Next Article in Journal
Regional Differences in Ca2+ Signaling and Transverse-Tubules across Left Atrium from Adult Sheep
Next Article in Special Issue
Evidence for Endogenous Opioid Dependence Related to Latent Sensitization in a Rat Model of Chronic Inflammatory Pain
Previous Article in Journal
Methyl Donors, Epigenetic Alterations, and Brain Health: Understanding the Connection
Previous Article in Special Issue
Ca2+-Permeable AMPA Receptors Contribute to Changed Dorsal Horn Neuronal Firing and Inflammatory Pain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microglial Cannabinoid CB2 Receptors in Pain Modulation

1
Department of Medical Neuroscience, Key University Laboratory of Metabolism and Health of Guangdong, SUSTech Center for Pain Medicine, School of Medicine, Southern University of Science and Technology, Shenzhen 518055, China
2
Department of Anesthesiology, Shenzhen People’s Hospital, The First Affiliated Hospital, Southern University of Science and Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(3), 2348; https://doi.org/10.3390/ijms24032348
Submission received: 22 December 2022 / Revised: 17 January 2023 / Accepted: 19 January 2023 / Published: 25 January 2023
(This article belongs to the Special Issue New Advance on Molecular Targets for the Treatment of Pain)

Abstract

:
Pain, especially chronic pain, can strongly affect patients’ quality of life. Cannabinoids ponhave been reported to produce potent analgesic effects in different preclinical pain models, where they primarily function as agonists of Gi/o protein-coupled cannabinoid CB1 and CB2 receptors. The CB1 receptors are abundantly expressed in both the peripheral and central nervous systems. The central activation of CB1 receptors is strongly associated with psychotropic adverse effects, thus largely limiting its therapeutic potential. However, the CB2 receptors are promising targets for pain treatment without psychotropic adverse effects, as they are primarily expressed in immune cells. Additionally, as the resident immune cells in the central nervous system, microglia are increasingly recognized as critical players in chronic pain. Accumulating evidence has demonstrated that the expression of CB2 receptors is significantly increased in activated microglia in the spinal cord, which exerts protective consequences within the surrounding neural circuitry by regulating the activity and function of microglia. In this review, we focused on recent advances in understanding the role of microglial CB2 receptors in spinal nociceptive circuitry, highlighting the mechanism of CB2 receptors in modulating microglia function and its implications for CB2 receptor- selective agonist-mediated analgesia.

1. Introduction

Chronic pain is a complex web of emotional experiences and subjective senses that brings patients enormous physical and psychological burdens. Due to the current unmet demand for pain relief, attention has been highly focused on effective analgesics without major central adverse effects. The endocannabinoid system serves as an important neuromodulator signaling, pathway that is deeply involved in the complex modulation of endogenous homeostasis and a variety of pathological processes. Currently, in parallel to the great interest in the endocannabinoid system, cannabinoid compounds, including endocannabinoids, plant-derived, and synthetic cannabinoids, are increasingly recognized as potential therapeutic alternatives for pathological pain, because of their vital roles in modulating nociceptive information processing [1,2].
The biological effects of these cannabinoids are mainly mediated by two endocannabinoid receptors, cannabinoid receptors 1 (CB1R) and 2 (CB2R). However, currently available cannabis-based medicines are largely targeting CB1R, which is detected abundantly in both the peripheral and central nervous systems (PNS and CNS, respectively). Their activation can result in a spectrum of adverse effects, such as the development of tolerance [3], addiction, and psychotomimetic effects [4], thus limiting their therapeutic potential. Subsequently, several strategies have been developed for reducing the central side effects of cannabinoid compounds acting as analgesics, including peripherally restricted CB1R agonists, topically applied cannabinoids, cannabinoid metabolic enzyme inhibitors, bifunctional cannabinoids ligands, as well as selective CB2R agonists. For example, previously reported VF13, a bifunctional chimeric peptide containing the pharmacophores of the endogenous cannabinoid peptide VD-hemopressin(α) and neuropeptide VF, produced non-tolerance forming antinociception in multiple pain models with reduced cannabinoid-related side effects [5]. In addition, since CB2R is primarily expressed in immune cells and relative paucity in neurons, the advantageous profile of CB2R agonists devoid of centrally CB1R-mediated negative side effects suggests that CB2R may represent an attractive target for alleviating pain [6,7]. A large number of CB2R selective agonists such as JWH-133, JWH-015, AM1241, and GW405833 have been demonstrated to exert significant antinociception in various preclinical pain model, including inflammatory pain [8,9], postoperative pain [10], neuropathic pain [8,9,11,12,13,14], and bone cancer pain [15,16], without eliciting tolerance [8,17], hypoactivity [18], hypothermia [18,19], catalepsy [19], psychotropic effects and physical dependence [20,21,22]. Therefore, amounts of research efforts focused on the discovery and optimization of synthetic high-affinity CB2R agonists to promote selective binding to the CB2R are ongoing, such as the acyl hydrazone derivative MDA7, a more promising CB2R agonist for severe pain therapy [23,24,25,26].
Historically, CB2R was thought to be mainly expressed in peripheral immune cells. However, with the increasing interest in CB2R-related research, converging evidence over the last couple of decades has indicated that CB2R is also expressed in microglia and some neurons in the CNS. Meanwhile, it is markedly upregulated in activated microglia in response to injury or chronic pain states [10,27,28,29,30,31,32]. Microglia are the macrophage-like cells in the CNS that providing immunomodulatory functions in response to injury or disease. In recent years, increasing studies have demonstrated that spinal microglial activation and microgliosis induced by some inflammatory mediators released from the central nerve terminals of nociceptor neurons play an active role in pathological pain processing [33,34]. Along with rapid morphological changes in microglia, numerous signaling molecules were also altered in microglia in response to microenvironment changes [33,34]. However, strong evidence now supports the hypothesis that the extent of the spinal CB2R marked increase in activated microglia may contribute to the shift of microglia to adopt an anti-inflammatory phenotype, with important consequences for the surrounding neural circuitry of nociceptive transmission [35,36,37].
Accumulating studies have shown that targeting spinal CB2R represents a promising strategy for pain relief without the classical cannabinoid side effects. However, the underlying mechanisms involved in microglial CB2R activation and its contributions to pain processing remain poorly understood. This review will focus on the current research on the roles of microglial CB2R in pain processing in vitro and in vivo. We emphasize the involvement of CB2R signaling cascades in spinal microglia in the pain processes, as well as the molecular mechanisms involved in mediating the analgesic effects of cannabinoids, especially synthetic CB2R agonists.

2. The Endocannabinoid System

The endocannabinoid system (ECS) is comprised of multiple receptors, endogenous lipid ligands, and various enzymes that regulate endocannabinoid synthesis and degradation, as summarized in Figure 1. As a neuromodulatory system widely involved in lipid signaling responses, the ECS has shown great potential in the development of therapeutic drugs targeting pain management [38]. The endocannabinoids are generated from the hydrolysis of membrane polyunsaturated fatty acids, which is mediated by different enzymes. Arachidonoyl ethanolamide (AEA) and 2-arachidonoyl-glycerol (2-AG) were the first discovered and extensively studied endocannabinoids. Most AEA appears to be derived from its membrane precursor, N-acyl phosphatidyl ethanolamine (NAPE), which is hydrolyzed by N-acyl-phosphatidyl ethanolamine-phospholipase D (NAPE-PLD) [39]. After its release and uptake, the degradation of AEA is primarily via the enzyme fatty acid amino hydrolase (FAAH) [40] or cyclooxygenase-2 (COX-2) [39]. The 2-AG is mainly produced from the hydrolysis of an arachidonoyl-containing phosphatidyl inositol bis-phosphate (PIP2) mediated by phosphoinositol-phospholipase C (PLC) and two diacylglycerol lipases α/β (DAGLα/β) [41]. Sequentially, 2-AG is hydrolyzed primarily by monoacylglycerol lipase (MAGL) and to a lesser extent by α/β domain hydrolases (ABDH) 6 and 12. Occasionally, FAAH and COX-2 also degrade 2-AG under some conditions [42,43,44]. In addition to the best-known 2-AG and AEA, some neuropeptides are also considered to be endocannabinoids and have been described to have CB1R and CB2R activity, such as hemopressin, a fragment derived from the α-chain of hemoglobin [45,46,47,48].
The CB1R and CB2R are the two identified and best-characterized cannabinoid receptor subtypes. Both are seven transmembrane inhibitory Gi/o protein-coupled receptors involved in classical signal transduction pathways, including inhibition of adenylyl cyclase (AC) activity, reduction of cyclic adenosine monophosphate (cAMP) production, as well as signal pathways associated with mitogen-activated protein kinases (MAPK) [49]. The 2-AG acts as a high efficacy full agonist for both CB1R and CB2R. However, AEA exhibits low efficacy and can act as a partial agonist preferentially for CB1R with 4-fold selectivity compared with CB2R [50,51,52]. In addition, previous studies demonstrated that the putative non-CB1R/CB2R orphan receptor GPR55 might represent the third cannabinoid receptor (CB3R), though GPR55 shares a low amino acid sequence with both CB1R (13.5%) and CB2R (14.4%) [53,54]. Subsequently, lysophosphatidylinositol (LPI) is identified as an endogenous agonist of GPR55 [55], while the physiological role and pharmacology of GPR55 remain unclear [56]. Recent studies in animals suggest that GPR55 also plays an important regulatory role in pain signal processing in the CNS [57].
It is acknowledged that CB1R is abundantly expressed in neurons of the brain cortex, cerebellum, basal ganglia, and hippocampus within CNS areas. These brain areas not only regulate movement and memory functions, but also intricate pain signaling, which accounts for the impact of CB1R agonists on psychoactivity, cognition, memory, and locomotive activity [57,58,59]. In the spinal cord, CB1R was found in the glutamatergic interneurons, glycinergic interneurons, GABAergic inhibitory interneurons, and a subset of spinal projection neurons [60,61]. The CB1R functions primarily by modulating neuropeptide and neurotransmitter release to inhibit synaptic transmission in the brain and spinal cord. Activation of CB1R results in the activation of inwardly rectifying K+ channels, promoting cell hyperpolarization, which leads to the inhibition of the presynaptic neurons. Meanwhile, activation of CB1R can inhibit Ca2+ channels to decrease neurotransmitter release [62] (Figure 1). Additionally, CB1R has been detected in the peripheral nervous system like the dorsal root ganglion (DRG) and other tissues, such as the spleen [63], leukocytes [64], tonsils [65], skin [66], testes, heart, lungs, small intestine, urinary bladder, prostate, uterus, ovary, and vas deferens [67,68,69]. Recently, CB1R has also been reported as functionally expressed in different glial cells, specifically at a low level in microglia and mainly in different neuronal subtypes. Whether and how CB1R regulates microglial function remains unclear. Moreover, CB1R is also expressed in astrocytes, shaping synaptic transmission and memory [35,70,71,72,73].
The CB2R shows a distinctly different distribution compared to the CB1R, which was initially identified predominantly at the periphery, mainly in immunocytes, such as lymphocytes, neutrophils, leukocytes CD8 and CD4, and macrophages [2,4,35,74,75,76,77] in various tissues, such as the liver [78], pancreas [79], bone [80], skin [66], spleen, and vascular elements [81]. Subsequent data indicated that CB2R can also be detected in nervous systems, but the expression level of CB2R is 100-fold lower than that of CB1R under normal physiological conditions, which also suggests that the activation of CB2R may have a different pharmacological profile and not be involved in the unwanted cannabis-like effects [82]. Within the CNS, the CB2R appears to be expressed by some neurons specifically within the cortex, striatum, hippocampus, amygdala, olfactory, cerebellum, thalamus, spinal nuclei, and substantia nigra [69,83,84]. Meanwhile, the accumulating reports showed that CB2R can also be found in brain glial cells, particularly microglia [85,86,87]. When compared with CB1R, CB2R exhibit one unique feature that is a dynamic expression manner under certain pathological conditions (e.g., inflammation, anxiety, epilepsy, and nerve injury) in neurons and microglia, suggesting the alteration of CB2R expression and function are closely related to these neurological diseases [9,32,83,84,88,89,90,91].
Notably, the spinal cord distribution of the CB2R (mRNA or protein) has been identified primarily in microglia, especially activated microglia, and poorly expressed in neurons by the preponderance of evidence using quantitative real-time PCR [6,9], in situ hybridization [28], western blotting [92,93], immunohistochemical staining [17], or single-cell sequencing [60]. Intriguingly, similar to peripheral CB2R, which is mainly detected in immune cells, central CB2R is predominantly expressed in microglia, which indicates that CB2R is critically involved in the modulation of immune-related responses, including immune-neuron interactions and the occurrence of neuroinflammatory and pathological pain [94].

3. Microglia Express the CB2R-Related Functional Endocannabinoid System

In acting as the central custodians protected by the blood-brain barrier, microglia are derived from primitive macrophages that emanate from the embryonic yolk sac and invade the CNS via the circulatory system during development [95,96]. Over recent decades, a large amount of strong evidence has supported the hypothesis that microglia-dependent phagocytosis/degradation and neuroimmune response critically contributes to the alterations in synaptic remodeling and signaling pathways for chronic pain development [33,34,97,98]. In order to maintain stability in an ever-changing environment, microglia adopt different phenotype states characterized by distinct morphological types to perform their functions.
In the healthy adult CNS, homeostatic microglia with a small soma and highly ramified processes classically display a “resting” phenotype to continuously palpate the environment [95]. In nerve injury or chronic pain conditions, spinal microglia may rapidly proliferate and become activated to a pro-inflammatory (M1, classically activated state) phenotype with shorter processes and hypertrophic soma to respond to some inflammatory mediators, including neurotransmitters (e.g., calcitonin gene-related peptide, CGRP), growth factors (e.g., colony-stimulating factor-1, CSF-1), adenosine triphosphate (ATP), chemokines (e.g., CX3CL1), and enzymes (e.g., Caspase-6), released from central nerve terminals of nociceptor neurons [99,100,101]. Meanwhile, microglia produce and secrete various pro-inflammatory mediators, including interleukin 1β (IL-1β), IL-6, interferon γ (IFN-γ), tumor necrosis factor α (TNF-α), and prostaglandin E2 (PGE2), which can promote immunological actions and can also act on neurons to lead to central sensitization of surrounding neuronal circuits, contributing to the manifestation of chronic pain [102].
In addition, after this M1 response, aimed at eliminating noxious stimuli and restraining the initial inflammation from the area, the resolution of the inflammatory process is essential to bypass neurotoxicity and chronic inflammation [103]. For this resolving phase to occur, a shift in microglia phenotype along a spectrum of activation states from pro-inflammatory to phagocytic anti-inflammatory (M2, alternative activation state) is typically observed, which could increase the production of anti-inflammatory mediators, including transforming growth factor (TGF-β), interleukin 10 (IL-10), IL-13, and IL-4 [104,105]. This process can dampen the pro-inflammatory cytokine levels and promote the resolution of inflammation and the recovery of tissue homeostasis [105,106].
Previous studies in CB2R−/− mice have demonstrated that an exacerbated microgliosis occurs in the spinal cord after peripheral nerve injury [29,107], which indicates the relevant role of CB2R in controlling microglial proliferation and reactivity. Indeed, accumulating reports have shown that microglia express functional endocannabinoid signaling, which could modulate the activity of microglia to restore CNS homeostasis during pathological neuroinflammatory conditions [108,109]. Firstly, classic studies have shown that cultured microglia from humans [103,110], rats [77,111], or mouse tissues [9,85] and the BV-2 microglia cell line [85], express large amounts of CB2R. As previously mentioned in Section 2, CB2R is predominantly expressed in activated microglia within the CNS, which is consistent with CB2R upregulation in activated macrophages in the peripheral [77,112]. By contrast, the expression of CB1R in microglia is ambiguous. Preliminary evidence from early studies suggested that CB1R is expressed in culture microglia prepared from mollusks and rodents [111,113], but not in humans [110]. However, it has recently been shown that, unlike Cnr2 (CB2R) transcripts, Cnr1 (CB1R) expression was not detected in naïve microglia by microfluidic RT-qPCR [114]. These results require further functional and anatomical evidence to investigate whether CB1R can regulate microglial function. To date, Agnès Nadjar et al. have provided the only evidence of CB1R functional expression on microglia [115].Their results showed that neither male nor female animals exhibited any peculiar behavioral phenotype when CB1R was selectively deleted in CX3CR1-positive cells, including motor activity, anxiety levels, learning, and memory. These results suggested that there is a possibility that low expression of CB1R in microglia has no clear role on microglia function. Futhermore, Maciej Pietr et al. found that GPR55 is significantly expressed in both the BV-2 cells and primary mouse microglia. Their expression pattern response to the state of microglia is very similar to that of CB2R, suggesting it might also be involved in the neuroinflammatory process [116].
Microglia not only express CB2R responding to cannabinoid ligands but also produce and inactivate the endocannabinoids AEA and 2-AG by the complex cellular machinery [111,117,118]. The endocannabinoids are highly lipophilic molecules, that are naturally synthesized from glia and neuron membrane phospholipids during physiological events in the CNS. As a result, endocannabinoids are released immediately after their production to stimulate CB1R and CB2R, and in turn, they are inactivated by re-uptake and the ensuing hydrolysis [119,120]. Notably, under neuroinflammatory conditions in vitro, microglia are capable of producing 20 times more endocannabinoids than other glial cells and neurons. Therefore, microglia may be the primary cellular source of endocannabinoids in vivo [35,85]. Note that the expression of AEA and 2-AG is significantly upregulated when microglia are activated and switch to a protective phenotype (M2) under pathological conditions [68,103,121]. Further studies have demonstrated that the production of 2-AG and AEA is dependent on sustained rises in intracellular Ca2+ concentration, which may be mediated by the activation of ionotropic P2X purinergic receptors P2X4 and P2X7 [118,122]. The molecular mechanism underlying this process involves the fact that an increase in intracellular Ca2+ can directly increase NAPE-PLD and DAGL activity while inhibiting MAGL activity in microglia [85,123,124]. The inverse sensitivity of DAGL and MAGL to Ca2+ constitutes a primitive and efficient modality for microglia to continuously accumulate 2-AG [118,125]. Additionally, microglia also express the metabolic enzyme FAAH, which is responsible for the degradation of 2-AG and AEA [123]. Several reports have detailed that under neuropathological conditions, activated spinal microglia may produce significantly more 2-AG and AEA by upregulating DAGL and NAPE-PLD, while downregulating FAAH [103,126,127]. Meanwhile, it has been suggested that the exposure of human or rat microglia to 2-AG and AEA at low concentrations increases the expression of the M2 microglial marker arginase 1 (Arg-1), together with other markers such as suppressor of cytokine signaling 3 (SOCS3) [103]. These results support that the production and actions of endocannabinoids are closely related to the phenotype of microglia, and the endocannabinoid system also plays an important role in microglial immunomodulation by inducing an M2 phenotype. It also explains that some of the neuroprotective effects of the endocannabinoid system in diverse pathological states of inflammatory models may be mediated by their immunomodulatory actions, such as disrupting pro-inflammatory processes [128,129,130].
The increased production of endocannabinoids may also upregulate the expression of microglial CB2R, which in turn may activate more CB2R to dampen nociceptive signaling cascades, amplifying the anti-inflammatory responses [103]. Notably, the activation of CB2R has been reported to transform microglia from the M1 to the M2 phenotype [131,132], suggesting that CB2R signaling is very important for microglia to polarize towards the M2 phenotype with phagocytic capacity by morphology alterations [103]. These findings are consistent with previous studies that showed that upregulation of CB2R in activated microglia has been associated with improvement of the disease consequences in specific neuroinflammatory conditions [126,133,134].
In summary, the overall consequence of CB2R activation on microglia by endocannabinoids AEA and 2-AG or by exogenous cannabinoids appears to be to exert beneficial properties of microglia, such as the release of anti-inflammatory mediators, by promoting the generation of neuroprotective microglia phenotype (M2). The M2 phenotype could reduce neuronal hyperexcitation causally involved in central sensitization, with the capacity of phagocytosis and reduction of releasing detrimental factors like pro-inflammatory cytokines and free radicals [15,124,127,135,136,137,138,139,140,141]. The expression profiles of CB2R and endocannabinoids in homeostatic and activated microglia are summarized in Figure 2.

4. The Role of Microglial CB2R in Pathological Pain

4.1. Antinociceptive Effects of Well-Characterized CB2R Selective Agonists

Currently, intense interest has been focused on the use of cannabinoid compounds typically acting upon the CB1R and CB2R for the treatment of pathological pain. These widely researched compounds include endocannabinoids like AEA and 2-AG, phytocannabinoids like Δ-9-tetrahydrocannabinol and cannabidiol, as well as a large number of synthetic cannabinoids [20,142,143,144,145,146]. However, in parallel with a deeper understanding of the endocannabinoid system’s underlying expression profiles and the physiological and pharmacological properties of cannabinoid receptors, CB2R selective agonist compounds are increasingly recognized as safer novel therapeutic candidates, with their properties to bypass certain centrally mediated unwanted effects associated with the activation of CB1R.
The following section summarizes the antinociceptive effects of various well-characterized CB2R selective agonists, including HU308, JWH-015, JWH-133, GW405833, AM1241, and MDA7, in different pain states to provide direct support for the hypothesis that CB2R can serve as a promising therapeutic target for pain relief (Table 1).
The HU308 is the first CB2R selective synthetic compound (Ki = 22.7 ± 3.9 nM) that exhibits low affinity for CB1R (Ki > 10 μM), which exerted anti-inflammatory and peripheral antinociceptive activities in an arachidonic acid-induced mouse inflammatory pain model and the late phase of the mouse formalin pain model. These activities were significantly inhibited by the use of selective CB2R antagonist SR144528 [147]. In the rat postoperative pain model, surgical incision-induced tactile allodynia was significantly suppressed by HU308 [148]. The topical HU308 also has been found to reduce corneal hyperalgesia and inflammation in wild-type mice, but not in CB2R−/− mice, further validating that CB2R is a drug target of this compound [149].
The JWH-015, one of the earliest discovered compounds, was found to have improved selectivity for the CB2R (Ki = 13.8 nM at CB2R and Ki = 383 nM at CB1R) from the aminoalkylindole classification of CB2R agonists, which is effective in alleviating pain-related behaviors and reducing inflammatory responses without inducing psychotropic effects even when intrathecally applied [21,52,150,151,152]. The intrathecal administration of JWH015 reduced the paw incision and caused postoperative hypersensitivity and microglial activation in the spinal cord without inducing behavioral side effects. This effect was prevented by intrathecal injection of the CB2R selective antagonist AM630 [10,153], indicating a CB2R-dependent mechanism of action. In spinal nerve ligation (SNL) or lumbar 5 nerve transection (L5NT) neuropathic pain models, intrathecal JWH015 treatment significantly reduced nerve injury-induced hypersensitivity, which can also be blocked by intrathecal AM630 [13,17]. Additionally, it has also been reported that intrathecal or intraperitoneal injection of JWH-015 displayed an analgesic effect to attenuate bone cancer-induced spontaneous pain and mechanical allodynia [15,92,154].
The JWH-133 is also a well-characterized CB2R agonist, one of the most highly selective ligands for the CB2R (Ki = 677 ± 132 nM at CB1R and Ki = 3.4 ± 1.0 nM at CB2R) [155,156], which inhibits both inflammatory and neuropathic hyperalgesia through a CB2R-selective mechanism. For example, spinal (i.t.) or local (i.pl.) administration of JWH-133 reduced noxious mechanical stimulation evoked responses in wide dynamic range neurons recorded in SNL neuropathic pain, carrageenan-induced inflammatory pain and osteoarthritis pain, in a manner that was prevented by SR144528 [157,158,159]. The JWH-133, administered systemically (s.c.), can increase weight bearing and decrease peripheral edema or allodynia in the carrageenan-inflamed paw and osteoarthritis pain [160]. However, the JWH-133 has been shown to functionally interact with opioids to modulate antinociception in the formalin test without inducing tolerance, and can also attenuate cross-tolerance with morphine [161]. Furthermore, recent studies using CB2R constitutive knockout and tissue-specific genetic deletion mice suggested that self-administration of JWH-133 not only attenuated spontaneous pain and anxiety-associated behavior in the partial sciatic nerve ligation (PSNL) induced neuropathic pain model but also void of reinforcing effects in animals without pain, indicating the absence of abuse liability [14].
The GW405833, another highly selective CB2R ligand, is also classified as an aminoalkylindole [162]. In addition, the GW405833 was determined to be a selective human CB2R agonist in a recombinant binding assay (Ki = 2043 ± 183 nM at CB1R and Ki = 14 ± 6 nM at CB2R), while its selectivity appeared to be lower for the rat CB2R (Ki = 273 ± 42.6 nM at CB1R and Ki = 3.6 ± 1.1 nM at CB2R) [163,164]. In both rats and mice, pharmacological characterization of GW405833 has been previously shown to elicit efficacious antihyperalgesic and anti-inflammatory effects in several pain models, including PSNL, hind paw incision, and complete Freund’s adjuvant (CFA)-induced inflammatory pain, without eliciting the centrally CB1R-mediated side effects [163,165]. Similarly, systemic administration of GW405833 reduced the late phase of formalin pain and allodynia elicited by SNL in a dose-dependent manner [9]. Additionally, hind paw incision, chronic constriction injury (CCI)-induced tactile allodynia, and carrageenan-evoked peripheral edema or weight bearing were relieved by GW405833 [148,166,167]. These above-mentioned effects were demonstrated to be dependent upon CB2R activation rather than the activation of CB1R or opioid receptors by the experiments performed in CB2R−/− mice or utilizing CB2R and opioid receptor selective antagonists. Interestingly, analgesic effects of high-dose GW405833 (i.p. 100 mg/kg) were also evident in the tail flick and hot plate tests in CB2R−/− mice, which might be attributed to both moderate affinities for CB1R and significant CNS penetration [163,165]. Moreover, in stark contrast to treatment with WIN55,212-22, a mixed CB1R/CB2R agonist, chronic repeated injection of GW405833 was able to provide sustained reversal of allodynia following SNL without tolerance development [168].
The AM1241, another agonist possessing a high affinity for the CB2R (Ki = 3.4 ± 0.5 nM at the CB2R and Ki = 280 ± 41 nM at CB1R), belong to the aminoalkylindole class [169]. There is growing evidence supporting the hypothesis that AM1241 produced antinociceptive effects in preclinical inflammatory and neuropathic pain models lacking CNS side effects in a tetrad of behavioral tests that is used to assess cardinal signs of central CB1R [169]. The systemic (i.p.) and local (i.pl.) administration of AM1241 exhibited a thermal antinociceptive effect in the acute pain model of rats, which was significantly blocked by the CB2R selective antagonist AM630 [19]. These actions of CB2R have been confirmed by further studies, in which acute nociception of AM1241 was lost in the tail flick and hot plate tests of CB2R−/− mice [22]. In the rat postoperative pain model, AM1241 also obviously suppressed tactile allodynia [148]. Moreover, systemic (i.p.) or local (i.pl.) administration of AM1241 suppressed allodynia, hyperalgesia, and peripheral edema in the carrageenan-evoked rat inflammatory pain model in a CB2R-dependent manner because SR144528 or AM630 specifically blocked these effects [170,171,172,173]. Similarly, intravenously administered AM1241 reduced the late phase of formalin pain, which also depends upon CB2R activation [9]. In the SNL or CCI of sciatic nerve-induced neuropathic pain models and the CFA-induced chronic inflammatory pain model, AM1241 (i.p., i.DRG., or i.t.) produced a significant reversal of established mechanical and thermal hypersensitivity in rats or CB1R−/− mice [6,9,169,174]. The AM1241 could also reduce pain symptoms in a CB2R dependent manner in the vincristine-induced neuropathic pain model and bone the cancer-induced pain model [93,175,176].
The MDA7, one of the acylhydrazone derivatives, is a more promising CB2R selective agonist (hCB1R Ki > 10,000 nM; hCB2R Ki = 422 nM; rCB1R Ki = 2565 nM; rCB2R Ki = 238 nM) for the treatment of pain [24,177]. It has been shown that systemic administration of MDA7 exhibits an attenuated SNL-induced tactile allodynia in rats in a dose-dependent way. The target specificity of MDA7 was confirmed by pretreatment with selective antagonists, while attenuation of the antiallodynic effects was mediated by AM630 but not either the CB1R selective antagonist AM251 or the opioid antagonist naloxone [24]. This molecule has also been shown to effectively suppress mechanical allodynia rats and mice in paclitaxel (PTX)-induced neuropathic pain models. In addition, MDA7 can produce a modest thermal antinociceptive effect in naive rats without affecting locomotor activity. These effects were blocked after pretreatment with AM630 in wild type mice or were absent in CB2R−/− mice, which indicates that the action of MDA7 directly involves the activation of CB2R [24,25,26].
As a result of the great potential of targeting CB2R, new potential drugs are constantly being developed. Some of them are being tested in clinical trials. For example, Olorinab, an oral and highly selective full agonist of CB2R, reached phase II trials for abdominal pain in Crohn’s disease and for irritable bowel syndrome [178]. However, there is still no CB2R selective agonist on the market as a new analgesic drug. This situation resulted from many reasons. Firstly, most CB2R ligands were highly lipophilic and, as such, not optimal for clinical application due to unfavorable physicochemical properties, which potentially contributed to modest or lack of clinical efficacy. Secondly, these compounds will be required to have high affinity and selectivity for CB2R to avoid the adverse effects of activating CB1R, while the target engagement of current CB2R ligands is poor. Aside from the development and optimization of CB2R ligands, the fact that human and rodent CB2R sequences have relatively low homology should be considered, which may give rise to differences in ligand engagement and efficacy [179]. Furthermore, preclinical pain models in animals might not fully and accurately reflect human pathological mechanisms, which may also affect the clinical translation of CB2R agonists. Nonetheless, substantial efforts to better optimize CB2R ligands for clinical application are ongoing, and many existing ligands have reached the most advanced phases, such as JBT-101 [180,181]. There is no doubt that specifically activating CB2R is considered a good strategy for developing new analgesic agents with fewer side effects.
Table 1. Efficacy of some well-characterized CB2R selective ligands in different pain models.
Table 1. Efficacy of some well-characterized CB2R selective ligands in different pain models.
AgonistIn vitro binding profilePain ModelRoute of AdministrationSpeciesEfficacyReference
CB1RCB2R
HU308Ki > 10 μM Rat brainKi = 22.7 ± 3.9 nM Transfected cellsFormalin testi.p., 50 mg/kgMouseAntinociception[148]
Arachidonic acid-induced ear Inflammationi.p., 50 mg/kgMouseReduce ear swelling[148]
Post-operative paini.p., 0.3–30 mg/kgRatAntiallodynic effect[149]
JWH-015Ki = 383 nMKi = 13.8 nMTail flick testi.p., 1–100 mg/kgMouseAntinociception[21]
Tail immersion test/Paw pressure testi.p., 5–20 mg/kgRatAntinociception[153]
Formalin testi.p., 0.1–100 mg/kgMouseAntinociception[153]
Post-operative paini.t., 2–10 μgRatAntiallodynic effect[10,21,154]
i.p., 1–10 mg/kgRatAntinociception
Inflammatory pain (CFA)i.p., 5–10 mg/kgRatAntinociception[153]
Neuropathic pain (SNI, SNL, L5NT, or bone cancer induced)i.p., 1–10 mg/kgRatAntinociception[13,15,17,21,92,155]
i.t., 0.4–50 μgRatAntiallodynic effect
JWH-133Ki = 677 ± 132 nM Rat brainKi = 3.4 ± 1.0 nM Human embryonic kidney 293 cellsFormalin testi.p., 0.1–10 mg/kgMouseAntinociception[162]
Inflammatory pain (Carrageenan or osteoarthritis induced)s.c. 10 mg/kgRatIncrease weight bearing[158,161]
i.pl., 5–15 μgRatInhibits mechanically evoked neuron responses
Neuropathic pain (SNL or PSNL induced)i.pl., 5–15 μgRatInhibits mechanically evoked neuron responses[14,158,159,160]
i.t., 8–486 ngRatInhibits mechanically evoked neuron responses
s.c. 1 mg/kgRatAntiallodynic effect
i.v., 0.15–0.3 mg/kgMouseAntinociception
GW405833Ki = 2043 ± 183 nM Cos-7 cells Ki = 273 ± 42.6 nM Rat brainKi = 14 ± 6 nM Cos-M6 cells Ki = 3.6 ± 1.1 nM Rat spleenHot plate test/Tail flick testi.p., 100 mg/kgMouseAntinociception[166]
Formalin testi.v., 3–10 mg/kgMouseAntinociception[9]
Post-operative paini.p., 0.3–30 mg/kgRatAntiallodynic effect[149,164]
Inflammatory pain (Carrageenan or CFA)i.p., 3–30 mg/kgMouseAntiallodynic effect[164,166,167]
i.p., 0.1–100 mg/kgRatAntinociception
Neuropathic pain model (PSNL, L5NT, or CCI)i.p., 3–30 mg/kgMouseAntiallodynic effect[9,164,166,168,169]
i.p., 0.01–30 mg/kgRatAntiallodynic effect
AM1241Ki = 280 ± 41 nM Rat brainKi = 3.4 ± 0.5 nM Mouse spleenHargreaves acute thermal stimulationi.p., 0.3–3 mg/kgMouseAntinociception[22]
i.p., 0.033–0.33 mg/kgRatAntinociception[19]
Hot plate test/Tail flick testi.p., 0.3–10 mg/kgMouseAntinociception[22]
Formalin testi.p., 3–10 mg/kgMouseAntinociception[9]
i.v., 0.3–3 mg/kgMouseAntinociception
Post-operative paini.p., 3–30 mg/kgRatAntiallodynic effect[149]
Inflammatory pain (Carrageenan, capsaicin, or CFA induced)i.p., 0.033–1 mg/kgRatAntinociception[6,171,172,173]
i.pl., 0.033–4mg/kgRatAntinociception/Reduce paw edema
i.DRG, 100 nmolRatAntinociception
i.t., 100 nmolRatAntinociception
Neuropathic pain (SNL, bone cancer, vincristine-induced )i.t., 0.03–0.3 μgMouseAntinociception[93,176,177]
[6,9,170,175]
i.p., 0.3–10 mg/kgMouseAntiallodynic effect
i.p., 0.1–3 mg/kgRatAntinociception
i.DRG, 100 nmolRatAntinociception
i.t., 0.01–10 μgRatAntinociception
i.v., 3–6 mg/kgRatAntiallodynic effect
MDA7hKi > 10,000 nM CHO-K1 cells rKi = 2565 nM CHO-K1 cellshKi = 422 nM CHO-K1 cells rKi = 238 nM CHO-K1 cellsHargreaves acute thermal stimulationi.p., 1–10 mg/kgRatAntinociception[24]
Neuropathic pain (SNL or PTX)i.p., 5–15 mg/kgRatAntiallodynic effect[24,25,26]
Abbreviations: CFA: complete Freund’s adjuvant; CCI: chronic constriction injury; i.t.: intratecal; i.p.: intraperitoneal; i.pl.: intraplantar; i.v.: intravenous; i.DRG.: intra Dorsal Root Ganglion, s.c.: subcutaneous; PSNL: partial sciatic nerve ligation; SNL: spinal nerve ligation; L5NT: L5 nerve transection.

4.2. Molecular Mechanisms Involved in the Action of Microglial CB2R in Pain Processing

The dorsal horn of the spinal cord is the vital site for controlling pain intensity because it is where efficient transmission of nociceptive information occurs between the central terminals of primary afferents and second-order interneurons. Additionally, the crosstalk between spinal neurons, astrocytes, oligodendrocytes, and microglia is indispensable to mediating central pain sensitization of neuronal circuits. A considerable amount of evidence has implicated the crucial role of selective agonists of CB2R in treating pathologic pain symptoms through the modulation of microglia in a CB2R dependent manner [17,25,26,29,177,182,183,184,185]. Therefore, microglial CB2R can serve as a promising therapeutic target for pain relief because of the important role of spinal microglia in regulating central sensitization. Later in the remainder of this review, we will focus on the molecular mechanism of targeting spinal cord CB2R to inhibit neuroinflammatory signaling pathways for pain relief with a microglial-centric view.
As described above, reactive microglia express CB2R [17,26,184]. The protein and mRNA levels of spinal cord CB2R were both significantly upregulated in chronic pain conditions, including CFA-induced inflammatory pain [106], SNI, SNL, CCI, and chemotherapy-induced neuropathic pain [29,30]. The above studies indicate that CB2R over-expression in activated microglia at the dorsal horn of the spinal cord under pathological pain conditions may occur as a result of specific neuroinflammatory responses. After that, modulation targeting CB2R may result in a neuroprotective effect [186]. Accordingly, the preponderance of evidence has implicated that activation of the CB2R system via spinal administration of CB2R agonists produces significant control over inflammatory and neuropathic pain in multiple models. For example, intrathecal administration of AM1241 attenuates allodynia or thermal hyperalgesia induced by CFA, SNL [6], CCI, or bone cancer models [93]. The therapeutic utility of other CB2R selective agonists that exhibit analgesic effects to treat different pains has been thoroughly described in Section 4.1.
Below, we will summarize our current understanding of the cellular and molecular mechanisms involved in the action of microglial CB2R in pain processing. However, several studies have demonstrated that spinal CB2R activation limits microglia activity to the ipsilateral dorsal horn, because constitutive knockout of CB2R results in a spread of microgliosis to the contralateral dorsal horn in an arthritis model or after sciatic nerve injury [187,188,189]. The possible mechanisms involved in this process are discussed in detail. Firstly, the activation of CB2R can inhibit the activities of adenylyl cyclase [190] and MAPK [174,191]. The CB2R selective agonists exerted anti-allodynic effects in rats by reducing MAPK (p38 and ERK1/2) phosphorylation and inducing MAPK phosphatases (MKP-1 and MKP-3, the major regulators of MAPKs) expression in the spinal dorsal horn [13]. The downregulation of the p38 MAPK pathway can lead to a reduction of the cytokines IL-1β, TNF-α, and brain-derived neurotrophic factor (BDNF) [174,192,193], and the suppression of ERK1/2 can decrease microglia proliferation [26,194,195]. Furthermore, the AMPK pathway is upregulated after CB2R activation, which can downregulate the synthesis of nitric oxide (NO) [196]. Actually, the activation of CB2R in microglial cells has been found to lead to spinal decreased iNOS, IL-6, BDNF, CCR2, and TNFα receptor expression during neuropathic pain [25,197], and increased release of anti-inflammatory cytokines, such as IL-10 [174,198,199]. Specifically, the activation of spinal CB2R by exercise-induced AEA release also reduces the production of IL-1β and TNFα in mice within a carrageenan-induced pain model [136]. This may also be closely related to the reduction of endocannabinoid degradative enzymes at the spinal cord level. It has been shown that intrathecal AM1241 not only modulates critical glial factors but also reduces the expression levels of MAGL, while not altering FAAH [174]. In addition, CB2R agonist treatment can reduce microglial purinergic receptor P2X4 upregulation [25], which may be another mechanism by which CB2R activation reduces microglial contributions to pain. The P2X4 was identified as a microglia-specific molecule that was activated and upregulated after peripheral nerve injury and also plays critical roles in processing nociceptive information and contributing to microglial-dependent central pain sensitization [200].
In addition to modulating microglial immune function by reducing the production of pro-inflammatory cytokines and increasing the release of anti-inflammatory cytokines involved in neuroinflammatory signaling pathways, activation of CB2R also can switch microglia into a more anti-inflammatory state by limiting migration and promoting phagocytic function. For example, it has been found that 2-AG can induce the recruitment of microglia partly by stimulating CB2R in BV-2 cells [85]. However, another study showed that CB2R activation in microglia stimulates MKP phosphatases, which can inhibit the ERK pathway, thus decreasing microglial chemotaxis/migration mediated by ADP [201]. Moreover, it has been found that the activation of CB2R is capable of inducing the removal of native beta-amyloid both in situ and in vitro by promoting the phagocytic function of macrophages [138]. On the contrary, some results also show that the activation of CB2R can inhibit the phagocytosis of microglia by activating ERK1/2/AKT-Nurr1 signal pathways [202]. These different results of microglial migration or phagocytosis mediated by CB2R activation may be due to the different stages of inflammation development. However, as described in Section 3 of this review, the CB2R activation can promote the shift of M1 to M2 microglia. All of this may enhance the beneficial properties of microglia and be associate with the restoration of microglial activity.
Overall, accumulating evidence reveals that activation of CB2R vitally regulates microglial immune function by blocking the normal inflammatory response with increased production of anti-inflammatory mediators and decreased production of proinflammatory mediators causally involved in central sensitization [29,31,127,135,203,204].

5. Conclusions

In this review article, we summarize the analgesic effects mediated by CB2R and the mechanisms involved in pain regulation. Firstly, it is well known that the endocannabinoid system exerts an important role in neuronal regulation. Within the CNS, CB2R mainly expresses in homeostatic microglia, while there is a unique feature that their expression is rapidly upregulated in activated microglia under certain pathological conditions. The CB2R might serve as an intriguing target for the development of drugs for the management of pain because of its ability to mediate analgesia with few psychoactive effects. Indeed, accumulating data have demonstrated that the CB2R agonists exert analgesic effects in various preclinical pain models, such as inflammatory and neuropathic pain. Additionally, spinal microglia can modulate the activity of spinal cord neurons and have a critical role in the development and maintenance of chronic pain. The activation of CB2R can reduce pain signaling by regulating the activity of spinal microglia and inhibiting neuroinflammation. Specifically, the CB2R activation has been reported to transform microglia from the pro-inflammatory M1 to the neuroprotective M2 phenotype by promoting the beneficial properties of microglia, such as the releasing of anti-inflammatory mediators, or the induction of phagocytosis, and reducing their ability to release pro-inflammatory cytokines involved in central sensitization. Overall, we provided an improved understanding of the underlying mechanisms involved in the action of microglial CB2R in pain processing. However, further studies are needed to dissect the specific role of CB2R expressed in different phenotype microglia to provide a better alternative to controlling pain by regulating CB2R.

Author Contributions

K.X.: writing—original draft preparation, visualization. and conceptualization; Y.W., Z.T., and Y.X.: writing—review and editing; C.W.: writing—review and editing, supervision, and project administration. Z.W.: writing—review and editing, conceptualization, visualization, supervision, and project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC), grant numbers 82101297 and 81471141; Shenzhen Science and Technology Program, grant number JCYJ20220818103206013.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this manuscript.

Acknowledgments

Figure 1 and Figure 2 were created with BioRender.com.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ACAdenylyl cyclase
AEAArachidonoyl ethanolamide
ABDH6/12α/β domain hydrolases 6/12
Arg-1Arginase 1
ATPAdenosine triphosphate
BDNFBrain-derived neurotrophic factor
cAMPcyclic Adenosine monophosphate
CB1RCannabinoid receptors 1
CB2RCannabinoid receptors 2
CCIChronic constriction injury
CFAComplete Freund’s adjuvant
CGRPCalcitonin gene-related peptide
CNSCentral nervous system
COX-2Cyclooxygenase-2
CSF-1Colony-stimulating factor-1
DAGLα/βDiacylglycerol lipases
DRGDorsal root ganglion
ECSEndocannabinoid system
FAAHFatty acid amino hydrolase
IFN-γInterferon γ
IL-1βInterleukin 1β
IL-10Interleukin 10
LPILysophosphatidylinositol
L5NTLumbar 5 nerve transection
MAPKMitogen-activated protein kinases
NONitric oxide
NAPEN-acyl phosphatidyl ethanolamine
NAPE-PLDN-acyl-phosphatidyl ethanolamine-phospholipase D
NATN-acyltransferase
PCPhosphatidylcholine
PEPhosphatidylethanolamine
PGE2Prostaglandin E2
PIP2Phosphatidyl inositol bis-phosphate
PLCPhospholipase C
PNSperipheral nervous systems
PSNLPartial sciatic nerve ligation
PTXPaclitaxel
SNLSpinal nerve ligation
SOCS3Suppressor of cytokine signaling 3
TGF-βTransforming growth factor
TNF-αTumor necrosis factor α
2-AG2-arachidonoyl-glycerol

References

  1. Pratt, M.; Stevens, A.; Thuku, M.; Butler, C.; Skidmore, B.; Wieland, L.S.; Clemons, M.; Kanji, S.; Hutton, B. Benefits and harms of medical cannabis: A scoping review of systematic reviews. Syst. Rev. 2019, 8, 320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Romero-Sandoval, E.A.; Kolano, A.L.; Alvarado-Vazquez, P.A. Cannabis and Cannabinoids for Chronic Pain. Curr. Rheumatol. Rep. 2017, 19, 67. [Google Scholar] [CrossRef] [PubMed]
  3. Blair, R.E.; Deshpande, L.S.; Sombati, S.; Elphick, M.R.; Martin, B.R.; DeLorenzo, R.J. Prolonged exposure to WIN55,212-2 causes downregulation of the CB1 receptor and the development of tolerance to its anticonvulsant effects in the hippocampal neuronal culture model of acquired epilepsy. Neuropharmacology 2009, 57, 208–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Davis, M.P. Cannabinoids in pain management: CB1, CB2 and non-classic receptor ligands. Expert Opin. Investig. Drugs 2014, 23, 1123–1140. [Google Scholar] [CrossRef]
  5. Xu, B.; Xiao, J.; Xu, K.; Zhang, Q.; Chen, D.; Zhang, R.; Zhang, M.; Zhu, H.; Niu, J.; Zheng, T.; et al. VF-13, a chimeric peptide of VD-hemopressin(alpha) and neuropeptide VF, produces potent antinociception with reduced cannabinoid-related side effects. Neuropharmacology 2020, 175, 108178. [Google Scholar] [CrossRef] [PubMed]
  6. Hsieh, G.C.; Pai, M.; Chandran, P.; Hooker, B.A.; Zhu, C.Z.; Salyers, A.K.; Wensink, E.J.; Zhan, C.; Carroll, W.A.; Dart, M.J.; et al. 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. 2011, 162, 428–440. [Google Scholar] [CrossRef] [Green Version]
  7. Guindon, J.; Hohmann, A.G. Cannabinoid CB2 receptors: A therapeutic target for the treatment of inflammatory and neuropathic pain. Br. J. Pharmacol. 2008, 153, 319–334. [Google Scholar] [CrossRef] [Green Version]
  8. Klauke, A.L.; Racz, I.; Pradier, B.; Markert, A.; Zimmer, A.M.; Gertsch, J.; Zimmer, A. The cannabinoid CB(2) receptor-selective phytocannabinoid beta-caryophyllene exerts analgesic effects in mouse models of inflammatory and neuropathic pain. Eur. Neuropsychopharmacol. 2014, 24, 608–620. [Google Scholar] [CrossRef] [Green Version]
  9. Beltramo, M.; Bernardini, N.; Bertorelli, R.; Campanella, M.; Nicolussi, E.; Fredduzzi, S.; Reggiani, A. CB2 receptor-mediated antihyperalgesia: Possible direct involvement of neural mechanisms. Eur. J. Neurosci. 2006, 23, 1530–1538. [Google Scholar] [CrossRef]
  10. Romero-Sandoval, A.; Eisenach, J.C. Spinal cannabinoid receptor type 2 activation reduces hypersensitivity and spinal cord glial activation after paw incision. Anesthesiology 2007, 106, 787–794. [Google Scholar] [CrossRef] [Green Version]
  11. Ikeda, H.; Ikegami, M.; Kai, M.; Ohsawa, M.; Kamei, J. Activation of spinal cannabinoid CB2 receptors inhibits neuropathic pain in streptozotocin-induced diabetic mice. Neuroscience 2013, 250, 446–454. [Google Scholar] [CrossRef] [PubMed]
  12. Wilkerson, J.L.; Gentry, K.R.; Dengler, E.C.; Wallace, J.A.; Kerwin, A.A.; Armijo, L.M.; Kuhn, M.N.; Thakur, G.A.; Makriyannis, A.; Milligan, E.D. Intrathecal cannabilactone CB(2)R agonist, AM1710, controls pathological pain and restores basal cytokine levels. Pain 2012, 153, 1091–1106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Landry, R.P.; Martinez, E.; DeLeo, J.A.; Romero-Sandoval, E.A. Spinal cannabinoid receptor type 2 agonist reduces mechanical allodynia and induces mitogen-activated protein kinase phosphatases in a rat model of neuropathic pain. J. Pain 2012, 13, 836–848. [Google Scholar] [CrossRef] [Green Version]
  14. Cabanero, D.; Ramirez-Lopez, A.; Drews, E.; Schmole, A.; Otte, D.M.; Wawrzczak-Bargiela, A.; Huerga Encabo, H.; Kummer, S.; Ferrer-Montiel, A.; Przewlocki, R.; et al. Protective role of neuronal and lymphoid cannabinoid CB2 receptors in neuropathic pain. Elife 2020, 9, e55582. [Google Scholar] [CrossRef] [PubMed]
  15. Lu, C.; Liu, Y.; Sun, B.; Sun, Y.; Hou, B.; Zhang, Y.; Ma, Z.; Gu, X. 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 2015, 38, 1880–1890. [Google Scholar] [CrossRef]
  16. Wang, C.; Xu, K.; Wang, Y.; Mao, Y.; Huang, Y.; Liang, Y.; Liu, Y.; Hao, J.; Gu, X.; Ma, Z.; et al. Spinal cannabinoid receptor 2 activation reduces hypersensitivity associated with bone cancer pain and improves the integrity of the blood-spinal cord barrier. Reg. Anesth. Pain Med. 2020, 45, 783–791. [Google Scholar] [CrossRef]
  17. Romero-Sandoval, A.; Nutile-McMenemy, N.; DeLeo, J.A. Spinal microglial and perivascular cell cannabinoid receptor type 2 activation reduces behavioral hypersensitivity without tolerance after peripheral nerve injury. Anesthesiology 2008, 108, 722–734. [Google Scholar] [CrossRef] [Green Version]
  18. Rahn, E.J.; Thakur, G.A.; Wood, J.A.; Zvonok, A.M.; Makriyannis, A.; Hohmann, A.G. Pharmacological characterization of AM1710, a putative cannabinoid CB2 agonist from the cannabilactone class: Antinociception without central nervous system side-effects. Pharmacol. Biochem. Behav. 2011, 98, 493–502. [Google Scholar] [CrossRef] [Green Version]
  19. Malan, P.T., Jr.; Ibrahim, M.M.; Deng, H.; Liu, Q.; Mata, H.P.; Vanderah, T.; Porreca, F.; Makriyannis, A. CB2 cannabinoid receptor-mediated peripheral antinociception. Pain 2001, 93, 239–245. [Google Scholar] [CrossRef]
  20. Soliman, N.; Haroutounian, S.; Hohmann, A.G.; Krane, E.; Liao, J.; Macleod, M.; Segelcke, D.; Sena, C.; Thomas, J.; Vollert, J.; et al. Systematic review and meta-analysis of cannabinoids, cannabis-based medicines, and endocannabinoid system modulators tested for antinociceptive effects in animal models of injury-related or pathological persistent pain. Pain 2021, 162, S26–S44. [Google Scholar] [CrossRef] [PubMed]
  21. Grenald, S.A.; Young, M.A.; Wang, Y.; Ossipov, M.H.; Ibrahim, M.M.; Largent-Milnes, T.M.; Vanderah, T.W. Synergistic attenuation of chronic pain using mu opioid and cannabinoid receptor 2 agonists. Neuropharmacology 2017, 116, 59–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ibrahim, M.M.; Rude, M.L.; Stagg, N.J.; Mata, H.P.; Lai, J.; Vanderah, T.W.; Porreca, F.; Buckley, N.E.; Makriyannis, A.; Malan, T.P., Jr. CB2 cannabinoid receptor mediation of antinociception. Pain 2006, 122, 36–42. [Google Scholar] [CrossRef] [PubMed]
  23. Astruc-Diaz, F.; McDaniel, S.W.; Xu, J.J.; Parola, S.; Brown, D.L.; Naguib, M.; Diaz, P. In vivo efficacy of enabling formulations based on hydroxypropyl-beta-cyclodextrins, micellar preparation, and liposomes for the lipophilic cannabinoid CB2 agonist, MDA7. J. Pharm. Sci. 2013, 102, 352–364. [Google Scholar] [CrossRef] [Green Version]
  24. Naguib, M.; Diaz, P.; Xu, J.J.; Astruc-Diaz, F.; Craig, S.; Vivas-Mejia, P.; Brown, D.L. MDA7: A novel selective agonist for CB2 receptors that prevents allodynia in rat neuropathic pain models. Br. J. Pharmacol. 2008, 155, 1104–1116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wu, J.; Hocevar, M.; Bie, B.; Foss, J.F.; Naguib, M. Cannabinoid Type 2 Receptor System Modulates Paclitaxel-Induced Microglial Dysregulation and Central Sensitization in Rats. J. Pain 2019, 20, 501–514. [Google Scholar] [CrossRef] [PubMed]
  26. Naguib, M.; Xu, J.J.; Diaz, P.; Brown, D.L.; Cogdell, D.; Bie, B.; Hu, J.; Craig, S.; Hittelman, W.N. Prevention of paclitaxel-induced neuropathy through activation of the central cannabinoid type 2 receptor system. Anesth. Analg. 2012, 114, 1104–1120. [Google Scholar] [CrossRef] [Green Version]
  27. Atwood, B.K.; Mackie, K. CB2: A cannabinoid receptor with an identity crisis. Br. J. Pharmacol. 2010, 160, 467–479. [Google Scholar] [CrossRef] [Green Version]
  28. Zhang, J.; Hoffert, C.; Vu, H.K.; Groblewski, T.; Ahmad, S.; O’Donnell, D. Induction of CB2 receptor expression in the rat spinal cord of neuropathic but not inflammatory chronic pain models. Eur. J. Neurosci. 2003, 17, 2750–2754. [Google Scholar] [CrossRef]
  29. Racz, I.; Nadal, X.; Alferink, J.; Banos, J.E.; Rehnelt, J.; Martin, M.; Pintado, B.; Gutierrez-Adan, A.; Sanguino, E.; Manzanares, J.; et al. Crucial role of CB(2) cannabinoid receptor in the regulation of central immune responses during neuropathic pain. J. Neurosci. 2008, 28, 12125–12135. [Google Scholar] [CrossRef] [Green Version]
  30. Brownjohn, P.W.; Ashton, J.C. Spinal cannabinoid CB2 receptors as a target for neuropathic pain: An investigation using chronic constriction injury. Neuroscience 2012, 203, 180–193. [Google Scholar] [CrossRef]
  31. Ashton, J.C.; Glass, M. The cannabinoid CB2 receptor as a target for inflammation-dependent neurodegeneration. Curr. Neuropharmacol. 2007, 5, 73–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Shiue, S.J.; Peng, H.Y.; Lin, C.R.; Wang, S.W.; Rau, R.H.; Cheng, J.K. Continuous Intrathecal Infusion of Cannabinoid Receptor Agonists Attenuates Nerve Ligation-Induced Pain in Rats. Reg. Anesth. Pain Med. 2017, 42, 499–506. [Google Scholar] [CrossRef]
  33. Chen, G.; Zhang, Y.Q.; Qadri, Y.J.; Serhan, C.N.; Ji, R.R. Microglia in Pain: Detrimental and Protective Roles in Pathogenesis and Resolution of Pain. Neuron 2018, 100, 1292–1311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Inoue, K.; Tsuda, M. Microglia in neuropathic pain: Cellular and molecular mechanisms and therapeutic potential. Nat. Rev. Neurosci. 2018, 19, 138–152. [Google Scholar] [CrossRef] [PubMed]
  35. Stella, N. Endocannabinoid signaling in microglial cells. Neuropharmacology 2009, 56 (Suppl. S1), 244–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Bie, B.; Wu, J.; Foss, J.F.; Naguib, M. An overview of the cannabinoid type 2 receptor system and its therapeutic potential. Curr. Opin. Anaesthesiol. 2018, 31, 407–414. [Google Scholar] [CrossRef]
  37. Komorowska-Muller, J.A.; Schmole, A.C. CB2 Receptor in Microglia: The Guardian of Self-Control. Int. J. Mol. Sci. 2020, 22, 19. [Google Scholar] [CrossRef]
  38. Lu, H.C.; Mackie, K. An Introduction to the Endogenous Cannabinoid System. Biol. Psychiatry 2016, 79, 516–525. [Google Scholar] [CrossRef] [Green Version]
  39. Schmid, P.C.; Reddy, P.V.; Natarajan, V.; Schmid, H.H. Metabolism of N-acylethanolamine phospholipids by a mammalian phosphodiesterase of the phospholipase D type. J. Biol. Chem. 1983, 258, 9302–9306. [Google Scholar] [CrossRef]
  40. Cravatt, B.F.; Giang, D.K.; Mayfield, S.P.; Boger, D.L.; Lerner, R.A.; Gilula, N.B. Molecular characterization of an enzyme that degrades neuromodulatory fatty-acid amides. Nature 1996, 384, 83–87. [Google Scholar] [CrossRef]
  41. Murataeva, N.; Straiker, A.; Mackie, K. Parsing the players: 2-arachidonoylglycerol synthesis and degradation in the CNS. Br. J. Pharmacol. 2014, 171, 1379–1391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Blankman, J.L.; Simon, G.M.; Cravatt, B.F. A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chem. Biol. 2007, 14, 1347–1356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hermanson, D.J.; Gamble-George, J.C.; Marnett, L.J.; Patel, S. Substrate-selective COX-2 inhibition as a novel strategy for therapeutic endocannabinoid augmentation. Trends Pharmacol. Sci. 2014, 35, 358–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wen, J.; Ribeiro, R.; Tanaka, M.; Zhang, Y. Activation of CB2 receptor is required for the therapeutic effect of ABHD6 inhibition in experimental autoimmune encephalomyelitis. Neuropharmacology 2015, 99, 196–209. [Google Scholar] [CrossRef] [Green Version]
  45. Heimann, A.S.; Dale, C.S.; Guimaraes, F.S.; Reis, R.A.M.; Navon, A.; Shmuelov, M.A.; Rioli, V.; Gomes, I.; Devi, L.L.; Ferro, E.S. Hemopressin as a breakthrough for the cannabinoid field. Neuropharmacology 2021, 183, 108406. [Google Scholar] [CrossRef] [PubMed]
  46. Han, Z.L.; Fang, Q.; Wang, Z.L.; Li, X.H.; Li, N.; Chang, X.M.; Pan, J.X.; Tang, H.Z.; Wang, R. Antinociceptive effects of central administration of the endogenous cannabinoid receptor type 1 agonist VDPVNFKLLSH-OH [(m)VD-hemopressin(alpha)], an N-terminally extended hemopressin peptide. J. Pharmacol. Exp. Ther. 2014, 348, 316–323. [Google Scholar] [CrossRef]
  47. Zheng, T.; Zhang, T.; Zhang, R.; Wang, Z.L.; Han, Z.L.; Li, N.; Li, X.H.; Zhang, M.N.; Xu, B.; Yang, X.L.; et al. Pharmacological characterization of rat VD-hemopressin(alpha), an alpha-hemoglobin-derived peptide exhibiting cannabinoid agonist-like effects in mice. Neuropeptides 2017, 63, 83–90. [Google Scholar] [CrossRef]
  48. Pan, J.X.; Wang, Z.L.; Li, N.; Han, Z.L.; Li, X.H.; Tang, H.H.; Wang, P.; Zheng, T.; Fang, Q.; Wang, R. Analgesic tolerance and cross-tolerance to the cannabinoid receptors ligands hemopressin, VD-hemopressin(alpha) and WIN55,212-2 at the supraspinal level in mice. Neurosci. Lett. 2014, 578, 187–191. [Google Scholar] [CrossRef]
  49. Ibsen, M.S.; Connor, M.; Glass, M. Cannabinoid CB1 and CB2 Receptor Signaling and Bias. Cannabis Cannabinoid Res. 2017, 2, 48–60. [Google Scholar] [CrossRef] [Green Version]
  50. Gonsiorek, W.; Lunn, C.; Fan, X.; Narula, S.; Lundell, D.; Hipkin, R.W. Endocannabinoid 2-arachidonyl glycerol is a full agonist through human type 2 cannabinoid receptor: Antagonism by anandamide. Mol. Pharmacol. 2000, 57, 1045–1050. [Google Scholar]
  51. Luk, T.; Jin, W.; Zvonok, A.; Lu, D.; Lin, X.Z.; Chavkin, C.; Makriyannis, A.; Mackie, K. Identification of a potent and highly efficacious, yet slowly desensitizing CB1 cannabinoid receptor agonist. Br. J. Pharmacol. 2004, 142, 495–500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Showalter, V.M.; Compton, D.R.; Martin, B.R.; Abood, M.E. Evaluation of binding in a transfected cell line expressing a peripheral cannabinoid receptor (CB2): Identification of cannabinoid receptor subtype selective ligands. J. Pharmacol. Exp. Ther. 1996, 278, 989–999. [Google Scholar] [PubMed]
  53. Ryberg, E.; Larsson, N.; Sjogren, S.; Hjorth, S.; Hermansson, N.O.; Leonova, J.; Elebring, T.; Nilsson, K.; Drmota, T.; Greasley, P.J. The orphan receptor GPR55 is a novel cannabinoid receptor. Br. J. Pharmacol. 2007, 152, 1092–1101. [Google Scholar] [CrossRef] [PubMed]
  54. Kotsikorou, E.; Madrigal, K.E.; Hurst, D.P.; Sharir, H.; Lynch, D.L.; Heynen-Genel, S.; Milan, L.B.; Chung, T.D.; Seltzman, H.H.; Bai, Y.; et al. Identification of the GPR55 agonist binding site using a novel set of high-potency GPR55 selective ligands. Biochemistry 2011, 50, 5633–5647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Oka, S.; Nakajima, K.; Yamashita, A.; Kishimoto, S.; Sugiura, T. Identification of GPR55 as a lysophosphatidylinositol receptor. Biochem. Biophys. Res. Commun. 2007, 362, 928–934. [Google Scholar] [CrossRef] [PubMed]
  56. Ross, R.A. The enigmatic pharmacology of GPR55. Trends Pharmacol. Sci. 2009, 30, 156–163. [Google Scholar] [CrossRef]
  57. Pertwee, R.G.; Howlett, A.C.; Abood, M.E.; Alexander, S.P.; Di Marzo, V.; Elphick, M.R.; Greasley, P.J.; Hansen, H.S.; Kunos, G.; Mackie, K.; et al. International Union of Basic and Clinical Pharmacology. LXXIX. Cannabinoid receptors and their ligands: Beyond CB(1) and CB(2). Pharmacol. Rev. 2010, 62, 588–631. [Google Scholar] [CrossRef] [Green Version]
  58. Kumar, R.N.; Chambers, W.A.; Pertwee, R.G. Pharmacological actions and therapeutic uses of cannabis and cannabinoids. Anaesthesia 2001, 56, 1059–1068. [Google Scholar] [CrossRef]
  59. Pertwee, R.G. Pharmacology of cannabinoid CB1 and CB2 receptors. Pharmacol. Ther. 1997, 74, 129–180. [Google Scholar] [CrossRef]
  60. La Manno, G.; Siletti, K.; Furlan, A.; Gyllborg, D.; Vinsland, E.; Mossi Albiach, A.; Mattsson Langseth, C.; Khven, I.; Lederer, A.R.; Dratva, L.M.; et al. Molecular architecture of the developing mouse brain. Nature 2021, 596, 92–96. [Google Scholar] [CrossRef]
  61. Zhang, Y.; Ke, J.; Zhou, Y.; Liu, X.; Huang, T.; Wang, F. Sex-specific characteristics of cells expressing the cannabinoid 1 receptor in the dorsal horn of the lumbar spinal cord. J. Comp. Neurol. 2022, 530, 2451–2473. [Google Scholar] [CrossRef] [PubMed]
  62. Morales, P.; Hurst, D.P.; Reggio, P.H. Molecular Targets of the Phytocannabinoids: A Complex Picture. Prog. Chem. Org. Nat. Prod. 2017, 103, 103–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kaminski, N.E.; Abood, M.E.; Kessler, F.K.; Martin, B.R.; Schatz, A.R. Identification of a functionally relevant cannabinoid receptor on mouse spleen cells that is involved in cannabinoid-mediated immune modulation. Mol. Pharmacol. 1992, 42, 736–742. [Google Scholar] [PubMed]
  64. Bouaboula, M.; Rinaldi, M.; Carayon, P.; Carillon, C.; Delpech, B.; Shire, D.; Le Fur, G.; Casellas, P. Cannabinoid-receptor expression in human leukocytes. Eur. J. Biochem. 1993, 214, 173–180. [Google Scholar] [CrossRef]
  65. Gerard, C.M.; Mollereau, C.; Vassart, G.; Parmentier, M. Molecular cloning of a human cannabinoid receptor which is also expressed in testis. Biochem. J. 1991, 279 Pt 1, 129–134. [Google Scholar] [CrossRef]
  66. Stander, S.; Schmelz, M.; Metze, D.; Luger, T.; Rukwied, R. Distribution of cannabinoid receptor 1 (CB1) and 2 (CB2) on sensory nerve fibers and adnexal structures in human skin. J. Dermatol. Sci. 2005, 38, 177–188. [Google Scholar] [CrossRef]
  67. Hill, K.P.; Palastro, M.D.; Johnson, B.; Ditre, J.W. Cannabis and Pain: A Clinical Review. Cannabis Cannabinoid Res. 2017, 2, 96–104. [Google Scholar] [CrossRef] [Green Version]
  68. Mitrirattanakul, S.; Ramakul, N.; Guerrero, A.V.; Matsuka, Y.; Ono, T.; Iwase, H.; Mackie, K.; Faull, K.F.; Spigelman, I. Site-specific increases in peripheral cannabinoid receptors and their endogenous ligands in a model of neuropathic pain. Pain 2006, 126, 102–114. [Google Scholar] [CrossRef] [Green Version]
  69. Svizenska, I.; Dubovy, P.; Sulcova, A. Cannabinoid receptors 1 and 2 (CB1 and CB2), their distribution, ligands and functional involvement in nervous system structures—A short review. Pharmacol. Biochem. Behav. 2008, 90, 501–511. [Google Scholar] [CrossRef]
  70. Stella, N. Cannabinoid and cannabinoid-like receptors in microglia, astrocytes, and astrocytomas. Glia 2010, 58, 1017–1030. [Google Scholar] [CrossRef] [Green Version]
  71. Mecha, M.; Carrillo-Salinas, F.J.; Feliu, A.; Mestre, L.; Guaza, C. Microglia activation states and cannabinoid system: Therapeutic implications. Pharmacol. Ther. 2016, 166, 40–55. [Google Scholar] [CrossRef] [PubMed]
  72. Han, J.; Kesner, P.; Metna-Laurent, M.; Duan, T.; Xu, L.; Georges, F.; Koehl, M.; Abrous, D.N.; Mendizabal-Zubiaga, J.; Grandes, P.; et al. Acute cannabinoids impair working memory through astroglial CB1 receptor modulation of hippocampal LTD. Cell 2012, 148, 1039–1050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Molina-Holgado, E.; Vela, J.M.; Arevalo-Martin, A.; Almazan, G.; Molina-Holgado, F.; Borrell, J.; Guaza, C. Cannabinoids promote oligodendrocyte progenitor survival: Involvement of cannabinoid receptors and phosphatidylinositol-3 kinase/Akt signaling. J. Neurosci. 2002, 22, 9742–9753. [Google Scholar] [CrossRef] [PubMed]
  74. Munro, S.; Thomas, K.L.; Abu-Shaar, M. Molecular characterization of a peripheral receptor for cannabinoids. Nature 1993, 365, 61–65. [Google Scholar] [CrossRef]
  75. Turcotte, D.; Le Dorze, J.A.; Esfahani, F.; Frost, E.; Gomori, A.; Namaka, M. Examining the roles of cannabinoids in pain and other therapeutic indications: A review. Expert. Opin. Pharmacother. 2010, 11, 17–31. [Google Scholar] [CrossRef]
  76. Ramirez, S.H.; Hasko, J.; Skuba, A.; Fan, S.; Dykstra, H.; McCormick, R.; Reichenbach, N.; Krizbai, I.; Mahadevan, A.; Zhang, M.; et al. Activation of cannabinoid receptor 2 attenuates leukocyte-endothelial cell interactions and blood-brain barrier dysfunction under inflammatory conditions. J. Neurosci. 2012, 32, 4004–4016. [Google Scholar] [CrossRef] [Green Version]
  77. Carlisle, S.J.; Marciano-Cabral, F.; Staab, A.; Ludwick, C.; Cabral, G.A. Differential expression of the CB2 cannabinoid receptor by rodent macrophages and macrophage-like cells in relation to cell activation. Int. Immunopharmacol. 2002, 2, 69–82. [Google Scholar] [CrossRef]
  78. Julien, B.; Grenard, P.; Teixeira-Clerc, F.; Van Nhieu, J.T.; Li, L.; Karsak, M.; Zimmer, A.; Mallat, A.; Lotersztajn, S. Antifibrogenic role of the cannabinoid receptor CB2 in the liver. Gastroenterology 2005, 128, 742–755. [Google Scholar] [CrossRef]
  79. Juan-Pico, P.; Fuentes, E.; Bermudez-Silva, F.J.; Javier Diaz-Molina, F.; Ripoll, C.; Rodriguez de Fonseca, F.; Nadal, A. Cannabinoid receptors regulate Ca(2+) signals and insulin secretion in pancreatic beta-cell. Cell Calcium 2006, 39, 155–162. [Google Scholar] [CrossRef]
  80. Idris, A.I.; van ‘t Hof, R.J.; Greig, I.R.; Ridge, S.A.; Baker, D.; Ross, R.A.; Ralston, S.H. Regulation of bone mass, bone loss and osteoclast activity by cannabinoid receptors. Nat. Med. 2005, 11, 774–779. [Google Scholar] [CrossRef] [Green Version]
  81. Abrams, D.I.; Guzman, M. Cannabis in cancer care. Clin. Pharmacol. Ther. 2015, 97, 575–586. [Google Scholar] [CrossRef] [PubMed]
  82. Wu, J. Cannabis, cannabinoid receptors, and endocannabinoid system: Yesterday, today, and tomorrow. Acta Pharmacol. Sin. 2019, 40, 297–299. [Google Scholar] [CrossRef] [PubMed]
  83. Van Sickle, M.D.; Duncan, M.; Kingsley, P.J.; Mouihate, A.; Urbani, P.; Mackie, K.; Stella, N.; Makriyannis, A.; Piomelli, D.; Davison, J.S.; et al. Identification and functional characterization of brainstem cannabinoid CB2 receptors. Science 2005, 310, 329–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Jordan, C.J.; Xi, Z.X. Progress in brain cannabinoid CB2 receptor research: From genes to behavior. Neurosci. Biobehav. Rev. 2019, 98, 208–220. [Google Scholar] [CrossRef] [PubMed]
  85. Walter, L.; Franklin, A.; Witting, A.; Wade, C.; Xie, Y.; Kunos, G.; Mackie, K.; Stella, N. Nonpsychotropic cannabinoid receptors regulate microglial cell migration. J. Neurosci. 2003, 23, 1398–1405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Chen, D.J.; Gao, M.; Gao, F.F.; Su, Q.X.; Wu, J. Brain cannabinoid receptor 2: Expression, function and modulation. Acta Pharmacol. Sin. 2017, 38, 312–316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Tanaka, M.; Sackett, S.; Zhang, Y. Endocannabinoid Modulation of Microglial Phenotypes in Neuropathology. Front. Neurol. 2020, 11, 87. [Google Scholar] [CrossRef]
  88. Cabral, G.A.; Griffin-Thomas, L. Emerging role of the cannabinoid receptor CB2 in immune regulation: Therapeutic prospects for neuroinflammation. Expert Rev. Mol. Med. 2009, 11, e3. [Google Scholar] [CrossRef] [Green Version]
  89. Viscomi, M.T.; Oddi, S.; Latini, L.; Pasquariello, N.; Florenzano, F.; Bernardi, G.; Molinari, M.; Maccarrone, M. Selective CB2 receptor agonism protects central neurons from remote axotomy-induced apoptosis through the PI3K/Akt pathway. J. Neurosci. 2009, 29, 4564–4570. [Google Scholar] [CrossRef] [Green Version]
  90. Ishiguro, H.; Horiuchi, Y.; Ishikawa, M.; Koga, M.; Imai, K.; Suzuki, Y.; Morikawa, M.; Inada, T.; Watanabe, Y.; Takahashi, M.; et al. Brain cannabinoid CB2 receptor in schizophrenia. Biol. Psychiatry 2010, 67, 974–982. [Google Scholar] [CrossRef]
  91. Maresz, K.; Carrier, E.J.; Ponomarev, E.D.; Hillard, C.J.; Dittel, B.N. Modulation of the cannabinoid CB2 receptor in microglial cells in response to inflammatory stimuli. J. Neurochem. 2005, 95, 437–445. [Google Scholar] [CrossRef] [PubMed]
  92. Lu, C.; Shi, L.; Sun, B.; Zhang, Y.; Hou, B.; Sun, Y.; Ma, Z.; Gu, X. A Single Intrathecal or Intraperitoneal Injection of CB2 Receptor Agonist Attenuates Bone Cancer Pain and Induces a Time-Dependent Modification of GRK2. Cell. Mol. Neurobiol. 2017, 37, 101–109. [Google Scholar] [CrossRef] [PubMed]
  93. Curto-Reyes, V.; Llames, S.; Hidalgo, A.; Menendez, L.; Baamonde, A. Spinal and peripheral analgesic effects of the CB2 cannabinoid receptor agonist AM1241 in two models of bone cancer-induced pain. Br. J. Pharmacol. 2010, 160, 561–573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Miller, A.M.; Stella, N. CB2 receptor-mediated migration of immune cells: It can go either way. Br. J. Pharmacol. 2008, 153, 299–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Nimmerjahn, A.; Kirchhoff, F.; Helmchen, F. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 2005, 308, 1314–1318. [Google Scholar] [CrossRef] [Green Version]
  96. Nayak, D.; Roth, T.L.; McGavern, D.B. Microglia development and function. Annu. Rev. Immunol. 2014, 32, 367–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kohno, K.; Shirasaka, R.; Yoshihara, K.; Mikuriya, S.; Tanaka, K.; Takanami, K.; Inoue, K.; Sakamoto, H.; Ohkawa, Y.; Masuda, T.; et al. A spinal microglia population involved in remitting and relapsing neuropathic pain. Science 2022, 376, 86–90. [Google Scholar] [CrossRef]
  98. Tansley, S.; Gu, N.; Guzman, A.U.; Cai, W.; Wong, C.; Lister, K.C.; Munoz-Pino, E.; Yousefpour, N.; Roome, R.B.; Heal, J.; et al. Microglia-mediated degradation of perineuronal nets promotes pain. Science 2022, 377, 80–86. [Google Scholar] [CrossRef]
  99. Gao, Y.J.; Ji, R.R. Chemokines, neuronal-glial interactions, and central processing of neuropathic pain. Pharmacol. Ther. 2010, 126, 56–68. [Google Scholar] [CrossRef] [Green Version]
  100. Guan, Z.; Kuhn, J.A.; Wang, X.; Colquitt, B.; Solorzano, C.; Vaman, S.; Guan, A.K.; Evans-Reinsch, Z.; Braz, J.; Devor, M.; et al. Injured sensory neuron-derived CSF1 induces microglial proliferation and DAP12-dependent pain. Nat. Neurosci. 2016, 19, 94–101. [Google Scholar] [CrossRef] [Green Version]
  101. Berta, T.; Park, C.K.; Xu, Z.Z.; Xie, R.G.; Liu, T.; Lu, N.; Liu, Y.C.; Ji, R.R. Extracellular caspase-6 drives murine inflammatory pain via microglial TNF-alpha secretion. J. Clin. Investig. 2014, 124, 1173–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Ji, R.R.; Berta, T.; Nedergaard, M. Glia and pain: Is chronic pain a gliopathy? Pain 2013, 154 (Suppl. S1), S10–S28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Mecha, M.; Feliu, A.; Carrillo-Salinas, F.J.; Rueda-Zubiaurre, A.; Ortega-Gutierrez, S.; de Sola, R.G.; Guaza, C. Endocannabinoids drive the acquisition of an alternative phenotype in microglia. Brain Behav. Immun. 2015, 49, 233–245. [Google Scholar] [CrossRef] [PubMed]
  104. Fenn, A.M.; Hall, J.C.; Gensel, J.C.; Popovich, P.G.; Godbout, J.P. IL-4 signaling drives a unique arginase+/IL-1beta+ microglia phenotype and recruits macrophages to the inflammatory CNS: Consequences of age-related deficits in IL-4Ralpha after traumatic spinal cord injury. J. Neurosci. 2014, 34, 8904–8917. [Google Scholar] [CrossRef] [PubMed]
  105. Hu, X.; Leak, R.K.; Shi, Y.; Suenaga, J.; Gao, Y.; Zheng, P.; Chen, J. Microglial and macrophage polarization-new prospects for brain repair. Nat. Rev. Neurol. 2015, 11, 56–64. [Google Scholar] [CrossRef] [PubMed]
  106. Saijo, K.; Glass, C.K. Microglial cell origin and phenotypes in health and disease. Nat. Rev. Immunol. 2011, 11, 775–787. [Google Scholar] [CrossRef] [PubMed]
  107. Nozaki, C.; Nent, E.; Bilkei-Gorzo, A.; Zimmer, A. Involvement of leptin signaling in the development of cannabinoid CB2 receptor-dependent mirror image pain. Sci. Rep. 2018, 8, 10827. [Google Scholar] [CrossRef] [Green Version]
  108. Lisboa, S.F.; Gomes, F.V.; Guimaraes, F.S.; Campos, A.C. Microglial Cells as a Link between Cannabinoids and the Immune Hypothesis of Psychiatric Disorders. Front. Neurol. 2016, 7, 5. [Google Scholar] [CrossRef] [Green Version]
  109. van den Hoogen, N.J.; Harding, E.K.; Davidson, C.E.D.; Trang, T. Cannabinoids in Chronic Pain: Therapeutic Potential Through Microglia Modulation. Front. Neural Circuits 2021, 15, 816747. [Google Scholar] [CrossRef]
  110. Klegeris, A.; Bissonnette, C.J.; McGeer, P.L. Reduction of human monocytic cell neurotoxicity and cytokine secretion by ligands of the cannabinoid-type CB2 receptor. Br. J. Pharmacol. 2003, 139, 775–786. [Google Scholar] [CrossRef] [Green Version]
  111. Facchinetti, F.; Del Giudice, E.; Furegato, S.; Passarotto, M.; Leon, A. Cannabinoids ablate release of TNFalpha in rat microglial cells stimulated with lypopolysaccharide. Glia 2003, 41, 161–168. [Google Scholar] [CrossRef] [PubMed]
  112. Du, Y.; Ren, P.; Wang, Q.; Jiang, S.K.; Zhang, M.; Li, J.Y.; Wang, L.L.; Guan, D.W. Cannabinoid 2 receptor attenuates inflammation during skin wound healing by inhibiting M1 macrophages rather than activating M2 macrophages. J. Inflamm. 2018, 15, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Waksman, Y.; Olson, J.M.; Carlisle, S.J.; Cabral, G.A. The central cannabinoid receptor (CB1) mediates inhibition of nitric oxide production by rat microglial cells. J. Pharmacol. Exp. Ther. 1999, 288, 1357–1366. [Google Scholar]
  114. Moreno-Garcia, A.; Bernal-Chico, A.; Colomer, T.; Rodriguez-Antiguedad, A.; Matute, C.; Mato, S. Gene Expression Analysis of Astrocyte and Microglia Endocannabinoid Signaling during Autoimmune Demyelination. Biomolecules 2020, 10, 1228. [Google Scholar] [CrossRef]
  115. De Meij, J.; Alfanek, Z.; Morel, L.; Decoeur, F.; Leyrolle, Q.; Picard, K.; Carrier, M.; Aubert, A.; Sere, A.; Lucas, C.; et al. Microglial Cannabinoid Type 1 Receptor Regulates Brain Inflammation in a Sex-Specific Manner. Cannabis Cannabinoid Res. 2021, 6, 488–507. [Google Scholar] [CrossRef]
  116. Pietr, M.; Kozela, E.; Levy, R.; Rimmerman, N.; Lin, Y.H.; Stella, N.; Vogel, Z.; Juknat, A. Differential changes in GPR55 during microglial cell activation. FEBS Lett. 2009, 583, 2071–2076. [Google Scholar] [CrossRef] [Green Version]
  117. Walter, L.; Stella, N. Cannabinoids and neuroinflammation. Br. J. Pharmacol. 2004, 141, 775–785. [Google Scholar] [CrossRef] [Green Version]
  118. Witting, A.; Walter, L.; Wacker, J.; Moller, T.; Stella, N. P2X7 receptors control 2-arachidonoylglycerol production by microglial cells. Proc. Natl. Acad. Sci. USA 2004, 101, 3214–3219. [Google Scholar] [CrossRef] [Green Version]
  119. Alger, B.E.; Kim, J. Supply and demand for endocannabinoids. Trends Neurosci. 2011, 34, 304–315. [Google Scholar] [CrossRef] [Green Version]
  120. Mechoulam, R.; Parker, L.A. The endocannabinoid system and the brain. Annu. Rev. Psychol. 2013, 64, 21–47. [Google Scholar] [CrossRef] [Green Version]
  121. Witting, A.; Chen, L.; Cudaback, E.; Straiker, A.; Walter, L.; Rickman, B.; Moller, T.; Brosnan, C.; Stella, N. Experimental autoimmune encephalomyelitis disrupts endocannabinoid-mediated neuroprotection. Proc. Natl. Acad. Sci. USA 2006, 103, 6362–6367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Di Marzo, V. Targeting the endocannabinoid system: To enhance or reduce? Nat. Rev. Drug Discov. 2008, 7, 438–455. [Google Scholar] [CrossRef]
  123. Araujo, D.J.; Tjoa, K.; Saijo, K. The Endocannabinoid System as a Window Into Microglial Biology and Its Relationship to Autism. Front. Cell Neurosci. 2019, 13, 424. [Google Scholar] [CrossRef]
  124. Carrier, E.J.; Kearn, C.S.; Barkmeier, A.J.; Breese, N.M.; Yang, W.; Nithipatikom, K.; Pfister, S.L.; Campbell, W.B.; Hillard, C.J. Cultured rat microglial cells synthesize the endocannabinoid 2-arachidonylglycerol, which increases proliferation via a CB2 receptor-dependent mechanism. Mol. Pharmacol. 2004, 65, 999–1007. [Google Scholar] [CrossRef] [Green Version]
  125. Maccarrone, M.; Rossi, S.; Bari, M.; De Chiara, V.; Fezza, F.; Musella, A.; Gasperi, V.; Prosperetti, C.; Bernardi, G.; Finazzi-Agro, A.; et al. Anandamide inhibits metabolism and physiological actions of 2-arachidonoylglycerol in the striatum. Nat. Neurosci. 2008, 11, 152–159. [Google Scholar] [CrossRef]
  126. Garcia-Ovejero, D.; Arevalo-Martin, A.; Petrosino, S.; Docagne, F.; Hagen, C.; Bisogno, T.; Watanabe, M.; Guaza, C.; Di Marzo, V.; Molina-Holgado, E. The endocannabinoid system is modulated in response to spinal cord injury in rats. Neurobiol. Dis. 2009, 33, 57–71. [Google Scholar] [CrossRef] [Green Version]
  127. Eljaschewitsch, E.; Witting, A.; Mawrin, C.; Lee, T.; Schmidt, P.M.; Wolf, S.; Hoertnagl, H.; Raine, C.S.; Schneider-Stock, R.; Nitsch, R.; et al. The endocannabinoid anandamide protects neurons during CNS inflammation by induction of MKP-1 in microglial cells. Neuron 2006, 49, 67–79. [Google Scholar] [CrossRef] [Green Version]
  128. Maroof, N.; Pardon, M.C.; Kendall, D.A. Endocannabinoid signalling in Alzheimer’s disease. Biochem. Soc. Trans. 2013, 41, 1583–1587. [Google Scholar] [CrossRef] [Green Version]
  129. Arevalo-Martin, A.; Garcia-Ovejero, D.; Gomez, O.; Rubio-Araiz, A.; Navarro-Galve, B.; Guaza, C.; Molina-Holgado, E.; Molina-Holgado, F. CB2 cannabinoid receptors as an emerging target for demyelinating diseases: From neuroimmune interactions to cell replacement strategies. Br. J. Pharmacol. 2008, 153, 216–225. [Google Scholar] [CrossRef] [Green Version]
  130. Murphy, N.; Cowley, T.R.; Blau, C.W.; Dempsey, C.N.; Noonan, J.; Gowran, A.; Tanveer, R.; Olango, W.M.; Finn, D.P.; Campbell, V.A.; et al. The fatty acid amide hydrolase inhibitor URB597 exerts anti-inflammatory effects in hippocampus of aged rats and restores an age-related deficit in long-term potentiation. J. Neuroinflamm. 2012, 9, 79. [Google Scholar] [CrossRef] [Green Version]
  131. Orihuela, R.; McPherson, C.A.; Harry, G.J. Microglial M1/M2 polarization and metabolic states. Br. J. Pharmacol. 2016, 173, 649–665. [Google Scholar] [CrossRef] [Green Version]
  132. Ma, L.; Jia, J.; Liu, X.; Bai, F.; Wang, Q.; Xiong, L. Activation of murine microglial N9 cells is attenuated through cannabinoid receptor CB2 signaling. Biochem. Biophys. Res. Commun. 2015, 458, 92–97. [Google Scholar] [CrossRef] [PubMed]
  133. Miller, L.K.; Devi, L.A. The highs and lows of cannabinoid receptor expression in disease: Mechanisms and their therapeutic implications. Pharmacol. Rev. 2011, 63, 461–470. [Google Scholar] [CrossRef] [PubMed]
  134. Lu, H.C.; Mackie, K. Review of the Endocannabinoid System. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2021, 6, 607–615. [Google Scholar] [CrossRef] [PubMed]
  135. Correa, F.; Docagne, F.; Mestre, L.; Clemente, D.; Hernangomez, M.; Loria, F.; Guaza, C. A role for CB2 receptors in anandamide signalling pathways involved in the regulation of IL-12 and IL-23 in microglial cells. Biochem. Pharmacol. 2009, 77, 86–100. [Google Scholar] [CrossRef] [Green Version]
  136. Dos Santos, R.S.; Sorgi, C.A.; Peti, A.P.F.; Veras, F.P.; Faccioli, L.H.; Galdino, G. Involvement of Spinal Cannabinoid CB2 Receptors in Exercise-Induced Antinociception. Neuroscience 2019, 418, 177–188. [Google Scholar] [CrossRef]
  137. Ramirez, B.G.; Blazquez, C.; Gomez del Pulgar, T.; Guzman, M.; de Ceballos, M.L. Prevention of Alzheimer’s disease pathology by cannabinoids: Neuroprotection mediated by blockade of microglial activation. J. Neurosci. 2005, 25, 1904–1913. [Google Scholar] [CrossRef] [Green Version]
  138. Tolon, R.M.; Nunez, E.; Pazos, M.R.; Benito, C.; Castillo, A.I.; Martinez-Orgado, J.A.; Romero, J. The activation of cannabinoid CB2 receptors stimulates in situ and in vitro beta-amyloid removal by human macrophages. Brain Res. 2009, 1283, 148–154. [Google Scholar] [CrossRef]
  139. Malek, N.; Popiolek-Barczyk, K.; Mika, J.; Przewlocka, B.; Starowicz, K. Anandamide, Acting via CB2 Receptors, Alleviates LPS-Induced Neuroinflammation in Rat Primary Microglial Cultures. Neural Plast. 2015, 2015, 130639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Spittau, B.; Wullkopf, L.; Zhou, X.; Rilka, J.; Pfeifer, D.; Krieglstein, K. Endogenous transforming growth factor-beta promotes quiescence of primary microglia in vitro. Glia 2013, 61, 287–300. [Google Scholar] [CrossRef]
  141. Rom, S.; Persidsky, Y. Cannabinoid receptor 2: Potential role in immunomodulation and neuroinflammation. J. Neuroimmune Pharmacol. 2013, 8, 608–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Guindon, J.; Desroches, J.; Beaulieu, P. The antinociceptive effects of intraplantar injections of 2-arachidonoyl glycerol are mediated by cannabinoid CB2 receptors. Br. J. Pharmacol. 2007, 150, 693–701. [Google Scholar] [CrossRef] [Green Version]
  143. Guindon, J.; Guijarro, A.; Piomelli, D.; Hohmann, A.G. Peripheral antinociceptive effects of inhibitors of monoacylglycerol lipase in a rat model of inflammatory pain. Br. J. Pharmacol. 2011, 163, 1464–1478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Sokal, D.M.; Elmes, S.J.; Kendall, D.A.; Chapman, V. Intraplantar injection of anandamide inhibits mechanically-evoked responses of spinal neurones via activation of CB2 receptors in anaesthetised rats. Neuropharmacology 2003, 45, 404–411. [Google Scholar] [CrossRef] [PubMed]
  145. Calignano, A.; La Rana, G.; Giuffrida, A.; Piomelli, D. Control of pain initiation by endogenous cannabinoids. Nature 1998, 394, 277–281. [Google Scholar] [CrossRef] [Green Version]
  146. Jaggar, S.I.; Hasnie, F.S.; Sellaturay, S.; Rice, A.S. The anti-hyperalgesic actions of the cannabinoid anandamide and the putative CB2 receptor agonist palmitoylethanolamide in visceral and somatic inflammatory pain. Pain 1998, 76, 189–199. [Google Scholar] [CrossRef]
  147. Hanus, L.; Breuer, A.; Tchilibon, S.; Shiloah, S.; Goldenberg, D.; Horowitz, M.; Pertwee, R.G.; Ross, R.A.; Mechoulam, R.; Fride, E. HU-308: A specific agonist for CB(2), a peripheral cannabinoid receptor. Proc. Natl. Acad. Sci. USA 1999, 96, 14228–14233. [Google Scholar] [CrossRef] [Green Version]
  148. LaBuda, C.J.; Koblish, M.; Little, P.J. Cannabinoid CB2 receptor agonist activity in the hindpaw incision model of postoperative pain. Eur. J. Pharmacol. 2005, 527, 172–174. [Google Scholar] [CrossRef]
  149. Thapa, D.; Cairns, E.A.; Szczesniak, A.M.; Toguri, J.T.; Caldwell, M.D.; Kelly, M.E.M. The Cannabinoids Delta(8)THC, CBD, and HU-308 Act via Distinct Receptors to Reduce Corneal Pain and Inflammation. Cannabis Cannabinoid Res. 2018, 3, 11–20. [Google Scholar] [CrossRef] [Green Version]
  150. Huffman, J.W.; Padgett, L.W. Recent developments in the medicinal chemistry of cannabimimetic indoles, pyrroles and indenes. Curr. Med. Chem. 2005, 12, 1395–1411. [Google Scholar] [CrossRef]
  151. Huffman, J.W.; Zengin, G.; Wu, M.J.; Lu, J.; Hynd, G.; Bushell, K.; Thompson, A.L.; Bushell, S.; Tartal, C.; Hurst, D.P.; et al. Structure-activity relationships for 1-alkyl-3-(1-naphthoyl)indoles at the cannabinoid CB(1) and CB(2) receptors: Steric and electronic effects of naphthoyl substituents. New highly selective CB(2) receptor agonists. Bioorg. Med. Chem. 2005, 13, 89–112. [Google Scholar] [CrossRef]
  152. Craft, R.M.; Greene, N.Z.; Wakley, A.A. Antinociceptive effects of JWH015 in female and male rats. Behav. Pharmacol. 2018, 29, 280–289. [Google Scholar] [CrossRef]
  153. Sun, Y.; Zhang, W.; Liu, Y.; Liu, X.; Ma, Z.; Gu, X. Intrathecal injection of JWH015 attenuates remifentanil-induced postoperative hyperalgesia by inhibiting activation of spinal glia in a rat model. Anesth. Analg. 2014, 118, 841–853. [Google Scholar] [CrossRef] [Green Version]
  154. Mao, Y.; Huang, Y.; Zhang, Y.; Wang, C.; Wu, H.; Tian, X.; Liu, Y.; Hou, B.; Liang, Y.; Rong, H.; et al. Cannabinoid receptor 2selective agonist JWH015 attenuates bone cancer pain through the amelioration of impaired autophagy flux induced by inflammatory mediators in the spinal cord. Mol. Med. Rep. 2019, 20, 5100–5110. [Google Scholar] [CrossRef] [Green Version]
  155. Huffman, J.W.; Liddle, J.; Yu, S.; Aung, M.M.; Abood, M.E.; Wiley, J.L.; Martin, B.R. 3-(1′,1′-Dimethylbutyl)-1-deoxy-delta8-THC and related compounds: Synthesis of selective ligands for the CB2 receptor. Bioorg. Med. Chem. 1999, 7, 2905–2914. [Google Scholar] [CrossRef]
  156. Jonsson, K.O.; Persson, E.; Fowler, C.J. The cannabinoid CB2 receptor selective agonist JWH133 reduces mast cell oedema in response to compound 48/80 in vivo but not the release of beta-hexosaminidase from skin slices in vitro. Life Sci. 2006, 78, 598–606. [Google Scholar] [CrossRef]
  157. Elmes, S.J.; Jhaveri, M.D.; Smart, D.; Kendall, D.A.; Chapman, V. Cannabinoid CB2 receptor activation inhibits mechanically evoked responses of wide dynamic range dorsal horn neurons in naive rats and in rat models of inflammatory and neuropathic pain. Eur. J. Neurosci. 2004, 20, 2311–2320. [Google Scholar] [CrossRef]
  158. Sagar, D.R.; Kelly, S.; Millns, P.J.; O’Shaughnessey, C.T.; Kendall, D.A.; Chapman, V. Inhibitory effects of CB1 and CB2 receptor agonists on responses of DRG neurons and dorsal horn neurons in neuropathic rats. Eur. J. Neurosci. 2005, 22, 371–379. [Google Scholar] [CrossRef]
  159. Burston, J.J.; Sagar, D.R.; Shao, P.; Bai, M.; King, E.; Brailsford, L.; Turner, J.M.; Hathway, G.J.; Bennett, A.J.; Walsh, D.A.; et al. Cannabinoid CB2 receptors regulate central sensitization and pain responses associated with osteoarthritis of the knee joint. PLoS ONE 2013, 8, e80440. [Google Scholar] [CrossRef] [Green Version]
  160. Elmes, S.J.R.; Winyard, L.A.; Medhurst, S.J.; Clayton, N.M.; Wilson, A.W.; Kendall, D.A.; Chapman, V. Activation of CB1 and CB2 receptors attenuates the induction and maintenance of inflammatory pain in the rat. Pain 2005, 118, 327–335. [Google Scholar] [CrossRef]
  161. Yuill, M.B.; Hale, D.E.; Guindon, J.; Morgan, D.J. Anti-nociceptive interactions between opioids and a cannabinoid receptor 2 agonist in inflammatory pain. Mol. Pain 2017, 13, 1744806917728227. [Google Scholar] [CrossRef] [Green Version]
  162. Huffman, J.W. The search for selective ligands for the CB2 receptor. Curr. Pharm. Des. 2000, 6, 1323–1337. [Google Scholar] [CrossRef]
  163. Valenzano, K.J.; Tafesse, L.; Lee, G.; Harrison, J.E.; Boulet, J.M.; Gottshall, S.L.; Mark, L.; Pearson, M.S.; Miller, W.; Shan, S.; et al. Pharmacological and pharmacokinetic characterization of the cannabinoid receptor 2 agonist, GW405833, utilizing rodent models of acute and chronic pain, anxiety, ataxia and catalepsy. Neuropharmacology 2005, 48, 658–672. [Google Scholar] [CrossRef]
  164. Thakur, G.A.; Tichkule, R.; Bajaj, S.; Makriyannis, A. Latest advances in cannabinoid receptor agonists. Expert Opin. Ther. Pat. 2009, 19, 1647–1673. [Google Scholar] [CrossRef] [PubMed]
  165. Whiteside, G.T.; Gottshall, S.L.; Boulet, J.M.; Chaffer, S.M.; Harrison, J.E.; Pearson, M.S.; Turchin, P.I.; Mark, L.; Garrison, A.E.; Valenzano, K.J. A role for cannabinoid receptors, but not endogenous opioids, in the antinociceptive activity of the CB2-selective agonist, GW405833. Eur. J. Pharmacol. 2005, 528, 65–72. [Google Scholar] [CrossRef] [PubMed]
  166. Clayton, N.; Marshall, F.H.; Bountra, C.; O’Shaughnessy, C.T. CB1 and CB2 cannabinoid receptors are implicated in inflammatory pain. Pain 2002, 96, 253–260. [Google Scholar] [CrossRef]
  167. Hu, B.; Doods, H.; Treede, R.D.; Ceci, A. Depression-like behaviour in rats with mononeuropathy is reduced by the CB2-selective agonist GW405833. Pain 2009, 143, 206–212. [Google Scholar] [CrossRef] [PubMed]
  168. Leichsenring, A.; Andriske, M.; Backer, I.; Stichel, C.C.; Lubbert, H. Analgesic and antiinflammatory effects of cannabinoid receptor agonists in a rat model of neuropathic pain. Naunyn Schmiedebergs Arch. Pharmacol. 2009, 379, 627–636. [Google Scholar] [CrossRef]
  169. Ibrahim, M.M.; Deng, H.; Zvonok, A.; Cockayne, D.A.; Kwan, J.; Mata, H.P.; Vanderah, T.W.; Lai, J.; Porreca, F.; Makriyannis, A.; et al. Activation of CB2 cannabinoid receptors by AM1241 inhibits experimental neuropathic pain: Pain inhibition by receptors not present in the CNS. Proc. Natl. Acad. Sci. USA 2003, 100, 10529–10533. [Google Scholar] [CrossRef] [Green Version]
  170. Nackley, A.G.; Makriyannis, A.; Hohmann, A.G. Selective activation of cannabinoid CB(2) receptors suppresses spinal fos protein expression and pain behavior in a rat model of inflammation. Neuroscience 2003, 119, 747–757. [Google Scholar] [CrossRef]
  171. Quartilho, A.; Mata, H.P.; Ibrahim, M.M.; Vanderah, T.W.; Porreca, F.; Makriyannis, A.; Malan, T.P., Jr. Inhibition of inflammatory hyperalgesia by activation of peripheral CB2 cannabinoid receptors. Anesthesiology 2003, 99, 955–960. [Google Scholar] [CrossRef]
  172. Gutierrez, T.; Farthing, J.N.; Zvonok, A.M.; Makriyannis, A.; Hohmann, A.G. Activation of peripheral cannabinoid CB1 and CB2 receptors suppresses the maintenance of inflammatory nociception: A comparative analysis. Br. J. Pharmacol. 2007, 150, 153–163. [Google Scholar] [CrossRef] [Green Version]
  173. Nackley, A.G.; Zvonok, A.M.; Makriyannis, A.; Hohmann, A.G. Activation of cannabinoid CB2 receptors suppresses C-fiber responses and windup in spinal wide dynamic range neurons in the absence and presence of inflammation. J. Neurophysiol. 2004, 92, 3562–3574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Wilkerson, J.L.; Gentry, K.R.; Dengler, E.C.; Wallace, J.A.; Kerwin, A.A.; Kuhn, M.N.; Zvonok, A.M.; Thakur, G.A.; Makriyannis, A.; Milligan, E.D. Immunofluorescent spectral analysis reveals the intrathecal cannabinoid agonist, AM1241, produces spinal anti-inflammatory cytokine responses in neuropathic rats exhibiting relief from allodynia. Brain Behav. 2012, 2, 155–177. [Google Scholar] [CrossRef] [PubMed]
  175. Rahn, E.J.; Makriyannis, A.; Hohmann, A.G. Activation of cannabinoid CB1 and CB2 receptors suppresses neuropathic nociception evoked by the chemotherapeutic agent vincristine in rats. Br. J. Pharmacol. 2007, 152, 765–777. [Google Scholar] [CrossRef] [Green Version]
  176. Lozano-Ondoua, A.N.; Wright, C.; Vardanyan, A.; King, T.; Largent-Milnes, T.M.; Nelson, M.; Jimenez-Andrade, J.M.; Mantyh, P.W.; Vanderah, T.W. A cannabinoid 2 receptor agonist attenuates bone cancer-induced pain and bone loss. Life Sci. 2010, 86, 646–653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Diaz, P.; Phatak, S.S.; Xu, J.; Fronczek, F.R.; Astruc-Diaz, F.; Thompson, C.M.; Cavasotto, C.N.; Naguib, M. 2,3-Dihydro-1-benzofuran derivatives as a series of potent selective cannabinoid receptor 2 agonists: Design, synthesis, and binding mode prediction through ligand-steered modeling. ChemMedChem 2009, 4, 1615–1629. [Google Scholar] [CrossRef] [Green Version]
  178. Arena Pharmaceuticals. Olorinab (APD371) [Arena Pharmaceuticals Pipeline]. Arena Pharmaceuticals. Available online: https://www.arenapharm.com/pipeline/apd371/ (accessed on 18 February 2021).
  179. Soethoudt, M.; Grether, U.; Fingerle, J.; Grim, T.W.; Fezza, F.; de Petrocellis, L.; Ullmer, C.; Rothenhausler, B.; Perret, C.; van Gils, N.; et al. Cannabinoid CB(2) receptor ligand profiling reveals biased signalling and off-target activity. Nat. Commun. 2017, 8, 13958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Spiera, R.; Hummers, L.; Chung, L.; Frech, T.M.; Domsic, R.; Hsu, V.; Furst, D.E.; Gordon, J.; Mayes, M.; Simms, R.; et al. Safety and Efficacy of Lenabasum in a Phase II, Randomized, Placebo-Controlled Trial in Adults With Systemic Sclerosis. Arthritis Rheumatol. 2020, 72, 1350–1360. [Google Scholar] [CrossRef]
  181. Spiera, R.; Khanna, D.; Kuwana, M.; Furst, D.E.; Frech, T.M.; Hummers, L.; Stevens, W.; Matucci-Cerinic, M.; Baron, M.; Distler, O.; et al. A randomised, double-blind, placebo-controlled phase 3 study of lenabasum in diffuse cutaneous systemic sclerosis: RESOLVE-1 design and rationale. Clin. Exp. Rheumatol. 2021, 39 (Suppl. S131), 124–133. [Google Scholar] [CrossRef]
  182. Kinsey, S.G.; Mahadevan, A.; Zhao, B.; Sun, H.; Naidu, P.S.; Razdan, R.K.; Selley, D.E.; Imad Damaj, M.; Lichtman, A.H. The CB2 cannabinoid receptor-selective agonist O-3223 reduces pain and inflammation without apparent cannabinoid behavioral effects. Neuropharmacology 2011, 60, 244–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Luongo, L.; Palazzo, E.; Tambaro, S.; Giordano, C.; Gatta, L.; Scafuro, M.A.; Rossi, F.S.; Lazzari, P.; Pani, L.; de Novellis, V.; et al. 1-(2′,4′-dichlorophenyl)-6-methyl-N-cyclohexylamine-1,4-dihydroindeno[1,2-c]pyraz ole-3-carboxamide, a novel CB2 agonist, alleviates neuropathic pain through functional microglial changes in mice. Neurobiol. Dis. 2010, 37, 177–185. [Google Scholar] [CrossRef] [PubMed]
  184. Xu, J.; Tang, Y.; Xie, M.; Bie, B.; Wu, J.; Yang, H.; Foss, J.F.; Yang, B.; Rosenquist, R.W.; Naguib, M. Activation of cannabinoid receptor 2 attenuates mechanical allodynia and neuroinflammatory responses in a chronic post-ischemic pain model of complex regional pain syndrome type I in rats. Eur. J. Neurosci. 2016, 44, 3046–3055. [Google Scholar] [CrossRef]
  185. Xu, J.J.; Diaz, P.; Bie, B.; Astruc-Diaz, F.; Wu, J.; Yang, H.; Brown, D.L.; Naguib, M. Spinal gene expression profiling and pathways analysis of a CB2 agonist (MDA7)-targeted prevention of paclitaxel-induced neuropathy. Neuroscience 2014, 260, 185–194. [Google Scholar] [CrossRef]
  186. Gomez-Canas, M.; Morales, P.; Garcia-Toscano, L.; Navarrete, C.; Munoz, E.; Jagerovic, N.; Fernandez-Ruiz, J.; Garcia-Arencibia, M.; Pazos, M.R. Biological characterization of PM226, a chromenoisoxazole, as a selective CB2 receptor agonist with neuroprotective profile. Pharmacol. Res. 2016, 110, 205–215. [Google Scholar] [CrossRef] [PubMed]
  187. Nent, E.; Nozaki, C.; Schmole, A.C.; Otte, D.; Zimmer, A. CB2 receptor deletion on myeloid cells enhanced mechanical allodynia in a mouse model of neuropathic pain. Sci. Rep. 2019, 9, 7468. [Google Scholar] [CrossRef] [Green Version]
  188. La Porta, C.; Bura, S.A.; Aracil-Fernandez, A.; Manzanares, J.; Maldonado, R. Role of CB1 and CB2 cannabinoid receptors in the development of joint pain induced by monosodium iodoacetate. Pain 2013, 154, 160–174. [Google Scholar] [CrossRef]
  189. Racz, I.; Nadal, X.; Alferink, J.; Banos, J.E.; Rehnelt, J.; Martin, M.; Pintado, B.; Gutierrez-Adan, A.; Sanguino, E.; Bellora, N.; et al. Interferon-gamma is a critical modulator of CB(2) cannabinoid receptor signaling during neuropathic pain. J. Neurosci. 2008, 28, 12136–12145. [Google Scholar] [CrossRef] [Green Version]
  190. Slipetz, D.M.; O’Neill, G.P.; Favreau, L.; Dufresne, C.; Gallant, M.; Gareau, Y.; Guay, D.; Labelle, M.; Metters, K.M. Activation of the human peripheral cannabinoid receptor results in inhibition of adenylyl cyclase. Mol. Pharmacol. 1995, 48, 352–361. [Google Scholar]
  191. Bouaboula, M.; Poinot-Chazel, C.; Marchand, J.; Canat, X.; Bourrie, B.; Rinaldi-Carmona, M.; Calandra, B.; Le Fur, G.; Casellas, P. Signaling pathway associated with stimulation of CB2 peripheral cannabinoid receptor. Involvement of both mitogen-activated protein kinase and induction of Krox-24 expression. Eur. J. Biochem. 1996, 237, 704–711. [Google Scholar] [CrossRef]
  192. Niu, J.; Huang, D.; Zhou, R.; Yue, M.; Xu, T.; Yang, J.; He, L.; Tian, H.; Liu, X.; Zeng, J. Activation of dorsal horn cannabinoid CB2 receptor suppresses the expression of P2Y12 and P2Y13 receptors in neuropathic pain rats. J. Neuroinflamm. 2017, 14, 185. [Google Scholar] [CrossRef] [Green Version]
  193. Ji, R.R.; Suter, M.R. p38 MAPK, microglial signaling, and neuropathic pain. Mol. Pain 2007, 3, 33. [Google Scholar] [CrossRef] [Green Version]
  194. Zhuang, Z.Y.; Gerner, P.; Woolf, C.J.; Ji, R.R. ERK is sequentially activated in neurons, microglia, and astrocytes by spinal nerve ligation and contributes to mechanical allodynia in this neuropathic pain model. Pain 2005, 114, 149–159. [Google Scholar] [CrossRef] [PubMed]
  195. Calvo, M.; Zhu, N.; Grist, J.; Ma, Z.; Loeb, J.A.; Bennett, D.L. Following nerve injury neuregulin-1 drives microglial proliferation and neuropathic pain via the MEK/ERK pathway. Glia 2011, 59, 554–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Giri, S.; Nath, N.; Smith, B.; Viollet, B.; Singh, A.K.; Singh, I. 5-aminoimidazole-4-carboxamide-1-beta-4-ribofuranoside inhibits proinflammatory response in glial cells: A possible role of AMP-activated protein kinase. J. Neurosci. 2004, 24, 479–487. [Google Scholar] [CrossRef] [Green Version]
  197. Burgos, E.; Gomez-Nicola, D.; Pascual, D.; Martin, M.I.; Nieto-Sampedro, M.; Goicoechea, C. 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. 2012, 682, 62–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Correa, F.; Hernangomez, M.; Mestre, L.; Loria, F.; Spagnolo, A.; Docagne, F.; Di Marzo, V.; Guaza, C. Anandamide enhances IL-10 production in activated microglia by targeting CB(2) receptors: Roles of ERK1/2, JNK, and NF-kappaB. Glia 2010, 58, 135–147. [Google Scholar] [CrossRef]
  199. Hernangomez, M.; Mestre, L.; Correa, F.G.; Loria, F.; Mecha, M.; Inigo, P.M.; Docagne, F.; Williams, R.O.; Borrell, J.; Guaza, C. CD200-CD200R1 interaction contributes to neuroprotective effects of anandamide on experimentally induced inflammation. Glia 2012, 60, 1437–1450. [Google Scholar] [CrossRef] [Green Version]
  200. Tsuda, M.; Shigemoto-Mogami, Y.; Koizumi, S.; Mizokoshi, A.; Kohsaka, S.; Salter, M.W.; Inoue, K. P2X4 receptors induced in spinal microglia gate tactile allodynia after nerve injury. Nature 2003, 424, 778–783. [Google Scholar] [CrossRef]
  201. Romero-Sandoval, E.A.; Horvath, R.; Landry, R.P.; DeLeo, J.A. Cannabinoid receptor type 2 activation induces a microglial anti-inflammatory phenotype and reduces migration via MKP induction and ERK dephosphorylation. Mol. Pain 2009, 5, 25. [Google Scholar] [CrossRef] [Green Version]
  202. Han, Q.W.; Shao, Q.H.; Wang, X.T.; Ma, K.L.; Chen, N.H.; Yuan, Y.H. CB2 receptor activation inhibits the phagocytic function of microglia through activating ERK/AKT-Nurr1 signal pathways. Acta Pharmacol. Sin. 2022, 43, 2253–2266. [Google Scholar] [CrossRef] [PubMed]
  203. Tao, Y.; Li, L.; Jiang, B.; Feng, Z.; Yang, L.; Tang, J.; Chen, Q.; Zhang, J.; Tan, Q.; Feng, H.; et al. Cannabinoid receptor-2 stimulation suppresses neuroinflammation by regulating microglial M1/M2 polarization through the cAMP/PKA pathway in an experimental GMH rat model. Brain Behav. Immun. 2016, 58, 118–129. [Google Scholar] [CrossRef] [PubMed]
  204. Ma, C.; Zhang, M.; Liu, L.; Zhang, P.; Liu, D.; Zheng, X.; Zhong, X.; Wang, G. Low-dose cannabinoid receptor 2 agonist induces microglial activation in a cancer pain-morphine tolerance rat model. Life Sci. 2021, 264, 118635. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Components of the endocannabinoid system are involved in the main routes of biosynthesis, action, and degradation of endocannabinoids in the nervous system. 2-AG is mainly produced from the hydrolysis of DAG, mediated by two diacylglycerol lipases DAGLα/β. DAG is derived from phosphatidylinositol trisphosphate (PIP2), hydrolyzed by PLC. Most AEA appears to be derived from its membrane precursor, NAPE, which is produced by N-acyltransferase (NAT) using phosphatidylethanolamine (PE) and phosphatidylcholine (PC). NAPE can be hydrolyzed by a specific phospholipase D (NAPE-PLD). Microglia may be the primary cellular source of 2-AG and AEA in neuroinflammatory conditions, as they are capable of producing 20 times more endocannabinoids than other glial cells and neurons. AEA and 2-AG benefit from their strong lipid solubility and can be released into the intercellular space through the cell membrane soon after production. AEA mainly plays a role by activating CB1R expressed on the presynaptic membrane and postsynaptic membrane. 2-AG can not only activate CB1R, but also activate CB2R expressed on microglia. After performing their functions, endocannabinoids undergo re-uptake into the neurons and microglia by membrane transporters and are hydrolyzed by different enzymes. 2-AG is degraded by MAGL, ABHD-6, ABHD-12, or COX-2 into arachidonic acid, ethanolamine, and glycerol, while AEA is mainly metabolized by FAAH or COX-2 into arachidonic acid and ethanolamine.
Figure 1. Components of the endocannabinoid system are involved in the main routes of biosynthesis, action, and degradation of endocannabinoids in the nervous system. 2-AG is mainly produced from the hydrolysis of DAG, mediated by two diacylglycerol lipases DAGLα/β. DAG is derived from phosphatidylinositol trisphosphate (PIP2), hydrolyzed by PLC. Most AEA appears to be derived from its membrane precursor, NAPE, which is produced by N-acyltransferase (NAT) using phosphatidylethanolamine (PE) and phosphatidylcholine (PC). NAPE can be hydrolyzed by a specific phospholipase D (NAPE-PLD). Microglia may be the primary cellular source of 2-AG and AEA in neuroinflammatory conditions, as they are capable of producing 20 times more endocannabinoids than other glial cells and neurons. AEA and 2-AG benefit from their strong lipid solubility and can be released into the intercellular space through the cell membrane soon after production. AEA mainly plays a role by activating CB1R expressed on the presynaptic membrane and postsynaptic membrane. 2-AG can not only activate CB1R, but also activate CB2R expressed on microglia. After performing their functions, endocannabinoids undergo re-uptake into the neurons and microglia by membrane transporters and are hydrolyzed by different enzymes. 2-AG is degraded by MAGL, ABHD-6, ABHD-12, or COX-2 into arachidonic acid, ethanolamine, and glycerol, while AEA is mainly metabolized by FAAH or COX-2 into arachidonic acid and ethanolamine.
Ijms 24 02348 g001
Figure 2. The expression profiles and possible molecular mechanisms of CB2R-related functional endocannabinoid system in homeostatsis and activated microglia in pain processing. When the primary afferent nerve is injured or in a state of chronic pain, the resting microglia will be activated by the mediator released from the central terminal of the primary afferent and transform into pro-inflammatory (M1) microglia. When ATP activates the increased expression of P2X4 and P2X7 on microglia, Ca2+ enters microglia and regulates the activities of MAGL, DAGL, and NAPE-PLD, which lead to increased production and relation of endocannabinoids such as AEA and 2-AG and pro-inflammatory mediators including IL-1β, IL-6, IL-12, IFN-γ, and TNF-α in reactive microglia. This transition was also accompanied by a distinct morphological change in the microglia, from a small soma with long, branched processes to a more amoeba-like shape. At the same time, endocannabinoid such as 2-AG or AEA and exogenous cannabinoids such as AM1241 can act on the increased expression of CB2R on microglia. Activation of CB2R can inhibit adenylate cyclase (AC), which results in a reduction of intracellular cAMP levels. Diminished cAMP level intracellularly suppresses the activity of PKA and changes the expression of respective ion channels such as P2X4 and P2X7 on microglia, leading to decreased cytosolic Ca2+ concentration. Changes in Ca2+ distribution upon CB2R stimulation can also regulate the activities and expressions of MAGL, DAGL, FAAH, and NAPE-PLD. Meanwhile, CB2R activation is also accompanied by downstream PLC activation through secondary messengers to regulate the activity of the members of the MAPK family, such as ERK1/2 and p38. As a final consequence, these processes can down-regulate the release of pro-inflammatory cytokines and up-regulate the release of anti-inflammatory cytokines such as IL-4, IL-10, and TGF-β by regulating the activity of different transcription factors, leading to a switch of microglia to an anti-inflammatory phenotype (M2).
Figure 2. The expression profiles and possible molecular mechanisms of CB2R-related functional endocannabinoid system in homeostatsis and activated microglia in pain processing. When the primary afferent nerve is injured or in a state of chronic pain, the resting microglia will be activated by the mediator released from the central terminal of the primary afferent and transform into pro-inflammatory (M1) microglia. When ATP activates the increased expression of P2X4 and P2X7 on microglia, Ca2+ enters microglia and regulates the activities of MAGL, DAGL, and NAPE-PLD, which lead to increased production and relation of endocannabinoids such as AEA and 2-AG and pro-inflammatory mediators including IL-1β, IL-6, IL-12, IFN-γ, and TNF-α in reactive microglia. This transition was also accompanied by a distinct morphological change in the microglia, from a small soma with long, branched processes to a more amoeba-like shape. At the same time, endocannabinoid such as 2-AG or AEA and exogenous cannabinoids such as AM1241 can act on the increased expression of CB2R on microglia. Activation of CB2R can inhibit adenylate cyclase (AC), which results in a reduction of intracellular cAMP levels. Diminished cAMP level intracellularly suppresses the activity of PKA and changes the expression of respective ion channels such as P2X4 and P2X7 on microglia, leading to decreased cytosolic Ca2+ concentration. Changes in Ca2+ distribution upon CB2R stimulation can also regulate the activities and expressions of MAGL, DAGL, FAAH, and NAPE-PLD. Meanwhile, CB2R activation is also accompanied by downstream PLC activation through secondary messengers to regulate the activity of the members of the MAPK family, such as ERK1/2 and p38. As a final consequence, these processes can down-regulate the release of pro-inflammatory cytokines and up-regulate the release of anti-inflammatory cytokines such as IL-4, IL-10, and TGF-β by regulating the activity of different transcription factors, leading to a switch of microglia to an anti-inflammatory phenotype (M2).
Ijms 24 02348 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xu, K.; Wu, Y.; Tian, Z.; Xu, Y.; Wu, C.; Wang, Z. Microglial Cannabinoid CB2 Receptors in Pain Modulation. Int. J. Mol. Sci. 2023, 24, 2348. https://doi.org/10.3390/ijms24032348

AMA Style

Xu K, Wu Y, Tian Z, Xu Y, Wu C, Wang Z. Microglial Cannabinoid CB2 Receptors in Pain Modulation. International Journal of Molecular Sciences. 2023; 24(3):2348. https://doi.org/10.3390/ijms24032348

Chicago/Turabian Style

Xu, Kangtai, Yifei Wu, Zhuangzhuang Tian, Yuanfan Xu, Chaoran Wu, and Zilong Wang. 2023. "Microglial Cannabinoid CB2 Receptors in Pain Modulation" International Journal of Molecular Sciences 24, no. 3: 2348. https://doi.org/10.3390/ijms24032348

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop