Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-19T13:46:28.664Z Has data issue: false hasContentIssue false

Endocannabinoid system and pain: an introduction

Published online by Cambridge University Press:  22 October 2013

James J. Burston*
Affiliation:
Arthritis Research UK Pain Centre, School of Life Sciences, University of Nottingham Medical School, Queen's Medical Centre, Nottingham NG7 2UH, UK
Stephen G. Woodhams
Affiliation:
Laboratory of Molecular Neurobiology, Hungarian Academy of Sciences Institute of Experimental Medicine, Budapest 1083, Szigony Utca 43, Hungary
*
*Corresponding author: J. J. Burston, fax +44 (0)115 823 0142, email james.burston@nottingham.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

The endocannabinoid (EC) system consists of two main receptors: cannabinoid type 1 receptor cannabinoid receptors are found in both the central nervous system (CNS) and periphery, whereas the cannabinoid type 2 receptor cannabinoid receptor is found principally in the immune system and to a lesser extent in the CNS. The EC family consists of two classes of well characterised ligands; the N-acyl ethanolamines, such as N-arachidonoyl ethanolamide or anandamide (AEA), and the monoacylglycerols, such as 2-arachidonoyl glycerol. The various synthetic and catabolic pathways for these enzymes have been (with the exception of AEA synthesis) elucidated. To date, much work has examined the role of EC in nociceptive processing and the potential of targeting the EC system to produce analgesia. Cannabinoid receptors and ligands are found at almost every level of the pain pathway from peripheral sites, such as peripheral nerves and immune cells, to central integration sites such as the spinal cord, and higher brain regions such as the periaqueductal grey and the rostral ventrolateral medulla associated with descending control of pain. EC have been shown to induce analgesia in preclinical models of acute nociception and chronic pain states. The purpose of this review is to critically evaluate the evidence for the role of EC in the pain pathway and the therapeutic potential of EC to produce analgesia. We also review the present clinical work conducted with EC, and examine whether targeting the EC system might offer a novel target for analgesics, and also potentially disease-modifying interventions for pathophysiological pain states.

Type
Conference on ‘PUFA mediators: implications for human health’
Copyright
Copyright © The Authors 2013 

From an evolutionary standpoint, pain can be considered a necessary evil, providing a potent warning system to protect an individual from present and future harm. However, not all pain is part of this adaptive response, e.g. persistent pain after injury healing (chronic pain) or pain arising from damage to nerve tissue (neuropathic pain). Pain is the most common complaint in those seeking a physician, and a recent study suggests that pain represents the greatest economic burden of any pathological condition in the USA, with an estimated annual cost of $565–635 billion( Reference Gaskin and Richard 1 ). Many chronic pain states are refractory to standard analgesics, and even in those which do respond, pain control can be incomplete or only short-term in nature. Of the imperfect currently available analgesics, the most efficacious are the opioids which exploit an endogenous pain control pathway within the central nervous system. However, opioids have significant issues with tolerance, dependence, respiratory depression and opioid-induced hyperalgesia( Reference Angst and Clark 2 ). Over the past few decades, the existence of a second endogenous anti-nociceptive pathway has been revealed: the endocannabinoid (EC) system.

Cannabinoids and the endocannabinoid system

Extracts from Cannabis sativa have been used as analgesics for centuries, but paucity of knowledge of the molecular pharmacology, combined with moral reservations about the psychotropic effects of this drug and its abuse for recreational purposes, led to its prohibition in the early 20th century. However, the identification of Δ9-tetrahydrocannabinol as the major psychoactive component of cannabis extracts( Reference Mechoulam and Gaoni 3 ), and the subsequent isolation of a cannabinoid receptor highly expressed in nervous tissue( Reference Devane, Dysarz and Johnson 4 Reference Herkenham, Lynn and Johnson 6 ), led to an explosion of research interest in this area. Since Devane's landmark papers, firstly identifying the cannabinoid receptor( Reference Devane, Dysarz and Johnson 4 ), and subsequently anandamide (AEA) the first endogenous ligand( Reference Devane, Hanus and Breuer 7 ) over 5000 articles featuring EC have been published. The extent of the literature signifies the breadth of pivotal roles this system plays in physiological and pathophysiological functioning, and it is now known that in neurones these are fundamentally predicated on the modulation of neuronal signalling via retrograde inhibition (for reviews, see( Reference Kano, Ohno-Shosaku and Hashimotodani 8 , Reference Katona and Freund 9 )).

The EC system is composed of two G protein-coupled cannabinoid receptors (cannabinoid type 1 receptor (CB1) and cannabinoid type 2 receptor (CB2)), the EC ligands that activate them, and their synthetic and catabolic enzymes. The EC system possesses several unique properties when compared with other neurotransmitter systems, and it is these properties that underlie its role in analgesia: (i) EC act in a retrograde manner at neuronal synapses, being synthesised in the post-synaptic cell and travelling back across the synapse to interact with the receptors on the pre-synaptic cell( Reference Katona, Sperlagh and Sik 10 , Reference Egertova and Elphick 11 ); (ii) EC are not stored in vesicles prior to release, but instead are produced through activity-driven ‘on demand’ synthesis following strong neuronal activation( Reference Ohno-Shosaku, Maejima and Kano 12 , Reference Wilson and Nicoll 13 ); (iii) The EC system involves a multitude of ligands acting at just two major receptors, in contrast to the single ligand/multiple receptor paradigm present in other systems (e.g. glutamate, γ-aminobutyric acid, 5-hydroxytryptamine, etc.)( Reference Di Marzo and De Petrocellis 14 ).

EC are thought to act as a brake on neuronal hyper-activity, being produced in response to high levels of stimulation and feeding back negatively on the circuit through interaction with pre-synaptic cannabinoid receptors. In pain pathways, these actions produce analgesia by inhibiting the transmission of pain signals.

Cannabinoid receptors

The cannabinoid receptors have divergent expression patterns underlying their separate physiological roles. CB1 is predominantly found on nerve cells, while CB2 is mostly expressed on cells of the immune system, with some evidence of limited neuronal expression (see discussion and references in( Reference Rani Sagar, Burston and Woodhams 15 )).

The CB1 receptor is predominantly found pre-synaptically on axon terminals( Reference Katona, Sperlagh and Sik 10 ), and is coupled with adenylate cyclase via Gi/o proteins. Activation leads to a reduction of pre-synaptic neurotransmitter release via inhibition of N- and P/Q-type calcium channels, and activation of potassium channels( Reference Galiegue, Mary and Marchand 16 ). The net result of these actions can be inhibition or excitation of neuronal circuits, depending on whether the pre-synaptic cell secretes excitatory or inhibitory neurotransmitters.

Endocannabinoids

The EC are lipid signalling molecules, produced on-demand by the activity-dependent enzymatic cleavage of membrane phospholipids( Reference Sagar, Gaw and Okine 17 ). The most widely investigated are the arachidonic acid derivatives AEA and 2-arachidonoyl glycerol (2-AG; structures shown in Fig. 1). AEA was the first to be discovered, and is thus the most studied, although it is present at far lower levels in tissue than 2-AG (∼170-fold lower in brain( Reference Stella, Schweitzer and Piomelli 18 )). Both AEA and 2-AG activate cannabinoid receptors; 2-AG is a full agonist at both CB1 and CB2 receptors( Reference Sugiura, Kondo and Kishimoto 19 ), whereas AEA shows slight selectivity for CB1 over CB2. AEA has also been shown to activate the ion channel receptor TRPV1( Reference Zygmunt, Petersson and Andersson 20 ) and as such can be considered an EC/endovanilloid substance (for review, see( Reference Starowicz and Przewlocka 21 )).

Fig. 1. (Colour online) Endocannabinoid (EC) signalling at a notional neuronal synapse. The major synthetic, signalling and catabolic pathways for anandamide (AEA) and 2-arachidonoylglycerol (2-AG) are shown. Alongside cannabinoid receptors (CB1) and (CB2) the other G protein coupled receptors (GPCR) may be involved in cannabinoid signalling. FLAT-1, a truncated form of fatty acid amide hydrolase (FAAH), and ECT are the putative EC transporters. MAGL, monoacylglycerol lipase; DAGL, diacyglycerol lipase; GPR, specific G protein-coupled receptor such as GPR55. Compounds in red are recognised enzyme inhibitors/receptor antagonists, which can modulate EC signals.

Endocannabinoids and nutrition

The EC are lipid-based, and as such are more susceptible to dietary-induced fluctuations in tissue levels than other transmitter substances( Reference Artmann, Petersen and Hellgren 22 , Reference Watanabe, Doshi and Hamazaki 23 ). Elevations in EC levels in epididymal fat and pancreas have been reported in a mouse model of diet-induced obesity( Reference Matias, Gonthier and Orlando 24 ), and these diet-induced elevations may alter the EC system functionally, since female mice on a high-fat diet show decreased sensitivity to Δ9-tetrahydrocannabinol( Reference Wiley, Jones and Wright 25 ). Recent clinical data suggests dietary modulation of EC also occurs in human subjects( Reference Berge, Piscitelli and Hoem 26 ), with an 84 % decrease in plasma AEA observed following chronic exposure to dietary DHA and EPA. Although the functional relevance of circulating EC levels remains unclear, this robust effect of fatty acid intake suggests that nutritional factors can influence the EC system, and may offer a means to modulate the analgesic potential discussed here.

Endocannabinoid synthesis and degradation

The major EC have distinct synthetic and degradative pathways, with the localisation of the synthetic and degradative machinery for each ligand determining its physiological effects (for review, see( Reference Luchicchi and Pistis 27 )). A schematic diagram of the actions of EC at a generic neuronal synapse can be seen in Fig. 1.

AEA was initially thought to be synthesised by a pathway involving the enzyme N-acyl-phosphatidylethanolamine-hydrolysing phospholipase D( Reference Di Marzo, Fontana and Cadas 28 , Reference Okamoto, Morishita and Tsuboi 29 ). However, generation of an N-acylphosphatidylethanolamine-hydrolysing phospholipase D knock-out mouse revealed normal brain levels of AEA( Reference Leung, Saghatelian and Simon 30 ), and additional synthetic pathways have since been identified( Reference Simon and Cravatt 31 , Reference Simon and Cravatt 32 ). It remains unclear which pathway predominates in the production of AEA in pain pathways. The major catabolic enzyme for AEA is fatty acid amide hydrolase (FAAH), although other catabolic pathways have been identified (for review see( Reference Ueda, Tsuboi and Uyama 33 )).

2-AG is produced from cleavage of membrane phospholipids by phospholipase C-β to produce diacylglycerol species, and subsequent cleavage to form 2-AG by the enzymes diacyglycerol lipase-α and β( Reference Bisogno, Howell and Williams 34 ), and metabolism is primarily via monoacylglycerol lipase (MAGL)( Reference Dinh, Carpenter and Leslie 35 ). Other enzymes participating in 2-AG metabolism have been identified( Reference Blankman, Simon and Cravatt 36 ), although their roles in termination of 2-AG signalling have yet to be fully elucidated.

Pain and nociception

Pain is an integrative experience, involving physiological, emotional and cognitive aspects (for a useful glossary of pain terminology, see( Reference Loeser and Treede 37 )). The subjective experience of pain varies significantly between individuals and cannot be reported by the non-human animals on which the majority of basic pain research is conducted. For the purposes of this review, we differentiate between the subjective experience of pain, and the measurable neuronal events which underlie it, hereafter referred to as nociception.

Nociceptive pathways begin with the transduction of a noxious stimulus, such as mechanical pressure, into action potentials by a specialised class of sensory afferent neurones in the periphery (e.g. mechanoreceptors in the skin). Action potentials travel via the axon of the primary afferent neurone, past the cell body located in a dorsal root ganglion, to a synapse in the superficial dorsal horn of the spinal cord. Following the integration of inputs from multiple cells types within the spinal cord, these action potentials will then pass up one of several ascending pathways to the brainstem, and subsequently to the thalamus, which then relays the signal to higher brain regions involved in the sensory (e.g. the somatosensory cortex) and emotional/affective (e.g. the amygdala and cingulate cortex) aspects of pain. There is significant cross-talk between supra-spinal nociceptive regions, and nociceptive signals can be amplified or dampened by descending modulatory pathways projecting from the brain to the spinal cord (pathways reviewed in( Reference Basbaum, Bautista and Scherrer 38 Reference Todd 40 )). Fig. 2 displays a schematic of a typical nociceptive pathway.

Fig. 2. (Colour online) Nociceptive pathways. Schematic of nociceptive pathways. Nociceptive stimuli are conducted from the periphery to the dorsal horn of the spinal cord, and then transmitted to the supra-spinal regions via the spinothalamic tract (STT, blue) and spinoparabrachial tract (SPBT, red). The major descending modulatory control pathway (DMCP, purple) is displayed on the right. This pathway crosses the midline at the level of the medulla. Coloured areas indicate the position of synapses in each pathway. The positions of laminae I–VI in the dorsal horn are indicated by dotted lines, while the black region in the brain represents the lateral ventricles. Thal., thalamus; VMH, ventromedial hypothalamus; PbN, parabrachial nucleus; PAG, periaqueductal grey matter; RVM, rostroventral medulla; Pyr., pyramidal tract; DRG, dorsal root ganglion. Adapted from( Reference Hunt and Mantyh 39 ).

Location of the endocannabinoid system in the pain pathway

The components of the EC system are expressed ubiquitously throughout the pain processing pathways, underlining its key modulatory role in nociception. Both ligands and receptors can be detected in the periphery, at the level of the spinal cord, and in nociceptive regions of the brain( Reference Herkenham, Lynn and Johnson 6 , Reference Sagar, Kendall and Chapman 41 Reference Calignano, La Rana and Giuffrida 71 ). CB1 is predominantly localised in neurones, while CB2 is found in immune cells, although there is some evidence for non-neuronal CB1 in B cells of the immune system( Reference Kaplan 47 Reference Pacher and Mechoulam 49 ) and astroglial cells of the central nervous system( Reference Salio, Doly and Fischer 59 , Reference Alkaitis, Solorzano and Landry 64 , Reference Stella 68 ). CB2 expression has been reported in microglia of the central nervous system( Reference Racz, Nadal and Alferink 62 Reference Alkaitis, Solorzano and Landry 64 ), with some evidence suggesting neuronal CB2 expression (see commentary in( Reference Atwood and Mackie 69 )), although this remains controversial.

Peripheral mechanisms

The peripheral compartment of the EC system plays a substantial role in cannabinoid-receptor-mediated anti-nociception, as demonstrated by the greatly reduced efficacy of locally and systemically administered cannabinoids following selective deletion of peripheral CB1 ( Reference Agarwal, Pacher and Tegeder 70 ). Evidence indicates that both neuronal and non-neuronal cannabinoid receptors contribute to the anti-nociceptive effects of peripheral EC, and the contributions of these and the ligands AEA, and 2-AG have been partially revealed via use of specific enzyme inhibitors and receptor antagonists in animal models of pain.

The most studied paradigm is that of inflammatory pain, in which application of an inflammatory substance to the rodent hindpaw elicits an oedemic response and measureable nociceptive behaviour. Peripheral administration of AEA in the formalin model temporarily reduced the nociceptive behaviour in a CB1-sensitive manner( Reference Calignano, La Rana and Giuffrida 71 ). Conversely, blocking CB1 and/or CB2 receptors prior to formalin administration increased nociceptive responses, suggesting an intrinsic role for EC in inflammatory pain. This may be restricted to early inflammatory pain states, since the hindpaw levels of EC are decreased at later time-points (see reviews in( Reference Sagar, Gaw and Okine 17 , Reference Maione, De Petrocellis and de Novellis 72 , Reference Beaulieu, Bisogno and Punwar 73 )). Similarly, intra-plantar administration of 2-AG blocked the second phase of formalin-evoked pain behaviour in rats( Reference Guindon, Desroches and Beaulieu 74 ), via a CB2-mediated mechanism.

Exogenously administered AEA and 2-AG are rapidly metabolised by FAAH and MAGL, respectively. Further studies have therefore focused on the use of FAAH and MAGL inhibitors to prolong the effects of endogenous EC actions. Systemic inactivation of FAAH via compounds such as URB597, OL135 and PF-3845 has been shown to be anti-nociceptive in models of acute and inflammatory pain( Reference Jayamanne, Greenwood and Mitchell 75 Reference Russo, Loverme and La Rana 80 ). Elevations in both AEA and 2-AG have been shown, as well as reduced carrageenan-induced hyperalgesia( Reference Jhaveri, Richardson and Robinson 81 ) and expansion of peripheral receptive field size of wide dynamic range neurons (a marker of central sensitisation) following intra-plantar URB597( Reference Sagar, Kendall and Chapman 82 ). Similarly, capsaicin-induced pain behaviour and thermal hypersensitivity were attenuated following blockade of peripheral MAGL via JZL184( Reference Spradley, Guindon and Hohmann 83 ), and peripheral administration of JZL184 produces local inhibition of MAGL activity, increased levels of 2-AG and anti-nociceptive effects in both phases of the formalin model through mechanisms involving both CB1 and CB2 ( Reference Guindon, Guijarro and Piomelli 84 ). A peripherally restricted FAAH inhibitor URB937( Reference Clapper, Moreno-Sanz and Russo 85 ), which cannot cross the blood–brain barrier, blocked the hypersensitivity induced by both inflammation and peripheral nerve injury via a CB1-mediated mechanism, without altering thermal nociceptive thresholds. These data reinforce the notion that peripheral EC tone is increased in pain states, and that actions of EC at peripheral CB1 receptors can reduce the transmission of nociceptive information. Further enhancing EC signalling may therefore prove an effective analgesic strategy, although it should be noted that FAAH and MAGL are not the only enzymes that can degrade EC( Reference Ueda, Tsuboi and Uyama 33 , Reference Blankman, Simon and Cravatt 36 ), and thus, the effects of specific enzyme inhibitors may not present the entire story.

Alongside direct anti-nociceptive effects at CB1 receptors on afferent neurones, EC can also act at peripheral CB2 on immune cells such as macrophages( Reference Han, Lim and Ryu 86 ), lymphocytes( Reference Cencioni, Chiurchiăą and Catanzaro 87 ) and mast cells( Reference Samson, Small-Howard and Shimoda 88 ). CB2 receptor activation inhibits the production and release of pro-inflammatory and pro-nociceptive mediators, such as reactive oxygen species( Reference Hao, Jiang and Fang 89 ) and cytokines( Reference Cencioni, Chiurchiăą and Catanzaro 87 ). In addition, metabolism of 2-AG produces arachidonic acid, a key precursor of pro-inflammatory PG, and disruption of 2-AG hydrolysis reduces the available pool of arachidonic acid and thus reduces inflammation( Reference Nomura, Morrison and Blankman 90 ). In summary, elevating levels of peripheral AEA produces anti-nociceptive effects in models of inflammatory and neuropathic pain, largely via the inhibitory actions of CB1 receptors on the primary neurones which transmit nociceptive signals. Similarly, increased 2-AG signalling at peripheral CB1 receptors is also anti-nociceptive, but there also appears to be a prominent CB2-mediated component, probably involving the inhibition of pro-nociceptive actions of immune cells.

Spinal mechanisms

Exogenous application of EC is anti-nociceptive at the level of the spinal cord( Reference Starowicz, Makuch and Osikowicz 91 , Reference Welch, Huffman and Lowe 92 ), while intra-thecal administration of a CB1 receptor antagonist produces hyperalgesia in mice( Reference Richardson, Aanonsen and Hargreaves 93 ), enhancing nociception-evoked firing of wide dynamic range neurones in the dorsal horn of the spinal cord( Reference Chapman 94 ). In addition, spinal EC are elevated in animal models of acute and chronic pain( Reference Sagar, Jhaveri and Richardson 95 Reference Okine, Norris and Woodhams 97 ). These data indicate a role for the spinal EC system in nociceptive transmission. In the surgical incision model of acute resolving pain in rats, spinal levels of AEA are reduced at early time-points coinciding with maximal mechanical hypersensitivity, returning to baseline as nociceptive behaviour subsides( Reference Alkaitis, Solorzano and Landry 64 ). In comparison, 2-AG levels were elevated at time-points coinciding with the appearance of glial cell activation and up-regulation of CB2 receptors, suggesting a temporal segregation of AEA and 2-AG signalling. In agreement with these data, spinal levels of AEA are significantly elevated at early time-points in the chronic constriction injury (neuropathic pain model) model of neuropathic pain in mice( Reference Petrosino, Palazzo and de Novellis 98 ). Spinal administration of URB597 reduced mechanically evoked responses of wide dynamic range neurones in rats that underwent spinal nerve ligation and this effect was blocked by pre-administration of a CB1 selective receptor inverse agonist/antagonist( Reference Jhaveri, Richardson and Kendall 99 ). Similarly, we have shown that spinal application of the MAGL selective inhibitor JZL184 produced a dose-related reduction in mechanically-evoked nociceptive neurotransmission in the spinal cord of naive rats, which was reversed in the presence of the CB1 selective antagonist, a CB1 selective receptor inverse agonist/antagonist( Reference Woodhams, Wong and Barrett 100 ). Spinally administered JZL184 was also able to prevent intra-plantar carrageenan-induced receptive field expansion of dorsal horn wide dynamic ranges, indicating that the inhibition of MAGL can block mechanisms underlying the development of central sensitisation following peripheral inflammation.

Further evidence of the involvement of the spinal EC system is provided by alterations seen in receptor expression in established pain states. CB1 expression is elevated in the spinal cord of neuropathic rats from 4 d post-injury, with levels continuing to rise until day 14( Reference Lim, Sung and Ji 101 ) while CB2 receptor up-regulation also occurs by day 4( Reference Zhang, Hoffert and Vu 102 ). Genetic deletion of CB2 receptors results in enhanced microglia and astrocytic expression in the contralateral spinal horn following nerve injury, accompanied by profound contralateral mechanical and thermal allodynia( Reference Racz, Nadal and Alferink 62 ). Conversely, overexpression of CB2 receptors protected against nerve injury-induced thermal and mechanical allodynia and prevented glial activation in the spinal cord. Numerous other studies have also implicated EC signalling in glial cell activation( Reference Alkaitis, Solorzano and Landry 64 , Reference Correa, Docagne and Mestre 103 , Reference Stella 104 ).

Exogenous 2-AG has been shown to stimulate microglial migration, whereas the CB2 receptor antagonist a CB2 selective receptor antagonist inhibits basal microglial migration( Reference Walter, Franklin and Witting 105 ). However, caution should be taken when interpreting these results as cell culture models might not reflect the in vivo situation. Nevertheless, these converging lines of research strongly suggest that the EC system, especially at the level of the spinal cord, is intimately involved in glial cell signalling. Further information regarding the links between the EC system and neuro-glial interactions can be found in the following review( Reference Bilkei-Gorzo 106 ). The analgesic actions of EC and cannabinoid receptor activation in the spinal cord suggest that targeting the EC system at this level could inhibit both neuronal hyper-excitability and glial cell activation. Thus, enhancing this endogenous pathway could prove to have a wide range of therapeutic applications in the treatment of multiple pain sates, including the underlying central sensitisation.

Supra-spinal mechanisms

The supra-spinal sites of cannabinoid anti-nociceptive action were first identified in rodents via microinjection of CB1 ligands into pain-associated regions including the rostroventral medulla, the dorsal raphe nucleus, the periaqueductal grey matter and the amygdala, prior to tests of acute nociception( Reference Martin, Patrick and Coffin 107 ). Later work revealed the role of endogenous ligands by demonstrating mobilisation of AEA( Reference Walker, Huang and Strangman 108 ), and CB1-mediated anti-nociception, following either electrical stimulation of these regions or a peripheral administration of formalin( Reference Rea, Roche and Finn 109 ). Enhancing AEA signalling in these areas via inhibition of FAAH activity is anti-nociceptive in acute pain( Reference Maione, Bisogno and de Novellis 110 ), probably via disinhibition of descending inhibitory inputs from the brainstem to the spinal cord, inhibiting spinal nociceptive signalling (for a review of the pathways involved, see( Reference Basbaum, Bautista and Scherrer 38 Reference Todd 40 )).

Supra-spinal EC are also responsible for a phenomenon known as stress-induced analgesia, in which brief exposure to environmental stress (e.g. immersion in cold water, or an electric shock to the paw) produces reduced nociceptive responses in a subsequent pain test. Detailed study of this effect revealed the mobilisation of both AEA and 2-AG in the periaqueductal grey matter( Reference Hohmann, Suplita and Bolton 111 ), and suggested that 2-AG acting at CB1 receptors was the predominant mechanism. Additional work indicated that further enhancing stress-induced EC signalling via FAAH or MAGL inhibition produces still greater anti-nociceptive effects( Reference Suplita, Farthing and Gutierrez 112 , Reference Suplita, Gutierrez and Fegley 113 ), and confirmed the pivotal role of 2-AG signalling in the periaqueductal grey matter( Reference Gregg, Jung and Spradley 114 ). The involvement of stress in human pain responses and the presence of an EC-mediated mechanism are now being studied clinically( Reference Vachon-Presseau, Martel and Roy 115 ). These data indicate a physiological role for supra-spinal EC in acute nociception, although agonism of supra-spinal CB1 receptors is an unappealing prospect due to the psychotropic side-effects. Instead, research has focused on modulating the altered EC signalling seen in pain states.

The EC system is plastic, with changes in levels of receptor expression, ligand concentrations and synthetic and catabolic enzymatic activity occurring in pain states (reviewed in( Reference Rani Sagar, Burston and Woodhams 116 )). Elevated levels of EC in the periaqueductal grey matter and dorsal raphe nucleus have been observed at 3 and 7 d in the chronic constriction injury model of neuropathic pain( Reference Petrosino, Palazzo and de Novellis 98 ). Interestingly, desensitisation of CB1 receptors in a pain-related cortical brain region has been described in this model at 10 d, when nociceptive behaviour is maximal( Reference Hoot, Sim-Selley and Poklis 117 ). CB1 receptor desensitisation is known to occur following chronic exposure to ligands( Reference Sim-Selley 118 ), and thus may reflect the result of chronically elevated EC levels in the brain. It remains to be seen whether pharmacologically enhancing EC signalling in pain-related brain regions is advantageous or detrimental under these conditions.

Supra-spinal EC can also modulate the affective (or emotional) aspects of pain, which are mediated by frontal and limbic brain regions, and can be dissected from the somatosensory aspects. A recent neuroimaging study of the analgesic effects of Δ9-tetrahydrocannabinol on capsaicin-evoked cutaneous pain in human subjects revealed no change in the intensity of pain sensation( Reference Lee, Ploner and Wiech 119 ). Instead, test subjects reported that Δ9-tetrahydrocannabinol reduced the unpleasantness of capsaicin-evoked cutaneous ongoing pain, rather than reduced sensation, concurrent with reduced activity in the anterior cingulate cortex and enhanced activity in the right amygdala compared with the control subjects. Whether modulation of the EC system can mimic these effects is as yet unknown. Ablation of the anterior cingulate cortex has been utilised as an effective last-resort treatment for intractable pain( Reference Foltz and White 120 ), and thus unlocking the potential of the EC system in this region could address the great unmet clinical need in chronic pain states.

Novel metabolic pathways

Modulating EC levels by FAAH and MAGL inhibition, as described earlier, can produce anti-nociceptive effects. However, bioactive lipids such as AEA and 2-AG are promiscuous, and can be metabolised by multiple enzymes( Reference Alexander and Kendall 121 ). Artificially elevating EC levels can unmask alternative metabolic routes, producing additional bioactive products. Interestingly, pathological conditions such as chronic pain states are associated with changes in levels of enzymes, such as cyclooxygenase (COX), lipoxygenase, αβ hydrolase and members of the cytochrome P450 family( Reference Fowler 122 Reference Snider, Nast and Tesmer 125 ), which can metabolise the EC to novel lipid signalling molecules. The physiological actions of these metabolic products are as yet unknown, but some preliminary investigations have been performed. The cytochrome P450 metabolite of AEA, 5,6-epoxyeicosatrienoic acid ethanolamide, has been shown to be a potent CB2 receptor ligand( Reference Snider, Nast and Tesmer 125 ), whereas the 15-lipoxygenase metabolite of 2-AG, 15-hydroxyeicosatetraenoic acid glyceryl ester, acts as a PPARα agonist( Reference Kozak, Crews and Morrow 123 ). COX-2 metabolites of AEA and 2-AG have been shown to have pro-nociceptive actions in the spinal cord. COX-2 metabolises AEA to prostamide F2α, whose spinal application increases the firing of nociceptive neurons and reduces paw withdrawal latencies, and levels of prostamide F2α are elevated in spinal cord tissue in mice with knee inflammation( Reference Gatta, Piscitelli and Giordano 124 ). Similarly, the COX-2 metabolite of 2-AG, PGE2 glycerol ester, is endogenously generated in rat tissue, and induces mechanical allodynia and thermal hyperalgesia following intraplantar administration( Reference Hu, Bradshaw and Chen 126 ). Based on these reports, it is clear that determining the levels of these potential ligands in pain states is of great interest, as many of these metabolites may have effects on pain processing. These findings may also limit the utility of FAAH and MAGL inhibitors as therapeutics in chronic pain states, as they may increase substrate levels for generation of alterative pro-nociceptive EC metabolites, and thus counteract the anti-nociceptive effects of AEA and 2-AG.

Cannabinoids and endocannabinoids in clinical trials

Despite the growing use of medicinal marijuana, and the development of licensed cannabinoid drugs such as Sativex for multiple sclerosis( Reference Garcia-Merino 127 ), concerns remain overdependence, tolerance and the cognitive side-effects produced by these medications. Despite the wealth of pre-clinical data on alternative EC-mediated compounds, the only major clinical trial conducted utilising an EC-directed compound looked at the ability of the selective FAAH inhibitor PF-04457845 to produce analgesia in an osteoarthritic patient population( Reference Huggins, Smart and Langman 128 ). Despite significant elevations in plasma AEA, no analgesic effect was observed. Although disappointing, perhaps, the negative outcome of this trial may indicate a limitation of elevating AEA to induce analgesia in pain sates. Previous work has shown that, in addition to being a CB1 receptor agonist, at higher concentrations AEA binds to and activates the pro-nociceptive TRPV1 channel( Reference Ross 129 ). Since AEA can be also converted into pro-nociceptive signalling molecules in the presence of COX-2 activity, then it is feasible that under pathological pain states where COX-2 activity is up-regulated (e.g. osteoarthritis), pro-nociceptive effects of AEA at TRPV1 may outweigh the anti-nociceptive actions. It has previously been suggested that the development of dual FAAH and COX-2 inhibitors( Reference Fowler, Bjorklund and Lichtman 130 ) or substrate-selective inhibitors of COX-2( Reference Windsor, Hermanson and Kingsley 131 ) would be advantageous in terms of uncoupling the pro- and anti-nociceptive actions of AEA, and producing compounds with superior analgesic profiles. An alternative approach that has been explored pre-clinically utilises N-arachidonoyl-serotonin, a dual FAAH inhibitor/TRPV1 antagonist. This compound has enhanced anti-nociceptive effects v. a FAAH inhibitor alone, in several models of pain( Reference Maione, De Petrocellis and de Novellis 72 , Reference Costa, Bettoni and Petrosino 132 , Reference de Novellis, Palazzo and Rossi 133 ). In addition, changes in AEA biotransformation in the aged patient population may contribute to the lack of analgesia following FAAH inhibition. A recent report describes greater susceptibility to chronic pain and decreased AEA-mediated anti-nociceptive effects in aged animals( Reference Bishay, Haussler and Lim 134 ).

An approach with potentially greater therapeutic appeal than FAAH inhibition involves the targeting of 2-AG signalling. The recent development of MAGL inhibitors such as JZL184 and KML 29( Reference Long, Li and Booker 135 , Reference Chang Jae, Niphakis Micah and Lum Kenneth 136 ), and subsequent preclinical studies suggest that low doses of MAGL inhibitors are devoid of analgesic tolerance( Reference Kinsey, Wise and Ramesh 137 ), and also decrease arachidonic acid pools that are required for the generation of pro-nociceptive molecules such as PGE2( Reference Nomura, Morrison and Blankman 90 ). This approach thus delivers a dual analgesic mechanism, elevating 2-AG and reducing pro-inflammatory PG levels. This may have particular utility in the treatment of inflammatory bowel diseases and the associated pain, especially as MAGL inhibitors have already been shown to attenuate non-steroidal anti-inflammatory drug-induced gastric haemorrhages in mice( Reference Kinsey, Nomura and O'Neal 138 ). Inflammatory bowel diseases, such as irritable bowel syndrome, result in significant pain and distress. Recently, several converging lines of evidence suggest that targeting the EC system may provide much sought after disease modifying therapeutics for these conditions( Reference Alhouayek, Lambert and Delzenne 139 ). At the present stage, little clinical work has been conducted to evaluate the efficacy of FAAH and/or MAGL inhibition in treating inflammatory bowel diseases and associated pain. Given the large unmet clinical need in this area, the disease-modifying potential of an EC therapy with a dual mechanism involving both CB1- and CB2-mediated analgesia, as well as a reduction of pathologically elevated levels of pro-inflammatory mediators, is very appealing.

Concluding remarks

Owing to the breadth and depth of the literature cited here, we have only presented a fraction of the excellent studies conducted in this area, though we hope this information is sufficient to demonstrate the significant potential of targeting the EC system to analgesic effect. It is our belief that such an approach can produce novel efficacious analgesic agents required to help fill the unmet clinical need in chronic pain states. However, to reach this goal, the current gap between the wealth of pre-clinical data and the paucity of clinical trials must be bridged. Translation from the laboratory to the clinic is fraught with difficulties, as evidenced by the failure of the FAAH inhibitor PF-04457845 in osteoarthritic patients, despite a seemingly well-designed study. Future attempts in this area should perhaps utilise patient stratification based on aspects of disease aetiology or stratification based on neuroimaging of the pain matrix in human subjects. Clinical work is also clearly needed that focuses on whether targeting the EC system in highly inflammatory conditions such as irritable bowel syndrome, may offer new analgesic treatments.

It remains to be seen whether medications producing chronic elevations of EC will suffer from similar side-effects to those seen with phytocannabinoids, or the issues of tolerance highlighted in recent animal studies. However, the recent success of Sativex clearly highlights the potential of this area of research.

Acknowledgements

We would like to thank Dr Devi Sagar, Professor Dave Kendall and Professor Victoria Chapman for helpful discussions and advice during the preparation of this work.

Financial support

J. J. B. is funded by Arthritis Research UK Pain Centre funding (grant no. 18769).

Conflicts of interest

None.

Authorship

Both the authors contributed equally to the work.

References

1. Gaskin, DJ & Richard, P (2012) The economic costs of pain in the united states. J Pain 13, 715724.CrossRefGoogle ScholarPubMed
2. Angst, MS & Clark, JD (2006) Opioid-induced hyperalgesia: a qualitative systematic review. Anesthesiology 104, 570587.CrossRefGoogle ScholarPubMed
3. Mechoulam, R & Gaoni, Y (1967) Absolute configuration of delta1-tetrahydrocannabinol major active constituent of hashish. Tetrahedr Lett 8, 1109.CrossRefGoogle Scholar
4. Devane, WA, Dysarz, FA, Johnson, MR et al. (1988) Determination and characterization of a cannabinoid receptor in rat-brain. Mol Pharmacol 34, 605613.Google ScholarPubMed
5. Herkenham, M, Lynn, AB, Little, MD et al. (1990) Cannabinoid receptor localization in brain. Proc Natl Acad Sci USA 87, 19321936.CrossRefGoogle ScholarPubMed
6. Herkenham, M, Lynn, AB, Johnson, MR et al. (1991) Characterization and localization of cannabinoid receptors in rat-brain - a quantitative in vitro autoradiographic study. J Neurosci 11, 563583.CrossRefGoogle ScholarPubMed
7. Devane, WA, Hanus, L, Breuer, A et al. (1992) Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science 258, 19461949.CrossRefGoogle Scholar
8. Kano, M, Ohno-Shosaku, T, Hashimotodani, Y et al. (2009) Endocannabinoid-mediated control of synaptic transmission. Physiol Rev 89, 309380.CrossRefGoogle ScholarPubMed
9. Katona, I & Freund, T (2012) Multiple functions of endocannabinoid signaling in the brain. Annu Rev Neurosci 35, 529558.CrossRefGoogle ScholarPubMed
10. Katona, I, Sperlagh, Bt, Sik, A et al. (1999) Presynaptically located CB1 cannabinoid receptors regulate GABA release from axon terminals of specific hippocampal interneurons. J Neurosci 19, 45444558.CrossRefGoogle ScholarPubMed
11. Egertova, M & Elphick, MR (2000) Localisation of cannabinoid receptors in the rat brain using antibodies to the intracellular C-terminal tail of CB1 . J Comp Neurol 422, 159171.3.0.CO;2-1>CrossRefGoogle Scholar
12. Ohno-Shosaku, T, Maejima, T & Kano, M (2001) Endogenous cannabinoids mediate retrograde signals from depolarized postsynaptic neurons to presynaptic terminals. Neuron 29, 729738.CrossRefGoogle ScholarPubMed
13. Wilson, RI & Nicoll, RA (2001) Endogenous cannabinoids mediate retrograde signalling at hippocampal synapses. Nature 410, 588592.CrossRefGoogle ScholarPubMed
14. Di Marzo, V & De Petrocellis, L (2012) Why do cannabinoid receptors have more than one endogenous ligand? Phil Trans R Soc Lond B Biol Sci 367, 32163228.CrossRefGoogle ScholarPubMed
15. Rani Sagar, D, Burston, JJ, Woodhams, SG et al. (2012) Dynamic changes to the endocannabinoid system in models of chronic pain. Phil Trans R Soc B Biol Sci 367, 33003311.CrossRefGoogle Scholar
16. Galiegue, S, Mary, S, Marchand, J et al. (1995) Expression of central and peripheral cannabinoid receptors in human immune tissues and leukocyte subpopulations. Eur J Biochem 232, 5461.CrossRefGoogle ScholarPubMed
17. Sagar, DR, Gaw, AG, Okine, BN et al. (2009) Dynamic regulation of the endocannabinoid system: implications for analgesia. Mol Pain 5, 59.CrossRefGoogle ScholarPubMed
18. Stella, N, Schweitzer, P & Piomelli, D (1997) A second endogenous cannabinoid that modulates long-term potentiation. Nature 388, 773778.CrossRefGoogle ScholarPubMed
19. Sugiura, T, Kondo, S, Kishimoto, S et al. (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, 605612.CrossRefGoogle ScholarPubMed
20. Zygmunt, PM, Petersson, J, Andersson, DA et al. (1999) Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature 400, 452457.CrossRefGoogle ScholarPubMed
21. Starowicz, K & Przewlocka, B (2012) Modulation of neuropathic-pain-related behaviour by the spinal endocannabinoid/endovanilloid system. Phil Trans R Soc B Biol Sci 367, 32863299.CrossRefGoogle ScholarPubMed
22. Artmann, A, Petersen, G, Hellgren, LI et al. (2008) Influence of dietary fatty acids on endocannabinoid and N-acylethanolamine levels in rat brain, liver and small intestine. Biochim Biophys Acta 1781, 200212.CrossRefGoogle ScholarPubMed
23. Watanabe, S, Doshi, M & Hamazaki, T (2003) n-3 Polyunsaturated fatty acid (PUFA) deficiency elevates and n-3 PUFA enrichment reduces brain 2-arachidonoylglycerol level in mice. Prostaglandins Leukot Essent Fatty Acids 69, 5159.CrossRefGoogle ScholarPubMed
24. Matias, I, Gonthier, MP, Orlando, P et al. (2006) Regulation, function, and dysregulation of endocannabinoids in models of adipose and beta-pancreatic cells and in obesity and hyperglycemia. J Clin Endocrinol Metab 91, 31713180.CrossRefGoogle ScholarPubMed
25. Wiley, JL, Jones, AR & Wright, MJ Jr (2011) Exposure to a high-fat diet decreases sensitivity to Delta9-tetrahydrocannabinol-induced motor effects in female rats. Neuropharmacology 60, 274283.CrossRefGoogle ScholarPubMed
26. Berge, K, Piscitelli, F, Hoem, N et al. (2013) Chronic treatment with krill powder reduces plasma triglyceride and anandamide levels in mildly obese men. Lipids Health Dis 12, 78.CrossRefGoogle ScholarPubMed
27. Luchicchi, A & Pistis, M (2012) Anandamide and 2-arachidonoylglycerol: pharmacological properties, functional features, and emerging specificities of the two major endocannabinoids. Mol Neurobiol 46, 374392.CrossRefGoogle ScholarPubMed
28. Di Marzo, V, Fontana, A, Cadas, H et al. (1994) Formation and inactivation of endogenous cannabinoid anandamide in central neurons. Nature 372, 686691.CrossRefGoogle ScholarPubMed
29. Okamoto, Y, Morishita, J, Tsuboi, K et al. (2004) Molecular characterization of a phospholipase D generating anandamide and its congeners. J Biol Chem 279, 52985305.CrossRefGoogle ScholarPubMed
30. Leung, D, Saghatelian, A, Simon, GM et al. (2006) Inactivation of N-acyl phosphatidylethanolamine phospholipase D reveals multiple mechanisms for the biosynthesis of endocannabinoids. Biochemistry 45, 47204726.CrossRefGoogle ScholarPubMed
31. Simon, GM & Cravatt, BF (2006) Endocannabinoid biosynthesis proceeding through glycerophospho-N-acyl ethanolamine and a role for alpha/beta-hydrolase 4 in this pathway. J Biol Chem 281, 2646526472.CrossRefGoogle Scholar
32. Simon, GM & Cravatt, BF (2008) Anandamide biosynthesis catalyzed by the phosphodiesterase GDE1 and detection of glycerophospho-N-acyl ethanolamine precursors in mouse brain. J Biol Chem 283, 93419349.CrossRefGoogle ScholarPubMed
33. Ueda, N, Tsuboi, K & Uyama, T (2010) N-acylethanolamine metabolism with special reference to N-acylethanolamine-hydrolyzing acid amidase (NAAA). Prog Lipid Res 49, 299315.CrossRefGoogle ScholarPubMed
34. Bisogno, T, Howell, F, Williams, G et al. (2003) Cloning of the first sn1-DAG lipases points to the spatial and temporal regulation of endocannabinoid signaling in the brain. J Cell Biol 163, 463468.CrossRefGoogle Scholar
35. Dinh, TP, Carpenter, D, Leslie, FM et al. (2002) Brain monoglyceride lipase participating in endocannabinoid inactivation. Proc Natl Acad Sci USA 99, 1081910824.CrossRefGoogle ScholarPubMed
36. Blankman, JL, Simon, GM & Cravatt, BF (2007) A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chem Biol 14, 13471356.CrossRefGoogle ScholarPubMed
37. Loeser, JD & Treede, RD (2008) The Kyoto protocol of IASP basic pain terminology. Pain 137, 473477.CrossRefGoogle ScholarPubMed
38. Basbaum, AI, Bautista, DM, Scherrer, G et al. (2009) Cellular and molecular mechanisms of pain. Cell 139, 267284.CrossRefGoogle ScholarPubMed
39. Hunt, SP & Mantyh, PW (2001) The molecular dynamics of pain control. Nat Rev Neurosci 2, 8391.CrossRefGoogle ScholarPubMed
40. Todd, AJ (2010) Neuronal circuitry for pain processing in the dorsal horn. Nat Rev Neurosci 11, 823836.CrossRefGoogle ScholarPubMed
41. Sagar, DR, Kendall, DA & Chapman, V (2008) Inhibition of fatty acid amide hydrolase produces PPAR-alpha-mediated analgesia in a rat model of inflammatory pain. Br J Pharmacol 155, 12971306.CrossRefGoogle Scholar
42. Sagar, DR, Staniaszek, LE, Okine, BN et al. (2010) Tonic modulation of spinal hyperexcitability by the endocannabinoid receptor system in a rat model of osteoarthritis pain. Arthritis Rheum 62, 36663676.CrossRefGoogle Scholar
43. Mitrirattanakul, S, Ramakul, N, Guerrero, AV et al. (2006) Site-specific increases in peripheral cannabinoid receptors and their endogenous ligands in a model of neuropathic pain. Pain 126, 102114.CrossRefGoogle Scholar
44. Chen, L, Zhang, J, Li, F et al. (2009) Endogenous anandamide and cannabinoid receptor-2 contribute to electroacupuncture analgesia in rats. J Pain 10, 732739.CrossRefGoogle ScholarPubMed
45. Muthian, S, Rademacher, DJ, Roelke, CT et al. (2004) Anandamide content is increased and CB1 cannabinoid receptor blockade is protective during transient, focal cerebral ischemia. Neuroscience 129, 743750.CrossRefGoogle ScholarPubMed
46. Sugiura, T, Kishimoto, S, Oka, S et al. (2006) Biochemistry, pharmacology and physiology of 2-arachidonoylglycerol, an endogenous cannabinoid receptor ligand. Prog Lipid Res 45, 405446.CrossRefGoogle ScholarPubMed
47. Kaplan, BL (2013) The role of CB1 in immune modulation by cannabinoids. Pharmacol Ther 137, 365374.CrossRefGoogle ScholarPubMed
48. Graham, ES, Angel, CE, Schwarcz, LE et al. (2010) Detailed characterisation of CB2 receptor protein expression in peripheral blood immune cells from healthy human volunteers using flow cytometry. Int J Immunopathol Pharmacol 23, 2534.CrossRefGoogle ScholarPubMed
49. Pacher, P & Mechoulam, R (2011) Is lipid signaling through cannabinoid 2 receptors part of a protective system? Prog Lipid Res 50, 193211.CrossRefGoogle ScholarPubMed
50. Griffin, G, Fernando, SR, Ross, RA et al. (1997) Evidence for the presence of CB2-like cannabinoid receptors on peripheral nerve terminals. Eur J Pharmacol 339, 5361.CrossRefGoogle ScholarPubMed
51. Stander, S, Schmelz, M, Metze, D et al. (2005) Distribution of cannabinoid receptor 1 (CB1) and 2 (CB2) on sensory nerve fibers and adnexal structures in human skin. J Dermatol Sci 38, 177188.CrossRefGoogle ScholarPubMed
52. Hohmann, AG & Herkenham, M (1999) Localization of central cannabinoid CB1 receptor messenger RNA in neuronal subpopulations of rat dorsal root ganglia: a double-label in situ hybridization study. Neuroscience 90, 923931.CrossRefGoogle ScholarPubMed
53. Hsieh, GC, Pai, M, Chandran, P et al. (2011) Central and peripheral sites of action for CB(2) receptor mediated analgesic activity in chronic inflammatory and neuropathic pain models in rats. Br J Pharmacol 162, 428440.CrossRefGoogle Scholar
54. Zhang, F, Hong, S, Stone, V et al. (2007) Expression of cannabinoid CB1 receptors in models of diabetic neuropathy. J Pharmacol Exp Ther 323, 508515.CrossRefGoogle ScholarPubMed
55. Schuelert, N, Zhang, C, Mogg, AJ et al. (2010) Paradoxical effects of the cannabinoid CB2 receptor agonist GW405833 on rat osteoarthritic knee joint pain. Osteoarthritis Cartilage 18, 15361543.CrossRefGoogle ScholarPubMed
56. Tsou, K, Brown, S, Sanudo-Pena, MC et al. (1998) Immunohistochemical distribution of cannabinoid CB1 receptors in the rat central nervous system. Neuroscience 83, 393411.CrossRefGoogle ScholarPubMed
57. Farquhar-Smith, WP, Egertova, M, Bradbury, EJ et al. (2000) Cannabinoid CB(1) receptor expression in rat spinal cord. Mol Cell Neurosci 15, 510521.CrossRefGoogle ScholarPubMed
58. Morisset, V & Urban, L (2001) Cannabinoid-induced presynaptic inhibition of glutamatergic EPSCs in substantia gelatinosa neurons of the rat spinal cord. J Neurophysiol 86, 4048.CrossRefGoogle ScholarPubMed
59. Salio, C, Doly, S, Fischer, J et al. (2002) Neuronal and astrocytic localization of the cannabinoid receptor-1 in the dorsal horn of the rat spinal cord. Neurosci Lett 329, 1316.CrossRefGoogle ScholarPubMed
60. Salio, C, Fischer, J, Franzoni, MF et al. (2002) Pre- and postsynaptic localizations of the CB1 cannabinoid receptor in the dorsal horn of the rat spinal cord. Neuroscience 110, 755764.CrossRefGoogle ScholarPubMed
61. Brownjohn, PW & Ashton, JC (2012) Spinal cannabinoid CB2 receptors as a target for neuropathic pain: an investigation using chronic constriction injury. Neuroscience 203, 180193.CrossRefGoogle ScholarPubMed
62. Racz, I, Nadal, X, Alferink, J et al. (2008) Crucial role of CB(2) cannabinoid receptor in the regulation of central immune responses during neuropathic pain. J Neurosci 28, 1212512135.CrossRefGoogle ScholarPubMed
63. Romero-Sandoval, A, Nutile-McMenemy, N & DeLeo, JA (2008) Spinal microglial and perivascular cell cannabinoid receptor type 2 activation reduces behavioral hypersensitivity without tolerance after peripheral nerve injury. Anesthesiology 108, 722734.CrossRefGoogle ScholarPubMed
64. Alkaitis, MS, Solorzano, C, Landry, RP et al. (2010) Evidence for a role of endocannabinoids, astrocytes and p38 phosphorylation in the resolution of postoperative pain. PLoS ONE 5, e10891.CrossRefGoogle ScholarPubMed
65. Berghuis, P, Dobszay, MB, Wang, X et al. (2005) Endocannabinoids regulate interneuron migration and morphogenesis by transactivating the TrkB receptor. Proc Natl Acad Sci USA 102, 1911519120.CrossRefGoogle ScholarPubMed
66. De March, Z, Zuccato, C, Giampa, C et al. (2008) Cortical expression of brain derived neurotrophic factor and type-1 cannabinoid receptor after striatal excitotoxic lesions. Neuroscience 152, 734740.CrossRefGoogle ScholarPubMed
67. Hu, H, Ho, W, Mackie, K et al. (2012) Brain CB(1) receptor expression following lipopolysaccharide-induced inflammation. Neuroscience 227, 211222.CrossRefGoogle ScholarPubMed
68. Stella, N (2010) Cannabinoid and cannabinoid-like receptors in microglia, astrocytes, and astrocytomas. Glia 58, 10171030.CrossRefGoogle ScholarPubMed
69. Atwood, BK & Mackie, K (2010) CB2: a cannabinoid receptor with an identity crisis. Br J Pharmacol 160, 467479.CrossRefGoogle ScholarPubMed
70. Agarwal, N, Pacher, P, Tegeder, I et al. (2007) Cannabinoids mediate analgesia largely via peripheral type 1 cannabinoid receptors in nociceptors. Nat Neurosci 10, 870879.CrossRefGoogle ScholarPubMed
71. Calignano, A, La Rana, G, Giuffrida, A et al. (1998) Control of pain initiation by endogenous cannabinoids. Nature 394, 277281.CrossRefGoogle ScholarPubMed
72. Maione, S, De Petrocellis, L, de Novellis, V et al. (2007) Analgesic actions of N-arachidonoyl-serotonin, a fatty acid amide hydrolase inhibitor with antagonistic activity at vanilloid TRPV1 receptors. Br J Pharmacol 150, 766781.CrossRefGoogle ScholarPubMed
73. Beaulieu, P, Bisogno, T, Punwar, S et al. (2000) Role of the endogenous cannabinoid system in the formalin test of persistent pain in the rat. Eur J Pharmacol 396, 8592.CrossRefGoogle Scholar
74. Guindon, J, Desroches, J & Beaulieu, P (2007) The antinociceptive effects of intraplantar injections of 2-arachidonoyl glycerol are mediated by cannabinoid CB2 receptors. Br J Pharmacol 150, 693701.CrossRefGoogle ScholarPubMed
75. Jayamanne, A, Greenwood, R, Mitchell, VA et al. (2006) Actions of the FAAH inhibitor URB597 in neuropathic and inflammatory chronic pain models. Br J Pharmacol 147, 281288.CrossRefGoogle ScholarPubMed
76. Chang, L, Luo, L, Palmer, JA et al. (2006) Inhibition of fatty acid amide hydrolase produces analgesia by multiple mechanisms. Br J Pharmacol 148, 102113.CrossRefGoogle ScholarPubMed
77. Lichtman, AH, Leung, D, Shelton, CC et al. (2004) Reversible inhibitors of fatty acid amide hydrolase that promote analgesia: evidence for an unprecedented combination of potency and selectivity. J Pharmacol Exp Ther 311, 441448.CrossRefGoogle ScholarPubMed
78. Kathuria, S, Gaetani, S, Fegley, D et al. (2003) Modulation of anxiety through blockade of anandamide hydrolysis. Nat Med 9, 7681.CrossRefGoogle ScholarPubMed
79. Fegley, D, Gaetani, S, Duranti, A et al. (2005) Characterization of the fatty acid amide hydrolase inhibitor cyclohexyl carbamic acid 3′-carbamoyl-biphenyl-3-yl ester (URB597): effects on anandamide and oleoylethanolamide deactivation. J Pharmacol Exp Ther 313, 352358.CrossRefGoogle ScholarPubMed
80. Russo, R, Loverme, J, La Rana, G et al. (2007) The fatty-acid amide hydrolase inhibitor URB597 (cyclohexyl carbamic acid 3′-carbamoyl-biphenyl-3-yl ester) reduces neuropathic pain after oral administration in mice. J Pharmacol Exp Ther 322, 236242.CrossRefGoogle ScholarPubMed
81. Jhaveri, MD, Richardson, D, Robinson, I et al. (2008) Inhibition of fatty acid amide hydrolase and cyclooxygenase-2 increases levels of endocannabinoid related molecules and produces analgesia via peroxisome proliferator-activated receptor-alpha in a model of inflammatory pain. Neuropharmacology 55, 8593.CrossRefGoogle Scholar
82. Sagar, DR, Kendall, DA & Chapman, V (2008) Inhibition of fatty acid amide hydrolase produces PPAR-alpha-mediated analgesia in a rat model of inflammatory pain. Br J Pharmacol 155, 12971306.CrossRefGoogle Scholar
83. Spradley, JM, Guindon, J & Hohmann, AG (2010) Inhibitors of monoacylglycerol lipase, fatty-acid amide hydrolase and endocannabinoid transport differentially suppress capsaicin-induced behavioral sensitization through peripheral endocannabinoid mechanisms. Pharmacol Res 62, 249258.CrossRefGoogle ScholarPubMed
84. Guindon, J, Guijarro, A, Piomelli, D et al. (2011) Peripheral antinociceptive effects of inhibitors of monoacylglycerol lipase in a rat model of inflammatory pain. Br J Pharmacol 163, 14641478.CrossRefGoogle Scholar
85. Clapper, JR, Moreno-Sanz, G, Russo, R et al. (2010) Anandamide suppresses pain initiation through a peripheral endocannabinoid mechanism. Nat Neurosci 13, 6.CrossRefGoogle ScholarPubMed
86. Han, KH, Lim, S, Ryu, J et al. (2009) CB1 and CB2 cannabinoid receptors differentially regulate the production of reactive oxygen species by macrophages. Cardiovasc Res 84, 378386.CrossRefGoogle ScholarPubMed
87. Cencioni, MT, Chiurchiăą, V, Catanzaro, G et al. (2010) Anandamide suppresses proliferation and cytokine release from primary human T-lymphocytes mainly via CB2 receptors. PLoS ONE 5, e8688.CrossRefGoogle ScholarPubMed
88. Samson, M-T, Small-Howard, A, Shimoda, LMN et al. (2003) Differential roles of CB1 and CB2 cannabinoid receptors in mast cells. J Immunol 170, 49534962.CrossRefGoogle ScholarPubMed
89. Hao, M-X, Jiang, L-S, Fang, N-Y et al. (2010) The cannabinoid WIN55,212-2 protects against oxidized LDL-induced inflammatory response in murine macrophages. J Lipid Res 51, 21812190.CrossRefGoogle ScholarPubMed
90. Nomura, DK, Morrison, BE, Blankman, JL et al. (2011) Endocannabinoid hydrolysis generates brain prostaglandins that promote neuroinflammation. Science 334, 809813.CrossRefGoogle ScholarPubMed
91. Starowicz, K, Makuch, W, Osikowicz, M et al. (2012) Spinal anandamide produces analgesia in neuropathic rats: possible CB(1)- and TRPV1-mediated mechanisms. Neuropharmacology 62, 17461755.CrossRefGoogle ScholarPubMed
92. Welch, SP, Huffman, JW & Lowe, J (1998) Differential blockade of the antinociceptive effects of centrally administered cannabinoids by SR141716A. J Pharmacol Exp Ther 286, 13011308.Google ScholarPubMed
93. Richardson, JD, Aanonsen, L & Hargreaves, KM (1997) SR 141716A, a cannabinoid receptor antagonist, produces hyperalgesia in untreated mice. Eur J Pharmacol 319, R3R4.CrossRefGoogle ScholarPubMed
94. Chapman, V (1999) The cannabinoid CB1 receptor antagonist, SR141716A, selectively facilitates nociceptive responses of dorsal horn neurones in the rat. Br J Pharmacol 127, 17651767.CrossRefGoogle ScholarPubMed
95. Sagar, DR, Jhaveri, M, Richardson, D et al. (2010) Endocannabinoid regulation of spinal nociceptive processing in a model of neuropathic pain. Eur J Neurosci 31, 8.CrossRefGoogle Scholar
96. Sagar, DR, Staniaszek, LE, Okine, BN et al. (2010) Tonic modulation of spinal hyperexcitability by the endocannabinoid receptor system in a rat model of osteoarthritis pain. Arthritis Rheum 62, 36663676.CrossRefGoogle Scholar
97. Okine, BN, Norris, LM, Woodhams, S et al. (2012) Lack of effect of chronic pre-treatment with the FAAH inhibitor URB597 on inflammatory pain behaviour: evidence for plastic changes in the endocannabinoid system. Br J Pharmacol 167, 627640.CrossRefGoogle ScholarPubMed
98. Petrosino, S, Palazzo, E, de Novellis, V et al. (2007) Changes in spinal and supraspinal endocannabinoid levels in neuropathic rats. Neuropharmacology 52, 415422.CrossRefGoogle ScholarPubMed
99. Jhaveri, MD, Richardson, D, Kendall, DA et al. (2006) Analgesic effects of fatty acid amide hydrolase inhibition in a rat model of neuropathic pain. J Neurosci 26, 1331813327.CrossRefGoogle Scholar
100. Woodhams, SG, Wong, A, Barrett, DA et al. (2012) Spinal administration of the monoacylglycerol lipase inhibitor JZL184 produces robust inhibitory effects on nociceptive processing and the development of central sensitization in the rat. Br J Pharmacol 167, 16091619.CrossRefGoogle ScholarPubMed
101. Lim, G, Sung, B, Ji, RR et al. (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, 275283.CrossRefGoogle ScholarPubMed
102. Zhang, J, Hoffert, C, Vu, HK et al. (2003) Induction of CB2 receptor expression in the rat spinal cord of neuropathic but not inflammatory chronic pain models. Eur J Neurosci 17, 27502754.CrossRefGoogle Scholar
103. Correa, F, Docagne, F, Mestre, L et al. (2009) A role for CB2 receptors in anandamide signalling pathways involved in the regulation of IL-12 and IL-23 in microglial cells. Biochem Pharmacol 77, 86100.CrossRefGoogle ScholarPubMed
104. Stella, N (2009) Endocannabinoid signaling in microglial cells. Neuropharmacology 56, Suppl. 1, 244253.CrossRefGoogle ScholarPubMed
105. Walter, L, Franklin, A, Witting, A et al. (2003) Nonpsychotropic cannabinoid receptors regulate microglial cell migration. J Neurosci 23, 13981405.CrossRefGoogle ScholarPubMed
106. Bilkei-Gorzo, A (2012) The endocannabinoid system in normal and pathological brain ageing. Phil Trans R Soc Lond B Biol Sci 367, 33263341.CrossRefGoogle ScholarPubMed
107. Martin, WJ, Patrick, SL, Coffin, PO et al. (1995) An examination of the central sites of action of cannabinoid-induced antinociception in the rat. Life Sci 56, 21032109.CrossRefGoogle ScholarPubMed
108. Walker, JM, Huang, SM, Strangman, NM et al. (1999) Pain modulation by release of the endogenous cannabinoid anandamide. Proc Natl Acad Sci USA 96, 1219812203.CrossRefGoogle ScholarPubMed
109. Rea, K, Roche, M & Finn, DP (2007) Supraspinal modulation of pain by cannabinoids: the role of GABA and glutamate. Br J Pharmacol 152, 633648.CrossRefGoogle ScholarPubMed
110. Maione, S, Bisogno, T, de Novellis, V et al. (2006) Elevation of endocannabinoid levels in the ventrolateral periaqueductal grey through inhibition of fatty acid amide hydrolase affects descending nociceptive pathways via both cannabinoid receptor type 1 and transient receptor potential vanilloid type-1 receptors. J Pharmacol Exper Ther 316, 969982.CrossRefGoogle ScholarPubMed
111. Hohmann, AG, Suplita, RL, Bolton, NM et al. (2005) An endocannabinoid mechanism for stress-induced analgesia. Nature 435, 11081112.CrossRefGoogle ScholarPubMed
112. Suplita, RL II, Farthing, JN, Gutierrez, T et al. (2005) Inhibition of fatty-acid amide hydrolase enhances cannabinoid stress-induced analgesia: sites of action in the dorsolateral periaqueductal gray and rostral ventromedial medulla. Neuropharmacology 49, 12011209.CrossRefGoogle ScholarPubMed
113. Suplita, RL II, Gutierrez, T, Fegley, D et al. (2006) Endocannabinoids at the spinal level regulate, but do not mediate, nonopioid stress-induced analgesia. Neuropharmacology 50, 372379.CrossRefGoogle Scholar
114. Gregg, LC, Jung, K-M, Spradley, JM et al. (2012) Activation of type 5 metabotropic glutamate receptors and diacylglycerol lipase-α initiates 2-arachidonoylglycerol formation and endocannabinoid-mediated analgesia. J Neurosci 32, 94579468.CrossRefGoogle Scholar
115. Vachon-Presseau, E, Martel, M-O, Roy, M et al. (2013) Acute stress contributes to individual differences in pain and pain-related brain activity in healthy and chronic pain patients. J Neurosci 33, 68266833.CrossRefGoogle ScholarPubMed
116. Rani Sagar, D, Burston, JJ, Woodhams, SG et al. (2012) Dynamic changes to the endocannabinoid system in models of chronic pain. Phil Trans R Soc Lond B Biol Sci 367, 33003311.CrossRefGoogle Scholar
117. Hoot, MR, Sim-Selley, LJ, Poklis, JL et al. (2010) Chronic constriction injury reduces cannabinoid receptor 1 activity in the rostral anterior cingulate cortex of mice. Brain Res 1339, 1825.CrossRefGoogle ScholarPubMed
118. Sim-Selley, LJ (2003) Regulation of cannabinoid CB1 receptors in the central nervous system by chronic cannabinoids. Crit Rev Neurobiol 15, 91119.CrossRefGoogle ScholarPubMed
119. Lee, MC, Ploner, M, Wiech, K et al. (2013) Amygdala activity contributes to the dissociative effect of cannabis on pain perception. Pain 154, 124134.CrossRefGoogle Scholar
120. Foltz, EL & White, LE (1962) Pain ‘Relief’ by frontal cingulumotomy. J Neurosurg 19, 89100.CrossRefGoogle ScholarPubMed
121. Alexander, SP & Kendall, DA (2007) The complications of promiscuity: endocannabinoid action and metabolism. Br J Pharmacol 152, 602623.CrossRefGoogle ScholarPubMed
122. Fowler, CJ (2007) The contribution of cyclooxygenase-2 to endocannabinoid metabolism and action. Br J Pharmacol 152, 594601.CrossRefGoogle ScholarPubMed
123. Kozak, KR, Crews, BC, Morrow, JD et al. (2002) Metabolism of the endocannabinoids, 2-arachidonylglycerol and anandamide, into prostaglandin, thromboxane, and prostacyclin glycerol esters and ethanolamides. J Biol Chem 277, 4487744885.CrossRefGoogle ScholarPubMed
124. Gatta, L, Piscitelli, F, Giordano, C et al. (2012) Discovery of prostamide F2alpha and its role in inflammatory pain and dorsal horn nociceptive neuron hyperexcitability. PLoS ONE 7, e31111.CrossRefGoogle ScholarPubMed
125. Snider, NT, Nast, JA, Tesmer, LA et al. (2009) A cytochrome P450-derived epoxygenated metabolite of anandamide is a potent cannabinoid receptor 2-selective agonist. Mol Pharmacol 75, 965972.CrossRefGoogle ScholarPubMed
126. Hu, SS, Bradshaw, HB, Chen, JS et al. (2008) Prostaglandin E2 glycerol ester, an endogenous COX-2 metabolite of 2-arachidonoylglycerol, induces hyperalgesia and modulates NFkappaB activity. Br J Pharmacol 153, 15381549.CrossRefGoogle ScholarPubMed
127. Garcia-Merino, A (2013) Endocannabinoid system modulator use in everyday clinical practice in the UK and Spain. Expert Rev Neurother 13, 913.CrossRefGoogle ScholarPubMed
128. Huggins, JP, Smart, TS, Langman, S et al. (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, 18371846.CrossRefGoogle ScholarPubMed
129. Ross, RA (2003) Anandamide and vanilloid TRPV1 receptors. Br J Pharmacol 140, 790801.CrossRefGoogle ScholarPubMed
130. Fowler, CJ, Bjorklund, E, Lichtman, AH et al. (2013) Inhibitory properties of ibuprofen and its amide analogues towards the hydrolysis and cyclooxygenation of the endocannabinoid anandamide. J Enzyme Inhib Med Chem 28, 172182.CrossRefGoogle ScholarPubMed
131. Windsor, MA, Hermanson, DJ, Kingsley, PJ et al. (2012) Substrate-selective inhibition of cyclooxygenase-2: development and evaluation of achiral profen probes. ACS Med Chem Lett 3, 759763.CrossRefGoogle ScholarPubMed
132. Costa, B, Bettoni, I, Petrosino, S et al. (2010) The dual fatty acid amide hydrolase/TRPV1 blocker, N-arachidonoyl-serotonin, relieves carrageenan-induced inflammation and hyperalgesia in mice. Pharmacol Res 61, 537546.CrossRefGoogle ScholarPubMed
133. de Novellis, V, Palazzo, E, Rossi, F et al. (2008) The analgesic effect of N-arachidonoyl-serotonin, a FAAH inhibitor and TRPV1 receptor antagonist, associated with changes in rostral ventromedial medulla and locus coeruleus cell activity in rats. Neuropharmacology 55, 11051113.CrossRefGoogle ScholarPubMed
134. Bishay, P, Haussler, A, Lim, H-Y et al. (2013) Anandamide deficiency and heightened neuropathic pain in aged mice. Neuropharmacology 71, 204215.CrossRefGoogle ScholarPubMed
135. Long, JZ, Li, WW, Booker, L et al. (2009) Selective blockade of 2-arachidonoylglycerol hydrolysis produces cannabinoid behavioral effects. Nat Chem Biol 5, 3744.CrossRefGoogle ScholarPubMed
136. Chang Jae, W, Niphakis Micah, J, Lum Kenneth, M et al. (2012) Highly selective inhibitors of monoacylglycerol lipase bearing a reactive group that is bioisosteric with endocannabinoid substrates. Chem Biol 19, 579588.CrossRefGoogle ScholarPubMed
137. Kinsey, SG, Wise, LE, Ramesh, D et al. (2013) Repeated low dose administration of the monoacylglycerol lipase inhibitor JZL184 retains CB1 receptor mediated antinociceptive and gastroprotective effects. J Pharmacol Exp Ther 345, 492501.CrossRefGoogle ScholarPubMed
138. Kinsey, SG, Nomura, DK, O'Neal, ST et al. (2011) Inhibition of monoacylglycerol lipase attenuates nonsteroidal anti-inflammatory drug-induced gastric hemorrhages in mice. J Pharmacol Exp Ther 338, 795802.CrossRefGoogle ScholarPubMed
139. Alhouayek, M, Lambert, DM, Delzenne, NM et al. (2011) Increasing endogenous 2-arachidonoylglycerol levels counteracts colitis and related systemic inflammation. FASEB J 25, 27112721.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. (Colour online) Endocannabinoid (EC) signalling at a notional neuronal synapse. The major synthetic, signalling and catabolic pathways for anandamide (AEA) and 2-arachidonoylglycerol (2-AG) are shown. Alongside cannabinoid receptors (CB1) and (CB2) the other G protein coupled receptors (GPCR) may be involved in cannabinoid signalling. FLAT-1, a truncated form of fatty acid amide hydrolase (FAAH), and ECT are the putative EC transporters. MAGL, monoacylglycerol lipase; DAGL, diacyglycerol lipase; GPR, specific G protein-coupled receptor such as GPR55. Compounds in red are recognised enzyme inhibitors/receptor antagonists, which can modulate EC signals.

Figure 1

Fig. 2. (Colour online) Nociceptive pathways. Schematic of nociceptive pathways. Nociceptive stimuli are conducted from the periphery to the dorsal horn of the spinal cord, and then transmitted to the supra-spinal regions via the spinothalamic tract (STT, blue) and spinoparabrachial tract (SPBT, red). The major descending modulatory control pathway (DMCP, purple) is displayed on the right. This pathway crosses the midline at the level of the medulla. Coloured areas indicate the position of synapses in each pathway. The positions of laminae I–VI in the dorsal horn are indicated by dotted lines, while the black region in the brain represents the lateral ventricles. Thal., thalamus; VMH, ventromedial hypothalamus; PbN, parabrachial nucleus; PAG, periaqueductal grey matter; RVM, rostroventral medulla; Pyr., pyramidal tract; DRG, dorsal root ganglion. Adapted from(39).