Skip to main content
International Journal of Neuropsychopharmacology logoLink to International Journal of Neuropsychopharmacology
. 2015 Sep 5;19(3):pyv095. doi: 10.1093/ijnp/pyv095

High Times for Painful Blues: The Endocannabinoid System in Pain-Depression Comorbidity

Marie Fitzgibbon 1, David P Finn 1, Michelle Roche 1,
PMCID: PMC4815466  PMID: 26342110

Abstract

Depression and pain are two of the most debilitating disorders worldwide and have an estimated cooccurrence of up to 80%. Comorbidity of these disorders is more difficult to treat, associated with significant disability and impaired health-related quality of life than either condition alone, resulting in enormous social and economic cost. Several neural substrates have been identified as potential mediators in the association between depression and pain, including neuroanatomical reorganization, monoamine and neurotrophin depletion, dysregulation of the hypothalamo-pituitary-adrenal axis, and neuroinflammation. However, the past decade has seen mounting evidence supporting a role for the endogenous cannabinoid (endocannabinoid) system in affective and nociceptive processing, and thus, alterations in this system may play a key role in reciprocal interactions between depression and pain. This review will provide an overview of the preclinical evidence supporting an interaction between depression and pain and the evidence supporting a role for the endocannabinoid system in this interaction.

Keywords: depression, pain, anandamide, cannabinoid, stress

Clinical Data Supporting Depression-Pain Comorbidity

Depression and pain are two of the most prevalent psychiatric and neurological disorders worldwide, and both are associated with significant disability, impaired health-related quality of life, and high mortality (Spitzer et al., 1995; Kvien, 2004; Scholich et al., 2012; Hassett et al., 2014). While each is considered a debilitating disorder in its own right, these disease entities frequently coexist, and it has been reported that this association may be as high as 80% of patients (Poole et al., 2009). For example, major depressive and bipolar disorder is associated with painful symptoms in up to 95% of patients (Grover et al., 2012; Maneeton et al., 2013; Nicholl et al., 2014). Similarly, patients suffering from inflammatory and neuropathic pain are up to 4.9 times more likely to develop depression or anxiety disorder than the general population (Hawker et al., 2011; Knaster et al., 2012; Emery et al., 2014; Lin et al., 2015). Patients exhibiting comorbid depression and pain do not respond as effectively to pharmacological treatment, and this comorbidity is more disabling and expensive to both patients and society than either condition alone (Emptage et al., 2005; Gameroff and Olfson, 2006). Furthermore, it has also been found that the severity of depression directly correlates with increased severity of pain symptomatology (Khongsaengdao et al., 2000). However, it should be noted that an intricate relationship exists between depression and pain such that although pain is commonly reported by depressed patients, examination of pain thresholds to various stimuli such as cold, heat, and pressure have been shown to be reduced, increased, or unchanged (Ben-Tovim and Schwartz, 1981; Lautenbacher et al., 1999; Gormsen et al., 2004; Bar et al., 2005; Boettger et al., 2013), effects which depend on the modality and intensity of the stimulus. Thus, given the complex interaction between affect and pain, and the high comorbidity of depression-pain, greater understanding of the neurobiological mechanisms underlying the association is warranted to develop more efficacious treatment strategies. There have been several studies investigating the role of neural substrates, including neuroanatomical organization, neurotransmission, neurotrophins, dysregulation of the hypothalamo-pituitary-adrenal (HPA) axis, and inflammation, to name but a few, in the interaction between affect and nociceptive processing (for review, see Blackburn-Munro, 2004; Goesling et al., 2013; Walker et al., 2014). A full review of the role of each of these substrates is beyond the scope of this review. However, increased evidence has indicated a role for a further substrate, the endogenous cannabinoid (endocannabinoid) system, in affective and nociceptive responding (for reviews, see Finn, 2010; Ashton and Moore, 2011; Gorzalka and Hill, 2011; Rani Sagar et al., 2012; Hillard and Liu, 2014; Jennings et al., 2014; Boychuk et al., 2015) and as such, alterations in this system may provide a common mechanism by which depression and pain coexist. Preclinical animal models provide a valuable means of investigating potential neurobiological substrates that may underlie the association between depression and pain. As such, this review will provide an overview of the preclinical evidence supporting an interaction between depression and pain, the evidence supporting a role for the endocannabinoid system in this interaction, and the potential mechanisms through which the endocannabinoid system may mediate effects on affect and nociceptive processing.

Preclinical Animal Models Support Depression-Pain Interactions

Animal Models of Depression Exhibit Altered Nociceptive Responding

Several animal models of depression based on genetics, stress, lesion, and pharmacological manipulation have been shown to exhibit alterations in nociceptive responding (for review, see Li, 2015), supporting the clinical finding of an association between depression and pain. For example, in rats, the chronic mild stress model of depression has been shown to display a reduced nociceptive threshold to cold (Bardin et al., 2009; Bravo et al., 2012; Bravo et al., 2014) and mechanical (Bardin et al., 2009; Imbe et al., 2012) stimuli and an increased threshold to noxious thermal stimuli (Shi et al., 2010). Furthermore, both inflammatory (Gameiro et al., 2005; Rivat et al., 2010; Wang et al., 2013) and neuropathic (Bravo et al., 2012) pain behavior are enhanced in chronic stress models of depression. Similarly, we and others have shown that the Wistar-Kyoto (WKY) rat, a stress hyperresponsive rat strain with a depressive-like phenotype, exhibits thermal hyperalgesia (Burke et al., 2010), visceral hyperalgesia to colorectal distension (Gibney et al., 2010; Gosselin et al., 2010; O’Malley et al., 2010), enhanced formalin-evoked inflammatory pain behavior (Burke et al., 2010; Rea et al., 2014), and enhanced mechanical allodynia following peripheral nerve injury (neuropathic pain) (Zeng et al., 2008; del Rey et al., 2011). Reserpine-induced monoamine depletion has long been known to result in depressive-like behavior, and recent evidence has demonstrated accompanying thermal allodynia (Liu et al., 2014), as well as pronounced and long-lasting mechanical hyperalgesia and allodynia, and cold allodynia (Nagakura et al., 2009; Arora et al., 2011). Thus, this model has been proposed as a possible rodent model of fibromyalgia (Nagakura et al., 2009). Furthermore, recent work from our group has demonstrated that the olfactory bulbectomised rat, a lesion model of depression, exhibits increased sensitivity to mechanical and thermal stimuli in the von Frey, acetone drop, hot plate, and tail flick tests (Burke et al., 2010, 2013), increased inflammatory pain responding in the formalin test (Burke et al., 2010), and enhanced neuropathic pain responding following spinal nerve ligation (Burke et al., 2013, 2014). Thus, taken together, several animal models of depression have been shown to exhibit altered nociceptive thresholds and enhanced inflammatory and neuropathic pain behavior, mimicking effects observed clinically.

Animal Models of Chronic Pain Exhibit Depressive-Like Behavior

Depressive- and anxiety-like behavior, as assessed by multiple paradigms, has been reported in a wide variety of preclinical models of chronic pain (for review, see Yalcin et al., 2014; Li, 2015). For example, peripheral or spared nerve injury in mice induces a pronounced mechanical allodynia accompanied by the development of depressive-like behavior as determined by enhanced immobility in the forced swim test (FST) (Goncalves et al., 2008; Norman et al., 2010; Wang et al., 2011). Similarly, rodents subjected to the chronic constriction injury model of neuropathic pain exhibit reduced sucrose preference (Dellarole et al., 2014) and increased immobility in the FST (Hu et al., 2009; Jesse et al., 2010; Fukuhara et al., 2012; Zhao et al., 2014), indicating the development of anhedonia and behavioral despair, hallmarks of depressive-like behavior. In the complete Freund’s adjuvant model of inflammatory pain, both mice and rats exhibited depression-like behavior in the FST (Maciel et al., 2013; Borges et al., 2014) and tail suspension test (Maciel et al., 2013) and anxiety-related behavior in the elevated plus maze, open field test, and social interaction test (Parent et al., 2012; Borges et al., 2014). Such changes in affective processing in chronic pain models have been shown to occur later than the development of enhanced somatosensory perception. For example, in neuropathic pain models, alterations in emotional behavior have been observed 4 to 8 weeks post nerve injury (Suzuki et al., 2007; Yalcin et al., 2011), but not prior to this (2-4weeks) when mechanical allodynia/hypersensitivity is observed (Kontinen et al., 1999; Hasnie et al., 2007). These studies highlight the development of depressive- and anxiety-like behavior in models of neuropathic or inflammatory pain and suggest that pathological alterations induced by persistent nociceptive input to brain regions that process both pain and affect may account, at least in part, for comorbid depressive-like behavioral changes.

Animal models that replicate the clinical scenario are important for in-depth investigation of the possible neurobiological substrates that may mediate the association between depression and chronic pain. Although the cause of this coexistence remains somewhat elusive, as highlighted earlier, common anatomical substrates and neurobiological mediators, including neurotransmitters, neurotrophins, neuroendocrine alterations, and inflammatory mediators, have been identified, any or all of which may alter neural functioning in key brain regions involved in regulating emotional and nociceptive processing (for review, see Blackburn-Munro, 2004; Maletic and Raison, 2009; Anderson et al., 2012; Goesling et al., 2013; Meerwijk et al., 2013; Jennings et al., 2014; Walker et al., 2014; Doan et al., 2015). Increasing evidence has highlighted an important role for the endocannabinoid system in modulating emotional and nociceptive processing, and thus this system may play a key role in the association between pain and depression.

The Endocannabinoid System

The plant Cannabis sativa has been used as a medicine throughout the world for several thousand years, with reports of its use in treating painful symptoms appearing as early as 2600 bc. The principal psychoactive ingredient of Cannabis sativa, delta-9-tetrahydrocannabinol (Δ9-THC), was first identified in 1964 (Gaoni and Mechoulam, 1964), and subsequent studies to understand its mechanism of action led to the discovery of the endogenous cannabinoid (endocannabinoid) system. This endocannabinoid system consists of the cannabinoid receptors (CB1 and CB2) (Devane et al., 1988; Matsuda et al., 1990; Munro et al., 1993), their naturally occurring endogenous ligands (the best characterized of which are N-arachidonoylethanolamine, or anandamide [AEA]) (Devane et al., 1992) and 2-arachidonylglycerol (2-AG) (Mechoulam et al., 1995), and the enzymes involved in their biosynthesis and degradation. Other endocannabinoid ligands have also been identified, including oleamide (Leggett et al., 2004), O-arachidonoyl ethanolamine (virodhamine) (Porter et al., 2002), 2-arachidonoyl glycerol ether (noladin ether) (Hanus et al., 2001), and N-arachidonoyl-dopamine (Huang et al., 2002), although their physiological role has not been examined in detail. Endocannabinoid biosynthesis occurs on demand via hydrolysis of cell membrane phospholipid precursors. AEA is formed from the precursor N-arachidonoylphosphatidylethanolamine due to the hydrolytic activity of the phospholipase D enzyme NAPE-PLD (Di Marzo et al., 1994; Sugiura et al., 1996), while fatty acid amide hydrolase (FAAH) is the primary enzyme responsible for the metabolism of this endocannabinoid (Cravatt et al., 1996). In comparison, the main biosynthetic pathway for 2-AG involves the hydrolysis of the membrane phospholipid phosphatidylinositol by phospholipase C, producing 1,2-diacylglycerol, which is then converted to 2-AG by diacylglycerol lipase (Prescott and Majerus, 1983, Sugiura et al., 1995). 2-AG is primarily metabolized by monoacylglycerol lipase (MAGL) (85%) (Blankman et al., 2007), although other enzymes including cyclooxygenase-2 (Yu et al., 1997; Kozak et al., 2000), lipoxygenase (van der Stelt et al., 2002), ABDH6 (serine hydrolase α/β-hydrolase domain), and ABDH12 (Blankman et al., 2007), have also been shown to play a role.

Upon release, endocannabinoids bind and activate the G-protein coupled receptors CB1 and/or CB2. CB1 receptors are highly expressed on presynaptic neurons throughout the human and rodent brain (Herkenham, 1991; Tsou et al., 1998; Mackie, 2008), the activation of which results in inhibition of cyclic AMP, activation of mitogen-activated protein kinase, inhibition of N- and P/Q-type voltage-activated Ca2+ channels, and induction of inwardly rectifying K+ currents, with the resultant inhibition of neurotransmitter release (Demuth and Molleman, 2006). CB1 receptors have also been shown to be expressed on glia and a wide range of peripheral tissues, though at lower levels than observed on neurons (Galiegue et al., 1995; Carlisle et al., 2002; Osei-Hyiaman et al., 2005; Cavuoto et al., 2007; Cota, 2007). In contrast, CB2 receptors are widely distributed in peripheral tissues and organs, with a particularly high density on immune cells and tissues (Munro et al., 1993; Berdyshev, 2000; Sugiura et al., 2000), including on glia within the brain, with enhanced expression observed under neuroinflammatory conditions (Carlisle et al., 2002; Nunez et al., 2004; Rock et al., 2007). Accumulating evidence has also indicated that the CB2 receptor is also expressed on subsets of neurons within the brain (Van Sickle et al., 2005; Gong et al., 2006; Baek et al., 2008; Zhang et al., 2014) and thus also modulates neurotransmission (Roche and Finn, 2010; Atwood et al., 2012; Kim and Li, 2015). Endocannabinoids also have affinity for and activity at other receptors, namely the transient receptor potential vanilloid 1, peroxisome proliferator-activated receptors, GPR55, and GPR119 (Huang et al., 2002; Overton et al., 2006; Sun et al., 2006; Ryberg et al., 2007). Activity at these receptors has been proposed to account, at least partially, for some of the differential effects observed with potent selective cannabinoid agonists and pharmacological modulators of endocannabinoid tone.

Because of the distribution of the endocannabinoid system throughout spinal and supraspinal regions, it is in a prime position to regulate neurophysiological activities such as affective and nociceptive processing. This has been a very active area of research over the past decade, with a number of excellent reviews synthesizing the data supporting a role for the endocannabinoid system in modulating mood and nociception (for review, see Ashton and Moore, 2011; Gorzalka and Hill, 2011; Rani Sagar et al., 2012; Hillard and Liu, 2014; Ulugol, 2014; Boychuk et al., 2015). However, no review to date has examined the evidence that may support a role for the endocannabinoid system as a link between depression and pain, and thus the remainder of this review aims to collate and synthesize these data.

The Role of the Endocannabinoid System In Depression-Pain Interactions

Clinical Evidence

Several lines of evidence have demonstrated alterations in the endocannabinoid system in chronic pain (Richardson et al., 2008; Kaufmann et al., 2009) and in psychiatric patients (Gobbi et al., 2005; Hill and Gorzalka, 2005; Koethe et al., 2007). For example, various polymorphisms of CB1 and CB2 receptors have been identified in patients with major depression and bipolar disorder (Monteleone et al., 2010; Minocci et al., 2011; Mitjans et al., 2012; Mitjans et al., 2013) with a single nucleotide polymorphism in the CB1 receptor reported to enhance the risk of treatment resistance in depression (Domschke et al., 2008) and the development of anhedonic depression following early life trauma (Agrawal et al., 2012). Similarly, genetic alterations in the CB1 receptor and FAAH have also been identified in patients with pain associated with migraine, Parkinson’s disease, and irritable bowel syndrome (Juhasz et al., 2009; Park et al., 2011; Greenbaum et al., 2012). In addition, serum levels of endocannabinoids have been reported to be reduced in both depressed patients (Hill et al., 2008c, 2009b) and chronic pain patients (Fichna et al., 2013). A recent study has reported enhanced plasma 2-AG levels and increased CB1 and CB2 mRNA expression on blood lymphocytes in osteoarthritic patients (La Porta et al., 2015). A significant positive correlation was observed between 2-AG levels, pain, and depression, and a negative correlation of 2-AG with quality of life and visual memory was observed (La Porta et al., 2015). In addition, CB1 receptor expression was positively correlated with depression scores, while CB2 receptor expression was correlated with pain scores. These data indicate that key components of the endocannabinoid system are upregulated in human osteoarthritis with significant correlations with pain and emotional symptoms. In addition to visual loss and sensory deficits, neuromyelitis optica is associated with significant pain (altered threshold responding and symptoms of neuropathic pain), depression, and increased plasma levels of 2-AG and AEA (Pellkofer et al., 2013). This study evaluated if a correlation existed between pain threshold and levels of endocannabinoids, demonstrating a considerable negative correlation between the plasma levels of 2-AG and mechanical pain thresholds in these patients, although this study did not evaluate if an association also existed with depressed mood. While the data suggest a possible association between pain, depression, and the endocannabinoid system in osteoarthritis and neuromyelitis optica patients, further clinical studies are required to determine if alterations in the genetics, levels, and activity of the endocannabinoid system exist in other patient groups exhibiting depression-pain comorbidity.

There has been a paucity of clinical studies directly investigating the role or activity of cannabinoids in depression-pain interactions; however, enhanced mood and improved quality of life have been reported in studies investigating the analgesic efficacy of cannabinoid-based therapies (Table 1). For instance, cannabis intake has been reported to improve muscle and nerve pain as well as depression and anxiety symptomatology in a group of HIV patients (Woolridge et al., 2005). Improvements in anxiety and overall distress have been reported in patients with advanced cancer in whom pain symptoms were managed by daily adjunctive administration of Cesamet (nabilone, a Δ9-THC analogue) for 30 days (Maida et al., 2008). Similarly, a randomized, double blind, placebo-controlled trial, which examined the therapeutic benefit of nabilone in terms of pain management and quality of life improvement in patients with fibromyalgia, identified significant pain relief and alleviation of anxiety symptoms after 4 weeks of therapy (Skrabek et al., 2008). In addition, a retrospective evaluation investigating the efficacy of nabilone for the management of concurrent disorders in seriously mentally ill correctional populations identified significant amelioration of symptoms related to posttraumatic stress disorder as well as a subjective improvement in chronic pain (Cameron et al., 2014). Furthermore, a multicenter retrospective survey of patients with chronic central neuropathic pain or fibromyalgia who were prescribed oral Δ9-THC (dronabinol), supplemental to existing medication, reported improved symptoms of both anxiety and depression after 7 months of treatment as assessed by the Hospital Anxiety and Depression Scale (Weber et al., 2009). Although Sativex (1:1 ratio of Δ9-THC:cannabidiol), indicated for resistant spasticity and pain in multiple sclerosis, has not yet been directly associated with significant mood changes, patients have reported improvements in overall quality of life following 16 weeks of treatment (Vermersch, 2011). In a separate randomized control clinical trial evaluating the effect of Sativex in patients with chronic painful diabetic neuropathy, patients with comorbid depression displayed significant improvements in total pain score in comparison with nondepressed counterparts (Selvarajah et al., 2010). Collectively, the above studies suggest that when coexistent, both depression/anxiety and pain respond to exogenously administered cannabinoids, although it remains to be determined if the effects are mediated by common or parallel mechanisms. Recent evidence has demonstrated enhanced amygdala activity and reduced functional connectivity between the amygdala and somatosensory cortex correlate with Δ9-THC–mediated reductions in the unpleasantness to ongoing pain (Lee et al., 2013), suggesting that the amygdala may provide a common neural circuit for the association between emotional responding and pain.

Table 1.

Clinical Studies Demonstrating Effects of Cannabinoid-Based Therapies on Symptoms of Comorbid Depression and Pain

Drug Pain Measurement Depression/
Anxiety Measurement
Outcomes in Pain Outcomes in Depression/
Anxiety
Reference
HIV Cannabis Pilot
questionnaire
Pilot Questionnaire muscle,
↓ nerve pain, headaches
↓ anxiety, depression Woolridge et al. (2005)
Cancer pain Nabilone (Cesamet®) ESAS MSE ESAS ↓ pain score,
MSE
↓ anxiety, overall stress Maida et al. (2008)
Fibromyalgia Nabilone VAS FIQ Anxiety ↓ pain ↓ anxiety Skrabek et al. (2008)
Mentally ill offenders Nabilone Self-reported
pain severity
PCL-C
GAF
↓ pain ↓ PTSD
symptoms
Cameron et al. (2014)
Multiple sclerosis- related resistant spasticity Sativex (Δ9-THC, cannabidiol) NRS spasticity score QOL ↓ spasticity ↑ QOL Vermersch (2011)
Chronic central neuropathic pain, fibromyalgia Δ9-THC VRS, NRS, PDI SF-12, QLIP, HADS, ↓ pain, pain intensity ↑ QOL, depression, ↓ anxiety Weber et al. (2009)
Painful diabetic peripheral neuropathy Sativex
9-THC, cannabidiol)
VAS HADS,
QOL
↓ pain (only in patients with baseline depression) ↑ QOL Selvarajah et al. (2010)

Abbreviations: ESAS, Edmonton symptom assessment system; FIQ, fibromyalgia impact questionnaire; GAF, Global Assessment of Functioning; HADS, hospital anxiety and depression scale; HIV, human immunodeficiency virus; MSE, morphine sulphate equivalent; NRS, numerical rating scale; PDI, pain disability index; PCL-C, Posttraumatic Checklist–Civilian version; PTSD, posttraumatic stress disorder; QLIP, quality of life; QOL, quality of life; SF-12, short form-12;VAS, visual analog scale; VRS, verbal rating scale.

Preclinical Evidence

Despite numerous reports of altered endocannabinoid signaling in various animal models of pain (Lim et al., 2003; Zhang et al., 2003; Walczak et al., 2005; Mitrirattanakul et al., 2006) and mood-related behavior (Vinod et al., 2012; Marco et al., 2014; Navarria et al., 2014), there is limited direct evidence available identifying alterations in endocannabinoid function in animal models of coexistent depressive and pain behavior (Tables 2 and 3). One of the first studies examining the role of the endocannabinoid system in the interaction between affect and pain was conducted by Takahashi and colleagues (2003). In this study, outbred Swiss-albino mice were stratified into groups of anxious and nonanxious animals as determined by behavioral responses in the elevated plus maze, before subsequent exposure to intraplantar formalin administration (Takahashi et al., 2003). Despite the hypothesis that the degree of anxiety may contribute to the perception of and response to the noxious stimulus, both anxious and nonanxious animals displayed comparable formalin-evoked biphasic nociceptive profiles. Systemic pretreatment with Δ9-THC elicited an analgesic effect in both groups of animals, an effect blocked by systemic pretreatment with a CB1 receptor antagonist, rimonabant (Takahashi et al., 2003). These data suggested that the endocannabinoid system (and in particular the CB1 receptor) may represent a potential treatment strategy for inflammatory pain in the presence and/or absence of anxiety and possibly other neuropsychiatric disorders. However, it should be noted that direct activation of central CB1 receptors is responsible for the psychoactive effects of potent synthetic or plant-derived cannabinoids; thus potent, direct agonism of this receptor is unlikely to be therapeutically viable for pain and/or psychiatric disorders. In comparison, modulation of endocannabinoid tone by inhibiting enzymes responsible for their metabolism has been proposed to confer improved efficacy and safety relative to direct cannabinoid agonists.

Table 2.

Endocannabinoid-Mediated Effects/Changes on Affective and Nociceptive Behavior in Animal Models

Depression/
Affective Model
Nociceptive Effects Cannabinoid-based drugs Endocannabinoid-related changes/effects Reference
Anxiety-stratified (EPM), mouse ↑ formalin-evoked
nociceptive responding in anxious and non- anxious
Δ9-THC Rimonabant CB1/2 agonist CB1 antagonist Δ9-THC ↓ nociception in both anxious and non-anxious mice,
Rimonabant blocked effects of Δ9-THC
Takahashdi et al. (2003)
WKY rat ↑ formalin-evoked
nociceptive responding
URB597
AM251
FAAH inhibitor
CB1 antagonist
Formalin-induced ↓ AEA in RVM,
No formalin-induced ↑ 2-AG,
NAPE-PLD or DAGL-α in RVM (compared with SD)
Systemic URB597 ↓ nociception Systemic AM251 ↑ nociception
AM251 within RVM blocked effect of URB597
Rea et al. (2014)
Repeated FST in SD and WKY rat Stress ↓ formalin- evoked nociceptive responding in SD
Stress ↓ formalin- evoked nociceptive responding in WKY
↑ MAGL mRNA in spinal cord of SD
↓ AEA in amygdala of SD
No change in MAGL mRNA in spinal cord of WKY
No change AEA in amygdala of WKY
Jennings et al. (2015)
CUS, mouse ↓ latency to respond
in HPT
URB597 JZL184 FAAH inhibitor, MAGL inhibitor URB597 ↓ anxiety (EPM, LD)
JZL184 ↓ anxiety (LD)
Both ↓ thermal hyperalgesia
Lomazzo et al. (2015)
CUS, mouse Chronic mechanical hyperalgesia following NGF URB597 JZL184 FAAH inhibitor, MAGL inhibitor URB597 ↓ hyperalgesia
No change with JZL184
Lomazzo et al. (2015)

Abbreviations: AEA, anandamide; 2-AG, 2-arachidonoylglycerol; CUS, chronic unpredicted stress; EPM, elevated plus maze; FAAH, fatty acid amino hydrolase; DAGL-α, diacylglycerol lipase-alpha; HPT, hot plate test; LD, light-dark box; MAGL, monoacylglycerol lipase; NAPE-PLD, N-acyl phosphatidylethanolamine-specific phospholipase D; NGF, nerve growth factor; RVM, rostral ventromedial medulla; SD, Sprague Dawley; WKY, Wistar-Kyoto.

Table 3.

Endocannabinoid-Mediated Effects/Changes on Affective and Nociceptive Behavior in Animal Models of Pain

Pain Model Depressive Effects Cannabinoid-Based Drugs Endocannabinoid-Related Changes/Effects Reference
PNL, mouse ↑ Anxiety in LD and
Zero Maze
↓ Sucrose Preference
in CB1 -/- mice only
Anxiety and depressive effects
only in CB1 -/- mice
Racz et al. (2015)
Monosodium iodoacetate, mouse ↑ Anxiety in EPM
Memory impairment in object recognition memory task
ACEA
JWH133
CB1 agonist
CB2 agonist
↑ anxiety in CB1 -/- mice
no anxiety in CB2 -/- mice
ACEA and JWH133 ↓ mechanical allodynia and anxiety
ACEA ↓ memory impairment
La Porta et al. (2015)
CCI, rat ↑ Immobility in FST GW405833 CB2 agonist GW405833 ↓ mechanical hyperalgesia
GW405833 ↓ immobility
Hu et al. (2009)
Acid-stimulated stretching, rat ↓ Food intake
↓ ICSS
Δ9-THC,
CP55940
CB1/2 agonist
CB1/2 agonist
Both blocked stretching
Both exacerbated ↓ ICSS
No effect on feeding
Kwilacz et al. (2012)
Acid-stimulated stretching, rat ↓ ICSS URB597
Rimonabant
SR144528
FAAH inhibitor
CB1 antagonist
CB2 antagonist
URB597 ↓ stretching; blocked by rimonabant,
URB597 induced delayed partial attenuation of ICSS - not attenuated by rimonabant or
SR144528
Kwilacz et al. (2014)

Abbreviations: CCI, chronic constrictive injury; CFA, complete Freud’s adjuvant; EPM, elevated plus maze; FAAH, fatty acid amino hydrolase; FST, forced swim test; ICSS, intracranial self-stimulation; LD, light-dark box; MBT, marble burying test; PNL, partial sciatic nerve ligation.

The WKY rat is a genetically stress-sensitive strain of rat that exhibits a depression- and anxiety-related phenotype (Pare and Redei, 1993) and heightened nociceptive behavioral responding in several paradigms (Zeng et al., 2008; Burke et al., 2010). Characterization of the endocannabinoid system in WKY rats has revealed higher levels of FAAH and CB1 receptor coupling and lower levels of AEA in the frontal cortex and hippocampus when compared with Wistar rats (Vinod et al., 2012). Furthermore, enhancing AEA tone by pharmacologically inhibiting FAAH activity resulted in an attenuation of depressive-like behavior (sucrose preference test and FST) in WKY rats (Vinod et al., 2012). Recent studies in our laboratory have identified alterations in the endocannabinoid system concurrent with enhanced formalin-evoked nociceptive behavior in the WKY rat (Rea et al., 2014) (Table 2). More specifically, we found that in WKY rats, intraplantar administration of the noxious inflammatory pain stimulus formalin resulted in a significant reduction in AEA in the rostral ventromedial medulla (RVM), a component of the descending pain pathway synonymous with pain facilitation and/or inhibition, an effect that was not observed in Sprague Dawley (SD) counterparts. Intraplantar administration of formalin increased levels of 2-AG in the RVM of SD rats, an effect not observed in WKY animals. Furthermore, exposure to formalin induced significant increases in mRNA expression of NAPE-PLD and diacylglycerol lipase-α, precursors of AEA and 2-AG, respectively, in the RVM of SD rats, an effect not observed in the WKY strain. Pharmacological studies were carried out to evaluate the functional significance of the alterations in the endocannabinoid system in WKY rats in response to formalin. Enhancing endogenous AEA tone following systemic administration of the FAAH inhibitor URB597 attenuated formalin-evoked hyperalgesic responding in WKY rats, while in comparison, CB1 receptor antagonism was associated with augmentation of nociceptive responding. Furthermore, CB1 receptor blockade within the RVM attenuated the reduction in nociceptive behavior induced by URB597 in WKY rats. In comparison, pharmacological manipulation of the endocannabinoid system in SD rats did not alter formalin-evoked nociceptive responding (Rea et al., 2014). These findings indicate a causative role of endocannabinoid dysregulation in hyperalgesic behavior associated with negative affect, and moreover identify a role for AEA-induced activation of CB1 receptors in the RVM as a mediator of pain suppression in animal subjects predisposed to anxiety and depression. In addition to its role in influencing responsivity of WKY rats to noxious stimuli in the absence of stress, we have also shown very recently that the endocannabinoid system may also play a role in the differential effects of repeated homotypic stress on inflammatory pain-related behavior in WKY vs SD rats (Jennings et al., 2015). Specifically, repeated forced swim stress exposure prolonged and attenuated formalin-evoked nociceptive behavior in SD and WKY rats, respectively. These behavioral alterations were accompanied by differential effects of stress on AEA levels in the amygdala and MAGL expression in the spinal cord between SD and WKY rats (Jennings et al., 2015). These data indicate that changes in the tone of the endocannabinoid system in the amygdala and spinal cord may underlie the differential effects of stress on inflammatory pain behavior between SD and WKY rats.

The chronic unpredictable stress (CUS) model of depression has been shown to exhibit thermal hyperalgesia in the hotplate test (Lomazzo et al., 2015), cold allodynia (Bravo et al., 2012), exacerbated trigeminovascular nociception (Zhang et al., 2013), inflammatory hyperalgesia in response to formalin administration (Shi et al., 2010), and persistent mechanical hyperalgesia following nerve growth factor administration (Lomazzo et al., 2015). Exposure to CUS has been shown to result in site-specific alterations in the endocannabinoid system, notably a downregulation of CB1 receptors, reduction in 2-AG levels and increased FAAH levels in the hippocampus (Hill et al., 2005; Reich et al., 2009), an increase in CB1 receptor mRNA expression in prefrontal cortex and decrease in expression in the midbrain (Bortolato et al., 2007), a decrease in CB1 receptor density in the hypothalamus and striatum and increased CB1 receptor density in the prefrontal cortex (Hill et al., 2008a; McLaughlin et al., 2013), a reduction in 2-AG–mediated retrograde synaptic transmission in the hippocampus (Zhong et al., 2014), and a reduction in AEA levels in the hypothalamus, prefrontal cortex, hippocampus, and striatum (Hill et al., 2008a). Depressive-like behaviors in the CUS model have been shown to be attenuated by endocannabinoid-modulating pharmacological agents, including the MAGL inhibitor JZL184 (Zhong et al., 2014; Zhang et al., 2015). Only one study to date has examined the effect of endocannabinoid modulation on affective and pain responding in CUS-exposed mice. Pretreatment with the FAAH inhibitor, URB597, or MAGL inhibitor, JZL184, which enhanced endogenous levels of AEA and 2-AG, respectively, significantly attenuated CUS-induced anxiety-related behavior in the light-dark box and concurrent thermal hyperalgesia (Lomazzo et al., 2015). Long-lasting widespread mechanical hyperalgesia, induced by intramuscular administration of nerve growth factor to CUS rats, was effectively reduced by URB597, but not JZL184 (Lomazzo et al., 2015). These data demonstrate an important role for AEA signaling in anxiety- and pain-related behavior in stress-exposed mice.

In addition to the evidence supporting a role for the endocannabinoid system in enhanced nociception observed in models of depression, alterations in endocannabinoid signaling have also been observed in animal models of chronic pain with comorbid alteration in affective responding (Table 3). A recent report by Racz and colleagues (2015) has revealed a prominent role of CB1-mediated events in affective behavior induced by neuropathic pain. In this study, partial sciatic nerve ligation (PNL) was employed to induce a model of neuropathic pain in wild-type and CB1 -/- mice. Wild-type and CB1 -/- mice exhibited mechanical allodynia following PNL. However, evaluation of anxiety- (light-dark test and the elevated zero-maze) and depressive-like (sucrose preference test) behavior 4 to 7 weeks following PNL revealed deficits in affective responding in CB1 -/-, but not wild-type, mice (Racz et al., 2015). Thus, these data demonstrate that functionally active CB1 receptors confer resilience to pain-related anxiety/depression, highlighting a protective role for CB1 receptors against the emotional consequences of neuropathic pain. In a similar fashion, La Porta et al. (2015) recently investigated the role of the endocannabinoid system in affective and cognitive manifestations in an animal model of osteoarthritis. This study revealed that the anxiety-related behavior of osteoarthritic mice, identified in the elevated plus maze, was enhanced in CB1 -/- and absent in CB2 -/- mice, indicating differential effects of CB1 and CB2 receptors in mediating the affective dimension of pain in the model. Similar to effects in a neuropathic model (Racz et al., 2015), the data would indicate that CB1 receptors confer resilience, while CB2 receptors confer susceptibility to the development of arthritis-related anxiety. The authors suggest and provide some support that the differential effects of CB1 and CB2 receptors may be mediated by alterations in HPA axis functionality and responses (La Porta et al., 2015). In addition, this study also demonstrated that acute pharmacological blockade of CB1 or CB2 receptors ameliorated both the nociceptive and affective dimension of pain in the model (La Porta et al., 2015). Taken together, the data suggest that cortico-limbic endocannabinoid signaling is a key modulator of different osteoarthritis pain manifestations.

Only one study to date has investigated the role of CB2 receptors in the interaction between neuropathic pain and affective behavior. The chronic constriction injury model of neuropathic pain results in mechanical hypersensitivity and depressive-like behavior (immobility in the FST) in mice (Hu et al., 2009). Both depressive-like behavior and mechanical hyperalgesia following constriction injury were significantly attenuated by systemic administration of the CB2 receptor agonist GW405833, effects that were not observed in sham-operated animals. Furthermore, such behavioral effects were superior to administration of a tricyclic antidepressant, first line treatment for depression and chronic pain (Hu et al., 2009). The precise mechanism by which CB2 receptor agonism may elicit analgesic and antidepressant-like effects was not evaluated; however, given the well-recognized role for inflammatory processes in mediating chronic pain, it is possible that CB2 receptor activation attenuates such responses, preventing the development of central sensitization and mechanical allodynia and the associated increase in neuronal input to affective supraspinal sites. Further studies are required to evaluate this theory.

Intraperitoneal administration of a dilute concentration of lactic or acetic acid has been shown to induce abdominal stretching/writhing (visceral pain behavior), an effect associated with a reduction in feeding and hedonic behaviors (pain-depressed behavior). Evaluation of the role of the endocannabinoid system in mediating pain-stimulated and pain-depressed/suppressed responses in this model has revealed that genetic antagonism of CB1 receptors enhances acid-induced writhing (visceral pain stimulated behavior) and augments acid-induced reductions in feeding (Miller et al., 2011). In contrast, CB1 receptor agonism using Δ9-THC and CP55940 dose dependently inhibits acid-stimulated stretching while eliciting either no effect (Miller et al., 2012) or exacerbating (Kwilasz and Negus, 2012) acid-induced depression of feeding and scheduled controlled/intracranial self-stimulation in rats. Thus, under these conditions, potent synthetic cannabinoids such as Δ9-THC and CP55940 may elicit differential effects on visceral pain (attenuated) and pain-related depressive (exacerbated) behavior. However, the FAAH inhibitor URB597 exhibits a dose-related and CB1 receptor-mediated decrease in acid-stimulated stretching and suppression of feeding (Miller et al., 2012; Kwilasz et al., 2014). Furthermore, URB597 also elicits a delayed but significant attenuation of acid-induced suppression of intracranial self stimulation, an effect occurring independent of CB1 or CB2 mediation (Kwilasz et al., 2014). Taken together, these data indicate a role for CB1 receptors in mediating acid-induced visceral pain, with a possible common and/or alternative endocannabinoid mechanism mediating the associated anhedonic/depressive-like behavior.

Overall, despite the limited data, evidence suggests a prominent role for the endocannabinoid system in the interaction between depression and pain, although whether the effects are mediated by the same or parallel neuroanatomical pathways remains to be determined.

Mechanisms By Which The Endocannabinoid System May Modulate Depression And Pain Interactions

While the exact mechanism(s) by which the endocannabinoid system may influence emotional and nociceptive processing remains undetermined, this system is known to elicit potent modulatory effects on neurotransmission, neuroendocrine, and inflammatory processes, all known to be altered in both depression and chronic pain. Presented here is an overview of how the endocannabinoid system may modulate affective and pain processing via interacting with these systems.

Neurotransmitters

GABA and Glutamate

GABA- and glutamatergic neurotransmission are well recognized as important mediators in affect and nociceptive processing, and alterations in these systems have been demonstrated in both depression and chronic pain (for review, see Kendell et al., 2005; Rea et al., 2007). There are an increasing number of studies demonstrating that glutamatergic and GABAergic signaling play an important role in mediating the depressive symptoms associated with chronic pain. For example, ketamine, an NMDA receptor antagonist, attenuated depressive-like behavior following spared nerve injury without altering injury-induced hypersensitivity (Wang et al., 2011), while AMPAkines (which augment AMPA receptor function) have been shown to attenuate both pain hypersensitivity and associated depressive-like behavior in models of chronic inflammatory and neuropathic pain (Le et al., 2014). Furthermore, facilitation of glutamatergic transmission through AMPA receptors in the nucleus accumbens, a brain region involved in reward, resulted in attenuation of depression-like behavior in an animal model of neuropathic pain (Goffer et al., 2013). Despite a recognized role for the GABAergic system in emotional and pain processes, to our knowledge few studies have investigated the role of this system in affect-pain interactions to date. One such study has reported that GABAA receptor activation in the RVM blocked formalin-induced hyperalgesia produced upon removal from an aversive elevated plus maze (stressor) (Cornelio et al., 2012). In addition, Quintero and colleagues (2011) have also identified that repeated forced swim stress-induced inflammatory hyperalgesia is initiated by decreased and delayed GABA release and GABAA receptor activation and maintained by increased glutamate release and NMDA activation at the spinal cord level (Suarez-Roca et al., 2008).

CB1 receptors are expressed at a particularly high density on presynaptic nerve terminals of GABAergic and glutamatergic synapses in cortical and limbic areas of the brain associated with stress, emotional response, and pain modulation (Herkenham et al., 1990; Katona et al., 2001; Domenici et al., 2006; Wittmann et al., 2007). Endocannabinoids have been shown to exert behavioral effects via CB1 receptor agonism and resultant presynaptic inhibition of GABAergic and glutamatergic transmission (Meng et al., 1998; Millan, 2002) in supraspinal and spinal regions (Ulugol, 2014). In addition, glutamatergic neurotransmission is known to enhance endocannabinoid formation and subsequent CB1 receptor activation (Galante and Diana, 2004). Thus, complex bidirectional interaction exists between the endocannabinoid-glutamatergic-GABAergic systems. Several studies have demonstrated that CB1 modulation of GABAergic signaling is important in nociceptive (Manning et al., 2003; Naderi et al., 2005; Pernia-Andrade et al., 2009) and emotional (Haller et al., 2007; Naderi et al., 2008; Rossi et al., 2010; Rey et al., 2012; Reich et al., 2013) processing. Although it would appear intuitive, it is unknown if endocannabinoid modulation of GABA and/or glutamate plays a role in depression-pain interactions. However, we have demonstrated that CB1 receptors play an important role in mediating analgesia in response to acute stress (contextual fear conditioning) (Finn et al., 2004; Butler et al., 2008) and demonstrated an important role for GABAergic and glutamatergic signaling in the basolateral amygdala in mediating this effect (Rea et al., 2014). Thus, endocannabinoid modulation of GABAergic and glutamatergic tone can mediate stress-pain interactions and therefore may play a prominent role in coexistent psychiatric and pain disorders.

Monoamines

In addition to the treatment of depression, monoamine-based antidepressants are now regarded as first-line therapy for fibromyalgia and neuropathic pain. The monoaminergic system has been proposed as a common neural substrate for depression-pain associations. In accordance, a number of preclinical studies have demonstrated beneficial effects of modulating monoaminergic tone on depression associated with chronic pain and vice versa. For example, recent studies have demonstrated that chronic administration of 3-(4-fluorophenylselenyl)-2,5-diphenylselenophene, which increases serotonergic neurotransmission by inhibiting presynaptic serotonin transport, attenuates mechanical allodynia and depressive-like behavior in an animal model of neuropathic pain (Gai et al., 2014). Furthermore, chronic imipramine treatment reduces depressive-like behavior, but not hyperalgesia, in a rat model of neuropathic pain, effects mediated by increasing the neurotrophin BDNF (Yasuda et al., 2014). In addition, direct administration of a BDNF inducer (4-MC) into the brain attenuated thermal hyperalgesia and depressive-like behavior following nerve injury (Fukuhara et al., 2012; Ishikawa et al., 2014). The spinal serotonergic system has been shown to participate in the thermal hyperalgesia response in an animal model of depression, the olfactory bulbectomized rat (Rodriguez-Gaztelumendi et al., 2014). We have shown recently chronic amitriptyline treatment elicits an antidepressant-like effect and potently attenuates nerve injury-induced mechanical and cold allodynia in the bulbectomy model of depression (Burke et al., 2015). Thus, enhancing monoaminergic tone, and consequently central BDNF expression, modulates pain-depression behaviors.

Several lines of evidence have demonstrated that chronic antidepressant administration modulates endocannabinoid signaling (Hill et al., 2006, 2008b; Mato et al., 2010; Smaga et al., 2014), which underlie at least in part the mechanism by which these pharmacological agents modulate affective and nociceptive processes. Furthermore, endocannabinoid-induced modulation of serotonergic, noradrenergic, and dopaminergic transmission has been thoroughly investigated in several excellent reviews (Haj-Dahmane and Shen, 2011; Melis and Pistis, 2012; Kirilly et al., 2013). CB1 receptors are highly expressed on serotonergic, noradrenergic, and dopaminergic neurons and play an important role in the regulation of monoaminergic activity. Local and systemic administration of exogenous CB1 receptor agonists significantly increases serotonin (Bambico et al., 2007), noradrenaline (Jentsch et al., 1997; Oropeza et al., 2005; Page et al., 2007, 2008), and dopamine (Cheer et al., 2004; Solinas et al., 2006) levels in discrete brain regions that mediate emotional and nociceptive processing. Increasing endogenous levels of AEA and 2-AG through systemic administration of FAAH or MAGL inhibitors, respectively, has also been shown to enhance serotonergic and dopaminergic activity (Gobbi et al., 2005; Seif et al., 2011), and FAAH inhibition in the PFC increases serotonergic neuronal firing in the dorsal raphe nucleus (McLaughlin et al., 2012). In addition, endocannabinoids can inhibit the activity of monoamine oxidase (Fisar, 2010), the enzyme responsible for the metabolism of monoamines, which may also contribute to the increasing synaptic availability of the monoamines. Endocannabinoid activation of the CB1 receptor has been shown to control the function and expression of specific serotonin receptors, namely 5-HT1A, 5-HT2A, and 5-HT2C, in discrete regions of the CNS (Aso et al., 2009; Moranta et al., 2009; Zavitsanou et al., 2010; Franklin et al., 2013). Furthermore, spinal noradrenergic depletion is associated with compromised analgesic effects of CB1 agonism on formalin-evoked inflammatory pain (Gutierrez et al., 2003), while CB1 receptor activation results in attenuation of enhanced serotonergic firing in the dorsal raphe following nerve injury, an effects accompanied by antinociception (Palazzo et al., 2006). Taken together, these data demonstrate that cannabinoids modulate nociceptive tone, in part through modulation of noradrenergic and serotonergic systems. Thus, endocannabinoid-induced enhancement of monoaminergic tone may modulate emotional and nociceptive processes and thus the interaction between depression and pain.

Opioids

Numerous studies have characterized the causative role and therapeutic potential of opioidergic signaling in affective and nociceptive processing (for review, see Maletic and Raison, 2009; Lutz and Kieffer, 2013). Furthermore, alterations in opioid signaling in key brain regions such as the amygdala have been reported in a chronic pain model that exhibits comorbid anxiety-related behavior (Narita et al., 2006). Acute morphine administration attenuated both mechanical allodynia and anxiety-related behavior in the complete Freund’s adjuvant model of inflammatory pain (Parent et al., 2012). A very substantial body of evidence is now available to suggest that the endocannabinoid and opioidergic systems interact in therapeutically beneficial ways. CB1 and µ-opioid receptors are highly colocalized on neurons in areas of the brain associated with emotional and pain processing such as the caudate putamen, periaqueductal gray, and spinal cord (Rodriguez et al., 2001; Salio et al., 2001; Wilson-Poe et al., 2012). Coadministration of opioids and cannabinoids results in synergistic and bidirectional antinociceptive effects in several animal models (Cichewicz et al., 1999; Cichewicz and McCarthy, 2003; Tham et al., 2005; Roberts et al., 2006; Smith et al., 2007; Wilson et al., 2008; Wilson-Poe et al., 2013). In addition, increasing endocannabinoid tone has been shown to attenuate withdrawal symptoms in morphine-dependent animals (Smith et al., 2007; Wilson et al., 2008; Shahidi and Hasanein, 2011). Interestingly, cross tolerance also exists between these neuromodulatory systems. For example, decreases in the analgesic effects of Δ9-THC have been identified in morphine-tolerant animals and vice versa (Thorat and Bhargava, 1994). Furthermore, inhibition of opioid signalling (via ĸ-opioid receptors) attenuates the antidepressant-like effect of rimonabant (CB1 receptor antagonist/inverse agonist) in the FST (Lockie et al., 2011), and conversely the antidepressant-like effects of ĸ-opioid receptor antagonism are attenuated by the CB1 receptor antagonist/inverse agonist AM251 (Braida et al., 2009). Although further studies are required, collectively these findings suggest a regulatory role of the endocannabinoid system on opioid transmission, which may underlie the maintenance of coexistent depression-pain processes.

Neuroendocrine Activity – HPA Axis

Dysregulation of the HPA axis has been implicated in the pathophysiology of both depression and pain disorders for decades (for review, see Bomholt et al., 2004; Vierck, 2006; Maric and Adzic, 2013; Belvederi Murri et al., 2014) and thus has also been proposed as a possible mediator in the depression-pain dyad (for review, see Blackburn-Munro, 2004). Several clinical studies have identified altered HPA axis activity in patients exhibiting symptoms of both depression and pain. For example, a cross-sectional study of patients with advanced breast cancer revealed increasing plasma cortisol levels that positively correlated with symptoms of depression and pain (Thornton et al., 2010). However in patients with fibromyalgia, enhanced cortisol release and dysregulation of HPA function associates with depressive, but not pain, symptoms, implying possible diverging mechanisms for both affective and nociceptive processing in this condition (Wingenfeld et al., 2010). Preclinical evidence of this reciprocity is also evident in an experimental model of gastritis, which is associated with gastrointestinal inflammation, pain, and anxiety- and depressive-like behaviors in rats (Luo et al., 2013). These behavioral alterations are accompanied by dysregulation of the HPA axis, characterized by increased expression of corticotrophin-releasing factor (CRF) mRNA and reduced expression of glucocorticoid receptor in the hypothalamus and increased plasma levels of corticosterone (Luo et al., 2013). Increased expression of CRF has also been reported in the paraventricular nucleus of the hypothalamus and dorsal raphe nucleus of WKY rats and animals preexposed to neonatal maternal separation, respectively (Bravo et al., 2011), two models of depression and associated hyperalgesia. Pharmacological blockade of CRF1 following intraamygdalar infusion of the CRF1 antagonist CP376395 inhibits hyperalgesic responding to colorectal distention in WKY rats, an effect not observed following glucocorticoid receptor or mineralocorticoid receptor antagonism (Johnson et al., 2012). Furthermore, the hyperalgesic visceromotor response to phasic colorectal distension following repeated water avoidance stress has been shown to be attenuated by CRF1 antagonism (Larauche et al., 2008). In addition, systemic or intraamygdalar injection of the CRF1 receptor antagonist NBI27914 blocks anxiety-related and nocifensive behavior in a rat model of arthritis (Ji et al., 2007). Thus, the CRF-HPA stress axis has been shown to play a key role in affective and/or nociceptive processing.

Several lines of evidence now support an important role for the endocannabinoid system as a modulator of HPA axis function and vice versa (for review, see Finn, 2010; Riebe and Wotjak, 2011; Hill and Tasker, 2012). The majority of evidence collated to date would suggest that basal HPA activity is under tonic inhibitory control by CB1 receptors. This has been shown in numerous reports where genetic deletion or pharmacological blockade of the CB1 receptor in vivo enhances expression of CRF and reduces glucocorticoid receptor expression in the hypothalamus and pituitary gland, respectively (Cota et al., 2007), and increases circulating levels of corticosterone and adrenocorticotropic hormone (Barna et al., 2004; Cota et al., 2007; Steiner et al., 2008). In addition, stress-induced increases in CRF expression in the paraventricular nucleus of the hypothalamus and the basolateral amygdala, as well as corticosterone secretion, are effectively blocked by pharmacological enhancement of endocannabinoid levels (Patel et al., 2004; Hill et al., 2009a; Bedse et al., 2014; Roberts et al., 2014), thus implying a role for endocannabinoid-CB1 receptor signaling in diminished hyperactivity of the HPA axis. Furthermore, recent evidence has shown that CRF1 activation in the amygdala induces FAAH and reduces AEA levels, an effect associated with anxiety-related behavior (Gray et al., 2015). Given the important role of the amygdala in affective modulation of pain, it is possible that CRF-mediated FAAH activation in this region may also modulate nociceptive possessing and associated emotional alterations. While there have been a few studies examining endocannabinoid-HPA axis effects in mediating the effects of the stress response (Hill et al., 2011; Roberts et al., 2014), to date no study has investigated if cannabinoid-mediated alterations of the HPA axis underlie alterations in nociceptive and/or affective behavior observed in depression-pain comorbidity.

Neuro-Inflammatory Processes

Increasing evidence indicates a potent and prominent interaction between inflammation, depression, and pain (for review, see Walker et al., 2014). For instance, there is a high prevalence of depression among patients with inflammatory pain disorders such as fibromyalgia, arthritis, and irritable bowel disorder (Kappelman et al., 2014; Scheidt et al., 2014; Lin et al., 2015). In addition, patients receiving cytokine therapy for specific cancers and malignancies also develop depressive and/or painful symptomatology (Capuron and Ravaud, 1999; Capuron et al., 2001; Nogueira et al., 2012). Furthermore, increases in serum and cerebrospinal fluid levels of proinflammatory cytokines have been widely reported in both depression (Tuglu et al., 2003; Knuth et al., 2014; Bay-Richter et al., 2015) and pain conditions (Koch et al., 2007; Ludwig et al., 2008; Kadetoff et al., 2012). Inflammatory processes have also been shown to underlie the interaction between depression and pain in several animal models. For instance, increased expression of proinflammatory cytokines, concomitant with depressive-like behavior, has been identified in animal models of inflammatory (Kim et al., 2012; Maciel et al., 2013) and neuropathic (Norman et al., 2010; Burke et al., 2014; Dellarole et al., 2014; Zhou et al., 2015) pain. The innate inflammatory cascade has been shown to increase glutamate neurotransmission, central sensitization, and excitotoxicity, reduce BDNF and neurogenesis, and activate neurodegenerative cascades, events observed in both depression and pain conditions (for review, see Dantzer et al., 2011; Maes et al., 2011; Song and Wang, 2011; Zunszain et al., 2013; Walker et al., 2014).

Over the past decade, a wealth of data has demonstrated an important role for the endocannabinoid system in modulating innate immune function and inflammatory processes (for review, see Alhouayek and Muccioli, 2012; Zajkowska et al., 2014; Henry et al., 2015). Interactions between the endocannabinoid system and inflammatory mediators has been shown to influence synaptic transmission and neuronal function (Rossi et al., 2014). Spinal cord injury has been shown to be associated with increased coexpression of CB1 receptors with chemokines CCL2, CCL3, and/or CCR2 in the hippocampus, thalamus, and periaqueductal grey, areas associated with affective pain responding (Knerlich-Lukoschus et al., 2011), and studies have also demonstrated that CB1-chemokine interactions in the periaqueductal grey can modulate nociceptive responding (Benamar et al., 2008). Pharmacological enhancement of endocannabinoid tone also modulates inflammatory effects in vivo. For example, the FAAH inhibitor URB597 and the MAGL inhibitor JZL184 attenuate inflammation-induced astrocyte and microglial activation (Katz et al., 2015) and neuroinflammatory processes (Kerr et al., 2012, 2013; Henry et al., 2014). Furthermore, Zoppi and colleagues (2011, 2014) have demonstrated that pharmacological activation of CB1 or CB2 receptors attenuates, while genetic deletion of these receptors augments, repeated stress-induced proinflammatory responses in the frontal cortex. In addition, several studies have demonstrated that the analgesic effects of cannabinoids in chronic inflammatory and neuropathic pain are at least partially mediated by modulation of inflammatory responses (Burgos et al., 2012; Wilkerson et al., 2012; Burston et al., 2013; Lu et al., 2015). While there are no studies to date investigating if cannabinoid modulation of inflammatory processes underlies coexistent depressive and pain behavior, the above findings suggest a potential role for cannabinoid-mediated immunomodulation in the pathogenesis and treatment of co-occurring depression and pain.

Conclusion and Future Directions

This review has provided an overview of the clinical and preclinical evidence supporting an association between depression and pain and vice versa. While a number of neural substrates have been proposed to underlie this association, this review provides a synthesis of the data supporting the contention that comorbid depression and pain may be mediated at least in part via dysregulation of the endocannabinoid system. Targeting the endocannabinoid system for therapeutic benefit has been an ever-expanding area of research with more than 150 clinical trials during the past decade evaluating the effects of cannabinoids in pain and psychiatric disorders (International Clinical Trials Registry Platform). While no study to date has specifically evaluated the effects of cannabinoids on depression-pain comorbidity, several have examined effects on mood and quality of life in patients receiving cannabinoid-based pharmaceuticals for analgesic purposes (Maida et al., 2008; Skrabek et al., 2008; Weber et al., 2009; Cameron et al., 2014), providing a basis for further study in this area. However, many of these cannabinoids are potent CB1 receptor agonists (Δ9-THC derivatives), an effect that may limit their usefulness in psychiatric conditions due to the associated adverse CNS effects. Cannabidiol has been shown to limit the adverse CNS effects associated with CB1 receptor agonists, and Sativex (1:1 Δ9-THC:cannabidiol) has been shown to reduce chronic pain and improve mood (Selvarajah et al., 2010; Vermersch, 2011). Thus, combination therapy may be a beneficial treatment strategy for depression-pain comorbidity. As highlighted throughout this review, modulation of endogenous cannabinoid tone provides an alternative to direct CB1 receptor agonism and although still in the early stages of clinical investigation, FAAH inhibitors such as PF-04457845 have demonstrated safety and tolerability in patients, although no effect on pain associated with osteoarthritis was reported (Huggins et al., 2012). However, this inhibitor is currently under clinical investigation for treatment of cannabis withdrawal, PTSD, and Tourette syndrome (International Clinical Trials Registry Platform), the results from which will provide important clinical data on the effect of FAAH inhibition on affective responding. Further clinical studies will provide greater insight into alterations and the role of the endocannabinoid system in the association between depression and pain.

Preclinical models that encapsulate the clinical scenario are particularly useful in gaining greater understanding of the neurobiology underlying depression-pain interactions. This review has presented the evidence to date demonstrating alterations in various components of the endocannabinoid system in models of depression-pain comorbidity and highlighted a particular role for AEA and CB1 receptors in mediating and modulating the affective and nociceptive processes in these models. However, research in this area is still in its infancy, and this review highlights the gaps in the knowledge and outstanding questions that remain to be addressed. For example, there have been limited data examining the role of other components of the endocannabinoid system on depression-pain interactions (2-AG, CB2, peroxisome proliferator-activated receptors, etc.), whether the endocannabinoid modulation of affect and nociception occur through the same or parallel pathways, and the mechanism by which the endocannabinoid system may mediate its effects (neurotransmitters, HPA axis, inflammation, or a combination). Furthermore, it is unknown if the role of the endocannabinoid system in chronic pain associated with depression is the same or different from altered affective processing associated with chronic pain conditions. Such studies are essential if we are to move towards a more comprehensive understanding of the neurobiology underlying the association between these pain and depression and fully explore the potential clinical efficacy of targeting the endocannabinoid system for resolution of these comorbid conditions.

Statement of Interest

None.

Acknowledgments

This work was supported by funding received from Molecular Medicine Ireland Clinical & Translational Research Scholars Programme.

References

  1. Agrawal A, Nelson EC, Littlefield AK, Bucholz KK, Degenhardt L, Henders AK, Madden PA, Martin NG, Montgomery GW, Pergadia ML, Sher KJ, Heath AC, Lynskey MT. (2012) Cannabinoid receptor genotype moderation of the effects of childhood physical abuse on anhedonia and depression. Arch Gen Psychiatry 69:732–740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alhouayek M, Muccioli GG. (2012) The endocannabinoid system in inflammatory bowel diseases: from pathophysiology to therapeutic opportunity. Trends Mol Med 18:615–625. [DOI] [PubMed] [Google Scholar]
  3. Anderson G, Maes M, Berk M. (2012) Inflammation-related disorders in the tryptophan catabolite pathway in depression and somatization. Adv Protein Chem Struct Biol 88:27–48. [DOI] [PubMed] [Google Scholar]
  4. Arora V, Kuhad A, Tiwari V, Chopra K. (2011) Curcumin ameliorates reserpine-induced pain-depression dyad: behavioral, biochemical, neurochemical and molecular evidences. Psychoneuroendocrinology 36:1570–1581. [DOI] [PubMed] [Google Scholar]
  5. Ashton CH, Moore PB. (2011) Endocannabinoid system dysfunction in mood and related disorders. Acta Psychiatr Scand 124:250–261. [DOI] [PubMed] [Google Scholar]
  6. Aso E, Renoir T, Mengod G, Ledent C, Hamon M, Maldonado R, Lanfumey L, Valverde O. (2009) Lack of CB1 receptor activity impairs serotonergic negative feedback. J Neurochem 109:935–944. [DOI] [PubMed] [Google Scholar]
  7. Atwood BK, Straiker A, Mackie K. (2012) CB(2) cannabinoid receptors inhibit synaptic transmission when expressed in cultured autaptic neurons. Neuropharmacology 63:514–523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Baek JH, Zheng Y, Darlington CL, Smith PF. (2008) Cannabinoid CB2 receptor expression in the rat brainstem cochlear and vestibular nuclei. Acta Otolaryngol 128:961–967. [DOI] [PubMed] [Google Scholar]
  9. Bambico FR, Katz N, Debonnel G, Gobbi G. (2007) Cannabinoids elicit antidepressant-like behavior and activate serotonergic neurons through the medial prefrontal cortex. J Neurosci 27:11700–11711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bar KJ, Brehm S, Boettger MK, Boettger S, Wagner G, Sauer H. (2005) Pain perception in major depression depends on pain modality. Pain 117:97–103. [DOI] [PubMed] [Google Scholar]
  11. Bardin L, Malfetes N, Newman-Tancredi A, Depoortere R. (2009) Chronic restraint stress induces mechanical and cold allodynia, and enhances inflammatory pain in rat: relevance to human stress-associated painful pathologies. Behav Brain Res 205:360–366. [DOI] [PubMed] [Google Scholar]
  12. Barna I, Zelena D, Arszovszki AC, Ledent C. (2004) The role of endogenous cannabinoids in the hypothalamo-pituitary-adrenal axis regulation: in vivo and in vitro studies in CB1 receptor knockout mice. Life Sci 75:2959–2970. [DOI] [PubMed] [Google Scholar]
  13. Bay-Richter C, Linderholm KR, Lim CK, Samuelsson M, Traskman-Bendz L, Guillemin GJ, Erhardt S, Brundin L. (2015) A role for inflammatory metabolites as modulators of the glutamate N-methyl-D-aspartate receptor in depression and suicidality. Brain Behav Immun 43:110–117. [DOI] [PubMed] [Google Scholar]
  14. Bedse G, Colangeli R, Lavecchia AM, Romano A, Altieri F, Cifani C, Cassano T, 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 24:1511–1523. [DOI] [PubMed] [Google Scholar]
  15. Belvederi Murri M, Pariante C, Mondelli V, Masotti M, Atti AR, Mellacqua Z, Antonioli M, Ghio L, Menchetti M, Zanetidou S, Innamorati M, Amore M. (2014) HPA axis and aging in depression: systematic review and meta-analysis. Psychoneuroendocrinology 41:46–62. [DOI] [PubMed] [Google Scholar]
  16. Benamar K, Geller EB, Adler MW. (2008) First in vivo evidence for a functional interaction between chemokine and cannabinoid systems in the brain. J Pharmacol Exp Ther 325:641–645. [DOI] [PubMed] [Google Scholar]
  17. Ben-Tovim DI, Schwartz MS. (1981) Hypoalgesia in depressive illness. Br J Psychiatry 138:37–39. [DOI] [PubMed] [Google Scholar]
  18. Berdyshev EV. (2000) Cannabinoid receptors and the regulation of immune response. Chem Phys Lipids 108:169–190. [DOI] [PubMed] [Google Scholar]
  19. Blackburn-Munro G. (2004) Hypothalamo-pituitary-adrenal axis dysfunction as a contributory factor to chronic pain and depression. Curr Pain Headache Rep 8:116–124. [DOI] [PubMed] [Google Scholar]
  20. Blankman JL, Simon GM, Cravatt BF. (2007) A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chem Biol 14:1347–1356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Boettger MK, Grossmann D, Bar KJ. (2013) Thresholds and perception of cold pain, heat pain, and the thermal grill illusion in patients with major depressive disorder. Psychosom Med 75:281–287. [DOI] [PubMed] [Google Scholar]
  22. Bomholt SF, Harbuz MS, Blackburn-Munro G, Blackburn-Munro RE. (2004) Involvement and role of the hypothalamo-pituitary-adrenal (HPA) stress axis in animal models of chronic pain and inflammation. Stress 7:1–14. [DOI] [PubMed] [Google Scholar]
  23. Borges G, Neto F, Mico JA, Berrocoso E. (2014) Reversal of monoarthritis-induced affective disorders by diclofenac in rats. Anesthesiology 120:1476–1490. [DOI] [PubMed] [Google Scholar]
  24. Bortolato M, Mangieri RA, Fu J, Kim JH, Arguello O, Duranti A, Tontini A, Mor M, Tarzia G, Piomelli D. (2007) Antidepressant-like activity of the fatty acid amide hydrolase inhibitor URB597 in a rat model of chronic mild stress. Biol Psychiatry 62:1103–1110. [DOI] [PubMed] [Google Scholar]
  25. Boychuk DG, Goddard G, Mauro G, Orellana MF. (2015) The effectiveness of cannabinoids in the management of chronic nonmalignant neuropathic pain: a systematic review. J Oral Facial Pain Headache 29:7–14. [DOI] [PubMed] [Google Scholar]
  26. Braida D, Capurro V, Zani A, Rubino T, Vigano D, Parolaro D, Sala M. (2009) Potential anxiolytic- and antidepressant-like effects of salvinorin A, the main active ingredient of Salvia divinorum, in rodents. Br J Pharmacol 157:844–853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Bravo JA, Dinan TG, Cryan JF. (2011) Alterations in the central CRF system of two different rat models of comorbid depression and functional gastrointestinal disorders. Int J Neuropsychopharmacol 14:666–683. [DOI] [PubMed] [Google Scholar]
  28. Bravo L, Mico JA, Rey-Brea R, Perez-Nievas B, Leza JC, Berrocoso E. (2012) Depressive-like states heighten the aversion to painful stimuli in a rat model of comorbid chronic pain and depression. Anesthesiology 117:613–625. [DOI] [PubMed] [Google Scholar]
  29. Bravo L, Torres-Sanchez S, Alba-Delgado C, Mico JA, Berrocoso E. (2014) Pain exacerbates chronic mild stress-induced changes in noradrenergic transmission in rats. Eur Neuropsychopharmacol 24:996–1003. [DOI] [PubMed] [Google Scholar]
  30. Burgos E, Gomez-Nicola D, Pascual D, Martin MI, Nieto-Sampedro M, Goicoechea C. (2012) Cannabinoid agonist WIN 55,212-2 prevents the development of paclitaxel-induced peripheral neuropathy in rats. Possible involvement of spinal glial cells. Eur J Pharmacol 682:62–72. [DOI] [PubMed] [Google Scholar]
  31. Burke NN, Hayes E, Calpin P, Kerr DM, Moriarty O, Finn DP, Roche M. (2010) Enhanced nociceptive responding in two rat models of depression is associated with alterations in monoamine levels in discrete brain regions. Neuroscience 171:1300–1313. [DOI] [PubMed] [Google Scholar]
  32. Burke NN, Geoghegan E, Kerr DM, Moriarty O, Finn DP, Roche M. (2013) Altered neuropathic pain behavior in a rat model of depression is associated with changes in inflammatory gene expression in the amygdala. Genes Brain Behav 12:705–713. [DOI] [PubMed] [Google Scholar]
  33. Burke NN, Kerr DM, Moriarty O, Finn DP, Roche M. (2014) Minocycline modulates neuropathic pain behavior and cortical M1-M2 microglial gene expression in a rat model of depression. Brain Behav Immun 42:147–156. [DOI] [PubMed] [Google Scholar]
  34. Burke NN, Finn DP, Roche M. (2015) Chronic administration of amitriptyline differentially alters neuropathic pain-related behavior in the presence and absence of a depressive-like phenotype. Behav Brain Res 278:193–201. [DOI] [PubMed] [Google Scholar]
  35. Burston JJ, Sagar DR, Shao P, Bai M, King E, Brailsford L, Turner JM, Hathway GJ, Bennett AJ, Walsh DA, Kendall DA, Lichtman A, Chapman V. (2013) Cannabinoid CB2 receptors regulate central sensitization and pain responses associated with osteoarthritis of the knee joint. PLoS One 8:e80440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Butler RK, Rea K, Lang Y, Gavin AM, Finn DP. (2008) Endocannabinoid-mediated enhancement of fear-conditioned analgesia in rats: opioid receptor dependency and molecular correlates. Pain 140:491–500. [DOI] [PubMed] [Google Scholar]
  37. 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 34:559–564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Capuron L, Ravaud A. (1999) Prediction of the depressive effects of interferon alfa therapy by the patient’s initial affective state. N Engl J Med 340:1370. [DOI] [PubMed] [Google Scholar]
  39. Capuron L, Ravaud A, Gualde N, Bosmans E, Dantzer R, Maes M, Neveu PJ. (2001) Association between immune activation and early depressive symptoms in cancer patients treated with interleukin-2-based therapy. Psychoneuroendocrinology 26:797–808. [DOI] [PubMed] [Google Scholar]
  40. Carlisle SJ, Marciano-Cabral F, Staab A, Ludwick C, Cabral GA. (2002) Differential expression of the CB2 cannabinoid receptor by rodent macrophages and macrophage-like cells in relation to cell activation. Int Immunopharmacol 2:69–82. [DOI] [PubMed] [Google Scholar]
  41. Cavuoto P, McAinch AJ, Hatzinikolas G, Janovska A, Game P, Wittert GA. (2007) The expression of receptors for endocannabinoids in human and rodent skeletal muscle. Biochem Biophys Res Commun 364:105–110. [DOI] [PubMed] [Google Scholar]
  42. Cheer JF, Wassum KM, Heien ML, Phillips PE, Wightman RM. (2004) Cannabinoids enhance subsecond dopamine release in the nucleus accumbens of awake rats. J Neurosci 24:4393–4400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Cichewicz DL, Martin ZL, Smith FL, Welch SP. (1999) Enhancement mu opioid antinociception by oral delta9-tetrahydrocannabinol: dose-response analysis and receptor identification. J Pharmacol Exp Ther 289:859–867. [PubMed] [Google Scholar]
  44. Cichewicz DL, McCarthy EA. (2003) Antinociceptive synergy between delta(9)-tetrahydrocannabinol and opioids after oral administration. J Pharmacol Exp Ther 304:1010–1015. [DOI] [PubMed] [Google Scholar]
  45. Cornelio AM, Nunes-de-Souza RL, Morgan MM. (2012) Contribution of the rostral ventromedial medulla to post-anxiety induced hyperalgesia. Brain Res 1450:80–86. [DOI] [PubMed] [Google Scholar]
  46. Cota D. (2007) CB1 receptors: emerging evidence for central and peripheral mechanisms that regulate energy balance, metabolism, and cardiovascular health. Diabetes Metab Res Rev 23:507–517. [DOI] [PubMed] [Google Scholar]
  47. Cota D, Steiner MA, Marsicano G, Cervino C, Herman JP, Grubler Y, Stalla J, Pasquali R, Lutz B, Stalla GK, Pagotto U. (2007) Requirement of cannabinoid receptor type 1 for the basal modulation of hypothalamic-pituitary-adrenal axis function. Endocrinology 148:1574–1581. [DOI] [PubMed] [Google Scholar]
  48. Cravatt BF, Giang DK, Mayfield SP, Boger DL, Lerner RA, Gilula NB. (1996) Molecular characterization of an enzyme that degrades neuromodulatory fatty-acid amides. Nature 384:83–87. [DOI] [PubMed] [Google Scholar]
  49. Dantzer R, O’Connor JC, Lawson MA, Kelley KW. (2011) Inflammation-associated depression: from serotonin to kynurenine. Psychoneuroendocrinology 36:426–436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. del Rey A, Yau HJ, Randolf A, Centeno MV, Wildmann J, Martina M, Besedovsky HO, Apkarian AV. (2011) Chronic neuropathic pain-like behavior correlates with IL-1beta expression and disrupts cytokine interactions in the hippocampus. Pain 152:2827–2835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Dellarole A, Morton P, Brambilla R, Walters W, Summers S, Bernardes D, Grilli M, Bethea JR. (2014) Neuropathic pain-induced depressive-like behavior and hippocampal neurogenesis and plasticity are dependent on TNFR1 signaling. Brain Behav Immun 41:65–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Demuth DG, Molleman A. (2006) Cannabinoid signalling. Life Sci 78:549–563. [DOI] [PubMed] [Google Scholar]
  53. Devane WA, Dysarz FA, 3rd, Johnson MR, Melvin LS, Howlett AC. (1988) Determination and characterization of a cannabinoid receptor in rat brain. Mol Pharmacol 34:605–613. [PubMed] [Google Scholar]
  54. Devane WA, Hanus L, Breuer A, Pertwee RG, Stevenson LA, Griffin G, Gibson D, Mandelbaum A, Etinger A, Mechoulam R. (1992) Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science 258:1946–1949. [DOI] [PubMed] [Google Scholar]
  55. Di Marzo V, Fontana A, Cadas H, Schinelli S, Cimino G, Schwartz JC, Piomelli D. (1994) Formation and inactivation of endogenous cannabinoid anandamide in central neurons. Nature 372:686–691. [DOI] [PubMed] [Google Scholar]
  56. Doan L, Manders T, Wang J. (2015) Neuroplasticity underlying the comorbidity of pain and depression. Neural Plast 2015:504691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Domenici MR, Azad SC, Marsicano G, Schierloh A, Wotjak CT, Dodt HU, Zieglgansberger W, Lutz B, Rammes G. (2006) Cannabinoid receptor type 1 located on presynaptic terminals of principal neurons in the forebrain controls glutamatergic synaptic transmission. J Neurosci 26:5794–5799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Domschke K, Dannlowski U, Ohrmann P, Lawford B, Bauer J, Kugel H, Heindel W, Young R, Morris P, Arolt V, Deckert J, Suslow T, Baune BT. (2008) Cannabinoid receptor 1 (CNR1) gene: impact on antidepressant treatment response and emotion processing in major depression. Eur Neuropsychopharmacol 18:751–759. [DOI] [PubMed] [Google Scholar]
  59. Emery PC, Wilson KG, Kowal J. (2014) Major depressive disorder and sleep disturbance in patients with chronic pain. Pain Res Manag 19:35–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Emptage NP, Sturm R, Robinson RL. (2005) Depression and comorbid pain as predictors of disability, employment, insurance status, and health care costs. Psychiatr Serv 56:468–474. [DOI] [PubMed] [Google Scholar]
  61. Fichna J, Wood JT, Papanastasiou M, Vadivel SK, Oprocha P, Salaga M, Sobczak M, Mokrowiecka A, Cygankiewicz AI, Zakrzewski PK, Malecka-Panas E, Krajewska WM, Koscielniak P, Makriyannis A, Storr MA. (2013) Endocannabinoid and cannabinoid-like fatty acid amide levels correlate with pain-related symptoms in patients with IBS-D and IBS-C: a pilot study. PLoS One 8:e85073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Finn DP, Beckett SR, Richardson D, Kendall DA, Marsden CA, Chapman V. (2004) Evidence for differential modulation of conditioned aversion and fear-conditioned analgesia by CB1 receptors. Eur J Neurosci 20:848–852. [DOI] [PubMed] [Google Scholar]
  63. Finn DP. (2010) Endocannabinoid-mediated modulation of stress responses: physiological and pathophysiological significance. Immunobiology 215:629–646. [DOI] [PubMed] [Google Scholar]
  64. Fisar Z. (2010) Inhibition of monoamine oxidase activity by cannabinoids. Naunyn Schmiedebergs Arch Pharmacol 381:563–572. [DOI] [PubMed] [Google Scholar]
  65. Franklin JM, Mathew M, Carrasco GA. (2013) Cannabinoid-induced upregulation of serotonin 2A receptors in the hypothalamic paraventricular nucleus and anxiety-like behaviors in rats. Neurosci Lett 548:165–169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Fukuhara K, Ishikawa K, Yasuda S, Kishishita Y, Kim HK, Kakeda T, Yamamoto M, Norii T, Ishikawa T. (2012) Intracerebroventricular 4-methylcatechol (4-MC) ameliorates chronic pain associated with depression-like behavior via induction of brain-derived neurotrophic factor (BDNF). Cell Mol Neurobiol 32:971–977. [DOI] [PubMed] [Google Scholar]
  67. Gai BM, Bortolatto CF, Bruning CA, Zborowski VA, Stein AL, Zeni G, Nogueira CW. (2014) Depression-related behavior and mechanical allodynia are blocked by 3-(4-fluorophenylselenyl)-2,5-diphenylselenophene in a mouse model of neuropathic pain induced by partial sciatic nerve ligation. Neuropharmacology 79:580–589. [DOI] [PubMed] [Google Scholar]
  68. Galante M, Diana MA. (2004) Group I metabotropic glutamate receptors inhibit GABA release at interneuron-Purkinje cell synapses through endocannabinoid production. J Neurosci 24:4865–4874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Galiegue S, Mary S, Marchand J, Dussossoy D, Carriere D, Carayon P, Bouaboula M, Shire D, Le Fur G, Casellas P. (1995) Expression of central and peripheral cannabinoid receptors in human immune tissues and leukocyte subpopulations. Eur J Biochem 232:54–61. [DOI] [PubMed] [Google Scholar]
  70. Gameiro GH, Andrade Ada S, de Castro M, Pereira LF, Tambeli CH, Veiga MC. (2005) The effects of restraint stress on nociceptive responses induced by formalin injected in rat’s TMJ. Pharmacol Biochem Behav 82:338–344. [DOI] [PubMed] [Google Scholar]
  71. Gameroff MJ, Olfson M. (2006) Major depressive disorder, somatic pain, and health care costs in an urban primary care practice. J Clin Psychiatry 67:1232–1239. [DOI] [PubMed] [Google Scholar]
  72. Gaoni Y, Mechoulam R. (1964) Isolation, structure, and partial synthesis of an active constituent of hashish. J Am Chem Soc 86:1646. [Google Scholar]
  73. Gibney SM, Gosselin RD, Dinan TG, Cryan JF. (2010) Colorectal distension–induced prefrontal cortex activation in the Wistar-Kyoto rat: implications for irritable bowel syndrome. Neuroscience 165:675–683. [DOI] [PubMed] [Google Scholar]
  74. Gobbi G, Bambico FR, Mangieri R, Bortolato M, Campolongo P, Solinas M, Cassano T, Morgese MG, Debonnel G, Duranti A, Tontini A, Tarzia G, Mor M, Trezza V, Goldberg SR, Cuomo V, Piomelli D. (2005) Antidepressant-like activity and modulation of brain monoaminergic transmission by blockade of anandamide hydrolysis. Proc Natl Acad Sci U S A 102:18620–18625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Goesling J, Clauw DJ, Hassett AL. (2013) Pain and depression: an integrative review of neurobiological and psychological factors. Curr Psychiatry Rep 15:421. [DOI] [PubMed] [Google Scholar]
  76. Goffer Y, Xu D, Eberle SE, D’Amour J, Lee M, Tukey D, Froemke RC, Ziff EB, Wang J. (2013) Calcium-permeable AMPA receptors in the nucleus accumbens regulate depression-like behaviors in the chronic neuropathic pain state. J Neurosci 33:19034–19044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Goncalves L, Silva R, Pinto-Ribeiro F, Pego JM, Bessa JM, Pertovaara A, Sousa N, Almeida A. (2008) Neuropathic pain is associated with depressive behavior and induces neuroplasticity in the amygdala of the rat. Exp Neurol 213:48–56. [DOI] [PubMed] [Google Scholar]
  78. Gong JP, Onaivi ES, Ishiguro H, Liu QR, Tagliaferro PA, Brusco A, Uhl GR. (2006) Cannabinoid CB2 receptors: immunohistochemical localization in rat brain. Brain Res 1071:10–23. [DOI] [PubMed] [Google Scholar]
  79. Gormsen L, Ribe AR, Raun P, Rosenberg R, Videbech P, Vestergaard P, Bach FW, Jensen TS. (2004) Pain thresholds during and after treatment of severe depression with electroconvulsive therapy. Eur J Pain 8:487–493. [DOI] [PubMed] [Google Scholar]
  80. Gorzalka BB, Hill MN. (2011) Putative role of endocannabinoid signaling in the etiology of depression and actions of antidepressants. Prog Neuropsychopharmacol Biol Psychiatry 35:1575–1585. [DOI] [PubMed] [Google Scholar]
  81. Gosselin RD, O’Connor RM, Tramullas M, Julio-Pieper M, Dinan TG, Cryan JF. (2010) Riluzole normalizes early-life stress-induced visceral hypersensitivity in rats: role of spinal glutamate reuptake mechanisms. Gastroenterology 138:2418–2425. [DOI] [PubMed] [Google Scholar]
  82. Gray JM, Vecchiarelli HA, Morena M, Lee TT, Hermanson DJ, Kim AB, McLaughlin RJ, Hassan KI, Kuhne C, Wotjak CT, Deussing JM, Patel S, Hill MN. (2015) Corticotropin-releasing hormone drives anandamide hydrolysis in the amygdala to promote anxiety. J Neurosci 35:3879–3892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Greenbaum L, Tegeder I, Barhum Y, Melamed E, Roditi Y, Djaldetti R. (2012) Contribution of genetic variants to pain susceptibility in Parkinson disease. Eur J Pain 16:1243–1250. [DOI] [PubMed] [Google Scholar]
  84. Grover S, Kumar V, Chakrabarti S, Hollikatti P, Singh P, Tyagi S, Kulhara P, Avasthi A. (2012) Prevalence and type of functional somatic complaints in patients with first-episode depression. East Asian Arch Psychiatry 22:146–153. [PubMed] [Google Scholar]
  85. Gutierrez T, Nackley AG, Neely MH, Freeman KG, Edwards GL, Hohmann AG. (2003) Effects of neurotoxic destruction of descending noradrenergic pathways on cannabinoid antinociception in models of acute and tonic nociception. Brain Res 987:176–185. [DOI] [PubMed] [Google Scholar]
  86. Haj-Dahmane S, Shen RY. (2011) Modulation of the serotonin system by endocannabinoid signaling. Neuropharmacology 61:414–420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Haller J, Matyas F, Soproni K, Varga B, Barsy B, Nemeth B, Mikics E, Freund TF, Hajos N. (2007) Correlated species differences in the effects of cannabinoid ligands on anxiety and on GABAergic and glutamatergic synaptic transmission. Eur J Neurosci 25:2445–2456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Hanus L, Abu-Lafi S, Fride E, Breuer A, Vogel Z, Shalev DE, Kustanovich I, Mechoulam R. (2001) 2-arachidonyl glyceryl ether, an endogenous agonist of the cannabinoid CB1 receptor. Proc Natl Acad Sci U S A 98:3662–3665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Hasnie FS, Wallace VC, Hefner K, Holmes A, Rice AS. (2007) Mechanical and cold hypersensitivity in nerve-injured C57BL/6J mice is not associated with fear-avoidance- and depression-related behavior. Br J Anaesth 98:816–822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Hassett AL, Aquino JK, Ilgen MA. (2014) The risk of suicide mortality in chronic pain patients. Curr Pain Headache Rep 18:436. [DOI] [PubMed] [Google Scholar]
  91. Hawker GA, Gignac MA, Badley E, Davis AM, French MR, Li Y, Perruccio AV, Power JD, Sale J, Lou W. (2011) A longitudinal study to explain the pain-depression link in older adults with osteoarthritis. Arthritis Care Res (Hoboken) 63:1382–1390. [DOI] [PubMed] [Google Scholar]
  92. Henry RJ, Kerr DM, Finn DP, Roche M. (2014) FAAH-mediated modulation of TLR3-induced neuroinflammation in the rat hippocampus. J Neuroimmunol 276:126–134. [DOI] [PubMed] [Google Scholar]
  93. Henry RJ, Kerr DM, Finn DP, Roche M. (2015) For whom the endocannabinoid tolls: modulation of innate immune function and implications for psychiatric disorders. Prog Neuropsychopharmacol Biol Psychiatry PMID: 25794989. [DOI] [PubMed] [Google Scholar]
  94. Herkenham M. (1991) Characterization and localization of cannabinoid receptors in brain: an in vitro technique using slide-mounted tissue sections. NIDA Res Monogr 112:129–145. [PubMed] [Google Scholar]
  95. Herkenham M, Lynn AB, Little MD, Johnson MR, Melvin LS, de Costa BR, Rice KC. (1990) Cannabinoid receptor localization in brain. Proc Natl Acad Sci U S A 87:1932–1936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Hill MN, Gorzalka BB. (2005) Is there a role for the endocannabinoid system in the etiology and treatment of melancholic depression? Behav Pharmacol 16:333–352. [DOI] [PubMed] [Google Scholar]
  97. Hill MN, Tasker JG. (2012) Endocannabinoid signaling, glucocorticoid-mediated negative feedback, and regulation of the hypothalamic-pituitary-adrenal axis. Neuroscience 204:5–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Hill MN, Patel S, Carrier EJ, Rademacher DJ, Ormerod BK, Hillard CJ, Gorzalka BB. (2005) Downregulation of endocannabinoid signaling in the hippocampus following chronic unpredictable stress. Neuropsychopharmacology 30:508–515. [DOI] [PubMed] [Google Scholar]
  99. Hill MN, Ho WS, Sinopoli KJ, Viau V, Hillard CJ, Gorzalka BB. (2006) Involvement of the endocannabinoid system in the ability of long-term tricyclic antidepressant treatment to suppress stress-induced activation of the hypothalamic-pituitary-adrenal axis. Neuropsychopharmacology 31:2591–2599. [DOI] [PubMed] [Google Scholar]
  100. Hill MN, Carrier EJ, McLaughlin RJ, Morrish AC, Meier SE, Hillard CJ, Gorzalka BB. (2008a) Regional alterations in the endocannabinoid system in an animal model of depression: effects of concurrent antidepressant treatment. J Neurochem 106:2322–2336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Hill MN, Ho WS, Hillard CJ, Gorzalka BB. (2008b) Differential effects of the antidepressants tranylcypromine and fluoxetine on limbic cannabinoid receptor binding and endocannabinoid contents. J Neural Transm 115:1673–1679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Hill MN, Miller GE, Ho WS, Gorzalka BB, Hillard CJ. (2008c) Serum endocannabinoid content is altered in females with depressive disorders: a preliminary report. Pharmacopsychiatry 41:48–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Hill MN, McLaughlin RJ, Morrish AC, Viau V, Floresco SB, Hillard CJ, Gorzalka BB. (2009a) Suppression of amygdalar endocannabinoid signaling by stress contributes to activation of the hypothalamic-pituitary-adrenal axis. Neuropsychopharmacology 34:2733–2745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Hill MN, Miller GE, Carrier EJ, Gorzalka BB, Hillard CJ. (2009b) Circulating endocannabinoids and N-acyl ethanolamines are differentially regulated in major depression and following exposure to social stress. Psychoneuroendocrinology 34:1257–1262. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Hill MN, McLaughlin RJ, Pan B, Fitzgerald ML, Roberts CJ, Lee TT, Karatsoreos IN, Mackie K, Viau V, Pickel VM, McEwen BS, Liu QS, Gorzalka BB, Hillard CJ. (2011) Recruitment of prefrontal cortical endocannabinoid signaling by glucocorticoids contributes to termination of the stress response. J Neurosci 31:10506–10515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Hillard CJ, Liu QS. (2014) Endocannabinoid signaling in the etiology and treatment of major depressive illness. Curr Pharm Des 20:3795–3811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Hu B, Doods H, Treede RD, Ceci A. (2009) Depression-like behavior in rats with mononeuropathy is reduced by the CB2-selective agonist GW405833. Pain 143:206–212. [DOI] [PubMed] [Google Scholar]
  108. Huang SM, Bisogno T, Trevisani M, Al-Hayani A, De Petrocellis L, Fezza F, Tognetto M, Petros TJ, Krey JF, Chu CJ, Miller JD, Davies SN, Geppetti P, Walker JM, Di Marzo V. (2002) An endogenous capsaicin-like substance with high potency at recombinant and native vanilloid VR1 receptors. Proc Natl Acad Sci U S A 99:8400–8405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Huggins JP, Smart TS, 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 153:1837–1846. [DOI] [PubMed] [Google Scholar]
  110. Imbe H, Kimura A, Donishi T, Kaneoke Y. (2012) Chronic restraint stress decreases glial fibrillary acidic protein and glutamate transporter in the periaqueductal gray matter. Neuroscience 223:209–218. [DOI] [PubMed] [Google Scholar]
  111. Ishikawa K, Yasuda S, Fukuhara K, Iwanaga Y, Ida Y, Ishikawa J, Yamagata H, Ono M, Kakeda T, Ishikawa T. (2014) 4-Methylcatechol prevents derangements of brain-derived neurotrophic factor and TrkB-related signaling in anterior cingulate cortex in chronic pain with depression-like behavior. Neuroreport 25:226–232. [DOI] [PubMed] [Google Scholar]
  112. Jennings EM, Okine BN, Olango WM, Roche M, Finn DP. (2015) Repeated forced swim stress differentially affects formalin-evoked nociceptive behavior and the endocannabinoid system in stress normo-responsive and stress hyper-responsive rat strains. Prog Neuropsychopharmacol Biol Psychiatry PMID: 25988529. [DOI] [PubMed] [Google Scholar]
  113. Jennings EM, Okine BN, Roche M, Finn DP. (2014) Stress-induced hyperalgesia. Prog Neurobiol 121:1–18. [DOI] [PubMed] [Google Scholar]
  114. Jentsch JD, Andrusiak E, Tran A, Bowers MB, Jr, Roth RH. (1997) Delta 9-tetrahydrocannabinol increases prefrontal cortical catecholaminergic utilization and impairs spatial working memory in the rat: blockade of dopaminergic effects with HA966. Neuropsychopharmacology 16:426–432. [DOI] [PubMed] [Google Scholar]
  115. Jesse CR, Wilhelm EA, Nogueira CW. (2010) Depression-like behavior and mechanical allodynia are reduced by bis selenide treatment in mice with chronic constriction injury: a comparison with fluoxetine, amitriptyline, and bupropion. Psychopharmacology (Berl) 212:513–522. [DOI] [PubMed] [Google Scholar]
  116. Ji G, Fu Y, Ruppert KA, Neugebauer V. (2007) Pain-related anxiety-like behavior requires CRF1 receptors in the amygdala. Mol Pain 3:13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Johnson AC, Tran L, Schulkin J, Greenwood-Van Meerveld B. (2012) Importance of stress receptor-mediated mechanisms in the amygdala on visceral pain perception in an intrinsically anxious rat. Neurogastroenterol Motil 24:479–486, e219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Juhasz G, Lazary J, Chase D, Pegg E, Downey D, Toth ZG, Stones K, Platt H, Mekli K, Payton A, Anderson IM, Deakin JF, Bagdy G. (2009) Variations in the cannabinoid receptor 1 gene predispose to migraine. Neurosci Lett 461:116–120. [DOI] [PubMed] [Google Scholar]
  119. Kadetoff D, Lampa J, Westman M, Andersson M, Kosek E. (2012) Evidence of central inflammation in fibromyalgia-increased cerebrospinal fluid interleukin-8 levels. J Neuroimmunol 242:33–38. [DOI] [PubMed] [Google Scholar]
  120. Kappelman MD, Long MD, Martin C, DeWalt DA, Kinneer PM, Chen W, Lewis JD, Sandler RS. (2014) Evaluation of the patient-reported outcomes measurement information system in a large cohort of patients with inflammatory bowel diseases. Clin Gastroenterol Hepatol 12:1315–1323 e1312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Katona I, Rancz EA, Acsady L, Ledent C, Mackie K, Hajos N, Freund TF. (2001) Distribution of CB1 cannabinoid receptors in the amygdala and their role in the control of GABAergic transmission. J Neurosci 21:9506–9518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Katz PS, Sulzer JK, Impastato RA, Teng SX, Rogers EK, Molina PE. (2015) Endocannabinoid degradation inhibition improves neurobehavioral function, blood-brain barrier integrity, and neuroinflammation following mild traumatic brain injury. J Neurotrauma 32:297–306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Kaufmann I, Hauer D, Huge V, Vogeser M, Campolongo P, Chouker A, Thiel M, Schelling G. (2009) Enhanced anandamide plasma levels in patients with complex regional pain syndrome following traumatic injury: a preliminary report. Eur Surg Res 43:325–329. [DOI] [PubMed] [Google Scholar]
  124. Kendell SF, Krystal JH, Sanacora G. (2005) GABA and glutamate systems as therapeutic targets in depression and mood disorders. Expert Opin Ther Targets 9:153–168. [DOI] [PubMed] [Google Scholar]
  125. Kerr DM, Burke NN, Ford GK, Connor TJ, Harhen B, Egan LJ, Finn DP, Roche M. (2012) Pharmacological inhibition of endocannabinoid degradation modulates the expression of inflammatory mediators in the hypothalamus following an immunological stressor. Neuroscience 204:53–63. [DOI] [PubMed] [Google Scholar]
  126. Kerr DM, Harhen B, Okine BN, Egan LJ, Finn DP, Roche M. (2013) The monoacylglycerol lipase inhibitor JZL184 attenuates LPS-induced increases in cytokine expression in the rat frontal cortex and plasma: differential mechanisms of action. Br J Pharmacol 169:808–819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Khongsaengdao B, Louthrenoo W, Srisurapanont M. (2000) Depression in Thai patients with rheumatoid arthritis. J Med Assoc Thai 83:743–747. [PubMed] [Google Scholar]
  128. Kim H, Chen L, Lim G, Sung B, Wang S, McCabe MF, Rusanescu G, Yang L, Tian Y, Mao J. (2012) Brain indoleamine 2,3-dioxygenase contributes to the comorbidity of pain and depression. J Clin Invest 122:2940–2954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Kim J, Li Y. (2015) Chronic activation of CB2 cannabinoid receptors in the hippocampus increases excitatory synaptic transmission. J Physiol 593:871–886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Kirilly E, Hunyady L, Bagdy G. (2013) Opposing local effects of endocannabinoids on the activity of noradrenergic neurons and release of noradrenaline: relevance for their role in depression and in the actions of CB(1) receptor antagonists. J Neural Transm 120:177–186. [DOI] [PubMed] [Google Scholar]
  131. Knaster P, Karlsson H, Estlander AM, Kalso E. (2012) Psychiatric disorders as assessed with SCID in chronic pain patients: the anxiety disorders precede the onset of pain. Gen Hosp Psychiatry 34:46–52. [DOI] [PubMed] [Google Scholar]
  132. Knerlich-Lukoschus F, Noack M, von der Ropp-Brenner B, Lucius R, Mehdorn HM, Held-Feindt J. (2011) Spinal cord injuries induce changes in CB1 cannabinoid receptor and C-C chemokine expression in brain areas underlying circuitry of chronic pain conditions. J Neurotrauma 28:619–634. [DOI] [PubMed] [Google Scholar]
  133. Knuth B, Radtke V, Rocha P, da Silva KS, Dalsoglio F, Gazal M, Jansen K, Souza DO, Portela LV, Kaster M, Oses JP. (2014) Prevalence of depression symptoms and serum levels of interleukin-6 in hemodialysis patients. Psychiatry Clin Neurosci 68:275–282. [DOI] [PubMed] [Google Scholar]
  134. Koch A, Zacharowski K, Boehm O, Stevens M, Lipfert P, von Giesen HJ, Wolf A, Freynhagen R. (2007) Nitric oxide and pro-inflammatory cytokines correlate with pain intensity in chronic pain patients. Inflamm Res 56:32–37. [DOI] [PubMed] [Google Scholar]
  135. Koethe D, Llenos IC, Dulay JR, Hoyer C, Torrey EF, Leweke FM, Weis S. (2007) Expression of CB1 cannabinoid receptor in the anterior cingulate cortex in schizophrenia, bipolar disorder, and major depression. J Neural Transm 114:1055–1063. [DOI] [PubMed] [Google Scholar]
  136. Kontinen VK, Kauppila T, Paananen S, Pertovaara A, Kalso E. (1999) Behavioral measures of depression and anxiety in rats with spinal nerve ligation-induced neuropathy. Pain 80:341–346. [DOI] [PubMed] [Google Scholar]
  137. Kozak KR, Rowlinson SW, Marnett LJ. (2000) Oxygenation of the endocannabinoid, 2-arachidonylglycerol, to glyceryl prostaglandins by cyclooxygenase-2. J Biol Chem 275:33744–33749. [DOI] [PubMed] [Google Scholar]
  138. Kvien TK. (2004) Epidemiology and burden of illness of rheumatoid arthritis. Pharmacoeconomics 22:1–12. [DOI] [PubMed] [Google Scholar]
  139. Kwilasz AJ, Abdullah RA, Poklis JL, Lichtman AH, Negus SS. (2014) Effects of the fatty acid amide hydrolase inhibitor URB597 on pain-stimulated and pain-depressed behavior in rats. Behav Pharmacol 25:119–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Kwilasz AJ, Negus SS. (2012) Dissociable effects of the cannabinoid receptor agonists Delta9-tetrahydrocannabinol and CP55940 on pain-stimulated versus pain-depressed behavior in rats. J Pharmacol Exp Ther 343:389–400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Larauche M, Bradesi S, Million M, McLean P, Tache Y, Mayer EA, McRoberts JA. (2008) Corticotropin-releasing factor type 1 receptors mediate the visceral hyperalgesia induced by repeated psychological stress in rats. Am J Physiol Gastrointest Liver Physiol 294:G1033–1040. [DOI] [PubMed] [Google Scholar]
  142. La Porta C, Bura AS, Llorente-Onaindia J, Pastor A, Navarrete F, Garcia-Gutierrez MS, De la Torre R, Manzanares J, Monfort J, Maldonado R. (2015) Role of the endocannabinoid system in the emotional manifestations of osteoarthritis pain. Pain PMID: 26067584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Lautenbacher S, Spernal J, Schreiber W, Krieg JC. (1999) Relationship between clinical pain complaints and pain sensitivity in patients with depression and panic disorder. Psychosom Med 61:822–827. [DOI] [PubMed] [Google Scholar]
  144. Le AM, Lee M, Su C, Zou A, Wang J. (2014) AMPAkines have novel analgesic properties in rat models of persistent neuropathic and inflammatory pain. Anesthesiology 121:1080–1090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Lee MC, Ploner M, Wiech K, Bingel U, Wanigasekera V, Brooks J, Menon DK, Tracey I. (2013) Amygdala activity contributes to the dissociative effect of cannabis on pain perception. Pain 154:124–134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Leggett JD, Aspley S, Beckett SR, D’Antona AM, Kendall DA. (2004) Oleamide is a selective endogenous agonist of rat and human CB1 cannabinoid receptors. Br J Pharmacol 141:253–262. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Li JX. (2015) Pain and depression comorbidity: a preclinical perspective. Behav Brain Res 276:92–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Lim G, Sung B, Ji RR, Mao J. (2003) Upregulation of spinal cannabinoid-1-receptors following nerve injury enhances the effects of Win 55,212-2 on neuropathic pain behaviors in rats. Pain 105:275–283. [DOI] [PubMed] [Google Scholar]
  149. Lin MC, Guo HR, Lu MC, Livneh H, Lai NS, Tsai TY. (2015) Increased risk of depression in patients with rheumatoid arthritis: a seven-year population-based cohort study. Clinics (Sao Paulo) 70:91–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Liu SB, Zhao R, Li XS, Guo HJ, Tian Z, Zhang N, Gao GD, Zhao MG. (2014) Attenuation of reserpine-induced pain/depression dyad by gentiopicroside through downregulation of GluN2B receptors in the amygdala of mice. Neuromolecular Med 16:350–359. [DOI] [PubMed] [Google Scholar]
  151. Lockie SH, Czyzyk TA, Chaudhary N, Perez-Tilve D, Woods SC, Oldfield BJ, Statnick MA, Tschop MH. (2011) CNS opioid signaling separates cannabinoid receptor 1-mediated effects on body weight and mood-related behavior in mice. Endocrinology 152:3661–3667. [DOI] [PubMed] [Google Scholar]
  152. Lomazzo E, Bindila L, Remmers F, Lerner R, Schwitter C, Hoheisel U, Lutz B. (2015) Therapeutic potential of inhibitors of endocannabinoid degradation for the treatment of stress-related hyperalgesia in an animal model of chronic pain. Neuropsychopharmacology 40:488–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Lu C, Liu Y, Sun B, Sun Y, Hou B, Zhang Y, Ma Z, Gu X. (2015) Intrathecal injection of JWH-015 attenuates bone cancer pain via time-dependent modification of pro-inflammatory cytokines expression and astrocytes activity in spinal cord. Inflammation PMID: 25896633. [DOI] [PubMed] [Google Scholar]
  154. Ludwig J, Binder A, Steinmann J, Wasner G, Baron R. (2008) Cytokine expression in serum and cerebrospinal fluid in non-inflammatory polyneuropathies. J Neurol Neurosurg Psychiatry 79:1268–1273. [DOI] [PubMed] [Google Scholar]
  155. Luo J, Wang T, Liang S, Hu X, Li W, Jin F. (2013) Experimental gastritis leads to anxiety- and depression-like behaviors in female but not male rats. Behav Brain Funct 9:46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Lutz PE, Kieffer BL. (2013) Opioid receptors: distinct roles in mood disorders. Trends Neurosci 36:195–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Maciel IS, Silva RB, Morrone FB, Calixto JB, Campos MM. (2013) Synergistic effects of celecoxib and bupropion in a model of chronic inflammation-related depression in mice. PLoS One 8:e77227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Mackie K. (2008) Cannabinoid receptors: where they are and what they do. J Neuroendocrinol 20 Suppl 1:10–14. [DOI] [PubMed] [Google Scholar]
  159. Maes M, Leonard BE, Myint AM, Kubera M, Verkerk R. (2011) The new ‘5-HT’ hypothesis of depression: cell-mediated immune activation induces indoleamine 2,3-dioxygenase, which leads to lower plasma tryptophan and an increased synthesis of detrimental tryptophan catabolites (TRYCATs), both of which contribute to the onset of depression. Prog Neuropsychopharmacol Biol Psychiatry 35:702–721. [DOI] [PubMed] [Google Scholar]
  160. Maida V, Ennis M, Irani S, Corbo M, Dolzhykov M. (2008) Adjunctive nabilone in cancer pain and symptom management: a prospective observational study using propensity scoring. J Support Oncol 6:119–124. [PubMed] [Google Scholar]
  161. Maletic V, Raison CL. (2009) Neurobiology of depression, fibromyalgia and neuropathic pain. Front Biosci (Landmark Ed) 14:5291–5338. [DOI] [PubMed] [Google Scholar]
  162. Maneeton N, Maneeton B, Srisurapanont M. (2013) Prevalence and predictors of pain in patients with major depressive disorder. Asian J Psychiatr 6:288–291. [DOI] [PubMed] [Google Scholar]
  163. Manning BH, Martin WJ, Meng ID. (2003) The rodent amygdala contributes to the production of cannabinoid-induced antinociception. Neuroscience 120:1157–1170. [DOI] [PubMed] [Google Scholar]
  164. Marco EM, Echeverry-Alzate V, Lopez-Moreno JA, Gine E, Penasco S, Viveros MP. (2014) Consequences of early life stress on the expression of endocannabinoid-related genes in the rat brain. Behav Pharmacol 25:547–556. [DOI] [PubMed] [Google Scholar]
  165. Maric NP, Adzic M. (2013) Pharmacological modulation of HPA axis in depression - new avenues for potential therapeutic benefits. Psychiatr Danub 25:299–305. [PubMed] [Google Scholar]
  166. Mato S, Vidal R, Castro E, Diaz A, Pazos A, Valdizan EM. (2010) Long-term fluoxetine treatment modulates cannabinoid type 1 receptor-mediated inhibition of adenylyl cyclase in the rat prefrontal cortex through 5-hydroxytryptamine 1A receptor-dependent mechanisms. Mol Pharmacol 77:424–434. [DOI] [PubMed] [Google Scholar]
  167. Matsuda LA, Lolait SJ, Brownstein MJ, Young AC, Bonner TI. (1990) Structure of a cannabinoid receptor and functional expression of the cloned cDNA. Nature 346:561–564. [DOI] [PubMed] [Google Scholar]
  168. McLaughlin RJ, Hill MN, Bambico FR, Stuhr KL, Gobbi G, Hillard CJ, Gorzalka BB. (2012) Prefrontal cortical anandamide signaling coordinates coping responses to stress through a serotonergic pathway. Eur Neuropsychopharmacol 22:664–671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. McLaughlin RJ, Hill MN, Dang SS, Wainwright SR, Galea LA, Hillard CJ, Gorzalka BB. (2013) Upregulation of CB(1) receptor binding in the ventromedial prefrontal cortex promotes proactive stress-coping strategies following chronic stress exposure. Behav Brain Res 237:333–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Mechoulam R, Ben-Shabat S, Hanus L, Ligumsky M, Kaminski NE, Schatz AR, Gopher A, Almog S, Martin BR, Compton DR, et al. (1995) Identification of an endogenous 2-monoglyceride, present in canine gut, that binds to cannabinoid receptors. Biochem Pharmacol 50:83–90. [DOI] [PubMed] [Google Scholar]
  171. Meerwijk EL, Ford JM, Weiss SJ. (2013) Brain regions associated with psychological pain: implications for a neural network and its relationship to physical pain. Brain Imaging Behav 7:1–14. [DOI] [PubMed] [Google Scholar]
  172. Melis M, Pistis M. (2012) Hub and switches: endocannabinoid signalling in midbrain dopamine neurons. Philos Trans R Soc Lond B Biol Sci 367:3276–3285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Meng ID, Manning BH, Martin WJ, Fields HL. (1998) An analgesia circuit activated by cannabinoids. Nature 395:381–383. [DOI] [PubMed] [Google Scholar]
  174. Millan MJ. (2002) Descending control of pain. Prog Neurobiol 66:355–474. [DOI] [PubMed] [Google Scholar]
  175. Miller LL, Picker MJ, Schmidt KT, Dykstra LA. (2011) Effects of morphine on pain-elicited and pain-suppressed behavior in CB1 knockout and wildtype mice. Psychopharmacology (Berl) 215:455–465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Miller LL, Picker MJ, Umberger MD, Schmidt KT, Dykstra LA. (2012) Effects of alterations in cannabinoid signaling, alone and in combination with morphine, on pain-elicited and pain-suppressed behavior in mice. J Pharmacol Exp Ther 342:177–187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Minocci D, Massei J, Martino A, Milianti M, Piz L, Di Bello D, Sbrana A, Martinotti E, Rossi AM, Nieri P. (2011) Genetic association between bipolar disorder and 524A>C (Leu133Ile) polymorphism of CNR2 gene, encoding for CB2 cannabinoid receptor. J Affect Disord 134:427–430. [DOI] [PubMed] [Google Scholar]
  178. Mitjans M, Gasto C, Catalan R, Fananas L, Arias B. (2012) Genetic variability in the endocannabinoid system and 12-week clinical response to citalopram treatment: the role of the CNR1, CNR2 and FAAH genes. J Psychopharmacol 26:1391–1398. [DOI] [PubMed] [Google Scholar]
  179. Mitjans M, Serretti A, Fabbri C, Gasto C, Catalan R, Fananas L, Arias B. (2013) Screening genetic variability at the CNR1 gene in both major depression etiology and clinical response to citalopram treatment. Psychopharmacology (Berl) 227:509–519. [DOI] [PubMed] [Google Scholar]
  180. Mitrirattanakul S, Ramakul N, Guerrero AV, Matsuka Y, Ono T, Iwase H, Mackie K, Faull KF, Spigelman I. (2006) Site-specific increases in peripheral cannabinoid receptors and their endogenous ligands in a model of neuropathic pain. Pain 126:102–114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Monteleone P, Bifulco M, Maina G, Tortorella A, Gazzerro P, Proto MC, Di Filippo C, Monteleone F, Canestrelli B, Buonerba G, Bogetto F, Maj M. (2010) Investigation of CNR1 and FAAH endocannabinoid gene polymorphisms in bipolar disorder and major depression. Pharmacol Res 61:400–404. [DOI] [PubMed] [Google Scholar]
  182. Moranta D, Esteban S, Garcia-Sevilla JA. (2009) Chronic treatment and withdrawal of the cannabinoid agonist WIN 55,212-2 modulate the sensitivity of presynaptic receptors involved in the regulation of monoamine syntheses in rat brain. Naunyn Schmiedebergs Arch Pharmacol 379:61–72. [DOI] [PubMed] [Google Scholar]
  183. Munro S, Thomas KL, Abu-Shaar M. (1993) Molecular characterization of a peripheral receptor for cannabinoids. Nature 365:61–65. [DOI] [PubMed] [Google Scholar]
  184. Naderi N, Shafaghi B, Khodayar MJ, Zarindast MR. (2005) Interaction between gamma-aminobutyric acid GABAB and cannabinoid CB1 receptors in spinal pain pathways in rat. Eur J Pharmacol 514:159–164. [DOI] [PubMed] [Google Scholar]
  185. Naderi N, Haghparast A, Saber-Tehrani A, Rezaii N, Alizadeh AM, Khani A, Motamedi F. (2008) Interaction between cannabinoid compounds and diazepam on anxiety-like behavior of mice. Pharmacol Biochem Behav 89:64–75. [DOI] [PubMed] [Google Scholar]
  186. Nagakura Y, Oe T, Aoki T, Matsuoka N. (2009) Biogenic amine depletion causes chronic muscular pain and tactile allodynia accompanied by depression: a putative animal model of fibromyalgia. Pain 146:26–33. [DOI] [PubMed] [Google Scholar]
  187. Narita M, Kaneko C, Miyoshi K, Nagumo Y, Kuzumaki N, Nakajima M, Nanjo K, Matsuzawa K, Yamazaki M, Suzuki T. (2006) Chronic pain induces anxiety with concomitant changes in opioidergic function in the amygdala. Neuropsychopharmacology 31:739–750. [DOI] [PubMed] [Google Scholar]
  188. Navarria A, Tamburella A, Iannotti FA, Micale V, Camillieri G, Gozzo L, Verde R, Imperatore R, Leggio GM, Drago F, Di Marzo V. (2014) The dual blocker of FAAH/TRPV1 N-arachidonoylserotonin reverses the behavioral despair induced by stress in rats and modulates the HPA-axis. Pharmacol Res 87:151–159. [DOI] [PubMed] [Google Scholar]
  189. Nicholl BI, Mackay D, Cullen B, Martin DJ, Ul-Haq Z, Mair FS, Evans J, McIntosh AM, Gallagher J, Roberts B, Deary IJ, Pell JP, Smith DJ. (2014) Chronic multisite pain in major depression and bipolar disorder: cross-sectional study of 149,611 participants in UK Biobank. BMC Psychiatry 14:350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Nogueira JB, Sena LC, Quintans Jde S, Almeida JR, Franca AV, Junior LJ. (2012) Side effects of the therapy with peginterferon and ribavirin in chronic hepatitis C: a small audit. J Pharm Pract 25:85–88. [DOI] [PubMed] [Google Scholar]
  191. Norman GJ, Karelina K, Zhang N, Walton JC, Morris JS, Devries AC. (2010) Stress and IL-1beta contribute to the development of depressive-like behavior following peripheral nerve injury. Mol Psychiatry 15:404–414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Nunez E, Benito C, Pazos MR, Barbachano A, Fajardo O, Gonzalez S, Tolon RM, Romero J. (2004) Cannabinoid CB2 receptors are expressed by perivascular microglial cells in the human brain: an immunohistochemical study. Synapse 53:208–213. [DOI] [PubMed] [Google Scholar]
  193. O’Malley D, Julio-Pieper M, Gibney SM, Dinan TG, Cryan JF. (2010) Distinct alterations in colonic morphology and physiology in two rat models of enhanced stress-induced anxiety and depression-like behavior. Stress 13:114–122. [DOI] [PubMed] [Google Scholar]
  194. Oropeza VC, Page ME, Van Bockstaele EJ. (2005) Systemic administration of WIN 55,212-2 increases norepinephrine release in the rat frontal cortex. Brain Res 1046:45–54. [DOI] [PubMed] [Google Scholar]
  195. Osei-Hyiaman D, DePetrillo M, Pacher P, Liu J, Radaeva S, Batkai S, Harvey-White J, Mackie K, Offertaler L, Wang L, Kunos G. (2005) Endocannabinoid activation at hepatic CB1 receptors stimulates fatty acid synthesis and contributes to diet-induced obesity. J Clin Invest 115:1298–1305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Overton HA, Babbs AJ, Doel SM, Fyfe MC, Gardner LS, Griffin G, Jackson HC, Procter MJ, Rasamison CM, Tang-Christensen M, Widdowson PS, Williams GM, Reynet C. (2006) Deorphanization of a G protein-coupled receptor for oleoylethanolamide and its use in the discovery of small-molecule hypophagic agents. Cell Metab 3:167–175. [DOI] [PubMed] [Google Scholar]
  197. Page ME, Oropeza VC, Sparks SE, Qian Y, Menko AS, Van Bockstaele EJ. (2007) Repeated cannabinoid administration increases indices of noradrenergic activity in rats. Pharmacol Biochem Behav 86:162–168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Page ME, Oropeza VC, Van Bockstaele EJ. (2008) Local administration of a cannabinoid agonist alters norepinephrine efflux in the rat frontal cortex. Neurosci Lett 431:1–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Palazzo E, de Novellis V, Petrosino S, Marabese I, Vita D, Giordano C, Di Marzo V, Mangoni GS, Rossi F, Maione S. (2006) Neuropathic pain and the endocannabinoid system in the dorsal raphe: pharmacological treatment and interactions with the serotonergic system. Eur J Neurosci 24:2011–2020. [DOI] [PubMed] [Google Scholar]
  200. Pare WP, Redei E. (1993) Depressive behavior and stress ulcer in Wistar Kyoto rats. J Physiol Paris 87:229–238. [DOI] [PubMed] [Google Scholar]
  201. Parent AJ, Beaudet N, Beaudry H, Bergeron J, Berube P, Drolet G, Sarret P, Gendron L. (2012) Increased anxiety-like behaviors in rats experiencing chronic inflammatory pain. Behav Brain Res 229:160–167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Park JM, Choi MG, Cho YK, Lee IS, Kim SW, Choi KY, Chung IS. (2011) Cannabinoid receptor 1 gene polymorphism and irritable bowel syndrome in the Korean population: a hypothesis-generating study. J Clin Gastroenterol 45:45–49. [DOI] [PubMed] [Google Scholar]
  203. Patel S, Roelke CT, Rademacher DJ, Cullinan WE, Hillard CJ. (2004) Endocannabinoid signaling negatively modulates stress-induced activation of the hypothalamic-pituitary-adrenal axis. Endocrinology 145:5431–5438. [DOI] [PubMed] [Google Scholar]
  204. Pellkofer HL, Havla J, Hauer D, Schelling G, Azad SC, Kuempfel T, Magerl W, Huge V. (2013) The major brain endocannabinoid 2-AG controls neuropathic pain and mechanical hyperalgesia in patients with neuromyelitis optica. PLoS One 8:e71500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. Pernia-Andrade AJ, Kato A, Witschi R, Nyilas R, Katona I, Freund TF, Watanabe M, Filitz J, Koppert W, Schuttler J, Ji G, Neugebauer V, Marsicano G, Lutz B, Vanegas H, Zeilhofer HU. (2009) Spinal endocannabinoids and CB1 receptors mediate C-fiber-induced heterosynaptic pain sensitization. Science 325:760–764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Poole H, White S, Blake C, Murphy P, Bramwell R. (2009) Depression in chronic pain patients: prevalence and measurement. Pain Pract 9:173–180. [DOI] [PubMed] [Google Scholar]
  207. Porter AC, Sauer JM, Knierman MD, Becker GW, Berna MJ, Bao J, Nomikos GG, Carter P, Bymaster FP, Leese AB, Felder CC. (2002) Characterization of a novel endocannabinoid, virodhamine, with antagonist activity at the CB1 receptor. J Pharmacol Exp Ther 301:1020–1024. [DOI] [PubMed] [Google Scholar]
  208. Prescott SM, Majerus PW. (1983) Characterization of 1,2-diacylglycerol hydrolysis in human platelets. Demonstration of an arachidonoyl-monoacylglycerol intermediate. J Biol Chem 258:764–769. [PubMed] [Google Scholar]
  209. Quintero L, Cardenas R, Suarez-Roca H. (2011) Stress-induced hyperalgesia is associated with a reduced and delayed GABA inhibitory control that enhances post-synaptic NMDA receptor activation in the spinal cord. Pain 152:1909–1922. [DOI] [PubMed] [Google Scholar]
  210. Racz I, Nent E, Erxlebe E, Zimmer A. (2015) CB1 receptors modulate affective behavior induced by neuropathic pain. Brain Res Bull 114:42–48. [DOI] [PubMed] [Google Scholar]
  211. Rani Sagar D, Burston JJ, Woodhams SG, Chapman V. (2012) Dynamic changes to the endocannabinoid system in models of chronic pain. Philos Trans R Soc Lond B Biol Sci 367:3300–3311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Rea K, Roche M, Finn DP. (2007) Supraspinal modulation of pain by cannabinoids: the role of GABA and glutamate. Br J Pharmacol 152:633–648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Rea K, Olango WM, Okine BN, Madasu MK, McGuire IC, Coyle K, Harhen B, Roche M, Finn DP. (2014) Impaired endocannabinoid signalling in the rostral ventromedial medulla underpins genotype-dependent hyper-responsivity to noxious stimuli. Pain 155:69–79. [DOI] [PubMed] [Google Scholar]
  214. Reich CG, Mihalik GR, Iskander AN, Seckler JC, Weiss MS. (2013) Adolescent chronic mild stress alters hippocampal CB1 receptor-mediated excitatory neurotransmission and plasticity. Neuroscience 253:444–454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Reich CG, Taylor ME, McCarthy MM. (2009) Differential effects of chronic unpredictable stress on hippocampal CB1 receptors in male and female rats. Behav Brain Res 203:264–269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Rey AA, Purrio M, Viveros MP, Lutz B. (2012) Biphasic effects of cannabinoids in anxiety responses: CB1 and GABA(B) receptors in the balance of GABAergic and glutamatergic neurotransmission. Neuropsychopharmacology 37:2624–2634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Richardson D, Pearson RG, Kurian N, Latif ML, Garle MJ, Barrett DA, Kendall DA, Scammell BE, Reeve AJ, Chapman V. (2008) Characterisation of the cannabinoid receptor system in synovial tissue and fluid in patients with osteoarthritis and rheumatoid arthritis. Arthritis Res Ther 10:R43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  218. Riebe CJ, Wotjak CT. (2011) Endocannabinoids and stress. Stress 14:384–397. [DOI] [PubMed] [Google Scholar]
  219. Rivat C, Becker C, Blugeot A, Zeau B, Mauborgne A, Pohl M, Benoliel JJ. (2010) Chronic stress induces transient spinal neuroinflammation, triggering sensory hypersensitivity and long-lasting anxiety-induced hyperalgesia. Pain 150:358–368. [DOI] [PubMed] [Google Scholar]
  220. Roberts CJ, Stuhr KL, Hutz MJ, Raff H, Hillard CJ. (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] [PMC free article] [PubMed] [Google Scholar]
  221. Roberts JD, Gennings C, Shih M. (2006) Synergistic affective analgesic interaction between delta-9-tetrahydrocannabinol and morphine. Eur J Pharmacol 530:54–58. [DOI] [PubMed] [Google Scholar]
  222. Roche M, Finn DP. (2010) Brain CB2 Receptors: implications for neuropsychiatric disorders. Pharmaceuticals 3:2517–2553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Rock RB, Gekker G, Hu S, Sheng WS, Cabral GA, Martin BR, Peterson PK. (2007) WIN55,212-2-mediated inhibition of HIV-1 expression in microglial cells: involvement of cannabinoid receptors. J Neuroimmune Pharmacol 2:178–183. [DOI] [PubMed] [Google Scholar]
  224. Rodriguez JJ, Mackie K, Pickel VM. (2001) Ultrastructural localization of the CB1 cannabinoid receptor in mu-opioid receptor patches of the rat Caudate putamen nucleus. J Neurosci 21:823–833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Rodriguez-Gaztelumendi A, Rojo ML, Pazos A, Diaz A. (2014) An altered spinal serotonergic system contributes to increased thermal nociception in an animal model of depression. Exp Brain Res 232:1793–1803. [DOI] [PubMed] [Google Scholar]
  226. Rossi S, De Chiara V, Musella A, Sacchetti L, Cantarella C, Castelli M, Cavasinni F, Motta C, Studer V, Bernardi G, Cravatt BF, Maccarrone M, Usiello A, Centonze D. (2010) Preservation of striatal cannabinoid CB1 receptor function correlates with the antianxiety effects of fatty acid amide hydrolase inhibition. Mol Pharmacol 78:260–268. [DOI] [PubMed] [Google Scholar]
  227. Rossi S, Motta C, Musella A, Centonze D. (2014) The interplay between inflammatory cytokines and the endocannabinoid system in the regulation of synaptic transmission. Neuropharmacology 96(A):105–112. [DOI] [PubMed] [Google Scholar]
  228. Ryberg E, Larsson N, Sjogren S, Hjorth S, Hermansson NO, Leonova J, Elebring T, Nilsson K, Drmota T, Greasley PJ. (2007) The orphan receptor GPR55 is a novel cannabinoid receptor. Br J Pharmacol 152:1092–1101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. Salio C, Fischer J, Franzoni MF, Mackie K, Kaneko T, Conrath M. (2001) CB1-cannabinoid and mu-opioid receptor co-localization on postsynaptic target in the rat dorsal horn. Neuroreport 12:3689–3692. [DOI] [PubMed] [Google Scholar]
  230. Scheidt CE, Mueller-Becsangele J, Hiller K, Hartmann A, Goldacker S, Vaith P, Waller E, Lacour M. (2014) Self-reported symptoms of pain and depression in primary fibromyalgia syndrome and rheumatoid arthritis. Nord J Psychiatry 68:88–92. [DOI] [PubMed] [Google Scholar]
  231. Scholich SL, Hallner D, Wittenberg RH, Hasenbring MI, Rusu AC. (2012) The relationship between pain, disability, quality of life and cognitive-behavioral factors in chronic back pain. Disabil Rehabil 34:1993–2000. [DOI] [PubMed] [Google Scholar]
  232. Seif T, Makriyannis A, Kunos G, Bonci A, Hopf FW. (2011) The endocannabinoid 2-arachidonoylglycerol mediates D1 and D2 receptor cooperative enhancement of rat nucleus accumbens core neuron firing. Neuroscience 193:21–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. Selvarajah D, Gandhi R, Emery CJ, Tesfaye S. (2010) Randomized placebo-controlled double-blind clinical trial of cannabis-based medicinal product (Sativex) in painful diabetic neuropathy: depression is a major confounding factor. Diabetes Care 33:128–130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Shahidi S, Hasanein P. (2011) Behavioral effects of fatty acid amide hydrolase inhibition on morphine withdrawal symptoms. Brain Res Bull 86:118–122. [DOI] [PubMed] [Google Scholar]
  235. Shi M, Wang JY, Luo F. (2010) Depression shows divergent effects on evoked and spontaneous pain behaviors in rats. J Pain 11:219–229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Skrabek RQ, Galimova L, Ethans K, Perry D. (2008) Nabilone for the treatment of pain in fibromyalgia. J Pain 9:164–173. [DOI] [PubMed] [Google Scholar]
  237. Smaga I, Bystrowska B, Gawlinski D, Pomierny B, Stankowicz P, Filip M. (2014) Antidepressants and changes in concentration of endocannabinoids and N-acylethanolamines in rat brain structures. Neurotox Res 26:190–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Smith PA, Selley DE, Sim-Selley LJ, Welch SP. (2007) Low dose combination of morphine and delta9-tetrahydrocannabinol circumvents antinociceptive tolerance and apparent desensitization of receptors. Eur J Pharmacol 571:129–137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  239. Solinas M, Justinova Z, Goldberg SR, Tanda G. (2006) Anandamide administration alone and after inhibition of fatty acid amide hydrolase (FAAH) increases dopamine levels in the nucleus accumbens shell in rats. J Neurochem 98:408–419. [DOI] [PubMed] [Google Scholar]
  240. Song C, Wang H. (2011) Cytokines mediated inflammation and decreased neurogenesis in animal models of depression. Prog Neuropsychopharmacol Biol Psychiatry 35:760–768. [DOI] [PubMed] [Google Scholar]
  241. Spitzer RL, Kroenke K, Linzer M, Hahn SR, Williams JB, deGruy FV, 3rd, Brody D, Davies M. (1995) Health-related quality of life in primary care patients with mental disorders. Results from the PRIME-MD 1000 Study. JAMA 274:1511–1517. [PubMed] [Google Scholar]
  242. Steiner MA, Marsicano G, Nestler EJ, Holsboer F, Lutz B, Wotjak CT. (2008) Antidepressant-like behavioral effects of impaired cannabinoid receptor type 1 signaling coincide with exaggerated corticosterone secretion in mice. Psychoneuroendocrinology 33:54–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Suarez-Roca H, Leal L, Silva JA, Pinerua-Shuhaibar L, Quintero L. (2008) Reduced GABA neurotransmission underlies hyperalgesia induced by repeated forced swimming stress. Behav Brain Res 189:159–169. [DOI] [PubMed] [Google Scholar]
  244. Sugiura T, Kondo S, Sukagawa A, Nakane S, Shinoda A, Itoh K, Yamashita A, Waku K. (1995) 2-Arachidonoylglycerol: a possible endogenous cannabinoid receptor ligand in brain. Biochem Biophys Res Commun 215:89–97. [DOI] [PubMed] [Google Scholar]
  245. Sugiura T, Kondo S, Sukagawa A, Tonegawa T, Nakane S, Yamashita A, Waku K. (1996) Enzymatic synthesis of anandamide, an endogenous cannabinoid receptor ligand, through N-acylphosphatidylethanolamine pathway in testis: involvement of Ca(2+)-dependent transacylase and phosphodiesterase activities. Biochem Biophys Res Commun 218:113–117. [DOI] [PubMed] [Google Scholar]
  246. Sugiura T, Kondo S, Kishimoto S, Miyashita T, Nakane S, Kodaka T, Suhara Y, Takayama H, Waku K. (2000) Evidence that 2-arachidonoylglycerol but not N-palmitoylethanolamine or anandamide is the physiological ligand for the cannabinoid CB2 receptor. Comparison of the agonistic activities of various cannabinoid receptor ligands in HL-60 cells. J Biol Chem 275:605–612. [DOI] [PubMed] [Google Scholar]
  247. Sun Y, Alexander SP, Kendall DA, Bennett AJ. (2006) Cannabinoids and PPARalpha signalling. Biochem Soc Trans 34:1095–1097. [DOI] [PubMed] [Google Scholar]
  248. Suzuki T, Amata M, Sakaue G, Nishimura S, Inoue T, Shibata M, Mashimo T. (2007) Experimental neuropathy in mice is associated with delayed behavioral changes related to anxiety and depression. Anesth Analg 104:1570–1577. [DOI] [PubMed] [Google Scholar]
  249. Takahashi RN, Ramos GA, Assini FL. (2003) Anxiety does not affect the antinociceptive effect of Delta 9-THC in mice: participation of cannabinoid and opioid receptors. Pharmacol Biochem Behav 75:763–768. [DOI] [PubMed] [Google Scholar]
  250. Tham SM, Angus JA, Tudor EM, Wright CE. (2005) Synergistic and additive interactions of the cannabinoid agonist CP55,940 with mu opioid receptor and alpha2-adrenoceptor agonists in acute pain models in mice. Br J Pharmacol 144:875–884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  251. Thorat SN, Bhargava HN. (1994) Evidence for a bidirectional cross-tolerance between morphine and delta 9-tetrahydrocannabinol in mice. Eur J Pharmacol 260:5–13. [DOI] [PubMed] [Google Scholar]
  252. Thornton LM, Andersen BL, Blakely WP. (2010) The pain, depression, and fatigue symptom cluster in advanced breast cancer: covariation with the hypothalamic-pituitary-adrenal axis and the sympathetic nervous system. Health Psychol 29:333–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  253. Tsou K, Brown S, Sanudo-Pena MC, Mackie K, Walker JM. (1998) Immunohistochemical distribution of cannabinoid CB1 receptors in the rat central nervous system. Neuroscience 83:393–411. [DOI] [PubMed] [Google Scholar]
  254. Tuglu C, Kara SH, Caliyurt O, Vardar E, Abay E. (2003) Increased serum tumor necrosis factor-alpha levels and treatment response in major depressive disorder. Psychopharmacology (Berl) 170:429–433. [DOI] [PubMed] [Google Scholar]
  255. Ulugol A. (2014) The endocannabinoid system as a potential therapeutic target for pain modulation. Balkan Med J 31:115–120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. van der Stelt M, van Kuik JA, Bari M, van Zadelhoff G, Leeflang BR, Veldink GA, Finazzi-Agro A, Vliegenthart JF, Maccarrone M. (2002) Oxygenated metabolites of anandamide and 2-arachidonoylglycerol: conformational analysis and interaction with cannabinoid receptors, membrane transporter, and fatty acid amide hydrolase. J Med Chem 45:3709–3720. [DOI] [PubMed] [Google Scholar]
  257. Van Sickle MD, Duncan M, Kingsley PJ, Mouihate A, Urbani P, Mackie K, Stella N, Makriyannis A, Piomelli D, Davison JS, Marnett LJ, Di Marzo V, Pittman QJ, Patel KD, Sharkey KA. (2005) Identification and functional characterization of brainstem cannabinoid CB2 receptors. Science 310:329–332. [DOI] [PubMed] [Google Scholar]
  258. Vermersch P. (2011) Sativex((R)) (tetrahydrocannabinol + cannabidiol), an endocannabinoid system modulator: basic features and main clinical data. Expert Rev Neurother 11:15–19. [DOI] [PubMed] [Google Scholar]
  259. Vierck CJ., Jr (2006) Mechanisms underlying development of spatially distributed chronic pain (fibromyalgia). Pain 124:242–263. [DOI] [PubMed] [Google Scholar]
  260. Vinod KY, Xie S, Psychoyos D, Hungund BL, Cooper TB, Tejani-Butt SM. (2012) Dysfunction in fatty acid amide hydrolase is associated with depressive-like behavior in Wistar Kyoto rats. PLoS One 7:e36743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  261. Walczak JS, Pichette V, Leblond F, Desbiens K, Beaulieu P. (2005) Behavioral, pharmacological and molecular characterization of the saphenous nerve partial ligation: a new model of neuropathic pain. Neuroscience 132:1093–1102. [DOI] [PubMed] [Google Scholar]
  262. Walker AK, Kavelaars A, Heijnen CJ, Dantzer R. (2014) Neuroinflammation and comorbidity of pain and depression. Pharmacol Rev 66:80–101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Wang J, Goffer Y, Xu D, Tukey DS, Shamir DB, Eberle SE, Zou AH, Blanck TJ, Ziff EB. (2011) A single subanesthetic dose of ketamine relieves depression-like behaviors induced by neuropathic pain in rats. Anesthesiology 115:812–821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Wang N, Shi M, Wang JY, Luo F. (2013) Brain-network mechanisms underlying the divergent effects of depression on spontaneous versus evoked pain in rats: a multiple single-unit study. Exp Neurol 250:165–175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Weber J, Schley M, Casutt M, Gerber H, Schuepfer G, Rukwied R, Schleinzer W, Ueberall M, Konrad C. (2009) Tetrahydrocannabinol (delta 9-THC) Treatment in chronic central neuropathic pain and fibromyalgia patients: results of a multicenter survey. Anesthesiol Res Pract 2009: Article ID 827290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  266. Wilkerson JL, Gentry KR, Dengler EC, Wallace JA, Kerwin AA, Armijo LM, Kuhn MN, Thakur GA, Makriyannis A, Milligan ED. (2012) Intrathecal cannabilactone CB(2)R agonist, AM1710, controls pathological pain and restores basal cytokine levels. Pain 153:1091–1106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  267. Wilson AR, Maher L, Morgan MM. (2008) Repeated cannabinoid injections into the rat periaqueductal gray enhance subsequent morphine antinociception. Neuropharmacology 55:1219–1225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  268. Wilson-Poe AR, Morgan MM, Aicher SA, Hegarty DM. (2012) Distribution of CB1 cannabinoid receptors and their relationship with mu-opioid receptors in the rat periaqueductal gray. Neuroscience 213:191–200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Wilson-Poe AR, Pocius E, Herschbach M, Morgan MM. (2013) The periaqueductal gray contributes to bidirectional enhancement of antinociception between morphine and cannabinoids. Pharmacol Biochem Behav 103:444–449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Wingenfeld K, Nutzinger D, Kauth J, Hellhammer DH, Lautenbacher S. (2010) Salivary cortisol release and hypothalamic pituitary adrenal axis feedback sensitivity in fibromyalgia is associated with depression but not with pain. J Pain 11:1195–1202. [DOI] [PubMed] [Google Scholar]
  271. Wittmann G, Deli L, Kallo I, Hrabovszky E, Watanabe M, Liposits Z, Fekete C. (2007) Distribution of type 1 cannabinoid receptor (CB1)-immunoreactive axons in the mouse hypothalamus. J Comp Neurol 503:270–279. [DOI] [PubMed] [Google Scholar]
  272. Woolridge E, Barton S, Samuel J, Osorio J, Dougherty A, Holdcroft A. (2005) Cannabis use in HIV for pain and other medical symptoms. J Pain Symptom Manage 29:358–367. [DOI] [PubMed] [Google Scholar]
  273. Yalcin I, Bohren Y, Waltisperger E, Sage-Ciocca D, Yin JC, Freund-Mercier MJ, Barrot M. (2011) A time-dependent history of mood disorders in a murine model of neuropathic pain. Biol Psychiatry 70:946–953. [DOI] [PubMed] [Google Scholar]
  274. Yalcin I, Barthas F, Barrot M. (2014) Emotional consequences of neuropathic pain: insight from preclinical studies. Neurosci Biobehav Rev 47:154–164. [DOI] [PubMed] [Google Scholar]
  275. Yasuda S, Yoshida M, Yamagata H, Iwanaga Y, Suenaga H, Ishikawa K, Nakano M, Okuyama S, Furukawa Y, Furukawa S, Ishikawa T. (2014) Imipramine ameliorates pain-related negative emotion via induction of brain-derived neurotrophic factor. Cell Mol Neurobiol 34:1199–1208. [DOI] [PubMed] [Google Scholar]
  276. Yu M, Ives D, Ramesha CS. (1997) Synthesis of prostaglandin E2 ethanolamide from anandamide by cyclooxygenase-2. J Biol Chem 272:21181–21186. [DOI] [PubMed] [Google Scholar]
  277. Zajkowska ZE, Englund A, Zunszain PA. (2014) Towards a personalized treatment in depression: endocannabinoids, inflammation and stress response. Pharmacogenomics 15:687–698. [DOI] [PubMed] [Google Scholar]
  278. Zavitsanou K, Wang H, Dalton VS, Nguyen V. (2010) Cannabinoid administration increases 5HT1A receptor binding and mRNA expression in the hippocampus of adult but not adolescent rats. Neuroscience 169:315–324. [DOI] [PubMed] [Google Scholar]
  279. Zeng Q, Wang S, Lim G, Yang L, Mao J, Sung B, Chang Y, Lim JA, Guo G. (2008) Exacerbated mechanical allodynia in rats with depression-like behavior. Brain Res 1200:27–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Zhang HY, Gao M, Liu QR, Bi GH, Li X, Yang HJ, Gardner EL, Wu J, Xi ZX. (2014) Cannabinoid CB2 receptors modulate midbrain dopamine neuronal activity and dopamine-related behavior in mice. Proc Natl Acad Sci U S A 111:E5007–5015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  281. Zhang J, Hoffert C, Vu HK, Groblewski T, Ahmad S, O’Donnell D. (2003) Induction of CB2 receptor expression in the rat spinal cord of neuropathic but not inflammatory chronic pain models. Eur J Neurosci 17:2750–2754. [DOI] [PubMed] [Google Scholar]
  282. Zhang M, Dai W, Liang J, Chen X, Hu Y, Chu B, Pan M, Dong Z, Yu S. (2013) Effects of UCMS-induced depression on nociceptive behaviors induced by electrical stimulation of the dura mater. Neurosci Lett 551:1–6. [DOI] [PubMed] [Google Scholar]
  283. Zhang Z, Wang W, Zhong P, Liu SJ, Long JZ, Zhao L, Gao HQ, Cravatt BF, Liu QS. (2015) Blockade of 2-arachidonoylglycerol hydrolysis produces antidepressant-like effects and enhances adult hippocampal neurogenesis and synaptic plasticity. Hippocampus 25:16–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Zhao X, Wang C, Zhang JF, Liu L, Liu AM, Ma Q, Zhou WH, Xu Y. (2014) Chronic curcumin treatment normalizes depression-like behaviors in mice with mononeuropathy: involvement of supraspinal serotonergic system and GABAA receptor. Psychopharmacology (Berl) 231:2171–2187. [DOI] [PubMed] [Google Scholar]
  285. Zhong P, Wang W, Pan B, Liu X, Zhang Z, Long JZ, Zhang HT, Cravatt BF, Liu QS. (2014) Monoacylglycerol lipase inhibition blocks chronic stress-induced depressive-like behaviors via activation of mTOR signaling. Neuropsychopharmacology 39:1763–1776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Zhou W, Dantzer R, Budac DP, Walker AK, Mao-Ying QL, Lee AW, Heijnen CJ, Kavelaars A. (2015) Peripheral indoleamine 2,3-dioxygenase 1 is required for comorbid depression-like behavior but does not contribute to neuropathic pain in mice. Brain Behav Immun 46:147–153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  287. Zoppi S, Madrigal JL, Caso JR, Garcia-Gutierrez MS, Manzanares J, Leza JC, Garcia-Bueno B. (2014) Regulatory role of the cannabinoid CB2 receptor in stress-induced neuroinflammation in mice. Br J Pharmacol 171:2814–2826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  288. Zoppi S, Perez Nievas BG, Madrigal JL, Manzanares J, Leza JC, Garcia-Bueno B. (2011) Regulatory role of cannabinoid receptor 1 in stress-induced excitotoxicity and neuroinflammation. Neuropsychopharmacology 36:805–818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  289. Zunszain PA, Hepgul N, Pariante CM. (2013) Inflammation and depression. Curr Top Behav Neurosci 14:135–151. [DOI] [PubMed] [Google Scholar]

Articles from International Journal of Neuropsychopharmacology are provided here courtesy of Oxford University Press

RESOURCES