTSD may benefit other fields. Clinical trials are desperately needed around the world for cannabinoid products, especially for cannabidiol and FAAH inhibitors, once safety trials have been successfully completed. This is increasingly easy to achieve in many countries as cannabinoids become legalised for medicinal purposes. However, given the weaknesses of current approaches, development, validation and safety testing of new pharmacological agents should also be a priority moving forward. Promising molecules such as HUF-101, KLS-13019 and novel FAAH inhibitors need to be extensively tested in animals, and anxiety and PTSD preclinical laboratories should consider using options such as these in their research.

Finally, there are important differences in endocannabinoid mobilisation and cannabinoid metabolism between the sexes that need to be considered before cannabinoids can be pursued as therapies for anxiety and PTSD, which are both at least twice as prevalent in women compared to men (Kessler et al., 2017; Kessler, Sonnega, Bromet, Hughes, &

Nelson, 1995; Ney et al., 2018). For instance, recent animal studies have begun to demonstrate that the endocannabinoid system is mobilised to a greater degree in males compared to females in response to stress (Wyrofsky, Reyes, Yu, Kirby, & Van Bockstaele, ACCEPTED MANUSCRIPT 2018; Xing et al., 2011; Zer-Aviv & Akirav, 2016). Similarly, men have been reported to have higher hair and plasma concentrations of endocannabinoids, especially 2-AG, than women (Fanelli et al., 2012; Mwanza et al., 2016), and men display stronger physiological effects to cannabis than women (Haney, 2007; Leatherdale, Hammond, Kaiserman, & Ahmed,

2007; Penetar et al., 2005). Therefore, future clinical trials will need to incorporate the

ACCEPTED MANUSCRIPT 39 increasingly evident modulatory role of sex hormones and sex differences on the effects of cannabinoids in humans (Ney, et al. 2018).

3.2 Conclusion

In conclusion, the endocannabinoid system is a promising novel target for reduction in anxiety and PTSD symptomology, or as an adjunct to enhance the effect of behaviourally- based psychological treatments (Berardi et al., 2016; Hill et al., 2018; Morena et al., 2016;

Patel et al., 2017). However, it is inconclusive what types of pharmacological agents may positively modulate the endocannabinoid system for reducing PTSD symptoms without aversive side effects. In this review, we have summarised recent studies using several of these methods whilst highlighting the challenges remaining for these approaches. Future research will need to be geared towards improving safety of FAAH inhibitors, as well as improving the potency of cannabidiol analogues and understanding both the therapeutic potential and mechanisms of actions of different drugs. Future research will also benefit from attending to the growing recognition of molecular interactions in both PTSD and the endocannabinoid system; here we have recommended several areas where this understanding may be beneficially improved.

References

Abraham, A. D., Neve, K. A., & Lattal, K. M. (2014). Dopamine and extinction: a convergence of theory withACCEPTED fear and reward circuitry. Neurobiol MANUSCRIPT Learn Mem, 108, 65-77. doi:10.1016/j.nlm.2013.11.007 Aghazadeh Tabrizi, M., Baraldi, P. G., Baraldi, S., Ruggiero, E., De Stefano, L., Rizzolio, F., . . . Tuccinardi, T. (2018). Discovery of 1,5-Diphenylpyrazole-3-Carboxamide Derivatives as Potent, Reversible, and Selective Monoacylglycerol Lipase (MAGL) Inhibitors. J Med Chem, 61(3), 1340-1354. doi:10.1021/acs.jmedchem.7b01845 Ahn, K., Smith, S. E., Liimatta, M. B., Beidler, D., Sadagopan, N., Dudley, D. T., . . . Cravatt, B. F. (2011). Mechanistic and pharmacological characterization of PF-04457845: a highly potent and selective fatty acid amide hydrolase inhibitor that reduces inflammatory and noninflammatory pain. J Pharmacol Exp Ther, 338(1), 114-124. doi:10.1124/jpet.111.180257

ACCEPTED MANUSCRIPT 40

Aisenberg, N., Serova, L., Sabban, E. L., & Akirav, I. (2017). The effects of enhancing endocannabinoid signaling and blocking corticotrophin releasing factor receptor in the amygdala and hippocampus on the consolidation of a stressful event. Eur Neuropsychopharmacol, 27(9), 913-927. doi:10.1016/j.euroneuro.2017.06.006 Aliczki, M., Zelena, D., Mikics, E., Varga, Z. K., Pinter, O., Bakos, N. V., . . . Haller, J. (2013). Monoacylglycerol lipase inhibition-induced changes in plasma corticosterone levels, anxiety and locomotor activity in male CD1 mice. Hormones and Behavior, 63(5), 752-758. doi:https://doi.org/10.1016/j.yhbeh.2013.03.017 Almeida-Santos, A. F., Moreira, F. A., Guimaraes, F. S., & Aguiar, D. C. (2017). 2-Arachidonoylglycerol endocannabinoid signaling coupled to metabotropic glutamate receptor type-5 modulates anxiety-like behavior in the rat ventromedial prefrontal cortex. Journal of Psychopharmacology, 31(6), 740-749. doi:10.1177/0269881117704986 American Psychiatric Association. (2013). Diagnostic and Statistical Manual of Mental Disorders (5th ed.). Washington DC: American Psychiatric Association. Back, F. P., & Carobrez, A. P. (2018). Periaqueductal gray glutamatergic, cannabinoid and vanilloid receptor interplay in defensive behavior and aversive memory formation. Neuropharmacology, 135, 399-411. doi:10.1016/j.neuropharm.2018.03.032 Balsevich, G., Petrie, G. N., & Hill, M. N. (2017). Endocannabinoids: Effectors of glucocorticoid signaling. Front Neuroendocrinol. doi:10.1016/j.yfrne.2017.07.005 Banister, S. D., Stuart, J., Kevin, R. C., Edington, A., Longworth, M., Wilkinson, S. M., . . . Kassiou, M. (2015). Effects of Bioisosteric Fluorine in Synthetic Cannabinoid Designer Drugs JWH-018, AM-2201, UR-144, XLR-11, PB-22, 5F-PB-22, APICA, and STS-135. ACS Chemical Neuroscience, 6(8), 1445-1458. doi:10.1021/acschemneuro.5b00107 Bedse, G., Bluett, R. J., Patrick, T. A., Romness, N. K., Gaulden, A. D., Kingsley, P. J., . . . Patel, S. (2018). Therapeutic endocannabinoid augmentation for mood and anxiety disorders: comparative profiling of FAAH, MAGL and dual inhibitors. Transl Psychiatry, 8(1), 92. doi:10.1038/s41398- 018-0141-7 Bedse, G., Colangeli, R., Lavecchia, A. M., Romano, A., Altieri, F., Cifani, C., . . . Gaetani, S. (2014). Role of the basolateral amygdala in mediating the effects of the fatty acid amide hydrolase inhibitor URB597 on HPA axis response to stress. Eur Neuropsychopharmacol, 24(9), 1511- 1523. doi:10.1016/j.euroneuro.2014.07.005 Bedse, G., Hartley, N. D., Neale, E., Gaulden, A. D., Patrick, T. A., Kingsley, P. J., . . . Patel, S. (2017). Functional Redundancy Between Canonical Endocannabinoid Signaling Systems in the Modulation of Anxiety. Biol Psychiatry, 82(7), 488-499. doi:10.1016/j.biopsych.2017.03.002 Beltramo, M., Stella, N., Calignano, A., Lin, S. Y., Makriyannis, A., & Piomelli, D. (1997). Functional Role of High-Affinity Transport, as Revealed by Selective Inhibition. Science, 277(5329), 1094-1097. doi:10.1126/science.277.5329.1094 Berardi, A., Schelling, G., & Campolongo, P. (2016). The endocannabinoid system and Post Traumatic Stress Disorder (PTSD): From preclinical findings to innovative therapeutic approaches in clinical settings. Pharmacol Res, 111, 668-678. doi:10.1016/j.phrs.2016.07.024 Bergamaschi, M. M.,ACCEPTED Queiroz, R. H., Crippa, J. A., MANUSCRIPT & Zuardi, A. W. (2011). Safety and Side Effects of Cannabidiol, a Cannabis sativa Constituent Current Drug Safety, 6(4), 237-249. doi:10.2174/157488611798280924 Bisogno, T., Hanus, L., de Petrocellis, L., Tchilibon, S., Ponde, D. E., Brandi, I., . . . di Marzo, V. (2001). Molecular targets for cannabidiol and its synthetic analogues: effect on vanilloid VR1 receptors and on the cellular uptake and enzymatic hydrolysis of anandamide. British Journal of Pharmacology, 134, 845-852. doi:10.1038/sj.bjp.0704327 Bisogno, T., & Maccarrone, M. (2013). Latest advances in the discovery of fatty acid amide hydrolase inhibitors. Expert Opin Drug Discov, 8(5), 509-522. doi:10.1517/17460441.2013.780021

ACCEPTED MANUSCRIPT 41

Bisogno, T., Melck, D., De Petrocellis, L., Bobrov, M. Y., Gretskaya, N. M., Bezuglov, V. V., . . . Di Marzo, V. (1998). Arachidonoylserotonin and Other Novel Inhibitors of Fatty Acid Amide Hydrolase. Biochem Biophys Res Commun, 248(3), 515-522. Bisson, J. I., Roberts, N. P., Andrew, M., Cooper, R., & Lewis, C. (2013). Psychological therapies for chronic post-traumatic stress disorder (PTSD) in adults. Cochrane Database Syst Rev(12), CD003388. doi:10.1002/14651858.CD003388.pub4 Bitencourt, R. M., Pamplona, F. A., & Takahashi, R. N. (2008). Facilitation of contextual fear memory extinction and anti-anxiogenic effects of AM404 and cannabidiol in conditioned rats. Eur Neuropsychopharmacol, 18(12), 849-859. doi:10.1016/j.euroneuro.2008.07.001 Blessing, E. M., Steenkamp, M. M., Manzanares, J., & Marmar, C. R. (2015). Cannabidiol as a Potential Treatment for Anxiety Disorders. Neurotherapeutics, 12(4), 825-836. doi:10.1007/s13311-015-0387-1 Bluett, R. J., Baldi, R., Haymer, A., Gaulden, A. D., Hartley, N. D., Parrish, W. P., . . . Patel, S. (2017). Endocannabinoid signalling modulates susceptibility to traumatic stress exposure. Nat Commun, 8, 14782. doi:10.1038/ncomms14782 Bosch-Bouju, C., Larrieu, T., Linders, L., Manzoni, O. J., & Laye, S. (2016). Endocannabinoid-Mediated Plasticity in Nucleus Accumbens Controls Vulnerability to Anxiety after Social Defeat Stress. Cell Rep, 16(5), 1237-1242. doi:10.1016/j.celrep.2016.06.082 Bouton, M. E. (2004). Context and behavioral processes in extinction. Learn Mem, 11(5), 485-494. doi:10.1101/lm.78804 Bowers, M. E., & Ressler, K. J. (2015). Interaction between the cholecystokinin and endogenous cannabinoid systems in cued fear expression and extinction retention. Neuropsychopharmacology, 40(3), 688-700. doi:10.1038/npp.2014.225 Breuer, A., Haj, C. G., Fogaca, M. V., Gomes, F. V., Silva, N. R., Pedrazzi, J. F., . . . Guimaraes, F. S. (2016). Fluorinated Cannabidiol Derivatives: Enhancement of Activity in Mice Models Predictive of Anxiolytic, Antidepressant and Antipsychotic Effects. PLoS One, 11(7), e0158779. doi:10.1371/journal.pone.0158779 Brewin, C. R. (2011). The nature and significance of memory disturbance in posttraumatic stress disorder. Annu Rev Clin Psychol, 7, 203-227. doi:10.1146/annurev-clinpsy-032210-104544 Brewin, C. R., Gregory, J. D., Lipton, M., & Burgess, N. (2010). Intrusive images in psychological disorders: characteristics, neural mechanisms, and treatment implications. Psychol Rev, 117(1), 210-232. doi:10.1037/a0018113 Bryant, R. A., McGrath, C., & Felmingham, K. L. (2013). The roles of noradrenergic and glucocorticoid activation in the development of intrusive memories. PLoS One, 8(4), e62675. doi:10.1371/journal.pone.0062675 Burman, M. A., Szolusha, K., Bind, R., Kerney, K., Boger, D. L., & Bilsky, E. J. (2016). FAAH inhibitor OL- 135 disrupts contextual, but not auditory, fear conditioning in rats. Behav Brain Res, 308, 1-5. doi:10.1016/j.bbr.2016.04.014 Burstein, O., Shoshan, N., Doron, R., & Akirav, I. (2018). Cannabinoids prevent depressive -like symptoms and alterations in BDNF expression in a rat model of PTSD. Progress in Neuro- PsychopharmacologyACCEPTED and Biological Psychiatry, MANUSCRIPT 84, 129-139. doi:https://doi.org/10.1016/j.pnpbp.2018.01.026 Bush, D. M., & Woodwell, D. A. (2013). Update: Drug-Related Emergency Department Visits Involving Synthetic Cannabinoids The CBHSQ Report (pp. 1-10). Rockville (MD): Substance Abuse and Mental Health Services Administration (US). Calabrese, E. J., & Rubio-Casillas, A. (2018). Biphasic effects of THC in memory and cognition. European Journal of Clinical Investigation, 48(5), e12920. doi:10.1111/eci.12920 Cameron, C., Watson, D., & Robinson, J. (2014). Use of a synthetic cannabinoid in a correctional population for posttraumatic stress disorder-related insomnia and nightmares, chronic pain, harm reduction, and other indications: a retrospective evaluation. J Clin Psychopharmacol, 34(5), 559-564. doi:10.1097/JCP.0000000000000180

ACCEPTED MANUSCRIPT 42

Campos, A. C., de Paula Soares, V., Carvalho, M. C., Ferreira, F. R., Vicente, M. A., Brandao, M. L., . . . Guimaraes, F. S. (2013). Involvement of serotonin-mediated neurotransmission in the dorsal periaqueductal gray matter on cannabidiol chronic effects in panic-like responses in rats. Psychopharmacology (Berl), 226(1), 13-24. doi:10.1007/s00213-012-2878-7 Campos, A. C., Ferreira, F. R., & Guimaraes, F. S. (2012). Cannabidiol blocks long-lasting behavioral consequences of predator threat stress: possible involvement of 5HT1A receptors. J Psychiatr Res, 46(11), 1501-1510. doi:10.1016/j.jpsychires.2012.08.012 Campos, A. C., & Guimaraes, F. S. (2008). Involvement of 5HT1A receptors in the anxiolytic-like effects of cannabidiol injected into the dorsolateral periaqueductal gray of rats. Psychopharmacology (Berl), 199(2), 223-230. doi:10.1007/s00213-008-1168-x Campos, A. C., & Guimarães, F. S. (2009). Evidence for a potential role for TRPV1 receptors in the dorsolateral periaqueductal gray in the attenuation of the anxiolytic effects of cannabinoids. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 33(8), 1517-1521. doi:https://doi.org/10.1016/j.pnpbp.2009.08.017 Campos, A. C., Moreira, F. A., Gomes, F. V., Del Bel, E. A., & Guimaraes, F. S. (2012). Multiple mechanisms involved in the large-spectrum therapeutic potential of cannabidiol in psychiatric disorders. Philos Trans R Soc Lond B Biol Sci, 367(1607), 3364-3378. doi:10.1098/rstb.2011.0389 Campos, A. C., Ortega, Z., Palazuelos, J., Fogaca, M. V., Aguiar, D. C., Diaz-Alonso, J., . . . Guimaraes, F. S. (2013). The anxiolytic effect of cannabidiol on chronically stressed mice depends on hippocampal neurogenesis: involvement of the endocannabinoid system. Int J Neuropsychopharmacol, 16(6), 1407-1419. doi:10.1017/S1461145712001502 Capasso, R., Borrelli, F., Aviello, G., Romano, B., Scalisi, C., Capasso, F., & Izzo, A. A. (2008). Cannabidiol, extracted from Cannabis sativa, selectively inhibits inflammatory hypermotility in mice. Br J Pharmacol, 154(5), 1001-1008. doi:10.1038/bjp.2008.177 Carlini, E. A., & Cunha, J. M. (1981). Hypnotic and Antiepileptic Effects of Cannabidiol. The Journal of Clinical Pharmacology, 21(S1), 417S-427S. doi:10.1002/j.1552-4604.1981.tb02622.x Carrier, E. J., Auchampach, J. A., & Hillard, C. J. (2006). Inhibition of an equilibrative nucleoside transporter by cannabidiol: a mechanism of cannabinoid immunosuppression. Proc Natl Acad Sci U S A, 103(20), 7895-7900. doi:10.1073/pnas.0511232103 Cassol-Jr, O. J., Comim, C. M., Silva, B. R., Hermani, F. V., Constantino, L. S., Felisberto, F., . . . Dal- Pizzol, F. (2010). Treatment with cannabidiol reverses oxidative stress parameters, cognitive impairment and mortality in rats submitted to sepsis by cecal ligation and puncture. Brain Research, 1348, 128-138. doi:https://doi.org/10.1016/j.brainres.2010.06.023 Castaneto, M. S., Gorelick, D. A., Desrosiers, N. A., Hartman, R. L., Pirard, S., & Huestis, M. A. (2014). Synthetic cannabinoids: epidemiology, pharmacodynamics, and clinical implications. Drug Alcohol Depend, 144, 12-41. doi:10.1016/j.drugalcdep.2014.08.005 Ceccarini, J., Kuepper, R., Kemels, D., van Os, J., Henquet, C., & Van Laere, K. (2015). [18F]MK-9470 PET measurement of cannabinoid CB1 receptor availability in chronic cannabis users. Addict Biol, 20(2), 357-367. doi:10.1111/adb.12116 Chagas, M. H. N., Eckeli,ACCEPTED A. L., Zuardi, A. W., Pena MANUSCRIPT-Pereira, M. A., Sobreira-Neto, M. A., Sobreira, E. T., . . . Crippa, J. A. S. (2014). Cannabidiol can improve complex sleep-related behaviours associated with rapid eye movement sleep behaviour disorder in Parkinson's disease patients: a case series. Journal of Clinical Pharmacy and Therapeutics, 39(5), 564-566. doi:10.1111/jcpt.12179 Chicca, A., Nicolussi, S., Bartholomäus, R., Blunder, M., Aparisi Rey, A., Petrucci, V., . . . Gertsch, J. (2017). Chemical probes to potently and selectively inhibit endocannabinoid cellular reuptake. Proceedings of the National Academy of Sciences, 114(25), E5006-E5015. doi:10.1073/pnas.1704065114

ACCEPTED MANUSCRIPT 43

Childs, E., Lutz, J. A., & de Wit, H. (2017). Dose-related effects of delta-9-THC on emotional responses to acute psychosocial stress. Drug Alcohol Depend, 177, 136-144. doi:10.1016/j.drugalcdep.2017.03.030 Chou, C. Y., La Marca, R., Steptoe, A., & Brewin, C. R. (2014). Biological responses to trauma and the development of intrusive memories: an analog study with the trauma film paradigm. Biol Psychol, 103, 135-143. doi:10.1016/j.biopsycho.2014.08.002 Compton, D. R., Gold, L. H., Ward, S. J., Balster, R. L., & Martin, B. R. (1992). Aminoalkylindole analogs: cannabimimetic activity of a class of compounds structurally distinct from delta 9- tetrahydrocannabinol. Journal of Pharmacology and Experimental Therapeutics, 263(3), 1118-1126. Consroe, P., Kennedy, K., & Schram, K. (1991). Assay of plasma cannabidiol by capillary gas chromatography/ion trap mass spectroscopy following high-dose repeated daily oral administration in humans. Pharmacology Biochemistry and Behavior, 40(3), 517-522. doi:https://doi.org/10.1016/0091-3057(91)90357-8 Danandeh, A., Vozella, V., Lim, J., Oveisi, F., Ramirez, G. L., Mears, D., . . . Piomelli, D. (2018). Effects of fatty acid amide hydrolase inhibitor URB597 in a rat model of trauma-induced long-term anxiety. Psychopharmacology (Berl), 235(11), 3211-3221. doi:10.1007/s00213-018-5020-7 Das, R. K., Kamboj, S. K., Ramadas, M., Yogan, K., Gupta, V., Redman, E., . . . Morgan, C. J. (2013). Cannabidiol enhances consolidation of explicit fear extinction in humans. Psychopharmacology (Berl), 226(4), 781-792. doi:10.1007/s00213-012-2955-y De Petrocellis, L., & Di Marzo, V. (2010). Non-CB1, non-CB2 receptors for endocannabinoids, plant cannabinoids, and synthetic cannabimimetics: focus on G-protein-coupled receptors and transient receptor potential channels. J Neuroimmune Pharmacol, 5(1), 103-121. doi:10.1007/s11481-009-9177-z Devane, W. A., Hanus, L., Breuer, A., Pertwee, R. G., Stevenson, L. A., Griffin, G., . . . Mechoulam, R. (1992). Isolation and Structure of a Brain Constituent That Binds to the Cannabinoid Receptor. Science, 258, 1946-1949. Di Marzo, V. (2008). Targeting the endocannabinoid system: to enhance or reduce? Nat Rev Drug Discov, 7(5), 438-455. doi:10.1038/nrd2553 Di Marzo, V. (2009). The endocannabinoid system: its general strategy of action, tools for its pharmacological manipulation and potential therapeutic exploitation. Pharmacol Res, 60(2), 77-84. doi:10.1016/j.phrs.2009.02.010 Di, S., Itoga, C. A., Fisher, M. O., Solomonow, J., Roltsch, E. A., Gilpin, N. W., & Tasker, J. G. (2016). Acute Stress Suppresses Synaptic Inhibition and Increases Anxiety via Endocannabinoid Release in the Basolateral Amygdala. J Neurosci, 36(32), 8461-8470. doi:10.1523/JNEUROSCI.2279-15.2016 Di Venere, A., Dainese, E., Fezza, F., Angelucci, B. C., Rosato, N., Cravatt, B. F., . . . Maccarrone, M. (2012). Rat and human fatty acid amide hydrolases: Overt similarities and hidden differences. Biochimica et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids, 1821(11), 1425- 1433. doi:https://doi.org/10.1016/j.bbalip.2012.07.021 Dincheva, I., Drysdale,ACCEPTED A. T., Hartley, C. A., Johnson, MANUSCRIPT D. C., Jing, D., King, E. C., . . . Lee, F. S. (2015). FAAH genetic variation enhances fronto-amygdala function in mouse and human. Nat Commun, 6, 6395. doi:10.1038/ncomms7395 Do Monte, F. H., Souza, R. R., Bitencourt, R. M., Kroon, J. A., & Takahashi, R. N. (2013). Infusion of cannabidiol into infralimbic cortex facilitates fear extinction via CB1 receptors. Behav Brain Res, 250, 23-27. doi:10.1016/j.bbr.2013.04.045 Doss, M. K., Weafer, J., Gallo, D. A., & de Wit, H. (2018). Δ9-Tetrahydrocannabinol at Retrieval Drives False Recollection of Neutral and Emotional Memories. Biological Psychiatry. doi:https://doi.org/10.1016/j.biopsych.2018.04.020 Duncan, L. E., Ratanatharathorn, A., Aiello, A. E., Almli, L. M., Amstadter, A. B., Ashley-Koch, A. E., . . . Koenen, K. C. (2018). Largest GWAS of PTSD (N=20 070) yields genetic overlap with

ACCEPTED MANUSCRIPT 44

schizophrenia and sex differences in heritability. Mol Psychiatry, 23(3), 666-673. doi:10.1038/mp.2017.77 Elms, L., Shannon, S., Hughes, S., & Lewis, N. (2018). Cannabidiol in the Treatment of Post-Traumatic Stress Disorder: A Case Series. J Altern Complement Med. doi:10.1089/acm.2018.0437 Englund, A., Morrison, P. D., Nottage, J., Hague, D., Kane, F., Bonaccorso, S., . . . Kapur, S. (2013). Cannabidiol inhibits THC-elicited paranoid symptoms and hippocampal-dependent memory impairment. J Psychopharmacol, 27(1), 19-27. doi:10.1177/0269881112460109 Evanson, N. K., Tasker, J. G., Hill, M. N., Hillard, C. J., & Herman, J. P. (2010). Fast feedback inhibition of the HPA axis by glucocorticoids is mediated by endocannabinoid signaling. Endocrinology, 151(10), 4811-4819. doi:10.1210/en.2010-0285 Fagherazzi, E. V., Garcia, V. A., Maurmann, N., Bervanger, T., Halmenschlager, L. H., Busato, S. B., . . . Schroder, N. (2012). Memory-rescuing effects of cannabidiol in an animal model of cognitive impairment relevant to neurodegenerative disorders. Psychopharmacology (Berl), 219(4), 1133-1140. doi:10.1007/s00213-011-2449-3 Fanelli, F., Di Lallo, V. D., Belluomo, I., De Iasio, R., Baccini, M., Casadio, E., . . . Pagotto, U. (2012). Estimation of reference intervals of five endocannabinoids and endocannabinoid related compounds in human plasma by two dimensional-LC/MS/MS. J Lipid Res, 53(3), 481-493. doi:10.1194/jlr.M021378 Faraji, N., Komaki, A., & Salehi, I. (2017). Interaction Between the Cannabinoid and Vanilloid Systems on Anxiety in Male Rats. Basic Clin Neurosci, 8(2), 129-137. doi:10.18869/nirp.bcn.8.2.129 Fattore, L. (2016). Synthetic Cannabinoids-Further Evidence Supporting the Relationship Between Cannabinoids and Psychosis. Biol Psychiatry, 79(7), 539-548. doi:10.1016/j.biopsych.2016.02.001 Fegley, D., Gaetani, S., Duranti, A., Tontini, A., Mor, M., Tarzia, G., & Piomelli, D. (2005). Characterization of the fatty acid amide hydrolase inhibitor cyclohexyl carbamic acid 3'- carbamoyl-biphenyl-3-yl ester (URB597): effects on anandamide and oleoylethanolamide deactivation. J Pharmacol Exp Ther, 313(1), 352-358. doi:10.1124/jpet.104.078980 Fidelman, S., Mizrachi Zer-Aviv, T., Lange, R., Hillard, C. J., & Akirav, I. (2018). Chronic treatment with URB597 ameliorates post-stress symptoms in a rat model of PTSD. Eur Neuropsychopharmacol, 28(5), 630-642. doi:10.1016/j.euroneuro.2018.02.004 Fitoussi, A., Zunder, J., Tan, H., & Laviolette, S. R. (2018). Delta-9-tetrahydrocannabinol potentiates fear memory salience through functional modulation of mesolimbic dopaminergic activity states. Eur Neuropsychopharmacol, 47(11), 1385-1400. doi:10.1111/ejn.13951 Forrester, M. B., Kleinschmidt, K., Schwarz, E., & Young, A. (2011). Synthetic cannabinoid exposures reported to Texas poison centers. J Addict Dis, 30(4), 351-358. doi:10.1080/10550887.2011.609807 Fraser, G. A. (2009). The use of a synthetic cannabinoid in the management of treatment-resistant nightmares in posttraumatic stress disorder (PTSD). CNS Neurosci Ther, 15(1), 84-88. doi:10.1111/j.1755-5949.2008.00071.x Galatzer-Levy, I. R., Ma, S., Statnikov, A., Yehuda, R., & Shalev, A. Y. (2017). Utilization of machine learning forACCEPTED prediction of post-traumatic MANUSCRIPT stress: a re-examination of cortisol in the prediction and pathways to non-remitting PTSD. Transl Psychiatry, 7(3), e0. doi:10.1038/tp.2017.38 Gamble-George, J. C., Baldi, R., Halladay, L., Kocharian, A., Hartley, N., Silva, C. G., . . . Patel, S. (2016). Cyclooxygenase-2 inhibition reduces stress-induced affective pathology. Elife, 5. doi:10.7554/eLife.14137 Germain, A. (2013). Sleep Disturbances as the Hallmark of PTSD: Where Are We Now? American Journal of Psychiatry: Official Journal of the American Psychiatric Association, 170(4), 372- 382. doi:10.1176/appi.ajp.2012.12040432 Ghasemi, M., Abrari, K., Goudarzi, I., & Rashidy-Pour, A. (2017). Effect of WIN55-212-2 and Consequences of Extinction Training on Conditioned Fear Memory in PTSD Male Rats. Basic Clin Neurosci, 8(6), 493-502. doi:10.29252/NIRP.BCN.8.6.493

ACCEPTED MANUSCRIPT 45

Gobira, P. H., Lima, I. V., Batista, L. A., de Oliveira, A. C., Resstel, L. B., Wotjak, C. T., . . . Moreira, F. A. (2017). N-arachidonoyl-serotonin, a dual FAAH and TRPV1 blocker, inhibits the retrieval of contextual fear memory: Role of the cannabinoid CB1 receptor in the dorsal hippocampus. J Psychopharmacol, 31(6), 750-756. doi:10.1177/0269881117691567 Gold, P. E. (1986). The Use of Avoidance Training in Studies of Modulation of Memory Storage. Behavioural and Neural Biology, 46, 87-98. doi:https://doi.org/10.1016/S0163- 1047(86)90927-1 Gomes, F. V., Reis, D. G., Alves, F. H., Correa, F. M., Guimaraes, F. S., & Resstel, L. B. (2012). Cannabidiol injected into the bed nucleus of the stria terminalis reduces the expression of contextual fear conditioning via 5-HT1A receptors. J Psychopharmacol, 26(1), 104-113. doi:10.1177/0269881110389095 Graham, B. M., Callaghan, B. L., & Richardson, R. (2014). Bridging the gap: Lessons we have learnt from the merging of psychology and psychiatry for the optimisation of treatments for emotional disorders. Behav Res Ther, 62, 3-16. doi:10.1016/j.brat.2014.07.012 Granchi, C., Caligiuri, I., Minutolo, F., Rizzolio, F., & Tuccinardi, T. (2017). A patent review of Monoacylglycerol Lipase (MAGL) inhibitors (2013-2017). Expert Opin Ther Pat, 27(12), 1341- 1351. doi:10.1080/13543776.2018.1389899 Granjeiro, E. M., Gomes, F. V., Guimaraes, F. S., Correa, F. M., & Resstel, L. B. (2011). Effects of intracisternal administration of cannabidiol on the cardiovascular and behavioral responses to acute restraint stress. Pharmacol Biochem Behav, 99(4), 743-748. doi:10.1016/j.pbb.2011.06.027 Gray, J. M., Vecchiarelli, H. A., Morena, M., Lee, T. T., Hermanson, D. J., Kim, A. B., . . . Hill, M. N. (2015). Corticotropin-releasing hormone drives anandamide hydrolysis in the amygdala to promote anxiety. J Neurosci, 35(9), 3879-3892. doi:10.1523/JNEUROSCI.2737-14.2015 Griebel, G., Pichat, P., Beeske, S., Leroy, T., Redon, N., Jacquet, A., . . . Escoubet, J. (2015). Selective blockade of the hydrolysis of the endocannabinoid 2-arachidonoylglycerol impairs learning and memory performance while producing antinociceptive activity in rodents. Sci Rep, 5, 7642. doi:10.1038/srep07642 Griebel, G., Stemmelin, J., Lopez-Grancha, M., Fauchey, V., Slowinski, F., Pichat, P., . . . Bergis, O. E. (2018). The selective reversible FAAH inhibitor, SSR411298, restores the development of maladaptive behaviors to acute and chronic stress in rodents. Scientific Reports, 8(1), 2416. doi:10.1038/s41598-018-20895-z Griffin, M. G., Resick, P. A., & Yehuda, R. (2005). Enhanced cortisol suppression following dexamethasone administration in domestic violence survivors. Am J Psychiatry, 162(6), 1192-1199. doi:10.1176/appi.ajp.162.6.1192 Gunduz-Cinar, O., MacPherson, K. P., Cinar, R., Gamble-George, J., Sugden, K., Williams, B., . . . Holmes, A. (2013). Convergent translational evidence of a role for anandamide in amygdala- mediated fear extinction, threat processing and stress-reactivity. Mol Psychiatry, 18(7), 813- 823. doi:10.1038/mp.2012.72 Guthrie, R. M., & Bryant, R. A. (2006). Extinction learning before trauma and subsequent posttraumaticACCEPTED stress. Psychosom Med, 68MANUSCRIPT(2), 307-311. doi:10.1097/01.psy.0000208629.67653.cc Hall, W., & Degenhardt, L. (2009). Adverse health effects of non- use. 374, 1383- 1391. Haney, M. (2007). Opioid antagonism of cannabinoid effects: differences between marijuana smokers and nonmarijuana smokers. Neuropsychopharmacology, 32(6), 1391-1403. doi:10.1038/sj.npp.1301243 Hanus, L. O., Tchilibon, S., Ponde, D. E., Breuer, A., Fride, E., & Mechoulam, R. (2005). Enantiomeric cannabidiol derivatives: synthesis and binding to cannabinoid receptors. Org Biomol Chem, 3(6), 1116-1123. doi:10.1039/b416943c

ACCEPTED MANUSCRIPT 46

Hartley, N. D., Gunduz-Cinar, O., Halladay, L., Bukalo, O., Holmes, A., & Patel, S. (2016). 2- arachidonoylglycerol signaling impairs short-term fear extinction. Transl Psychiatry, 6, e749. doi:10.1038/tp.2016.26 Hawksworth, G., & McArdle, K. (2004). Metabolism and pharmacokinetics of cannabinoids. The Medicinal Uses of Cannabis and Cannabinoids. (pp. 205-228). London: Pharmaceutical Press. Heinz, D. E., Genewsky, A., & Wotjak, C. T. (2017). Enhanced anandamide signaling reduces flight behavior elicited by an approaching robo-beetle. Neuropharmacology, 126, 233-241. doi:10.1016/j.neuropharm.2017.09.010 Hill, M. N., Bierer, L. M., Makotkine, I., Golier, J. A., Galea, S., McEwen, B. S., . . . Yehuda, R. (2013). Reductions in circulating endocannabinoid levels in individuals with post-traumatic stress disorder following exposure to the World Trade Center attacks. Psychoneuroendocrinology, 38(12), 2952-2961. doi:10.1016/j.psyneuen.2013.08.004 Hill, M. N., Campolongo, P., Yehuda, R., & Patel, S. (2018). Integrating Endocannabinoid Signaling and Cannabinoids into the Biology and Treatment of Posttraumatic Stress Disorder. Neuropsychopharmacology, 43(1), 80-102. doi:10.1038/npp.2017.162 Hill, M. N., McLaughlin, R. J., Morrish, A. C., Viau, V., Floresco, S. B., Hillard, C. J., & Gorzalka, B. B. (2009). Suppression of amygdalar endocannabinoid signaling by stress contributes to activation of the hypothalamic-pituitary-adrenal axis. Neuropsychopharmacology, 34(13), 2733-2745. doi:10.1038/npp.2009.114 Hill, M. N., McLaughlin, R. J., Pan, B., Fitzgerald, M. L., Roberts, C. J., Lee, T. T., . . . Hillard, C. J. (2011). Recruitment of prefrontal cortical endocannabinoid signaling by glucocorticoids contributes to termination of the stress response. J Neurosci, 31(29), 10506-10515. doi:10.1523/JNEUROSCI.0496-11.2011 Hirvonen, J., Goodwin, R. S., Li, C. T., Terry, G. E., Zoghbi, S. S., Morse, C., . . . Innis, R. B. (2012). Reversible and regionally selective downregulation of brain cannabinoid CB1 receptors in chronic daily cannabis smokers. Mol Psychiatry, 17(6), 642-649. doi:10.1038/mp.2011.82 Hogestatt, E. D., Jonsson, B. A., Ermund, A., Andersson, D. A., Bjork, H., Alexander, J. P., . . . Zygmunt, P. M. (2005). Conversion of acetaminophen to the bioactive N-acylphenolamine AM404 via fatty acid amide hydrolase-dependent arachidonic acid conjugation in the nervous system. J Biol Chem, 280(36), 31405-31412. doi:10.1074/jbc.M501489200 Howes, O. D., & Kapur, S. (2009). The dopamine hypothesis of schizophrenia: version III--the final common pathway. Schizophrenia Bulletin, 35(3), 549-562. doi:10.1093/schbul/sbp006 Howlett, A. C., Barth, F., Bonner, T. I., Cabral, G., Casellas, P., Devane, W. A., . . . Pertwee, R. G. (2002). International Union of Pharmacology. XXVII. Classification of Cannabinoid Receptors. Pharmacological Reviews, 54(2), 161-202. Howlett, A. C., Champion, T. M., Wilken, G. H., & Mechoulam, R. (1990). Stereochemical effects of 11-OH-Δ8-tetrahydrocannabinol-dimethylheptyl to inhibit adenylate cyclase and bind to the cannabinoid receptor. Neuropharmacology, 29(2), 161-165. doi:https://doi.org/10.1016/0028-3908(90)90056-W Hoyte, C. O., Jacob, J., Monte, A. A., Al-Jumaan, M., Bronstein, A. C., & Heard, K. J. (2012). A characterizationACCEPTED of synthetic cannabinoid MANUSCRIPT exposures reported to the National Poison Data System in 2010. Ann Emerg Med, 60(4), 435-438. doi:10.1016/j.annemergmed.2012.03.007 Huestis, M. A. (2007). Human cannabinoid pharmacokinetics. Chem Biodivers, 4(8), 1770-1804. doi:10.1002/cbdv.200790152 Huggins, J. P., Smart, T. S., Langman, S., Taylor, L., & Young, T. (2012). An efficient randomised, placebo-controlled clinical trial with the irreversible fatty acid amide hydrolase-1 inhibitor PF-04457845, which modulates endocannabinoids but fails to induce effective analgesia in patients with pain due to osteoarthritis of the knee. PAIN®, 153(9), 1837-1846. doi:https://doi.org/10.1016/j.pain.2012.04.020 Iannotti, F. A., Hill, C. L., Leo, A., Alhusaini, A., Soubrane, C., Mazzarella, E., . . . Stephens, G. J. (2014). Nonpsychotropic Plant Cannabinoids, (CBDV) and Cannabidiol (CBD), Activate

ACCEPTED MANUSCRIPT 47

and Desensitize Transient Receptor Potential Vanilloid 1 (TRPV1) Channels in Vitro: Potential for the Treatment of Neuronal Hyperexcitability. ACS Chemical Neuroscience, 5(11), 1131- 1141. doi:10.1021/cn5000524 Iffland, K., & Grotenhermen, F. (2017). An Update on Safety and Side Effects of Cannabidiol: A Review of Clinical Data and Relevant Animal Studies. Cannabis Cannabinoid Res, 2(1), 139- 154. doi:10.1089/can.2016.0034 Izzo, A. A., Borrelli, F., Capasso, R., Di Marzo, V., & Mechoulam, R. (2009). Non-psychotropic plant cannabinoids: new therapeutic opportunities from an ancient herb. Trends Pharmacol Sci, 30(10), 515-527. doi:10.1016/j.tips.2009.07.006 Jetly, R., Heber, A., Fraser, G., & Boisvert, D. (2015). The efficacy of nabilone, a synthetic cannabinoid, in the treatment of PTSD-associated nightmares: A preliminary randomized, double-blind, placebo-controlled cross-over design study. Psychoneuroendocrinology, 51, 585-588. doi:10.1016/j.psyneuen.2014.11.002 John, C. S., & Currie, P. J. (2012). N-Arachidonoyl-serotonin in the basolateral amygdala increases anxiolytic behavior in the elevated plus maze. Behavioural Brain Research, 233(2), 382-388. doi:https://doi.org/10.1016/j.bbr.2012.05.025 Johnson, D. S., Stiff, C., Lazerwith, S. E., Kesten, S. R., Fay, L. K., Morris, M., . . . Ahn, K. (2011). Discovery of PF-04457845: A Highly Potent, Orally Bioavailable, and Selective Urea FAAH Inhibitor. ACS Medicinal Chemistry Letters, 2(2), 91-96. doi:10.1021/ml100190t Johnson, M. J., Pierce, J. D., Mavandadi, S., Klaus, J., Defelice, D., Ingram, E., & Oslin, D. W. (2016). Mental health symptom severity in cannabis using and non-using Veterans with probable PTSD. Journal of Affective Disorders, 190, 439-442. doi:10.1016/j.jad.2015.10.048 Kalebasi, N., Kuelen, E., Schnyder, U., Schumacher, S., Mueller-Pfeiffer, C., Wilhelm, F. H., . . . Martin- Soelch, C. (2015). Blunted responses to reward in remitted post-traumatic stress disorder. Brain and Behavior, 5(8), e00357. doi:10.1002/brb3.357 Kano, M., Ohno-Shosaku, T., Hashimotodani, Y., Uchigashima, M., & Watanabe, M. (2009). Endocannabinoid-mediated control of synaptic transmission. Physiol Rev, 89(1), 309-380. doi:10.1152/physrev.00019.2008 Kathuria, S., Gaetani, S., Fegley, D., Valino, F., Duranti, A., Tontini, A., . . . Piomelli, D. (2003). Modulation of anxiety through blockade of anandamide hydrolysis. Nat Med, 9(1), 76-81. doi:10.1038/nm803 Keith, J. M., Jones, W. M., Tichenor, M., Liu, J., Seierstad, M., Palmer, J. A., . . . Breitenbucher, J. G. (2015). Preclinical Characterization of the FAAH Inhibitor JNJ-42165279. ACS Med Chem Lett, 6(12), 1204-1208. doi:10.1021/acsmedchemlett.5b00353 Kerbrat, A., Ferre, J. C., Fillatre, P., Ronziere, T., Vannier, S., Carsin-Nicol, B., . . . Edan, G. (2016). Acute Neurologic Disorder from an Inhibitor of Fatty Acid Amide Hydrolase. N Engl J Med, 375(18), 1717-1725. doi:10.1056/NEJMoa1604221 Kessler, R. C., Aguilar-Gaxiola, S., Alonso, J., Benjet, C., Bromet, E. J., Cardoso, G., . . . On behalf of the, W. H. O. W. M. H. S. C. (2017). Trauma and PTSD in the WHO World Mental Health Surveys. European Journal of Psychotraumatology, 8(sup5), 1353383. doi:10.1080/20008198.2017.1353383ACCEPTED MANUSCRIPT Kessler, R. C., Sonnega, A., Bromet, E., Hughes, M., & Nelson, C. B. (1995). Posttraumatic Stress Disorder in the National Comorbidity Survey. Archives of General Psychiatry, 52(12), 1048- 1060. doi:10.1002/1099-1298(200011/12)10:6<475::AID-CASP578>3.0.CO;2- Kinney, W. A., McDonnell, M. E., Zhong, H. M., Liu, C., Yang, L., Ling, W., . . . Brenneman, D. E. (2016). Discovery of KLS-13019, a Cannabidiol-Derived Neuroprotective Agent, with Improved Potency, Safety, and Permeability. ACS Med Chem Lett, 7(4), 424-428. doi:10.1021/acsmedchemlett.6b00009 Klumpers, F., Denys, D., Kenemans, J. L., Grillon, C., van der Aart, J., & Baas, J. M. (2012). Testing the effects of Delta9-THC and D-cycloserine on extinction of conditioned fear in humans. J Psychopharmacol, 26(4), 471-478. doi:10.1177/0269881111431624

ACCEPTED MANUSCRIPT 48

Korem, N., Lange, R., Hillard, C. J., & Akirav, I. (2017). Role of beta-catenin and endocannabinoids in the nucleus accumbens in extinction in rats exposed to shock and reminders. Neuroscience, 357, 285-294. doi:10.1016/j.neuroscience.2017.06.015 Lazary, J., Eszlari, N., Juhasz, G., & Bagdy, G. (2016). Genetically reduced FAAH activity may be a risk for the development of anxiety and depression in persons with repetitive childhood trauma. Eur Neuropsychopharmacol, 26(6), 1020-1028. doi:10.1016/j.euroneuro.2016.03.003 Leatherdale, S. T., Hammond, D. G., Kaiserman, M., & Ahmed, R. (2007). Marijuana and use among young adults in Canada: are they what we think they are smoking? Cancer Causes Control, 18(4), 391-397. doi:10.1007/s10552-006-0103-x LeDoux, J. E. (2014). Coming to terms with fear. Proc Natl Acad Sci U S A, 111(8), 2871-2878. doi:10.1073/pnas.1400335111 LeDoux, J. E., Iwata, J., Cicchetti, P., & Reis, D. J. (1988). Different Projections of the Central Amygdaloid Nucleus Mediate Autonomic and Behavioral Correlates of Conditioned Fear The Journal of Neuroscience, 8(7), 2517-2529. Lee, J. C., Wang, L. P., & Tsien, J. Z. (2016). Dopamine Rebound-Excitation Theory: Putting Brakes on PTSD. Front Psychiatry, 7, 163. doi:10.3389/fpsyt.2016.00163 Lee, J. L. C., Bertoglio, L. J., Guimaraes, F. S., & Stevenson, C. W. (2017). Cannabidiol regulation of emotion and emotional memory processing: relevance for treating anxiety-related and substance abuse disorders. Br J Pharmacol, 174(19), 3242-3256. doi:10.1111/bph.13724 Liang, T., Neumann, C. N., & Ritter, T. (2013). Introduction of fluorine and fluorine-containing functional groups. Angew Chem Int Ed Engl, 52(32), 8214-8264. doi:10.1002/anie.201206566 Lichtman, A. H., Leung, D., Shelton, C. C., Saghatelian, A., Hardouin, C., Boger, D. L., & Cravatt, B. F. (2004). Reversible Inhibitors of Fatty Acid Amide Hydrolase That Promote Analgesia: Evidence for an Unprecedented Combination of Potency and Selectivity. Journal of Pharmacology and Experimental Therapeutics, 311(2), 441-448. doi:10.1124/jpet.104.069401 Ligresti, A., De Petrocellis, L., & Di Marzo, V. (2016). From Phytocannabinoids to Cannabinoid Receptors and Endocannabinoids: Pleiotropic Physiological and Pathological Roles Through Complex Pharmacology. Physiol Rev, 96(4), 1593-1659. doi:10.1152/physrev.00002.2016 Lisboa, S. F., Niraula, A., Resstel, L. B., Guimaraes, F. S., Godbout, J. P., & Sheridan, J. F. (2018). Repeated social defeat-induced neuroinflammation, anxiety-like behavior and resistance to fear extinction were attenuated by the cannabinoid receptor agonist WIN55,212-2. Neuropsychopharmacology, 43(9), 1924-1933. doi:10.1038/s41386-018-0064-2 Llorente-Berzal, A., Terzian, A. L. B., Di Marzo, V., Micale, V., Viveros, M. P., & Wotjak, C. T. (2015). 2- AG promotes the expression of conditioned fear via cannabinoid receptor type 1 on GABAergic neurons. Psychopharmacology, 232(15), 2811-2825. doi:10.1007/s00213-015- 3917-y Lommen, M. J., Engelhard, I. M., Sijbrandij, M., van den Hout, M. A., & Hermans, D. (2013). Pre - trauma individual differences in extinction learning predict posttraumatic stress. Behav Res Ther, 51(2), 63-67. doi:10.1016/j.brat.2012.11.004 Long, J. Z., Nomura,ACCEPTED D. K., Vann, R. E., Walentiny, MANUSCRIPT D. M., Booker, L., Jin, X., . . . Cravatt, B. F. (2009). Dual blockade of FAAH and MAGL identifies behavioral processes regulated by endocannabinoid crosstalk in vivo. Proceedings of the National Academy of Sciences of the United States of America, 106(48), 20270-20275. doi:10.1073/pnas.0909411106 McGaugh, J. L., & Roozendaal, B. (2002). Role of adrenal stress hormones in forming lasting memories in the brain. Current Opinion in Neurobiology, 12, 205-210. McKinney, M. K., & Cravatt, B. F. (2005). Structure and function of fatty acid amide hydrolase. Annu Rev Biochem, 74, 411-432. doi:10.1146/annurev.biochem.74.082803.133450 McLaughlin, R. J., Verlezza, S., Gray, J. M., Hill, M. N., & Walker, C. D. (2016). Inhibition of anandamide hydrolysis dampens the neuroendocrine response to stress in neonatal rats

ACCEPTED MANUSCRIPT 49

subjected to suboptimal rearing conditions. Stress, 19(1), 114-124. doi:10.3109/10253890.2015.1117448 McPartland, J. M., Duncan, M., Di Marzo, V., & Pertwee, R. G. (2015). Are cannabidiol and Delta(9) - negative modulators of the endocannabinoid system? A systematic review. Br J Pharmacol, 172(3), 737-753. doi:10.1111/bph.12944 McPartland, J. M., Guy, G. W., & Di Marzo, V. (2014). Care and feeding of the endocannabinoid system: a systematic review of potential clinical interventions that upregulate the endocannabinoid system. PLoS One, 9(3), e89566. doi:10.1371/journal.pone.0089566 Mechoulam, R., Ben-Shabat, S., Hanus, L., Ligumsky, M., Kaminski, N. E., Schatz, A. R., . . . Vogel, Z. (1995). Identification of an Endogenous 2-Monoglyceride, Present in Canine Gut, that Binds to Cannbinoid Receptors. Biochemical Pharmacology, 50(1), 83-90. Mechoulam, R., Kogan, N. M., Gallily, R., & Breuer, A. (2008). Washington, DC Patent No. WO2008/107879. Mechoulam, R., Lander, N., Srebnik, M., Breuer, A., Segal, M., Feigenbaum, J., . . . Consroe, P. (1987). Stereochemical requirements for cannabimimetic activity (Vol. 79). Mechoulam, R., & Parker, L. A. (2013). The endocannabinoid system and the brain. Annu Rev Psychol, 64, 21-47. doi:10.1146/annurev-psych-113011-143739 Mechoulam, R., Peters, M., Murillo-Rodriguez, E., & Hanus, L. (2007). Cannabidiol – Recent Advances. Chemistry & Biodiversity, 4(8), 1678-1692. doi:10.1002/cbdv.200790147 Micale, V., Cristino, L., Tamburella, A., Petrosino, S., Leggio, G. M., Drago, F., & Di Marzo, V. (2009). Anxiolytic effects in mice of a dual blocker of fatty acid amide hydrolase and transient receptor potential vanilloid type-1 channels. Neuropsychopharmacology, 34(3), 593-606. doi:10.1038/npp.2008.98 Micale, V., Di Marzo, V., Sulcova, A., Wotjak, C. T., & Drago, F. (2013). Endocannabinoid system and mood disorders: priming a target for new therapies. Pharmacol Ther, 138(1), 18-37. doi:10.1016/j.pharmthera.2012.12.002 Micale, V., Stepan, J., Jurik, A., Pamplona, F. A., Marsch, R., Drago, F., . . . Wotjak, C. T. (2017). Extinction of avoidance behavior by safety learning depends on endocannabinoid signaling in the hippocampus. J Psychiatr Res, 90, 46-59. doi:10.1016/j.jpsychires.2017.02.002 Milad, M. R., Pitman, R. K., Ellis, C. B., Gold, A. L., Shin, L. M., Lasko, N. B., . . . Rauch, S. L. (2009). Neurobiological basis of failure to recall extinction memory in posttraumatic stress disorder. Biol Psychiatry, 66(12), 1075-1082. doi:10.1016/j.biopsych.2009.06.026 Milad, M. R., Quinn, B. T., Pitman, R. K., Orr, S. P., Fischl, B., & Rauch, S. L. (2005). Thickness of ventromedial prefrontal cortex in humans is correlated with extinction memory. PNAS, 102(30), 10706-10711. doi:10.1073/pnas.0502441102 Milad, M. R., & Quirk, G. J. (2012). Fear extinction as a model for translational neuroscience: ten years of progress. Annu Rev Psychol, 63, 129-151. doi:10.1146/annurev.psych.121208.131631 Milad, M. R., Wright, C. I., Orr, S. P., Pitman, R. K., Quirk, G. J., & Rauch, S. L. (2007). Recall of fear extinction in humans activates the ventromedial prefrontal cortex and hippocampus in concert. BiolACCEPTED Psychiatry, 62(5), 446-454. MANUSCRIPTdoi:10.1016/j.biopsych.2006.10.011 Moghaddam, B., & Javitt, D. (2012). From revolution to evolution: the glutamate hypothesis of schizophrenia and its implication for treatment. Neuropsychopharmacology : official publication of the American College of Neuropsychopharmacology, 37(1), 4-15. doi:10.1038/npp.2011.181 Moore, T. H. M., Zammit, S., Lingford-Hughes, A., Barnes, T. R. E., Jones, P. B., Burke, M., & Lewis, G. (2007). Cannabis use and risk of psychotic or affective mental health outcomes: a systematic review. The Lancet, 370(9584), 319-328. doi:10.1016/s0140-6736(07)61162-3 Morales, P., Hurst, D. P., & Reggio, P. H. (2017). Molecular Targets of the Phytocannabinoids: A Complex Picture. Prog Chem Org Nat Prod, 103, 103-131. doi:10.1007/978-3-319-45541-9_4

ACCEPTED MANUSCRIPT 50

Morales, P., Reggio, P. H., & Jagerovic, N. (2017). An Overview on Medicinal Chemistry of Synthetic and Natural Derivatives of Cannabidiol. Front Pharmacol, 8, 422. doi:10.3389/fphar.2017.00422 Moreira, F. A., Grieb, M., & Lutz, B. (2009). Central side-effects of therapies based on CB1 cannabinoid receptor agonists and antagonists: focus on anxiety and depression. Best Pract Res Clin Endocrinol Metab, 23(1), 133-144. doi:10.1016/j.beem.2008.09.003 Morena, M., Berardi, A., Colucci, P., Palmery, M., Trezza, V., Hill, M. N., & Campolongo, P. (2017). Enhancing Endocannabinoid Neurotransmission Augments The Efficacy of Extinction Training and Ameliorates Traumatic Stress-Induced Behavioral Alterations in Rats. Neuropsychopharmacology. doi:10.1038/npp.2017.305 Morena, M., & Campolongo, P. (2014). The endocannabinoid system: an emotional buffer in the modulation of memory function. Neurobiol Learn Mem, 112, 30-43. doi:10.1016/j.nlm.2013.12.010 Morena, M., De Castro, V., Gray, J. M., Palmery, M., Trezza, V., Roozendaal, B., . . . Campolongo, P. (2015). Training-Associated Emotional Arousal Shapes Endocannabinoid Modulation of Spatial Memory Retrieval in Rats. J Neurosci, 35(41), 13962-13974. doi:10.1523/JNEUROSCI.1983-15.2015 Morena, M., Patel, S., Bains, J. S., & Hill, M. N. (2016). Neurobiological Interactions Between Stress and the Endocannabinoid System. Neuropsychopharmacology, 41(1), 80-102. doi:10.1038/npp.2015.166 Morena, M., Roozendaal, B., Trezza, V., Ratano, P., Peloso, A., Hauer, D., . . . Campolongo, P. (2014). Endogenous cannabinoid release within prefrontal-limbic pathways affects memory consolidation of emotional training. Proc Natl Acad Sci U S A, 111(51), 18333-18338. doi:10.1073/pnas.1420285111 Muller, K., Faeh, C., & Diederich, F. (2007). Fluorine in pharmaceuticals: looking beyond intuition. Science, 317(5846), 1881-1886. doi:10.1126/science.1131943 Murataeva, N., Straiker, A., & Mackie, K. (2014). Parsing the players: 2-arachidonoylglycerol synthesis and degradation in the CNS. British Journal of Pharmacology, 171(6), 1379–1391. doi:10.1111/bph.2014.171.issue-6 Mwanza, C., Chen, Z., Zhang, Q., Chen, S., Wang, W., & Deng, H. (2016). Simultaneous HPLC-APCI- MS/MS quantification of endogenous cannabinoids and glucocorticoids in hair. J Chromatogr B Analyt Technol Biomed Life Sci, 1028, 1-10. doi:10.1016/j.jchromb.2016.06.002 Nasehi, M., Farrahizadeh, M., Ebrahimi-Ghiri, M., & Zarrindast, M. R. (2016). Modulation of cannabinoid signaling by hippocampal 5-HT4 serotonergic system in fear conditioning. J Psychopharmacol, 30(9), 936-944. doi:10.1177/0269881116652584 Nasehi, M., Gerami-Majd, F., Khakpai, F., & Zarrindast, M. R. (2018). Dorsal hippocampal cannabinergic and GABAergic systems modulate memory consolidation in passive avoidance task. Brain Res Bull, 137, 197-203. doi:10.1016/j.brainresbull.2017.11.017 Nasehi, M., Shahbazzadeh, S., Ebrahimi-Ghiri, M., & Zarrindast, M. R. (2018). Bidirectional influence of amygdala beta1-adrenoceptors blockade on cannabinoid signaling in contextual and auditory fearACCEPTED memory. J Psychopharmacol, MANUSCRIPT 32(8), 932-942. doi:10.1177/0269881118760654 Natividad, L. A., Buczynski, M. W., Herman, M. A., Kirson, D., Oleata, C. S., Irimia, C., . . . Parsons, L. H. (2017). Constitutive Increases in Amygdalar Corticotropin-Releasing Factor and Fatty Acid Amide Hydrolase Drive an Anxious Phenotype. Biol Psychiatry, 82(7), 500-510. doi:10.1016/j.biopsych.2017.01.005 Nazario, L. R., Antonioli, R. J., Capiotti, K. M., Hallak, J. E., Zuardi, A. W., Crippa, J. A., . . . da Silva, T. R. (2015). Caffeine protects against memory loss induced by high and non-anxiolytic dose of cannabidiol in adult zebrafish (Danio rerio). Pharmacol Biochem Behav, 135, 210-216. doi:10.1016/j.pbb.2015.06.008

ACCEPTED MANUSCRIPT 51

Neumeister, A., Seidel, J., Ragen, B. J., & Pietrzak, R. H. (2015). Translational evidence for a role of endocannabinoids in the etiology and treatment of posttraumatic stress disorder. Psychoneuroendocrinology, 51, 577-584. doi:10.1016/j.psyneuen.2014.10.012 Ney, L. J., Matthews, A., Bruno, R., & Felmingham, K. L. (2018). Modulation of the endocannabinoid system by sex hormones: Implications for Posttraumatic Stress Disorder. Neurosci Biobehav Rev, 94, 302-320. doi:10.1016/j.neubiorev.2018.07.006 Nicholson, A. N., Turner, C., Stone, B. M., & Robson, P. J. (2004). Effect of Delta-9- Tetrahydrocannabinol and Cannabidiol on Nocturnal Sleep and Early-Morning Behavior in Young Adults. Journal of Clinical Psychopharmacology, 24(3), 305-313. doi:10.1097/01.jcp.0000125688.05091.8f Nicholson, E. L., Bryant, R. A., & Felmingham, K. L. (2014). Interaction of noradrenaline and cortisol predicts negative intrusive memories in posttraumatic stress disorder. Neurobiol Learn Mem, 112, 204-211. doi:10.1016/j.nlm.2013.11.018 Nicolussi, S., & Gertsch, J. (2015). Endocannabinoid Transport Revisited. Vitamins and hormones, 98, 441-485. doi:10.1016/bs.vh.2014.12.011 Norrholm, S. D., Jovanovic, T., Olin, I. W., Sands, L. A., Karapanou, I., Bradley, B., & Ressler, K. J. (2011). Fear extinction in traumatized civilians with posttraumatic stress disorder: relation to symptom severity. Biol Psychiatry, 69(6), 556-563. doi:10.1016/j.biopsych.2010.09.013 Norris, C., Loureiro, M., Kramar, C., Zunder, J., Renard, J., Rushlow, W., & Laviolette, S. R. (2016). Cannabidiol Modulates Fear Memory Formation Through Interactions with Serotonergic Transmission in the Mesolimbic System. Neuropsychopharmacology, 41(12), 2839-2850. doi:10.1038/npp.2016.93 Pace-Schott, E. F., Germain, A., & Milad, M. R. (2015). Effects of sleep on memory for conditioned fear and fear extinction. Psychol Bull, 141(4), 835-857. doi:10.1037/bul0000014 Papini, S., Ruglass, L. M., Lopez-Castro, T., Powers, M. B., Smits, J. A., & Hien, D. A. (2017). Chronic cannabis use is associated with impaired fear extinction in humans. J Abnorm Psychol, 126(1), 117-124. doi:10.1037/abn0000224 Parsons, L. H., & Hurd, Y. L. (2015). Endocannabinoid signaling in reward and addiction. Nature reviews. Neuroscience, 16(10), 579-594. doi:10.1038/nrn4004 Passie, T., Emrich, H. M., Karst, M., Brandt, S. D., & Halpern, J. H. (2012). Mitigation of post-traumatic stress symptoms by Cannabis resin: a review of the clinical and neurobiological evidence. Drug Test Anal, 4(7-8), 649-659. doi:10.1002/dta.1377 Patel, S., Hill, M. N., Cheer, J. F., Wotjak, C. T., & Holmes, A. (2017). The endocannabinoid system as a target for novel anxiolytic drugs. Neurosci Biobehav Rev, 76(Pt A), 56-66. doi:10.1016/j.neubiorev.2016.12.033 Patel, S., & Hillard, C. J. (2006). Pharmacological evaluation of cannabinoid receptor ligands in a mouse model of anxiety: further evidence for an anxiolytic role for endogenous cannabinoid signaling. J Pharmacol Exp Ther, 318(1), 304-311. doi:10.1124/jpet.106.101287 Penetar, D. M., Kouri, E. M., Gross, M. M., McCarthy, E. M., Rhee, C. K., Peters, E. N., & Lukas, S. E. (2005). Transdermal alters some of marihuana's effects in male and female volunteers.ACCEPTED Drug Alcohol Depend, 79(2), MANUSCRIPT 211-223. doi:10.1016/j.drugalcdep.2005.01.008 Pertwee, R. G. (2006). Cannabinoid pharmacology: the first 66 years. Br J Pharmacol, 147 Suppl 1, S163-171. doi:10.1038/sj.bjp.0706406 Pertwee, R. G., Ross, R. A., Craib, S. J., & Thomas, A. (2002). (−)-Cannabidiol antagonizes cannabinoid receptor agonists and noradrenaline in the mouse vas deferens. European Journal of Pharmacology, 456(1), 99-106. doi:https://doi.org/10.1016/S0014-2999(02)02624-9 Phelps, E. A., Delgado, M. R., Nearing, K. I., & LeDoux, J. E. (2004). Extinction learning in humans: role of the amygdala and vmPFC. Neuron, 43(6), 897-905. doi:10.1016/j.neuron.2004.08.042 Piscitelli, F., & Di Marzo, V. (2012). "Redundancy" of endocannabinoid inactivation: new challenges and opportunities for pain control. ACS Chem Neurosci, 3(5), 356-363. doi:10.1021/cn300015x

ACCEPTED MANUSCRIPT 52

Pitman, R. K., & Delahanty, D. L. (2005). Conceptually Driven Pharmacological Approaches to Acute Trauma. CNS Spectrums, 10(2), 99-106. doi:https://doi.org/10.1017/S109285290001943X Pitman, R. K., Rasmusson, A. M., Koenen, K. C., Shin, L. M., Orr, S. P., Gilbertson, M. W., . . . Liberzon, I. (2012). Biological studies of post-traumatic stress disorder. Nat Rev Neurosci, 13(11), 769- 787. doi:10.1038/nrn3339 Postnov, A., Schmidt, M. E., Pemberton, D. J., de Hoon, J., van Hecken, A., van den Boer, M., . . . van Laere, K. (2018). Fatty Acid Amide Hydrolase Inhibition by JNJ-42165279: A Multiple- Ascending Dose and a Positron Emission Tomography Study in Healthy Volunteers. Clin Transl Sci, 11(4), 397-404. doi:10.1111/cts.12548 Rabinak, C. A., Angstadt, M., Lyons, M., Mori, S., Milad, M. R., Liberzon, I., & Phan, K. L. (2014). Cannabinoid modulation of prefrontal-limbic activation during fear extinction learning and recall in humans. Neurobiol Learn Mem, 113, 125-134. doi:10.1016/j.nlm.2013.09.009 Rabinak, C. A., Angstadt, M., Sripada, C. S., Abelson, J. L., Liberzon, I., Milad, M. R., & Phan, K. L. (2013). Cannabinoid facilitation of fear extinction memory recall in humans. Neuropharmacology, 64, 396-402. doi:10.1016/j.neuropharm.2012.06.063 Ratano, P., Palmery, M., Trezza, V., & Campolongo, P. (2017). Cannabinoid Modulation of Memory Consolidation in Rats: Beyond the Role of Cannabinoid Receptor Subtype 1. Front Pharmacol, 8, 200. doi:10.3389/fphar.2017.00200 Resstel, L. B., Tavares, R. F., Lisboa, S. F., Joca, S. R., Correa, F. M., & Guimaraes, F. S. (2009). 5-HT1A receptors are involved in the cannabidiol-induced attenuation of behavioural and cardiovascular responses to acute restraint stress in rats. Br J Pharmacol, 156(1), 181-188. doi:10.1111/j.1476-5381.2008.00046.x Rey, A. A., Purrio, M., Viveros, M. P., & Lutz, B. (2012). Biphasic effects of cannabinoids in anxiety responses: CB1 and GABA(B) receptors in the balance of GABAergic and glutamatergic neurotransmission. Neuropsychopharmacology, 37(12), 2624-2634. doi:10.1038/npp.2012.123 Reynoso-Moreno, I., Chicca, A., Flores-Soto, M. E., Viveros-Paredes, J. M., & Gertsch, J. (2018). The Endocannabinoid Reuptake Inhibitor WOBE437 Is Orally Bioavailable and Exerts Indirect Polypharmacological Effects via Different Endocannabinoid Receptors. Front Mol Neurosci, 11, 180. doi:10.3389/fnmol.2018.00180 Roberts, C. J., Stuhr, K. L., Hutz, M. J., Raff, H., & Hillard, C. J. (2014). Endocannabinoid signaling in hypothalamic-pituitary-adrenocortical axis recovery following stress: effects of indirect agonists and comparison of male and female mice. Pharmacol Biochem Behav, 117, 17-24. doi:10.1016/j.pbb.2013.11.026 Rodrigues, S. M., LeDoux, J. E., & Sapolsky, R. M. (2009). The influence of stress hormones on fear circuitry. Annu Rev Neurosci, 32, 289-313. doi:10.1146/annurev.neuro.051508.135620 Roitman, P., Mechoulam, R., Cooper-Kazaz, R., & Shalev, A. (2014). Preliminary, Open-Label, Pilot Study of Add-On Oral Δ9-Tetrahydrocannabinol in Chronic Post-Traumatic Stress Disorder. Clinical Drug Investigation, 34(8), 587-591. doi:10.1007/s40261-014-0212-3 Roozendaal, B., & McGaugh, J. L. (2011). Memory modulation. Behav Neurosci, 125(6), 797-824. doi:10.1037ACCEPTED/a0026187 MANUSCRIPT Rossignoli, M. T., Lopes-Aguiar, C., Ruggiero, R. N., Do Val da Silva, R. A., Bueno-Junior, L. S., Kandratavicius, L., . . . Romcy-Pereira, R. N. (2017). Selective post-training time window for memory consolidation interference of cannabidiol into the prefrontal cortex: Reduced dopaminergic modulation and immediate gene expression in limbic circuits. Neuroscience, 350, 85-93. doi:10.1016/j.neuroscience.2017.03.019 Ruhl, T., Zeymer, M., & von der Emde, G. (2017). Cannabinoid modulation of zebrafish fear learning and its functional analysis investigated by c-Fos expression. Pharmacol Biochem Behav, 153, 18-31. doi:10.1016/j.pbb.2016.12.005 Russo, E. B., Burnett, A., Hall, B., & Parker, K. K. (2005). Agonistic Properties of Cannabidiol at 5-HT1a Receptors. Neurochemical Research, 30(8), 1037-1043. doi:10.1007/s11064-005-6978-1

ACCEPTED MANUSCRIPT 53

Sachser, R. M., Crestani, A. P., Quillfeldt, J. A., Souza, T. M., & de Oliveira Alvares, L. (2015). The cannabinoid system in the retrosplenial cortex modulates fear memory consolidation, reconsolidation, and extinction. Learn Mem, 22(12). doi:10.1101/lm.039891.115 Sales, A. J., Crestani, C. C., Guimaraes, F. S., & Joca, S. R. L. (2018). Antidepressant-like effect induced by Cannabidiol is dependent on brain serotonin levels. Prog Neuropsychopharmacol Biol Psychiatry, 86, 255-261. doi:10.1016/j.pnpbp.2018.06.002 Sanofi. (2013). An eight-week, multicenter, randomized, double-blind, placebo-controlled dose- finding study, with escitalopram (10 mg daily) as active control, to evaluate the efficacy, safety and tolerability of three fixed doses of SSR411298 (10, 50, or 200 mg daily) in elderly patients with Major Depressive Disorder (FIDELIO). Retrieved from https://en.sanofi.com/img/content/study/dfi10560_summary.pdf Santana, F., Sierra, R. O., Haubrich, J., Crestani, A. P., Duran, J. M., de Freitas Cassini, L., . . . Quillfeldt, J. A. (2016). Involvement of the infralimbic cortex and CA1 hippocampal area in reconsolidation of a contextual fear memory through CB1 receptors: Effects of CP55,940. Neurobiol Learn Mem, 127, 42-47. doi:10.1016/j.nlm.2015.11.016 Sartim, A. G., Moreira, F. A., & Joca, S. R. L. (2017). Involvement of CB1 and TRPV1 receptors located in the ventral medial prefrontal cortex in the modulation of stress coping behavior. Neuroscience, 340, 126-134. doi:10.1016/j.neuroscience.2016.10.031 Sbarski, B., & Akirav, I. (2018). Chronic exposure to cannabinoids before an emotional trauma may have negative effects on emotional function. Eur Neuropsychopharmacol, 28(8), 955-969. doi:10.1016/j.euroneuro.2018.05.008 Scuderi, C., Filippis, D. D., Iuvone, T., Blasio, A., Steardo, A., & Esposito, G. (2009). Cannabidiol in medicine: a review of its therapeutic potential in CNS disorders. Phytother Res, 23(5), 597- 602. doi:10.1002/ptr.2625 Segev, A., & Akirav, I. (2016). Cannabinoids and Glucocorticoids in the Basolateral Amygdala Modulate Hippocampal-Accumbens Plasticity After Stress. Neuropsychopharmacology, 41(4), 1066-1079. doi:10.1038/npp.2015.238 Segev, A., Korem, N., Zer-Aviv, T., Abush, H., Lange, R., Sauber, G., . . . Akirav, I. (2018). Role of endocannabinoids in the hippocampus and amygdala in emotional memory and plasticity. Neuropsychopharmacology. doi:10.1038/s41386-018-0135-4 Shannon, S., Lewis, N., Lee, H., & Hughes, S. (2019). Cannabidiol in Anxiety and Sleep: A Large Case Series. The Permanente journal, 23, 18-041. doi:10.7812/TPP/18-041 Shannon, S., & Opila-Lehman, J. (2016). Effectiveness of Cannabidiol Oil for Pediatric Anxiety and Insomnia as Part of Posttraumatic Stress Disorder: A Case Report. Perm J, 20(4), 108-111. doi:10.7812/TPP/16-005 Sharma, C. V., Long, J. H., Shah, S., Rahman, J., Perrett, D., Ayoub, S. S., & Mehta, V. (2017). First evidence of the conversion of paracetamol to AM404 in human cerebrospinal fluid. J Pain Res, 10, 2703-2709. doi:10.2147/JPR.S143500 Shechner, T., Hong, M., Britton, J. C., Pine, D. S., & Fox, N. A. (2014). Fear conditioning and extinction across development: evidence from human studies and animal models. Biol Psychol, 100, 1- 12. doi:10.1016/j.biopsycho.2014.04.001ACCEPTED MANUSCRIPT Shoshan, N., & Akirav, I. (2017). The effects of cannabinoid receptors activation and glucocorticoid receptors deactivation in the amygdala and hippocampus on the consolidation of a traumatic event. Neurobiol Learn Mem, 144, 248-258. doi:10.1016/j.nlm.2017.08.004 Shoshan, N., Segev, A., Abush, H., Mizrachi Zer-Aviv, T., & Akirav, I. (2017). Cannabinoids prevent the differential long-term effects of exposure to severe stress on hippocampal- and amygdala- dependent memory and plasticity. Hippocampus, 27(10), 1093-1109. doi:10.1002/hipo.22755 Showalter, V. M., Compton, D. R., Martin, B. R., & Abood, M. E. (1996). Evaluation of binding in a transfected cell line expressing a peripheral cannabinoid receptor (CB2): identification of cannabinoid receptor subtype selective ligands. J Pharmacol Exp Ther, 278(3), 989-999.

ACCEPTED MANUSCRIPT 54

Shvil, E., Sullivan, G. M., Schafer, S., Markowitz, J. C., Campeas, M., Wager, T. D., . . . Neria, Y. (2014). Sex differences in extinction recall in posttraumatic stress disorder: a pilot fMRI study. Neurobiol Learn Mem, 113, 101-108. doi:10.1016/j.nlm.2014.02.003 Silva, N. R., Gomes, F. V., Fonseca, M. D., Mechoulam, R., Breuer, A., Cunha, T. M., & Guimaraes, F. S. (2017). Antinociceptive effects of HUF-101, a fluorinated cannabidiol derivative. Prog Neuropsychopharmacol Biol Psychiatry, 79(Pt B), 369-377. doi:10.1016/j.pnpbp.2017.07.012 Singewald, N., Schmuckermair, C., Whittle, N., Holmes, A., & Ressler, K. J. (2015). Pharmacology of cognitive enhancers for exposure-based therapy of fear, anxiety and trauma-related disorders. Pharmacol Ther, 149, 150-190. doi:10.1016/j.pharmthera.2014.12.004 Soares, V. d. P., Campos, A. C., Bortoli, V. C. d., Zangrossi, H., Guimarães, F. S., & Zuardi, A. W. (2010). Intra-dorsal periaqueductal gray administration of cannabidiol blocks panic-like response by activating 5-HT1A receptors. Behavioural Brain Research, 213(2), 225-229. doi:https://doi.org/10.1016/j.bbr.2010.05.004 Song, C., Stevenson, C. W., Guimaraes, F. S., & Lee, J. L. (2016). Bidirectional Effects of Cannabidiol on Contextual Fear Memory Extinction. Front Pharmacol, 7, 493. doi:10.3389/fphar.2016.00493 Southwick, S. M., Bremner, J. D., Rasmusson, A., Morgan, C. A., 3rd, Arnsten, A., & Charney, D. S. (1999). Role of norepinephrine in the pathophysiology and treatment of posttraumatic stress disorder. Biol Psychiatry, 46(9), 1192-1204. Spaderna, M., Addy, P. H., & D'Souza, D. C. (2013). Spicing things up: synthetic cannabinoids. Psychopharmacology, 228(4), 525-540. doi:10.1007/s00213-013-3188-4 Spagnolo, P. A., Ramchandani, V. A., Schwandt, M. L., Kwako, L. E., George, D. T., Mayo, L. M., . . . Heilig, M. (2016). FAAH Gene Variation Moderates Stress Response and Symptom Severity in Patients with Posttraumatic Stress Disorder and Comorbid Alcohol Dependence. Alcohol Clin Exp Res, 40(11), 2426-2434. doi:10.1111/acer.13210 Steenkamp, M. M., Blessing, E. M., Galatzer-Levy, I. R., Hollahan, L. C., & Anderson, W. T. (2017). Marijuana and other cannabinoids as a treatment for posttraumatic stress disorder: A literature review. Depress Anxiety, 34(3), 207-216. doi:10.1002/da.22596 Stein, D. J., Ipser, J. C., & Seedat, S. (2006). Pharmacotherapy for post traumatic stress disorder (PTSD). Cochrane Database Syst Rev(1), CD002795. doi:10.1002/14651858.CD002795.pub2 Stein, M. B., Yehuda, R., Koverola, C., & Hanna, C. (1997). Enhanced Dexamethasone Suppression of Plasma Cortisol in Adult Women Traumatized by Childhood Sexual Abuse. Biological Psychiatry, 42(8), 680-686. doi:https://doi.org/10.1016/S0006-3223(96)00489-1 Stern, C. A., Gazarini, L., Vanvossen, A. C., Zuardi, A. W., Galve-Roperh, I., Guimaraes, F. S., . . . Bertoglio, L. J. (2015). Delta9-Tetrahydrocannabinol alone and combined with cannabidiol mitigate fear memory through reconsolidation disruption. Eur Neuropsychopharmacol, 25(6), 958-965. doi:10.1016/j.euroneuro.2015.02.001 Stern, C. A. J., da Silva, T. R., Raymundi, A. M., de Souza, C. P., Hiroaki-Sato, V. A., Kato, L., . . . Bertoglio, L. J. (2017). Cannabidiol disrupts the consolidation of specific and generalized fear memories via dorsal hippocampus CB1 and CB2 receptors. Neuropharmacology, 125, 220- 230. doi:10.1016/j.neuropharm.2017.07.024ACCEPTED MANUSCRIPT Stern, C. A. J., de Carvalho, C. R., Bertoglio, L. J., & Takahashi, R. N. (2018). Effects of Cannabinoid Drugs on Aversive or Rewarding Drug-Associated Memory Extinction and Reconsolidation. Neuroscience, 370, 62-80. doi:10.1016/j.neuroscience.2017.07.018 Suguira, T., Kondo, S., Sukawaga, A., Nakane, S., Shinoda, A., Itoh, K., . . . Waku, K. (1995). 2- Arachidonoylgylcerol: A possible endogenous cannabinoid receptor ligand in the brain. Biochemical and Biophysical Research Communications, 215(1), 89-97. Sun, H., Su, R., Zhang, X., Wen, J., Yao, D., Gao, X., . . . Li, H. (2017). Hippocampal GR- and CB1- mediated mGluR5 differentially produces susceptibility and resilience to acute and chronic mild stress in rats. Neuroscience, 357, 295-302. doi:10.1016/j.neuroscience.2017.06.017

ACCEPTED MANUSCRIPT 55

Takeda, S., Usami, N., Yamamoto, I., & Watanabe, K. (2009). Cannabidiol-2',6'-dimethyl ether, a cannabidiol derivative, is a highly potent and selective 15-lipoxygenase inhibitor. Drug Metab Dispos, 37(8), 1733-1737. doi:10.1124/dmd.109.026930 Todd, S. M., & Arnold, J. C. (2016). Neural correlates of interactions between cannabidiol and Delta(9) -tetrahydrocannabinol in mice: implications for medical cannabis. Br J Pharmacol, 173(1), 53- 65. doi:10.1111/bph.13333 Trezza, V., & Campolongo, P. (2013). The endocannabinoid system as a possible target to treat both the cognitive and emotional features of post-traumatic stress disorder (PTSD). Front Behav Neurosci, 7, 100. doi:10.3389/fnbeh.2013.00100 Tyndale, R. F., Payne, J. I., Gerber, A. L., & Sipe, J. C. (2007). The fatty acid amide hydrolase C385A (P129T) missense variant in cannabis users: Studies of drug use and dependence in caucasians. American Journal of Medical Genetics Part B: Neuropsychiatric Genetics, 144B(5), 660-666. doi:10.1002/ajmg.b.30491 U.S. Food and Drug Administration. (2004). Marinol (Dronabinol) Capsules. Marietta, GA Solvay Pharmaceuticals, Inc. . U.S. Food and Drug Administration. (2006). Cesamat (nabilone) Capsules for Oral Administration. Costa Mesa, CA Valeant Pharmaceuticals International. U.S. Food and Drug Administration. (2016). FDA finds drugs under investigation in the U.S. related to French BIA 10-2474 drug do not pose similar safety risks. Retrieved from https://www.fda.gov/Drugs/DrugSafety/ucm482740.htm Uhernik, A. L., Montoya, Z. T., Balkissoon, C. D., & Smith, J. P. (2018). Learning and memory is modulated by cannabidiol when administered during trace fear-conditioning. Neurobiol Learn Mem, 149, 68-76. doi:10.1016/j.nlm.2018.02.009 van Esbroeck, A. C. M., Janssen, A. P. A., Cognetta, A. B., Ogasawara, D., Shpak, G., van der Kroeg, M., . . . van der Stelt, M. (2017). Activity-based protein profiling reveals off-target proteins of the FAAH inhibitor BIA 10-2474. Science, 356(6342), 1084-1087. doi:10.1126/science.aaf7497 Wang, M., Duan, F., Wu, J., Min, Q., Huang, Q., Luo, M., & He, Z. (2018). Effect of cyclooxygenase2 inhibition on the development of posttraumatic stress disorder in rats. Mol Med Rep, 17(4), 4925-4932. doi:10.3892/mmr.2018.8525 Ware, M. A., Martel, M. O., Jovey, R., Lynch, M. E., & Singer, J. (2018). A prospective observational study of problematic oral cannabinoid use. Psychopharmacology (Berl), 235(2), 409-417. doi:10.1007/s00213-017-4811-6 Wenzel, J. M., Oleson, E. B., Gove, W. N., Cole, A. B., Gyawali, U., Dantrassy, H. M., . . . Cheer, J. F. (2018). Phasic Dopamine Signals in the Nucleus Accumbens that Cause Active Avoidance Require Endocannabinoid Mobilization in the Midbrain. Curr Biol, 28(9), 1392-1404.e1395. doi:10.1016/j.cub.2018.03.037 Wilker, S., Pfeiffer, A., Elbert, T., Ovuga, E., Karabatsiakis, A., Krumbholz, A., . . . Kolassa, I. T. (2016). Endocannabinoid concentrations in hair are associated with PTSD symptom severity. Psychoneuroendocrinology, 67, 198-206. doi:10.1016/j.psyneuen.2016.02.010 Wilkinson, S. T., Stefanovics,ACCEPTED E., & Rosenheck, R. MANUSCRIPT A. (2015). Marijuana use is associated with worse outcomes in symptom severity and violent behavior in patients with posttraumatic stress disorder. J Clin Psychiatry, 76(9), 1174-1180. doi:10.4088/JCP.14m09475 World Health Organisation. (2017). Agenda Item 5.2: Cannabidiol (CBD): Pre-Review Report. Retrieved from Geneva: Wright, M. J., Vandewater, S. A., & Taffe, M. A. (2013). Cannabidiol attenuates deficits of visuospatial associative memory induced by D9tetrahydrocannabinol. Br J Pharmacol, 170(7), 1365-1373. doi:10.1111/bph.12199 Wyrofsky, R. R., Reyes, B. A. S., Yu, D., Kirby, L. G., & Van Bockstaele, E. J. (2018). Sex differences in the effect of cannabinoid type 1 receptor deletion on locus coeruleus-norepinephrine

ACCEPTED MANUSCRIPT 56

neurons and corticotropin releasing factor-mediated responses. Eur J Neurosci, 48(5), 2118- 2138. doi:10.1111/ejn.14103 Xing, G., Carlton, J., Zhang, L., Jiang, X., Fullerton, C., Li, H., & Ursano, R. (2011). Cannabinoid receptor expression and phosphorylation are differentially regulated between male and female cerebellum and brain stem after repeated stress: Implication for PTSD and drug abuse. Neuroscience Letters, 502(1), 5-9. doi:10.1016/j.neulet.2011.05.013 Yehuda, R. (2009). Status of glucocorticoid alterations in post-traumatic stress disorder. Ann N Y Acad Sci, 1179, 56-69. doi:10.1111/j.1749-6632.2009.04979.x Yehuda, R., Hoge, C. W., McFarlane, A. C., Vermetten, E., Lanius, R. A., Nievergelt, C. M., . . . Hyman, S. E. (2015). Post-traumatic stress disorder. Nature Reviews Disease Primers, 1, 15057. doi:10.1038/nrdp.2015.57 Zer-Aviv, T. M., & Akirav, I. (2016). Sex differences in hippocampal response to endocannabinoids after exposure to severe stress. Hippocampus, 26(7), 947-957. doi:10.1002/hipo.22577 Zer-Aviv, T. M., Segev, A., & Akirav, I. (2016). Cannabinoids and post-traumatic stress disorder: clinical and preclinical evidence for treatment and prevention. Behav Pharmacol, 27(7), 561- 569. doi:10.1097/FBP.0000000000000253 Zhang, D., Saraf, A., Kolasa, T., Bhatia, P., Zheng, G. Z., Patel, M., . . . Surowy, C. S. (2007). Fatty acid amide hydrolase inhibitors display broad selectivity and inhibit multiple carboxylesterases as off-targets. Neuropharmacology, 52(4), 1095-1105. doi:https://doi.org/10.1016/j.neuropharm.2006.11.009 Zhornitsky, S., & Potvin, S. (2012). Cannabidiol in humans-the quest for therapeutic targets. Pharmaceuticals (Basel), 5(5), 529-552. doi:10.3390/ph5050529 Zuardi, A. W. (2008). Cannabidiol: from an inactive cannabinoid to a drug with wide spectrum of action. Revista Brasileira de Psiquiatria, 30(3), 271-280. Zuardi, A. W., Morais, S. L., Guimaraes, F. S., & Mechoulam, R. (1995). Antipsychotic effect of cannabidiol. Journal of Clinical Psychiatry, 56(10), 485-486. Zuardi, A. W., Rodrigues, N. P., Silva, A. L., Bernardo, S. A., Hallak, J. E. C., Guimaraes, F. S., & Crippa, J. A. S. (2017). Inverted U-Shaped Dose-Response Curve of the Anxiolytic Effect of Cannabidiol during Public Speaking in Real Life. Front Pharmacol, 8, 259. doi:10.3389/fphar.2017.00259 Zubedat, S., & Akirav, I. (2017). The involvement of cannabinoids and mTOR in the reconsolidation of an emotional memory in the hippocampal-amygdala-insular circuit. Eur Neuropsychopharmacol, 27(4), 336-349. doi:10.1016/j.euroneuro.2017.01.011 Zuj, D. V., Palmer, M. A., Hsu, C. M., Nicholson, E. L., Cushing, P. J., Gray, K. E., & Felmingham, K. L. (2016). Impaired Fear Extinction Associated with Ptsd Increases with Hours-since-Waking. Depress Anxiety, 33(3), 203-210. doi:10.1002/da.22463 Zuj, D. V., Palmer, M. A., Lommen, M. J., & Felmingham, K. L. (2016). The centrality of fear extinction in linking risk factors to PTSD: A narrative review. Neurosci Biobehav Rev, 69, 15-35. doi:10.1016/j.neubiorev.2016.07.014

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 57

Figure Captions

Figure 1. Chemical structures of cannabinoid receptor agonists

Figure 2. Chemical structures of endocannabinoid catabolic and reuptake enzyme inhibitors

Figure 3. Chemical structures of cannabidiol and notable cannabidiol derivatives

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT 58

Highlights: Cannabinoid interventions for PTSD: Where to next?

 Cannabinoids are efficacious for PTSD in animal models of the disorder.

 Cannabinoid agents work through multiple pathways, such as serotonin and BDNF, to

modulate PTSD symptomology.

 Different cannabinoid agents are reviewed, with strengths, weaknesses and current

challenges of each approach summarised and compared.

 Clinical trials for PTSD should be constructed based on current directions in

pharmacological development of cannabinoids.

ACCEPTED MANUSCRIPT

Figure 1 Figure 2 Figure 3