Next Article in Journal
Simulated Microgravity Remodels Extracellular Matrix of Osteocommitted Mesenchymal Stromal Cells
Next Article in Special Issue
Alpha-Synuclein as a Prominent Actor in the Inflammatory Synaptopathy of Parkinson’s Disease
Previous Article in Journal
A Review on Ionic Liquids-Based Membranes for Middle and High Temperature Polymer Electrolyte Membrane Fuel Cells (PEM FCs)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuroprotective and Immunomodulatory Action of the Endocannabinoid System under Neuroinflammation

1
Otto Loewi Research Center, Division of Pharmacology, Medical University of Graz, 8010 Graz, Austria
2
Department of Anatomy and Structural Biology, Albert Einstein College of Medicine, Bronx, NY 10461, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(11), 5431; https://doi.org/10.3390/ijms22115431
Submission received: 27 April 2021 / Revised: 17 May 2021 / Accepted: 18 May 2021 / Published: 21 May 2021
(This article belongs to the Special Issue Molecular and Cellular Mechanisms in Inflammatory Synaptopathy)

Abstract

:
Endocannabinoids (eCBs) are lipid-based retrograde messengers with a relatively short half-life that are produced endogenously and, upon binding to the primary cannabinoid receptors CB1/2, mediate multiple mechanisms of intercellular communication within the body. Endocannabinoid signaling is implicated in brain development, memory formation, learning, mood, anxiety, depression, feeding behavior, analgesia, and drug addiction. It is now recognized that the endocannabinoid system mediates not only neuronal communications but also governs the crosstalk between neurons, glia, and immune cells, and thus represents an important player within the neuroimmune interface. Generation of primary endocannabinoids is accompanied by the production of their congeners, the N-acylethanolamines (NAEs), which together with N-acylneurotransmitters, lipoamino acids and primary fatty acid amides comprise expanded endocannabinoid/endovanilloid signaling systems. Most of these compounds do not bind CB1/2, but signal via several other pathways involving the transient receptor potential cation channel subfamily V member 1 (TRPV1), peroxisome proliferator-activated receptor (PPAR)-α and non-cannabinoid G-protein coupled receptors (GPRs) to mediate anti-inflammatory, immunomodulatory and neuroprotective activities. In vivo generation of the cannabinoid compounds is triggered by physiological and pathological stimuli and, specifically in the brain, mediates fine regulation of synaptic strength, neuroprotection, and resolution of neuroinflammation. Here, we review the role of the endocannabinoid system in intrinsic neuroprotective mechanisms and its therapeutic potential for the treatment of neuroinflammation and associated synaptopathy.

Graphical Abstract

1. Highlights

  • Retrograde endocannabinoid signaling provides a mechanism by which neurons can rapidly regulate the strength of their synaptic inputs.
  • Stimulation of postsynaptic neurotransmitter receptors and sustained Ca2+ influx is a potent trigger for the production of endocannabinoids (eCBs) and their congeners.
  • Neuroinflammation and alterations in endocannabinoid signaling is implicated in multiple neurological disorders.
  • The activity-dependent flow of glutamate and eCBs from synapses controls microglial attraction, secretion of pro-inflammatory and pro-survival factors, and defines the synapse stability under inflammation and excitotoxicity.
  • The pharmacological inhibition of eCB degradation exerts a primary effect in injured sites, where these mediators are actively produced de novo.
  • The endocannabinoid system mediates communication within the tripartite synapse during the development and resolution of neuroinflammation.

2. Introduction

Neuroinflammation is widely regarded as inflammation of the central nervous system (CNS) comprising the brain and spinal cord. The inflammatory process is driven by the release of pro-inflammatory mediators such as cytokines, prostaglandins, and reactive oxygen and nitrogen species by activated endothelial and glial cells with subsequent infiltration of peripheral inflammatory cells into the CNS. As a consequence, neuroinflammation can lead to edema, tissue damage, and loss of neuronal functions as well as accelerate and cause cognitive impairment and neurodegenerative diseases. Common triggers of chronic neuroinflammation include toxic metabolites, harmful self-proteins during autoimmunity, aging, bacterial and viral infections, as well as traumatic brain, and spinal cord injury.
The endocannabinoid system (ECS), consisting of the cannabinoid receptors CB1 and CB2, their major endogenous ligands 2-arachidonylglycerol (2-AG) and arachidonoylethanolamide (AEA or anandamide), and their synthesizing and degrading enzymes, plays a crucial role in the intrinsic response to neuroinflammation, brain injury, and neurodegenerative diseases [1,2,3]. This system is now being expanded with the ligands that do not show affinity for CB1/2 but display cannabimimetic activity and are able to modulate the actions of true eCBs. N-acylethaolamines generated as AEA congeners together with lipoaminoacids and acyl conjugates of neurotransmitters exert their biological activities through different receptors such as peroxisome proliferator-activated receptor (PPAR)-α [4], transient receptor potential cation channel subfamily V member 1 (TRPV1) and non-cannabinoid G-protein coupled receptors (GPRs). These compounds as well as true eCBs, their molecular targets, overall interplay, and metabolism comprise the endocannabinoidome having profound role in homeostatic response to noxious endogenous and exogenous stimuli.
The existence of multiple targets and routes for ligand synthesis and degradation within the endocannabinoidome enables complex scanning of cellular states and allows a tightly regulated response to relevant changes. Of particular interest is the involvement of this system in inherent protective responses against inflammatory and neurodegenerative processes in the brain as targeted manipulation of this system has high therapeutic potential. In this review, we provide an overview of the immune-regulatory and neuroprotective potential of the endocannabinoidome during neuroinflammation.

3. The Endocannabinoid System

The ECS is a complex biological network involved in the maintenance of homeostasis. It consists of a family of naturally occurring signaling lipids, the endocannabinoids (eCBs), their synthesizing and degrading enzymes, specific transmembrane eCB transporters, and receptors. Components of the ECS are found throughout the body: in the CNS, in cells of the immune system, the liver, the reproductive system, the respiratory system, the gastrointestinal tract, the cardiovascular system, and in skeletal muscles.

3.1. Metabolism of Endocannabinoids and Related Compounds

Levels of eCBs vary among specific regions of the brain and differ between animal species. In rat brain, basal 2-AG levels were estimated at 2–10 nmol/g [5], whereas AEA was detected at the range of 3–30 pmol/g [6] and typically comprises only 1–3% of generated NAEs [7]. Although both, 2-AG and AEA, are referred to as eCBs their metabolic pathways are completely different. 2-AG is formed from arachidonic acid-containing membrane phospholipids via three major pathways: * from diacylglycerol (DAG) via DAG lipase, ** from 2-acyl lysophosphatidic acid (LPA) via 2-LPA phosphatase, and *** from 2-acyl lysophosphatidylinositol (LPI) via lyso-phospholipase C. Similarly, AEA and related NAEs are produced on demand in response to certain stimuli [8,9] from membrane glycerophospholipids via the transacylation–phosphodiesterase pathway comprising two main enzymes, Ca2+-dependent N-acyltransferase and N-acyl-phosphatidylethanolamine-hydrolyzing phospholipase D (NAPE-PLD) (Figure 1). Lysophospholipids, generated in the course of 2-AG synthesis and produced at higher levels under inflammatory conditions such as neuroinflammation [10], possess biological activity themselves. For instance, LPI and its metabolite LPA activate respectively GPR55, GPR119, and LPA receptors to modulate cell proliferation, migration, cytokine and chemokine secretion, apoptosis/survival, and tumorigenesis.
The major degradation pathway of 2-AG is considered to be the hydrolysis to arachidonic acid and glycerol, which can be catalyzed by multiple enzymes, including monoacylglycerol lipase (MAGL), fatty acid amide hydrolase (FAAH), α/β-hydrolase domain containing (ABHD) 6, and ABHD12 proteins (reviewed in [4,11]). Furthermore, the arachidonoyl moiety of 2-AG can be oxygenated by cyclooxygenase (COX)-2 and lipoxygenases, resulting in the formation of pro-inflammatory glyceryl prostaglandins, such as prostaglandin E2–glycerol ester (PGE2-G), or hydroperoxy derivatives of 2-AG, respectively [12,13]. Although physiological significance of the oxygenation pathways remains unclear, biological activities of glyceryl prostaglandins have been reported [12]. The essential degradation pathway of NAEs is their hydrolysis to free fatty acids and ethanolamine by FAAH. Alternatively, NAE-hydrolyzing acid amidase (NAAA), a lysosomal enzyme, that is structurally similar to acid ceramidase and hydrolyzes NAEs at acidic pH, was described [14]. Both, specific FAAH and NAAA inhibitors, represent promising therapeutic tools for the treatment of neurological and inflammatory disorders as they raise the endogenous levels of bioactive NAEs [15,16]. In addition to hydrolytic degradation, polyunsaturated NAEs like AEA, DHEA, and EPEA are oxygenated by COX-2, lipoxygenases and cytochrome P450s and converted into prostaglandin ethanolamides (prostamides) or ω-3 endocannabinoid epoxides, which exhibit unique biological activities themselves [17,18]. The synthesis and degradation pathways of eCBs are summarized in Figure 1.
Lipoaminoacids and N-acylneurotransmitters are synthesized as a result of the conjugation of an amino acid/neurotransmitter with free fatty acids or acyl-CoA. Additionally, N-arachidonoylglycine can be generated from AEA in sequential steps catalyzed by alcohol dehydrogenase and aldehyde dehydrogenase. N-arachidonoyldopamine (NADA) biosynthesis in brain occurs through direct conjugation of dopamine with arachidonic acid. In this process FAAH is either a rate-limiting enzyme that liberates arachidonic acid from AEA, a conjugation enzyme, or both [19]. N-acylneurotransmitters and lipoaminoacids are degraded by FAAH or through modification of acyl- or amino acid/neurotransmitter residues.

3.2. Molecular Targets of the Endocannabinoid System

The main eCBs 2-AG and AEA bind to the central and peripheral G protein-coupled cannabinoid receptors CB1 and CB2. 2-AG behaves as a full agonist at CB1 and CB2, while AEA is a partial agonist of CB1, but almost inactive at CB2. Both CB1 and CB2 are G protein coupled receptors that have the ability to simultaneously activate multiple pools of G proteins. First, both receptors were shown to be associated with G proteins of the Gi/o family, while later additional signaling via Gs and Gq has been reported. Gi/o subunits inhibit adenylyl cyclase or couple to the mitogen-activated protein kinase (MAPK) pathway [20,21] Gs stimulates adenylyl cyclase [22], Gq couples to phospholipase C to promote the release of intracellular calcium [23,24] and Gβγ subunits derived from Gi/o inhibit voltage-gated calcium channels (VGCC) [25]. Further, coupling of arrestins with both CB1 and CB2 is implicated in the receptor desensitization, internalization, and G protein–independent signaling [26,27,28].
CB1 receptors are expressed predominantly in the nervous system, with enrichment in GABAergic axon terminals, and are among the most abundant GPCR in the brain, whereas CB2 receptors are mainly expressed in cells of the immune system and also detected in microglia. The neuronal expression of CB2 receptors, which for a long time remained controversial, is now shown on brainstem neurons and midbrain dopaminergic neurons [29,30,31]. The important features of neuronal CB2 are their low basal expression compared to CB1, and high inducibility under relevant stimuli like inflammation or addiction [32,33]. The functional contribution of CB1/2 (as receptors of true eCBs, 2-AG and AEA) in different brain cells is shown in Table 1.
Besides CB1 and CB2, a range of NAEs and monoacylglycerols also interact with several other receptors (Figure 2). AEA activates the transient receptor potential cation channel subfamily V member 1 (TRPV1), implicated in synaptic transmission and pain sensation, and seems to be a partial agonist for the non-CB receptor GPR55 [35]. Moreover, the interaction of 2-AG and non-CB receptors has emerged recently. Several studies have suggested that in addition to CB1 and CB2, there are non-CB1 and non-CB2 cannabinoid-related orphan GPCRs including GPR18, GPR55, and GPR119 [36]. It was demonstrated that 2-AG binds to GABAA receptors and can modulate the action of neurosteroids at GABAA receptors [37]. Moreover, endocannabinoids and some of their metabolites bind and activate peroxisome proliferator-activated receptors (PPARs) [38].
Saturated and monounsaturated NAEs do not show ligand affinity for cannabinoid receptors, but rather exert their biological activities through different receptors and pathways. PEA (C16:0 N-acylethanolamine) has been shown to promote anti-inflammatory, analgesic, anti-oxidative, and neuroprotective actions [39,40]. Most of its effects are mediated by PPAR-α [41]. Moreover, it has been reported that PEA activates TRPV1 [42] and GPR55 [43], and enhances 2-AG and AEA levels via inhibition of FAAH expression [44]. OEA (C18:1 N-acylethanolamine) is mainly known for its anorexic activity in experimental models, but has also been shown to mediate anti-inflammatory, analgesic, and antioxidative effects. OEA binds with high affinity to PPAR-α [45]) and interacts with TRPV1 [46]. Recent evidence supports the assumption that SEA (C18:0 N-acylethanolamine) mediates anti-inflammatory and neuroprotective activities [47], although there is no direct data on its molecular targets. In the tetrad of behavioral tests considered to be highly predictive for cannabimimetic compounds, SEA behaves similarly to AEA. This may be a result of so-called “entourage” effect of SEA, which potentiates the effects of AEA by inhibiting its degradation [48]. SEA demonstrated neuroprotective properties, decreased the onset of inflammation, and restricted leukocyte infiltration into the brain parenchyma in a mouse model of systemic inflammation. It seems that there is a certain therapeutic window for SEA treatment following the onset of inflammation. Mice treated with SEA had higher levels of 2-AG or its stable isomer 1-AG, in the prefrontal cortex and hippocampus. This rise of 2-AG/1-AG levels preceded the increase of neuronal CB1/2 expression and suggests the interference of SEA with the eCB system [47]. Moreover, SEA has been shown to protect from oxidative stress and to have analgesic properties. Similar to AEA, PEA, and OEA, it was demonstrated that SEA, together with linoleylethanolamide (LEA), eicosapentanoylethanolamide (EPEA), and docosahexaenoylethanolamine (DHEA) are activators of PPAR-α [49]. LEA was shown to activate TRPV1 [50], GPR119, and to inhibit AEA hydrolysis by FAAH. Pretreatment with LEA prior to ischemia/reperfusion injury significantly reduced cortical infarct volume and neurological deficit [51]. N-docosahexaenoylethanolamide (DHEA, synaptamide) promotes neurogenesis, neuritogenesis, and synaptogenesis [52] as an endogenous ligand of GPR110 [53], a member of the adhesion-GPR family which is highly expressed in fetal brain. Under systemic inflammation, immune-regulatory functions of GPR110 contribute to the anti-inflammatory action of synaptamide [54]. A summary of the known signaling pathways and anti-inflammatory actions of the major endogenously produced NAEs is presented in Figure 3.
N-acylneurotransmitters and lipoaminoacids represent a separate cluster within the endocannabinoidome, that bears the potential for identification of novel receptor targets in this system. Among these compounds N-acyldopamines are CB1 and TRPV1 agonists, while N-acylserotonines are TRPV1 antagonists and N-arachidonoyl-γ-aminobutyric acid (NAGABA) activates GPR92 receptor. Liberation of free neurotransmitter following degradation of N-acylneurotransmitters makes their action more complex.

3.3. Involvement of Endocannabinoid System in Response to Neuropathology

Endocannabinoids and related NAEs are produced on demand and play a crucial regulatory role in metabolic processes, behavior, and immunity. Under healthy conditions these lipid mediators are most abundant in the brain and barely found in circulation and peripheral tissues [55]. During various pathological conditions of the CNS the profiles of eCBs and their congeners undergo significant changes, which is associated with the inflammation-modulating, analgesic, and neuroprotective activity of these compounds (summarized in Table 2).

4. Glutamate Receptor-Mediated Neurotoxicity

4.1. Glutamate as a Major Excitatory Neurotransmitter in Mammals and Potential Neurotoxin

Activation of postsynaptic neurotransmitter receptors and Ca2+ influx into the postsynaptic terminal induce the synthesis of eCBs and related compounds. This activity-dependent production of eCBs is essential for the fine regulation of neurotransmission. In the mammalian brain, glutamate is the main excitatory neurotransmitter implicated in learning and memory formation. Glutamatergic neurotransmission mediates synaptic plasticity, whereby ionotropic and metabotropic (mGluRs) glutamate receptors play a primary role. Between the quantal neurotransmitter releases, the level of glutamate in the synaptic cleft is estimated to be <1 μM. This low basal level is maintained by rapid reuptake of glutamate from the extracellular space into the cytosol by high-affinity glutamate transporters EAATs (excitatory amino acid transporters). EAATs are localized on neurons (primarily EAAT4 and EAAT3 (EAAC1, Excitatory Amino Acid Carrier)) and astrocytes (primarily glutamate transporter GLT-1 and glutamate-aspartate transporter GLAST), and co-transport one molecule of L-glutamate (or L-/D-aspartate) with 3Na+ and 1H+ in exchange of 1K+ [66]. The dependence of this transport system on Na+ and K+ gradients across the plasma membrane makes it highly vulnerable to ATP depletion with subsequent inhibition of glutamate uptake or reversal of transporters [67]. Another factor affecting the efficiency of glutamate removal from the synaptic cleft is translational control of EAATs or post-translational modification of the transporter molecules, which affects the level of active transporters. Glutamate is further accumulated in the synaptic vesicles via vesicular glutamate transporters VGLUTs using the ΔμH+ gradient.
Considering the high energy-dependent compartmentalization of glutamate and its gradient across the synaptic bouton, i.e., from synaptic vesicles (~200 mM [68]) to the synaptic cleft (<1 μM between release events, around 1 mM during the peak of SV release [69]), any factors affecting the efficiency of high-affinity glutamate uptake represent a potential risk of neurotoxic neuronal damage. The pathophysiological conditions underlying the long-term glutamate rise in the synaptic cleft and extrasynaptic glutamate spillover are traumatic brain injury, ischemia, and other causes of hypoxia, stroke, and oxidative stress leading to transition of the significant portions of EAATs to the reverse mode, when glutamate is released from the cytosol to the extracellular space. Glutamate-mediated neurotoxicity originates from overstimulation of ionotropic glutamate receptors, primarily N-methyl-D-aspartate (NMDA) receptors, and massive Ca2+ flux to the postsynaptic terminal. High extracellular concentrations of glutamate lead to the prolonged co-activation of synaptic and extrasynaptic (localized to non-synaptic sites) NMDA receptors, glutamate-mediated neurotoxicity [70], and are involved in the pathogenesis of Alzheimer’s disease, amyotrophic lateral sclerosis (ALS), and Huntington’s disease.
Due to Ca2+ permeability and high affinity for glutamate, NMDA receptors are among the primary molecular targets implicated in the pathogenesis of excitotoxicity. At resting membrane potential, the current through channels of NMDA receptors is almost fully blocked by Mg2+ preventing the conductance between the stimuli. During quantal neurotransmitter release two conditions for Ca2+ influx through NMDA receptors are met: * glutamate concentrations rise rapidly, and ** the depolarization of synaptic membranes removes the Mg2+ block from NMDA receptor channels. Therapeutic concentrations (1–10 μM) of the NMDA receptor antagonist memantine, used for treatment of Alzheimer’s disease, preferentially block extrasynaptic rather than synaptic currents through NMDA receptors in the same neuron [71]. The mode of memantine action enables effective prevention of excessive extrasynaptic NMDA receptor stimulation, with much less effect on NMDA receptor-mediated synaptic activity, when glutamate is elevated for only milliseconds [72].
Under prominent rise of intracellular Ca2+ levels, vesicular glutamate release is another factor contributing to elevated extracellular glutamate concentration and excitotoxic damage. In rats and mice, ischemic conditions, followed by release of axonal vesicular glutamate into the peri-axonal space under the myelin sheath, trigger activation of myelinic GluN2C/D-containing NMDA receptors [73], which are generally extrasynaptic [74].

4.2. Excitotoxicity as a Prerequisite and Consequence of Neuroinflammation and Neurodegeneration

Due to the ability of Ca2+ to activate a range of enzymes, glutamate receptor-mediated excitotoxicity provokes necrotic and apoptotic neuronal death. Massive influx of Ca2+ overloads the intracellular buffer systems for this ion, provokes mitochondrial dysfunction, and activation of a range of proteases, including caspases and calpain, leading to the subsequent degradation of components of the neuronal cytoskeleton and the release of apoptotic factors.
One example of active involvement of the ECS in mediating cellular communication is the functional coupling of microglia and synapses during normal synaptic activity as well as excitotoxic injury. As previously suggested, microglia are a crucial source of de novo produced AEA and 2-AG under basal conditions and during neuroinflammation [75,76,77]; however, high glutamate application induces a prominent 2-AG overproduction in neurons [76,78]. Under these conditions the neuronal production of AEA increases only slightly, while the production of two putative endocannabinoids, homo-gamma-linolenylethanolamide and docosatetraenylethanolamide remains unchanged [76]. LPS-induced systemic inflammation in mice is accompanied by the increase in basal glutamate levels in the prefrontal cortex [47]. Elevated glutamate may originate from both inflammation-associated decrease in uptake, and neuronal and non-neuronal (from astrocytes and microglia) glutamate release. Glutamate flow favors the spatial cooperation between dendritic spines and ramified microglial cells and induces microglial process extension toward neurons (Figure 4). 2-AG induces chemokinesis (random motion increased by a chemical stimulus) and chemotaxis, (directed cell migration along a chemical gradient) in microglia cells [76,79]. In line with this, activated microglia express CB2 receptors at the leading edge of their motile protrusions [76]. Arachidonylcyclopropylamide (ACPA)-induced migration of BV-2 microglia could be blocked by the highly selective CB2 antagonist SR145528 [80]. Similarly, the migratory responses towards 2-AG and the synthetic cannabinoid CP 55,940 were inhibited by CB2 receptor antagonism [76,79]. We hypothesize that eCBs released at sites of synaptic activity (or injury) may act as a chemoattractants to recruit microglia in a CB2-dependent manner, toward neuroinflammatory lesion sites. Moreover, in organotypic hippocampal slice cultures, 2-AG mediated neuroprotection against NMDA-induced excitotoxicity by acting explicitly on abnormal-cannabidiol (abn-CBD)-sensitive receptor, putative GPR18, on microglial cells [81].
There is an activity-dependent modification of microglia–synapse contacts in vivo. Ischemic brain is characterized by the markedly prolonged contact time between microglial processes and synaptic structures and wrapping of microglial processes around the synapse, followed by the disappearance of presynaptic boutons [82].
One of the mechanisms by which microglia eliminate presynaptic boutons and axons is trogocytosis [83], a process described in the immune system as a non-apoptotic mechanism for the capture of membrane components that differs from phagocytosis and involves the engulfment and clearance of cellular structures larger than 1 µm [84]. In a mouse model of cortical multiple sclerosis in vivo imaging demonstrated that cortical inflammation disrupts circuit activity, which coincides with a widespread, but reversible, loss of dendritic spines. Under these circumstances, spines displaying local calcium accumulations are eliminated by invading macrophages or resident activated microglia [85].
Acute microglia activation is accompanied by the release of glutamate, quinolinic acid, proinflammatory cytokines (IL-1β, TNF-α, IL-2, IL-6), chemokines—macrophage inflammatory protein-1α (MIP-1α) and monocyte chemoattractant protein-1 (MCP-1), and free arachidonic acid. Quinolinic acid produced exclusively in activated microglia and macrophages, is a NMDA receptor agonist and mediates excitotoxicity during immune response. By contributing to destabilization of the cytoskeleton in astrocytes and endothelial cells, quinolinic acid decreases the integrity of the neurovascular unit and increases the influx of BBB impermeable quinolinic acid from the periphery [86]. Produced excitotoxic molecules and proinflammatory cytokines intensify free radical generation and lipid peroxidation, which provoke mitochondrial dysfunction and further exacerbate the excitotoxicity.
Microglia largely define the fate of damaged synaptic contacts and cells and promote the resolution of neuroinflammation and regeneration by releasing brain-derived neurotrophic factor (BDNF) and cytokines with dual (pro- and anti-inflammatory) potential, like TGF-β and IL-10.

5. The Role of Retrograde Endocannabinoid Signaling in the Tuning of Synaptic Strength

5.1. Synaptic Plasticity in Glutamatergic Synapses

When glutamate levels reach a certain concentration in the synaptic cleft, it binds to AMPA receptors and induces Na+ influx, which is registered as excitatory postsynaptic potentials (EPSP) of certain amplitudes. Due to the presence of the GluR2 subunit the majority of AMPA receptors in the CNS are impermeable to Ca2+ [87] and postsynaptic Ca2+ influx triggered by glutamate is mainly mediated by NMDA receptors.
Influx of Ca2+ through NMDA receptor channels activates a range of kinases, primarily Ca2+/calmodulin-dependent protein kinase II (CaMKII) [88,89], which in turn activates Rho GTPases, Cdc42, and RhoA [90]. This reorganizes the postsynaptic density via * remodeling of the actin cytoskeleton and transient (~5 min) enlargement of the spine (Figure 5); ** enhanced trafficking of AMPA receptors to post-synaptic sites as a result of their redistribution from recycling endosome to the plasma membrane [91,92], and *** increased single-channel conductance of AMPA receptors as a result of direct phosphorylation [93].
Synaptic recruitment of Ca2+-permeable AMPA receptors via CaMKI is also suggested to contribute to signaling pathways that drive the spine enlargement via actin polymerization [94]. Thus, following the repeated cycles of activation, the amplitude of evoked EPSC increases, i.e., is potentiated (long-term potentiation, LTP). This effect is typical for excitatory neurotransmission and persists in synapses depending on the stimulus mode.
In contrast, long-term depression (LTD) is a long-lasting drop in the efficiency of synaptic transmission as a result of a decrease in postsynaptic receptor density and/or presynaptic neurotransmitter release. While for development of LTP the activation of certain protein kinases is essential, LTD induction is dependent on protein phosphatase activity and target dephosphorylation. Prolonged 1 Hz stimulation leads to Ca2+ rise and calmodulin-dependent activation of calcineurin (protein phosphatase 2B, PP2B) [95], which via serine/threonine protein phosphatases PP1 or PP2A, results in the dephosphorylation of AMPA receptors [96], decrease of their channel conductance, and arrest of their recycling [97]. Dephosphorylation of the transcription factor cAMP response element binding protein (CREB) in the hippocampal area CA1 in vivo is suggested to be one of the mechanisms through which these protein phosphatases contribute to the prolonged maintenance of LTD [98].
The development of LTP or LTD depends on whether the frequency of the stimulation is higher than threshold frequency [99], with the postsynaptic rise of Ca2+ as a main determinant of the development of LTD or LTP. LTD and LTP are well-characterized molecular mechanisms underlying learning and memory formation and are induced in experiments with high- or low-frequency stimulations (repetition of hundreds of pre- or postsynaptic spikes). In the striatum of rodents, spike-timing-dependent potentiation (STDP), a phenomenon describing the dependence of strength of synaptic transmission on the timing between the neuron’s output and input action potentials (spikes), is observed for 75–100 pairings, disappears for 25–50 pairings and re-emerges for 5–10 pairings. STDP that is induced by very few pairings is independent from NMDA receptors but mediated by 2-AG and AEA, acting on both CB1 and TRPV1 [100].

5.2. Endocannabinoid-Mediated Synaptic Plasticity

Stimulation of postsynaptic neurotransmitter receptors and sustained Ca2+ influx is the potent trigger for the production of eCBs and their congeners [78,101] in a neuronal activity-dependent manner. The released neurotransmitter activates ionotropic neurotransmitter receptors, Gq-coupled metabotropic receptors (group I of metabotropic glutamate receptors, mGluRs, dopamine receptors D2, M1, and M3 muscarinic acetylcholine receptors, M1/M3 mAChRs), and/or voltage-gated calcium channels. This induces a burst of Ca2+ in the postsynapse and the production of eCBs, primarily AEA and 2-AG, which via CB1 and CB2 receptors modulate the presynaptic release of neurotransmitters. Thus, generation of AEA and 2-AG in the brain shows spatial variations and depends on complement of neurotransmitter receptors on certain neurons. Endocannabinoid-mediated synaptic potentiation can be realized in neurons receiving/sending inputs via taken activated synapses (homosynaptic) or in other neurons that do not directly contact taken activated synapses (heterosynaptic).
Endocannabinoid-mediated long-term depression (eCB-LTD). CB1 and CB2 are Gi/0 protein-coupled receptors that, upon ligand binding, inhibit adenylate cyclase activity and cAMP production, negatively regulate voltage-gated calcium channels (VGCC), and activate inwardly rectifying potassium (Kir) channels and MAP kinase. This decreases the Ca2+ influx into the presynaptic terminal, lowers the probability of Ca2+-dependent fusion of synaptic vesicles and attenuates the presynaptic neurotransmitter release. Synaptic vesicle recycling is a highly dynamic multistep process mediated by SNARE-proteins and other regulatory proteins, with Ca2+ influx being a key factor for docked and primed synaptic vesicles to enter the fusion step [102]. Together with the activity-dependent production of eCBs and activation of CB1, presynaptic activity is essential for this type of plasticity as it determines the afferents in which eCB-LTD will be induced [103]. Presynaptic activity dependence is mediated by presynaptic NMDA autoreceptors that detect the release of glutamate. Thus, timing-dependent-LTD is induced only under coincident activation of presynaptic NMDA and CB1 receptors [104]. Unlike LTP, which lasts from minutes to hours, short-term plasticity (STP) is maintained for tens of milliseconds to a few minutes at a maximum.
Depolarization-induced suppression of inhibition (DSI)/depolarization-induced suppression of excitation (DSE). DSI/DSE is a form of STP realized in the inhibitory (GABAergic) and excitatory (glutamatergic) synapses, respectively, and is induced by depolarization of the postsynaptic terminal and Ca2+ influx through VGCC. Pharmacological experiments favor a role for 2-AG rather than AEA or noladin ether (2-arachidonyl glyceryl ether) as the relevant endocannabinoid to elicit DSE [105]. Glutamate spillover may profoundly affect network excitability by shifting the duration of eCB-mediated inhibition at GABA synapses. Metabotropic glutamate receptors are involved in the control of the duration of DSI, most likely through heterologous desensitization of CB1 [106]. Thus, glutamate-mediated excitotoxicity can significantly modify establishment of various forms of synaptic plasticity and balance in glutamate-/GABAergic signaling [47].
Synaptically-evoked (or metabotropic-induced) suppression of inhibition/excitation (SSE/SSI). SSE/SSI is a form of STP, driven by activation of postsynaptic metabotropic (Gq/11-coupled) neurotransmitter receptors, subsequent activation of phospholipase C (PLC) and DAGL with generation of 2-AG and suppression of neurotransmitter release via presynaptic CB1 receptors [107]. If the synaptic stimulation is profound, produced eCBs can reach more distant sites and mediate the plasticity heterosynaptically.
TRPV1-mediated synaptic plasticity. TRPV1-mediated synaptic plasticity demonstrates the interplay between the eCB and endovanilloid system. This type of synaptic plasticity is mediated by binding of AEA to TRPV1 at the postsynapse and results in Ca2+-calcineurin and clathrin-dependent internalization of AMPA receptors, which provoke LTD of excitatory transmission [108,109]. TRPV1 integrates sensation of physical and chemical stimuli and is activated by temperature greater than 43 °C, acidic conditions, vanilloids like capsaicin, or endocannabinoids such as AEA, N-arachidonoyl dopamine, and N-oleoyl dopamine. TRPV1 is well known for its role in transmission of neuropathic (inflammatory) pain. At the same time TRPV1 mediates LTD in the hippocampus and 12-(S)-hydroperoxyeicosatetraenoic acid (12-(S)-HPETE), an endogenous eicosanoid released during synaptic stimulation, acts at TRPV1 receptors to trigger LTD [110].
It is hypothesized that eCB-mediated LTP is induced only when massive Ca2+ rise is observed and high levels of 2-AG are produced (i.e., following simultaneous activation of several postsynaptic neurotransmitter receptors, TRPV1, VGCC) [100]. The observations that2-AG and AEA mediate different forms of plasticity with the involvement of CB1, TRPV1, and mGluRs receptors depending on the brain region, characterize the eCB system as a polymodal signal integrator that allows the diversification of synaptic plasticity in a single neuron [111].
The interference of NAEs and other eCB congeners with enzymatic degradation or endocannabinoid signaling suggests their role in tuning the activity of primary eCBs. NAEs, monoacylglycerols, and certain N-acylneurotransmitters compete with AEA or 2-AG for FAAH-mediated degradation, thus extending their lifetimes [112] and their ability to interact with cannabinoid receptors. This “entourage effect” of the non-cannabinoid 2-acylglycerols and NAEs may serve as an additional fine regulator of cannabinoid activity. The overall interplay and metabolism of endogenous ligands of CB1/2, TRPV1, GPR55, and GPR18 is now integrated in the “endocannabinoidome” and demonstrates that this system is an essential player not only in many aspects of behavior, cognition, and memory, but also mediates inherent protective mechanisms of the neuro-immune interface.

6. Neuroinflammation-Induced Synaptopathy and Neurodegenerative Diseases

Neuroinflammation is a common feature of acute and chronic neurodegenerative disorders such as Alzheimer’s and Parkinson’s disease, viral infections of the CNS, stroke, paraneoplastic disorders, traumatic brain injury, and multiple sclerosis. Neuroinflammation is typically characterized by activation of immunocompetent glia cells (microglia and astroglia), release of cytokines, prostaglandins, and reactive oxygen species, the impairment of the blood–brain-barrier (BBB) integrity and resultant infiltration of peripheral immune cells.

6.1. Microglia

Microglia are residential innate immune cells that perform primary immune surveillance and macrophage-like activities of the CNS. In a non-stimulated state, microglia contribute to CNS development and maintain tissue homeostasis by supporting neuronal survival, cell death, and synaptogenesis [113]. However, microglial cells can be activated by various pathological stimuli during infections, brain trauma, stroke, and neurodegeneration [114]. Activated microglia are characterized by increased proliferation and the production and secretion of a wide spectrum of immune mediators such as cytokines, chemokines, prostaglandins, and reactive oxygen intermediates [115,116,117,118]. The production of cytokines and chemokines can facilitate the recruitment of peripheral leukocytes into the brain [119]. During neuroinflammation, activated microglia migrate to the site of injury or infection and perform pivotal immunological functions such as phagocytosis of invading microorganisms and removal of dead or damaged cells [120,121]. However, chronic activation of microglia is generally considered to be detrimental for neuronal health [122].
In response to activation, microglia can polarize to either a pro-inflammatory M1 phenotype or an anti-inflammatory M2 phenotype, although microglial activation states have been recognized to be more complex [123]. Various stimuli induce the classical M1 activation state of microglia such as LPS, interferon (IFN)-γ, amyloid β (Aβ), and α-synuclein [118,124,125,126]. Toll-like receptor 4 (TLR4), a member of the pattern recognition receptor family, mediates innate immunity and is abundantly expressed in microglia [127]. In fact, TLR4-dependent microglial activation has been observed in various neurodegenerative diseases like Alzheimer’s disease (AD) and Parkinson disease (PD) [128,129]. In addition, TLR4 is also responsible for chronic neuroinflammation after stroke and spinal cord injury leading to brain damage [130]. TLR4 can be activated by multiple pathogen-associated molecular patterns (PAMPs), such as LPS which is a major component of the outer membrane of Gram-negative bacteria [130]. LPS is one of the most extensively studied TLR4 ligands to understand the mechanism of microglial activation in neurodegeneration [127]. In addition to the above-mentioned pro-inflammatory response, microglia can also adopt an anti-inflammatory M2 phenotype. After M1 microglia attack invading organisms to limit tissue damage, anti-inflammatory M2 microglia are involved in phagocytosis of cellular debris and wound healing [123]. Another more immuno-suppressive phenotype is induced by the cytokines IL-10 and TGF-β, or by apoptotic cells [123].
Microglia are a crucial source of AEA and 2-AG under basal conditions and during neuroinflammation [75,76,77]. When stimulated with ATP (released from damaged tissue) microglia produce 2-AG themselves [77]. CB2 receptor expression is upregulated in microglia stimulated with pro-inflammatory cytokines [131], indicating a significant role of CB2 in the regulation of neuroinflammatory states. Further, it was demonstrated that 2-AG and PEA affect microglial cells and lead to a decrease in the number of damaged neurons after excitotoxical lesion in organotypic hippocampal slice cultures. 2-AG activated the abnormal cannabidiol (abn-CBD) receptor and PEA was shown to mediate neuroprotection via PPAR-α [132]. The CB2 agonist AM1241 has also been shown to attenuate microglia activation by reducing the expression of the inducible nitric oxide synthase and shifting their phenotype from M1 to M2 [133].
Microglia constantly scan their environment and due to the long processes directed towards synapses can monitor and respond to the functional status of synapses. Microglia contribute to neuronal circuit maturation and are involved in both synapse induction and elimination. During neuroinflammation, prolonged activation of microglial cells attracted to lesion sites can exacerbate neuronal damage. It is anticipated that 2-AG, produced by overstimulated neurons, induces microglial migration [76] and proliferation [134].

6.2. Astrocytes

Astrocytes are the largest and most abundant group of glial cells in the CNS and play a vital role in regulating CNS homeostasis, synaptic transmission and plasticity, and neuroprotective effects. Glia respond to neuronal activity with the elevation of their internal Ca2+ concentration, which triggers the release of mediators of glial origin. The term ‘tripartite synapse’ describes the bidirectional communication between astrocytes and neurons, where nearby astrocytes respond to synaptic activity and, vice versa, regulate synaptic transmission and plasticity [135]. Perisynaptic Schwann cells and synaptically associated astrocytes are viewed as integral modulatory elements of tripartite synapses.
CB1 activation in astrocytes amplifies Ca2+ influx and promotes the release of gliotransmitters, like glutamate (gliotransmission), which modulate the target response at pre- and postsynaptic sites [24]. Impairment of spatial working memory and in vivo long-term depression (LTD) of synaptic strength at hippocampal CA3-CA1 synapses, induced by an acute exposure of exogenous cannabinoids, is due to the activation of astroglial CB1 and is associated with astroglia-dependent hippocampal LTD [136].
Due to their proximity to blood vessels and other resident cells within the CNS such as neurons, microglia, and oligodendrocytes, astrocytes play a crucial role in BBB maintenance and permeability. In addition, astrocytes are involved in modulating the innate immune response by regulating inflammatory factors, such as cytokines, chemokines, complement fragments, reactive oxygen, or reactive nitrogen species. However, dysfunctional astrocytes seem to play an important role in the onset of neurodegenerative diseases such as AD and ALS (reviewed in [137]).
Astrocytes become activated or reactivated during various pathological conditions such as stroke, trauma, tumor growth, and neurodegenerative diseases. Following the M1/M2 phenotype classification of microglia and macrophages, neuroinflammation can induce two types of reactive astrocytes, termed A1 and A2 [138]. In A1 reactive astrocytes the pro-inflammatory NF-κB pathway is upregulated leading to the release of complement factors [139] that are destructive to synapses and to the secretion of neurotoxins and pro- and anti-inflammatory mediators such as PGD2, IFN-γ, TNF-α, IL-1-β, and TGF-β, respectively [140,141,142]. In A2 reactive astrocytes enhanced STAT3 activity has been observed [143]. Moreover, A2 reactive astrocytes can upregulate many neurotrophic factors such as thrombospondins [144] and brain-derived neurotrophic factor (BDNF) [145,146], which promote either survival and growth of neurons or synaptic repair.
Accumulating evidence supports the role of astrocytes as a source of eCBs. It has been shown that astrocytes have the potential to produce 2-AG in response to ATP [147], endothelin [148], and CB1 receptor activation [75,149]. Furthermore, the secretion of AEA, homo-gamma-linolenylethanolamide (HEA), and docosatetraenoylethanolamide (DEA) by activated mouse astrocytes has been confirmed [75].
Of note, not only microglial cannabinoid receptors but also astroglial CB1 and CB2 receptors play critical roles in the response to neuroinflammation [33]. Activation of astroglial CB1 receptors protects against ceramide-induced oxidative stress and apoptosis [150,151] and activation of both CB1 and CB2 receptors seems to prevent LPS-induced nitric oxide (NO) release by cultured astrocytes [152]. Accordingly, 2-AG has been shown to maintain glutamine synthase expression in astrocytes in a MAPK-dependent manner and thus to protect astrocytes from LPS exposure [153]. 2-AG also seems to reduce the astrocytic production of chondroitin sulfate proteoglycan, that accumulates in MS lesions and is thought to be linked to the failure to regenerate, impeding oligodendrocyte precursor cell differentiation, and neuronal growth [154]. Moreover, 2-AG protects astrocytes exposed to oxygen-glucose deprivation through a blockade of NDRG2 signaling and STAT3 phosphorylation [155]. 2-AG and PEA have been shown to attenuate amyloid β-induced astrocyte activation and PEA increased 2-AG production in astrocytes [156,157,158]. PEA is also suggested to improve neuronal survival by possibly counteracting reactive astrogliosis [159,160]. PEA- and OEA-mediated inhibition of astrocyte activation seems to involve PPAR-α [161,162]. Furthermore, AEA has been shown to elicit glutamate release through astrocytic CB1 receptor activation in the core of nucleus accumbens in rats [163].
Astrocytes isolated from mice with acute experimental autoimmune encephalomyelitis (EAE) exhibited reductions in all endocannabinoid metabolism-associated genes, with the exception of Faah, which persisted in chronic disease and was associated with reduced Cnr1 transcript levels both at acute and recovery phases [164]. Astrocytic- together with neuronal MAGL seems to be responsible for converting 2-AG to prostaglandins and thus protects the nervous system from excessive CB1 receptor activation [165].

6.3. Cytokines Involved in Neuroinflammation

Besides microglia and astrocytes, endothelial cells and other glial cells, may produce cytokines and chemokines. Common cytokines which are produced in response to brain injury or during neurodegenerative diseases able to induce neuronal cytotoxicity are IL-6, IL-1β, and TNF-α [166]. Moreover, sustained release of these cytokines leads to a compromised BBB [167]. Subsequently, peripheral immune cells such as macrophages, neutrophils, monocytes, T cells, and B cells are able to migrate into the brain. This process exacerbates and contributes to chronic neuroinflammation and neurodegeneration. For instance, following traumatic brain injury, IL-1β induces neuronal apoptosis, BBB breakdown, recruitment of immune cells, as well as the production of pro-inflammatory mediators [168]. Moreover, during spinal cord injury, the secretion of pro-inflammatory mediators including IL-1β, inducible nitric oxide synthase (iNOS), IFN-γ, IL-6, IL-23, and TNF-α is followed by the activation of local microglia and attraction of various immune cells such as naive bone-marrow derived macrophages [169]. Upon infiltration of the injured site, macrophages undergo phenotype switching from M2 phenotype to M1-like phenotype. Noteworthy, normal aging is often associated with an increased number of activated microglia in the brain which are involved in altered synaptic plasticity mechanisms in the hippocampus, including LTP and thereby reduce memory performance [170]. Moreover, aged brains show homeostatic imbalance between anti-inflammatory and pro-inflammatory cytokines increasing the risk for neurodegenerative diseases such as AD. Although the pro-inflammatory cytokines may cause cell death and tissue damage, they are also involved in tissue repair [171]. For example, TNF-α causes neurotoxicity at early stage, but contributes to tissue growth at later stages of neuroinflammation.
Several studies have highlighted the regulatory effects of the ECS on neuroinflammatory conditions by modulating the production of cytokines. For instance, 2-AG was shown to prevent the overexpression of TNF-α, IL-1β, and iNOS in a murine model of SO2-induced brain inflammation [172]. MAGL-deficiency leading to increased 2-AG levels also reduced brain PGE2 and pro-inflammatory cytokine levels following peripheral LPS administration in mice [173]. Selective pharmacologic inhibition of ABHD6 diminished cytokine and chemokine production in a murine model of neuropathic pain [174], and the MAGL inhibitor CPD-4645 significantly reduced IL-1β and IL-6 brain levels after systemic LPS challenge [175]. The FAAH inhibitor URB597 attenuated increased TNF-α and IL-1β levels in the hippocampi of aged mice [176] and decreased Iba-1, TNF-α, IL-6, and monocyte chemoattractant protein-1 (MCP-1) levels in the hippocampus of ethanol-exposed rats [177]. The FAAH inhibitor PF3845 increased levels of AEA, OEA, and PEA in the frontal cortex and hippocampus of rats [178]. Furthermore, this increase in FAAH substrate levels was associated with a robust attenuation in TNF-α, IL-6, and IL-1β levels in the prefrontal cortex. PEA has been shown to reduce pro-inflammatory cytokines after traumatic spinal cord [179] and brain injury [180,181], in a model of sciatic nerve crush [182], in Parkinson’s disease models [183,184] and in MS patients [185]. Further, OEA administration significantly reduced plasma and brain TNF-α levels after LPS application [186] and SEA was recently shown to suppress increased TNF-α and TGF-β1 levels in the prefrontal cortex of LPS challenged mice [47].
CB2 receptor agonism has been shown to decrease brain levels of pro-inflammatory cytokines induced by LPS application [187], intracerebral hemorrhages [188], or surgery [189] as well as in a model of PD [190]. At the same time the CB1 receptor inverse agonist SR141716A (rimonabant) and the CB2 receptor antagonist SR144528 significantly reduced LPS-induced IL-1β production in the brain [191] whereas SR141716A was also shown to increase pro-inflammatory cytokines in an EAE model [192]. Neuroprotective effect of SR141716A was shown in the retinal degeneration model [193] and in permanent photothrombotic cerebral ischemia [194]. These findings indicate that manipulation of CB1 or CB2 receptors may have therapeutic value in neuroinflammation; however, due to the complexity of the ECS, this concept remains to be carefully considered.

6.4. Effects of Endocannabinoids and Related Compounds on Neurodegenerative Diseases

Not only inflammatory disorders or tissue injury, but also neurodegenerative diseases are accompanied or caused by neuroinflammation. Neurodegeneration also refers to chronic and progressive loss of neuronal functions in the brain and spinal cord. Particularly, AD is characterized by amyloid β plaques and neurofibrillary tangles that cause a decline in memory and cognitive abilities. In addition to neuroinflammation, systemic infection and inflammation, characterized by a substantial amount of proinflammatory mediators in the circulation, have also been correlated to increased risks of developing AD [195]. Chronic inflammation is also a hallmark of PD, which is characterized by the loss of dopaminergic neurons and the presence of α-synuclein-containing aggregates in the substantia nigra pars compacta [196]. Huntington’s disease (HD) is a devastating neurodegenerative genetic disorder associated with progressive loss of a specific type of neurons found in the striatum and cortex [197]. Unfortunately, the relationship between neuroinflammation markers and the disease pathology is still poorly understood [198]. Amyotrophic lateral sclerosis (ALS) is indicated by the degeneration of motor neurons [199] and characterized by the occurrence of a neuroinflammatory reaction consisting of activated glial cells, mainly microglia and astrocytes, and T cells [200]. Of the neurodegenerative diseases, multiple sclerosis (MS) is an exception because neuronal death in MS is considered to be secondary to the initiating activity of autoreactive T cells that target myelin [201]. While neuroinflammation in MS involves infiltration of peripheral immune cells, breakdown of the BBB, and activation of CNS-resident glial cells, neuroinflammation in the other neurodegenerative diseases is more restricted to glial cell activation and inflammatory reactions in the parenchyma [201].
Several animal models and human studies have demonstrated that the ECS significantly influences the development of neuroinflammation and the progression of brain injury and neurodegenerative diseases [1,2,3]. Using a model of cerebral focal ischemia, it was shown that exogenously administered AEA and 2-AG in combination reduced infarct size in rats, but with no facilitatory effects beyond AEA or 2-AG alone [202]. Other studies reported neuroprotective effects of exogenous AEA [203] and 2-AG [56] under traumatic brain injury (TBI). Several studies documented the positive effects of PEA in experimental and clinical studies of TBI, spinal cord injuries, pain, cerebral ischemia, PD, and AD [179,180,181,183,184,185,204,205,206,207,208]. Moreover, the stimulation of CB1 and CB2 has been shown to be beneficial in neurodegenerative disorders such as AD and PD (reviewed in [209,210]). In contrast, the CB1 inverse agonist SR141716A (rimonabant) showed promising results in PD preclinical studies [211,212]; however, clinical studies with SR141716A failed to improve motor disability in PD patients [213]. Importantly, SR141716A, which has also been approved as an effective anti-obesity drug, shows a serious psychiatric side effect profile, and thus has been withdrawn from the pharmaceutical market worldwide. These contradictory data, indicate the interference of CB1 inverse agonists/antagonists with a basal ECS tone, prevailing in healthy conditions and essentially involved in the homeostatic regulation of brain function.
It has been observed that CB2 receptors and FAAH are selectively overexpressed in neuritic plaque-associated glia in AD [214], especially in reactive astrocytes and activated microglial cells [215]. In this regard, the use of hydrolase inhibitors has been mentioned as a promising therapeutic option for neuroinflammatory and neurodegenerative diseases. Treatment with MAGL and FAAH inhibitors, which have the capacity to increase the level of eCBs indirectly, leads to anxiolytic, antidepressant, and anti-inflammatory effects, and reduces amyloid β deposition and inhibition of the death of dopaminergic neurons, which are associated with the pathogenesis of AD and PD, respectively (reviewed in [216,217]). An overview of in vivo studies reviewing the effects of MAGL and FAAH inhibitors in neuroinflammation and neurodegenerative diseases, is given in Table 3. Several FAAH and MAGL inhibitors entered clinical phase I and II studies for neurological disorders such as pain, anxiety, Tourette syndrome, and cannabis withdrawal. Moreover, dual FAAH/cholinesterase inhibitors which might be beneficial for neurodegenerative diseases are currently under development [218]. In addition, an increasing number of studies have characterized the beneficial effects of NAAA inhibitors, which prevent the degradation of NAEs, in preclinical studies of pain and (neuro-) inflammation [219,220,221] (Table 3). However, in a phase I study, the FAAH inhibitor BIA 10-2474 resulted in severe adverse events such as lethal toxic cerebral syndrome [222], leading to the discontinuation of several clinical studies employing FAAH inhibitors. Thus, there is an urgent need of further research to develop highly specific, short-acting, indirect cannabinoid therapies with better safety profiles.

7. The Role of the Blood–Brain Barrier Integrity in Restriction of Systemic Inflammation

CNS dysfunction associated with systemic infection is common and includes symptoms such as sickness behavior and delirium. In the context of sepsis, CNS dysfunction is known as septic encephalopathy. A key step in the pathogenesis is the systemic production of pro-inflammatory cytokines such as TNF-α and IL-1β, which then act on the brain. Cytokine transport systems of the BBB are likely to play a role in permitting the passage of these signals [241]. Moreover, AD, MS, and CNS dysfunction in systemic infection are examples of conditions which are primarily neurodegenerative, neuroinflammatory, or systemic. In many cases it is not clear whether BBB changes are the cause or consequence of neuropathology, and it is possible that BBB changes and neuropathology drive each other in a self-perpetuating manner, contributing to disease progression.

7.1. Structure of the Blood–Brain Barrier

Histologically the BBB is a specialized multi-layered unit composed of a thick continuous glycocalyx, non-fenestrated endothelial cells with reduced vesicular activity and linked by tight junctions, two basement membranes (vascular basement membrane and glia limitans), and astrocytic end-feet. All elements of this ‘neurovascular unit’ contribute to the functional BBB. At the molecular level, there are ectoenzymes, receptors and transporters which regulate or reverse traffic across the BBB. Together, these components enable a stable CNS environment to minimize traffic of inflammatory cells and molecules, local inflammation, and prevent potential neuronal damage.

7.2. The BBB in Systemic Inflammation

Systemic inflammation, as induced by LPS, can lead to disruptive and non-disruptive BBB changes. Disruptive BBB change is accompanied by endothelial cell damage or tight junctional modifications, while non-disruptive change occurs at a molecular level. Identified mechanisms of LPS-induced disruptive BBB change include modification of tight junctions, endothelial damage and apoptosis, degradation of glycocalyx, breakdown of glia limitans, and astrocyte alteration (reviewed in [242]). Moreover, several reports demonstrate that systemic inflammation upregulates several endothelial cell receptors and transporters, induces cytokine production by endothelial cells, modulates astrocyte function, and enhances pathogen neuroinvasion without any visible changes in the BBB architecture [242].

7.3. The Role of the ECS in the Maintenance of the Blood Brain Barrier

2-AG is the most abundant endocannabinoid in the CNS and is elevated after brain injury and during neuroinflammation. Because of its rapid hydrolysis, however, the compensatory and neuroprotective effect of 2-AG is short-term. It has been shown previously that 2-AG decreases BBB permeability and inhibits the acute expression of the main proinflammatory cytokines TNF-α, IL-1β, and IL-6 [243]. Moreover, several reports demonstrated that inhibition of 2-AG and AEA degradation supports the BBB integrity in experimental traumatic brain injury [244,245] and ischemic insults [175]. Accordingly, CB2 agonists prevents BBB damage in several experimental models of brain injury and neurodegenerative disease [188,246,247,248,249,250,251,252,253,254,255,256,257]. Previously, Hind and colleagues suggested that AEA, OEA, and PEA may play an important modulatory role in normal BBB physiology, and afford protection to the BBB during ischemic stroke [258]. In addition, it has been shown that PPAR-α is involved in the protective effects of OEA and OEA analogues against ischemic brain injury, particularly in terms of BBB disruption [259,260]. A study from Mestre and colleagues suggested that CB1 receptor dependent inhibition of vascular cell adhesion molecule (VCAM) 1 is a novel mechanism for AEA-induced leukocyte transmigration trough the BBB [261]. Moreover, CB1 receptor blockade reduced leukocyte adhesion to intestinal microvasculature in a mouse model of systemic sepsis [213], potentially also affecting immune cell transmigration at the BBB. The major endogenously produced NAE and one of the most stable among these compounds, SEA prevents leukocyte, especially neutrophil, infiltration into the brain in a murine model of LPS-induced systemic inflammation [47].

7.4. Leukocyte Recruitment

Leukocytes entering the brain from peripheral circulation must pass through the BBB, the choroid plexus that forms the blood–cerebrospinal fluid barrier, and through post-capillary venules at the pial surface into subarachnoid and Virchow-Robin perivascular spaces [262,263]. These routes typically operate together and both the paracellular (junctional) and transcellular (non-junctional) mechanisms can be involved in leukocyte trafficking across the neurovascular unit [264]. Although cellular influx into the CNS is physiologically non-disruptive, it may result in disruptive BBB change. Leukocyte recruitment across the BBB in response to systemic inflammation has been demonstrated for lymphocytes [265], neutrophils [266], and monocytes [267]. Mechanistically, systemic inflammation can promote leukocyte transmigration at various points during the two-step passage through the endothelium and glia limitans.
On circulatory immune cells, CB1 and CB2 receptors are expressed at low levels in healthy human donors. Highest expression levels of CB2 were observed on NK-cell, B-cells and monocytes, while low levels could be found in T cells and neutrophils isolated from peripheral blood of healthy donors [268]. Stimulation with pro-inflammatory cytokines, such as IL-6 and TNF-α, enhanced transcription of both CB receptors, while the effect was more pronounced for CB2 [269]. In line with this, pro-inflammatory cytokine stimulation of murine bone marrow-derived macrophages increased CB2 expression [270], rendering inflammatory macrophages more susceptible to CB signaling. A detailed overview of the reported effects of CB1 and CB2 receptor activation in leukocytes is given in Table 4.

8. Conclusions

Neuroinflammation is caused and/or accompanied by the infiltration of immune cells through the BBB and secretion of a range of pro-inflammatory cytokines and other molecules with neurotoxic potential. These changes together with glutamate receptor-mediated neurotoxicity and neurodegenerative processes underlie the pathogenesis of several neuropathologies, among them are Alzheimer’s disease, amyotrophic lateral sclerosis, stroke, multiple sclerosis, and Parkinson’s disease. Bacterial and viral infection, traumatic injury, and autoimmune disease compromise the integrity of BBB and favor the transition of systemic inflammation to neuroinflammation. Maintenance or restauration of the selective permeability of the blood–brain barrier is one of the therapeutic strategies under systemic inflammation to prevent systemic inflammation from spreading to the CNS.
The interplay of true eCB and ligands that now belong to expanded ECS allow taking into the consideration all relevant molecular targets for therapy of neuroinflammation-associated neuropathologies. The interference of eCB congeners with enzymatic degradation or endocannabinoid signaling suggests their role in tuning the activity of primary eCBs. The ‘entourage effect’ of the produced non-cannabinoid 2-acylglycerols, NAEs and N-acylneurotransmitters may serve as an additional fine regulator of cannabinoid activity.
Resident microglia and astrocytes are tightly coupled to functions of active synaptic contacts and are highly involved in inflammatory progression, pro-survival changes and resolution of neuroinflammation. These cells promote neuronal survival, synaptogenesis, spine induction, and illumination and protect neurons from toxic metabolites. The contribution of eCB signaling into the functional coupling of neurons, astrocytes, and microglia, suggests that in line with conception of tripartite synapses, microglial cells are equal participants in such communication. Being activated during the immune response, microglial cells contribute to the resolution of neuroinflammation; however, chronic activation of microglia is detrimental to neurons and contributes to the development of synaptopathy in various neurodegenerative diseases. Components of ECS play an active role in the reactivity of these cells during inflammation, attenuate the production of pro-inflammatory cytokines, and mediate neuroprotection against glutamate-receptor mediated excitotoxicity, ischemia, and oxidative stress.
Normal synaptic activity as well as pathological overstimulation of postsynaptic neurotransmitter receptors is a potent trigger for the production of eCBs and non-cannabinoid NAEs. Glutamate-induced endocannabinoids [78] flown from active synapses and injured sites might attract resident microglial cells [76,79]. Tight structural and functional cooperation of synaptic contacts, astrocytes, and microglia enables highly dynamic response to synaptic events.
Retrograde endocannabinoid signaling is implicated in several forms of short- and long-term synaptic plasticity. These lipid mediators reach the presynaptic sites of the same or other synaptic contacts and by binding to CB1/2 inhibit the synaptic vesicle fusion and neurotransmitter release. This is how synaptic contacts dynamically tune their own strength and can potentiate/depress the response depending on present inputs.
Released eCBs have a restricted area of action due to short half-lives and differences in CB receptor expression on cells in close vicinity. Thus, the concentration gradient of eCBs is formed on the site of their synthesis. This makes the pharmacological inhibition of eCB degradation primarily effective in injured sites, where they are actively produced. Novel, highly selective inhibitors of FAAH and MAGL with a good safety profile may become prospective agents with anti-nociceptive, anxiolytic, and anti-inflammatory activity and targeted action.
The overall interplay and metabolism of endogenous ligands of CB1/2, TRPV1, GPR55, and GPR18 is now integrated in the “endocannabinoidome”, which is actively involved in the intrinsic response to inflammation and neuroinflammation. The polymodality of this system provides a wide field for development of highly efficient neuroprotective agents for the therapy of inflammation-associated synaptopathy.

Author Contributions

Conceptualization, E.M.S. and L.A.K.; writing—original draft preparation, E.M.S., L.A.K., and S.R.; writing—review and editing, E.M.S., L.A.K., and S.R.; supervision, E.M.S. and L.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Austrian Science Fund (FWF) grant W1241, the NIH grant R35 GM122567 and by the OeAD grant UA 09/2017.

Acknowledgments

Open Access Funding by the Austrian Science Fund (FWF). Figures were created with Biorender.com (accessed on 7 April 2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Basavarajappa, B.S.; Shivakumar, M.; Joshi, V.; Subbanna, S. Endocannabinoid system in neurodegenerative disorders. J. Neurochem. 2017, 142, 624–648. [Google Scholar] [CrossRef] [PubMed]
  2. Schurman, L.D.; Lichtman, A.H. Endocannabinoids: A Promising Impact for Traumatic Brain Injury. Front. Pharmacol. 2017, 8, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ranieri, R.; Laezza, C.; Bifulco, M.; Marasco, D.; Malfitano, A.M. Endocannabinoid System in Neurological Disorders. Recent Patents CNS Drug Discov. 2016, 10, 90–112. [Google Scholar] [CrossRef] [PubMed]
  4. Tsuboi, K.; Uyama, T.; Okamoto, Y.; Ueda, N. Endocannabinoids and related N-acylethanolamines: Biological activities and metabolism. Inflamm. Regen. 2018, 38, 28. [Google Scholar] [CrossRef] [PubMed]
  5. Bisogno, T.; Berrendero, F.; Ambrosino, G.; Cebeira, M.; Ramos, J.; Fernandez-Ruiz, J.; Di Marzo, V. Brain Regional Distribution of Endocannabinoids: Implications for Their Biosynthesis and Biological Function. Biochem. Biophys. Res. Commun. 1999, 256, 377–380. [Google Scholar] [CrossRef]
  6. Felder, C.C.; Nielsen, A.; Briley, E.M.; Palkovits, M.; Priller, J.; Axelrod, J.; Nguyen, D.N.; Richardson, J.M.; Riggin, R.M.; Koppel, G.A.; et al. Isolation and measurement of the endogenous cannabinoid receptor agonist, anandamide, in brain and peripheral tissues of human and rat. FEBS Lett. 1996, 393, 231–235. [Google Scholar] [CrossRef] [Green Version]
  7. Schmid, H.H.; Schmid, P.C.; Berdyshev, E.V. Cell signaling by endocannabinoids and their congeners: Questions of selectivity and other challenges. Chem. Phys. Lipids 2002, 121, 111–134. [Google Scholar] [CrossRef]
  8. Natarajan, V.; Reddy, P.; Schmid, P.; Schmid, H. On the biosynthesis and metabolism of N-acylethanolamine phospholipids in infarcted dog heart. Biochim. Biophys. Acta BBA Lipids Lipid Metab. 1981, 664, 445–448. [Google Scholar] [CrossRef]
  9. Gulaya, N.M.; Volkov, G.L.; Klimashevsky, V.M.; Govseeva, N.N.; Melnik, A.A. Changes in lipid composition of neuro-blastoma C1300 N18 cell during differentiation. Neuroscience 1989, 30, 153–164. [Google Scholar] [CrossRef]
  10. Kasatkina, L.A.; Tarasenko, A.S.; Krupko, O.O.; Kuchmerovska, T.M.; Lisakovska, O.O.; Trikash, I.O. Vitamin D deficiency induces the excitation/inhibition brain imbalance and the proinflammatory shift. Int. J. Biochem. Cell Biol. 2020, 119, 105665. [Google Scholar] [CrossRef]
  11. 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] [Green Version]
  12. Kozak, K.; Marnett, L. Oxidative metabolism of endocannabinoids. Prostaglandins Leukot. Essent. Fat. Acids 2002, 66, 211–220. [Google Scholar] [CrossRef] [PubMed]
  13. Rouzer, C.A.; Marnett, L.J. Non-redundant Functions of Cyclooxygenases: Oxygenation of Endocannabinoids. J. Biol. Chem. 2008, 283, 8065–8069. [Google Scholar] [CrossRef] [Green Version]
  14. Tsuboi, K.; Sun, Y.-X.; Okamoto, Y.; Araki, N.; Tonai, T.; Ueda, N. Molecular Characterization of N-Acylethanolamine-hydrolyzing Acid Amidase, a Novel Member of the Choloylglycine Hydrolase Family with Structural and Functional Similarity to Acid Ceramidase. J. Biol. Chem. 2005, 280, 11082–11092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Iannotti, F.A.; Di Marzo, V.; Petrosino, S. Endocannabinoids and endocannabinoid-related mediators: Targets, metabolism and role in neurological disorders. Prog. Lipid Res. 2016, 62, 107–128. [Google Scholar] [CrossRef]
  16. Solorzano, C.; Zhu, C.; Battista, N.; Astarita, G.; Lodola, A.; Rivara, S.; Mor, M.; Russo, R.; Maccarrone, M.; Antonietti, F.; et al. Selective N-acylethanolamine-hydrolyzing acid amidase inhibition reveals a key role for endogenous palmitoylethanolamide in inflammation. Proc. Natl. Acad. Sci. USA 2009, 106, 20966–20971. [Google Scholar] [CrossRef] [Green Version]
  17. Urquhart, P.; Nicolaou, A.; Woodward, D. Endocannabinoids and their oxygenation by cyclo-oxygenases, lipoxygenases and other oxygenases. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2015, 1851, 366–376. [Google Scholar] [CrossRef] [PubMed]
  18. McDougle, D.R.; Watson, J.E.; Abdeen, A.A.; Adili, R.; Caputo, M.P.; Krapf, J.E.; Johnson, R.W.; Kilian, K.A.; Holinstat, M.; Das, A. Anti-inflammatory omega-3 endocannabinoid epoxides. Proc. Natl. Acad. Sci. USA 2017, 114, E6034–E6043. [Google Scholar] [CrossRef] [Green Version]
  19. Hu, S.S.-J.; Bradshaw, H.B.; Benton, V.M.; Chen, J.S.-C.; Huang, S.M.; Minassi, A.; Bisogno, T.; Masuda, K.; Tan, B.; Roskoski, R.; et al. The biosynthesis of N-arachidonoyl dopamine (NADA), a putative endocannabinoid and endovanilloid, via conjugation of arachidonic acid with dopamine. Prostaglandins Leukot. Essent. Fat. Acids 2009, 81, 291–301. [Google Scholar] [CrossRef] [Green Version]
  20. Bouaboula, M.; Poinot-Chazel, C.; Bourrié, B.; Canat, X.; Calandra, B.; Rinaldi-Carmona, M.; Le Fur, G.; Casellas, P. Activation of mitogen-activated protein kinases by stimulation of the central cannabinoid receptor CB1. Biochem. J. 1995, 312, 637–641. [Google Scholar] [CrossRef] [Green Version]
  21. Gao, F.; Zhang, L.H.; Su, T.F.; Li, L.; Zhou, R.; Peng, M.; Wu, C.H.; Yuan, X.C.; Sun, N.; Meng, X.F.; et al. Signaling Mechanism of Cannabinoid Receptor-2 Activation-Induced beta-Endorphin Release. Mol. Neurobiol. 2016, 53, 3616–3625. [Google Scholar] [CrossRef]
  22. Glass, M.; Felder, C.C. Concurrent Stimulation of Cannabinoid CB1 and Dopamine D2 Receptors Augments cAMP Accumulation in Striatal Neurons: Evidence for a Gs Linkage to the CB1 Receptor. J. Neurosci. 1997, 17, 5327–5333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lauckner, J.E.; Hille, B.; Mackie, K. The cannabinoid agonist WIN55,212-2 increases intracellular calcium via CB1 receptor coupling to Gq/11 G proteins. Proc. Natl. Acad. Sci. USA 2005, 102, 19144–19149. [Google Scholar] [CrossRef] [Green Version]
  24. Navarrete, M.; Araque, A. Endocannabinoids Mediate Neuron-Astrocyte Communication. Neuron 2008, 57, 883–893. [Google Scholar] [CrossRef] [Green Version]
  25. Guo, J.; Ikeda, S.R. Endocannabinoids modulate N-type calcium channels and G-protein-coupled inwardly rectifying potas-sium channels via CB1 cannabinoid receptors heterologously expressed in mammalian neurons. Mol. Pharmacol. 2004, 65, 665–674. [Google Scholar] [CrossRef] [PubMed]
  26. Ibsen, M.S.; Finlay, D.B.; Patel, M.; Javitch, J.A.; Glass, M.; Grimsey, N.L. Cannabinoid CB1 and CB2 Receptor-Mediated Arrestin Translocation: Species, Subtype, and Agonist-Dependence. Front. Pharmacol. 2019, 10, 350. [Google Scholar] [CrossRef] [PubMed]
  27. McGuinness, D.; Malikzay, A.; Visconti, R.; Lin, K.; Bayne, M.; Monsma, F.; Lunn, C.A. Characterizing cannabinoid CB2 receptor ligands using DiscoveRx PathHunter beta-arrestin assay. J. Biomol. Screen 2009, 14, 49–58. [Google Scholar] [CrossRef] [Green Version]
  28. Jin, W.; Brown, S.; Roche, J.P.; Hsieh, C.; Celver, J.P.; Kovoor, A.; Chavkin, C.; Mackie, K. Distinct Domains of the CB1 Cannabinoid Receptor Mediate Desensitization and Internalization. J. Neurosci. 1999, 19, 3773–3780. [Google Scholar] [CrossRef] [Green Version]
  29. 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] [Green Version]
  30. Zhang, H.-Y.; Gao, M.; Liu, Q.-R.; Bi, G.-H.; Li, X.; Yang, H.-J.; Gardner, E.L.; Wu, J.; Xi, Z.-X. Cannabinoid CB2 receptors modulate midbrain dopamine neuronal activity and dopamine-related behavior in mice. Proc. Natl. Acad. Sci. USA 2014, 111, E5007–E5015. [Google Scholar] [CrossRef] [Green Version]
  31. Onaivi, E.S.; Ishiguro, H.; Gong, J.-P.; Patel, S.; Meozzi, P.A.; Myers, L.; Perchuk, A.; Mora, Z.; Tagliaferro, P.A.; Gardner, E.; et al. Brain Neuronal CB2 Cannabinoid Receptors in Drug Abuse and Depression: From Mice to Human Subjects. PLoS ONE 2008, 3, e1640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. 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]
  33. Haspula, D.; Clark, M.A. Cannabinoid Receptors: An Update on Cell Signaling, Pathophysiological Roles and Therapeutic Opportunities in Neurological, Cardiovascular, and Inflammatory Diseases. Int. J. Mol. Sci. 2020, 21, 7693. [Google Scholar] [CrossRef] [PubMed]
  34. Stempel, A.V.; Stumpf, A.; Zhang, H.-Y.; Özdoğan, T.; Pannasch, U.; Theis, A.-K.; Otte, D.-M.; Wojtalla, A.; Rácz, I.; Ponomarenko, A.; et al. Cannabinoid Type 2 Receptors Mediate a Cell Type-Specific Plasticity in the Hippocampus. Neuron 2016, 90, 795–809. [Google Scholar] [CrossRef] [Green Version]
  35. Sharir, H.; Console-Bram, L.; Mundy, C.; Popoff, S.N.; Kapur, A.; Abood, M.E. The Endocannabinoids Anandamide and Virodhamine Modulate the Activity of the Candidate Cannabinoid Receptor GPRJ. Neuroimmune Pharmacol. 2012, 7, 856–865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Irving, A.; Abdulrazzaq, G.; Chan, S.L.; Penman, J.; Harvey, J.; Alexander, S.P. Cannabinoid Receptor-Related Orphan G Protein-Coupled Receptors. Rapid Acting Antidepress. 2017, 80, 223–247. [Google Scholar] [CrossRef]
  37. Brown, R.C.; Cascio, C.; Papadopoulos, V. Pathways of neurosteroid biosynthesis in cell lines from human brain: Regulation of dehydroepiandrosterone formation by oxidative stress and beta-amyloid peptide. J. Neurochem. 2000, 74, 847–859. [Google Scholar] [CrossRef]
  38. O’Sullivan, S.E. An update on PPAR activation by cannabinoids. Br. J. Pharmacol. 2016, 173, 1899–1910. [Google Scholar] [CrossRef] [Green Version]
  39. D’Agostino, G.; La Rana, G.; Russo, R.; Sasso, O.; Iacono, A.; Esposito, E.; Mattace Raso, G.; Cuzzocrea, S.; Loverme, J.; Piomelli, D.; et al. Central administration of palmitoylethanolamide reduces hyperalgesia in mice via inhibition of NF-kappaB nuclear signalling in dorsal root ganglia. Eur. J. Pharmacol. 2009, 613, 54–59. [Google Scholar] [CrossRef]
  40. Beggiato, S.; Tomasini, M.C.; Cassano, T.; Ferraro, L. Chronic Oral Palmitoylethanolamide Administration Rescues Cognitive Deficit and Reduces Neuroinflammation, Oxidative Stress, and Glutamate Levels in A Transgenic Murine Model of Alzheimer’s Disease. J. Clin. Med. 2020, 9, 428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Rankin, L.; Fowler, C.J. The basal pharmacology of palmitoylethanolamide. Int. J. Mol. Sci. 2020, 21, 7942. [Google Scholar] [CrossRef] [PubMed]
  42. Ambrosino, P.; Soldovieri, M.V.; Russo, C.; Taglialatela, M. Activation and desensitization of TRPV1 channels in sensory neurons by the PPARalpha agonist palmitoylethanolamide. Br. J. Pharmacol. 2013, 168, 1430–1444. [Google Scholar] [CrossRef] [Green Version]
  43. Ryberg, E.; Larsson, N.; Sjögren, 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]
  44. Di Marzo, V.; Melck, D.; Orlando, P.; Bisogno, T.; Zagoory, O.; Bifulco, M.; Vogel, Z.; De Petrocellis, L. Palmitoylethanolamide inhibits the expression of fatty acid amide hydrolase and enhances the anti-proliferative effect of anandamide in human breast cancer cells. Biochem. J. 2001, 358, 249–255. [Google Scholar] [CrossRef]
  45. TuTunchi, H.; Saghafi-Asl, M.; Ostadrahimi, A. A systematic review of the effects of oleoylethanolamide, a high-affinity endogenous ligand of PPAR-α, on the management and prevention of obesity. Clin. Exp. Pharmacol. Physiol. 2020, 47, 543–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. González-Aparicio, R.; Moratalla, R. Oleoylethanolamide reduces L-DOPA-induced dyskinesia via TRPV1 receptor in a mouse model of Parkinson´s disease. Neurobiol. Dis. 2014, 62, 416–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Kasatkina, L.A.; Heinemann, A.; Hudz, Y.A.; Thomas, D.; Sturm, E.M. Stearoylethanolamide interferes with retrograde endocannabinoid signalling and supports the blood-brain barrier integrity under acute systemic inflammation. Biochem. Pharmacol. 2020, 174, 113783. [Google Scholar] [CrossRef]
  48. Maccarrone, M.; Cartoni, A.; Parolaro, D.; Margonelli, A.; Massi, P.; Bari, M.; Battista, N.; Finazzi-Agro, A. Cannabimimetic activity, binding, and degradation of stearoylethanolamide within the mouse central nervous system. Mol. Cell Neurosci. 2002, 21, 126–140. [Google Scholar] [CrossRef] [PubMed]
  49. Artmann, A.; Petersen, G.; Hellgren, L.I.; Boberg, J.; Skonberg, C.; Nellemann, C.; Hansen, S.H.; Hansen, H.S. Influence of dietary fatty acids on endocannabinoid and N-acylethanolamine levels in rat brain, liver and small intestine. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2008, 1781, 200–212. [Google Scholar] [CrossRef] [PubMed]
  50. Movahed, P.; Jönsson, B.A.G.; Birnir, B.; Wingstrand, J.A.; Jørgensen, T.D.; Ermund, A.; Sterner, O.; Zygmunt, P.M.; Högestätt, E.D. Endogenous Unsaturated C18 N-Acylethanolamines Are Vanilloid Receptor (TRPV1) Agonists. J. Biol. Chem. 2005, 280, 38496–38504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Garg, P.; Duncan, R.S.; Kaja, S.; Zabaneh, A.; Chapman, K.D.; Koulen, P. Lauroylethanolamide and linoleoylethanolamide improve functional outcome in a rodent model for stroke. Neurosci. Lett. 2011, 492, 134–138. [Google Scholar] [CrossRef] [Green Version]
  52. Kim, H.-Y.; Spector, A.A.; Xiong, Z.-M. A synaptogenic amide N-docosahexaenoylethanolamide promotes hippocampal development. Prostaglandins Other Lipid Mediat. 2011, 96, 114–120. [Google Scholar] [CrossRef] [Green Version]
  53. Lee, J.-W.; Huang, B.X.; Kwon, H.; Rashid, A.; Kharebava, G.; Desai, A.; Patnaik, S.; Marugan, J.; Kim, H.-Y. Orphan GPR110 (ADGRF1) targeted by N-docosahexaenoylethanolamine in development of neurons and cognitive function. Nat. Commun. 2016, 7, 13123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Park, T.; Chen, H.; Kim, H.-Y. GPR110 (ADGRF1) mediates anti-inflammatory effects of N-docosahexaenoylethanolamine. J. Neuroinflammation 2019, 16, 1–13. [Google Scholar] [CrossRef] [PubMed]
  55. Berman, P.; Sulimani, L.; Gelfand, A.; Amsalem, K.; Lewitus, G.M.; Meiri, D. Cannabinoidomics—An analytical approach to understand the effect of medical Cannabis treatment on the endocannabinoid metabolome. Talanta 2020, 219, 121336. [Google Scholar] [CrossRef] [PubMed]
  56. Panikashvili, D.; Simeonidou, C.; Ben-Shabat, S.; Hanus, L.; Breuer, A.; Mechoulam, R.; Shohami, E. An endogenous canna-binoid (2-AG) is neuroprotective after brain injury. Nature 2001, 413, 527–531. [Google Scholar] [CrossRef] [PubMed]
  57. 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] [PubMed] [Green Version]
  58. Hohmann, A.G.; Suplita, R.L.; Bolton, N.M.; Neely, M.H.; Fegley, D.; Mangieri, R.; Krey, J.F.; Walker, J.M.; Holmes, P.V.; Crystal, J.D.; et al. An endocannabinoid mechanism for stress-induced analgesia. Nat. Cell Biol. 2005, 435, 1108–1112. [Google Scholar] [CrossRef]
  59. Centonze, D.; Bari, M.; Rossi, S.; Prosperetti, C.; Furlan, R.; Fezza, F.; De Chiara, V.; Battistini, L.; Bernardi, G.; Bernardini, S.; et al. The endocannabinoid system is dysregulated in multiple sclerosis and in experimental autoimmune encephalomyelitis. Brain 2007, 130, 2543–2553. [Google Scholar] [CrossRef] [Green Version]
  60. Muthian, S.; Rademacher, D.; Roelke, C.; Gross, G.; Hillard, C. Anandamide content is increased and CB1 cannabinoid receptor blockade is protective during transient, focal cerebral ischemia. Neuroscience 2004, 129, 743–750. [Google Scholar] [CrossRef]
  61. Lawton, S.K.; Xu, F.; Tran, A.; Wong, E.; Prakash, A.; Schumacher, M.; Hellman, J.; Wilhelmsen, K. N-Arachidonoyl Do-pamine Modulates Acute Systemic Inflammation via Nonhematopoietic TRPV1. J. Immunol. 2017, 199, 1465–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Bisogno, T.; Martire, A.; Petrosino, S.; Popoli, P.; Di Marzo, V. Symptom-related changes of endocannabinoid and pal-mitoylethanolamide levels in brain areas of R6/2 mice, a transgenic model of Huntington’s disease. Neurochem. Int. 2008, 52, 307–313. [Google Scholar] [CrossRef] [PubMed]
  63. De Filippis, D.; Luongo, L.; Cipriano, M.; Palazzo, E.; Cinelli, M.P.; de Novellis, V.; Maione, S.; Iuvone, T. Palmitoylethano-lamide reduces granuloma-induced hyperalgesia by modulation of mast cell activation in rats. Mol. Pain 2011, 7, 3. [Google Scholar] [CrossRef] [Green Version]
  64. Costa, B.; Comelli, F.; Bettoni, I.; Colleoni, M.; Giagnoni, G. The endogenous fatty acid amide, palmitoylethanolamide, has anti-allodynic and anti-hyperalgesic effects in a murine model of neuropathic pain: Involvement of CB(1), TRPV1 and PPARgamma receptors and neurotrophic factors. Pain 2008, 139, 541–550. [Google Scholar] [CrossRef] [PubMed]
  65. Ahmad, A.; Crupi, R.; Impellizzeri, D.; Campolo, M.; Marino, A.; Esposito, E.; Cuzzocrea, S. Administration of pal-mitoylethanolamide (PEA) protects the neurovascular unit and reduces secondary injury after traumatic brain injury in mice. Brain Behav. Immun. 2012, 26, 1310–1321. [Google Scholar] [CrossRef] [PubMed]
  66. Zerangue, N.; Kavanaugh, M.P. Flux coupling in a neuronal glutamate transporter. Nat. Cell Biol. 1996, 383, 634–637. [Google Scholar] [CrossRef]
  67. Rossi, D.J.; Oshima, T.; Attwell, D. Glutamate release in severe brain ischaemia is mainly by reversed uptake. Nat. Cell Biol. 2000, 403, 316–321. [Google Scholar] [CrossRef]
  68. Riveros, N.; Fiedler, J.; Lagos, N.; Munoz, C.; Orrego, F. Glutamate in rat brain cortex synaptic vesicles: Influence of the vesicle isolation procedure. Brain Res. 1986, 386, 405–408. [Google Scholar] [CrossRef]
  69. Clements, J.D.; Lester, R.A.; Tong, G.; Jahr, C.E.; Westbrook, G.L. The time course of glutamate in the synaptic cleft. Science 1992, 258, 1498–1501. [Google Scholar] [CrossRef]
  70. Zhou, X.; Hollern, D.; Liao, J.; Andrechek, E.; Wang, H. NMDA receptor-mediated excitotoxicity depends on the coactivation of synaptic and extrasynaptic receptors. Cell Death Dis. 2013, 4, e560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Johnson, J.W.; Kotermanski, S.E. Mechanism of action of memantine. Curr. Opin. Pharmacol. 2006, 6, 61–67. [Google Scholar] [CrossRef]
  72. Xia, P.; Chen, H.-S.V.; Zhang, D.; Lipton, S.A. Memantine Preferentially Blocks Extrasynaptic over Synaptic NMDA Receptor Currents in Hippocampal Autapses. J. Neurosci. 2010, 30, 11246–11250. [Google Scholar] [CrossRef] [Green Version]
  73. Doyle, S.; Hansen, D.B.; Vella, J.; Bond, P.; Harper, G.; Zammit, C.; Valentino, M.; Fern, R. Vesicular glutamate release from central axons contributes to myelin damage. Nat. Commun. 2018, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  74. Brickley, S.G.; Misra, C.; Mok, M.H.S.; Mishina, M.; Cull-Candy, S.G. NR2B and NR2D Subunits Coassemble in Cerebellar Golgi Cells to Form a Distinct NMDA Receptor Subtype Restricted to Extrasynaptic Sites. J. Neurosci. 2003, 23, 4958–4966. [Google Scholar] [CrossRef] [PubMed]
  75. Walter, L.; Franklin, A.; Witting, A.; Moller, T.; Stella, N. Astrocytes in culture produce anandamide and other acylethano-lamides. J. Biol. Chem. 2002, 277, 20869–20876. [Google Scholar] [CrossRef] [Green Version]
  76. 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] [Green Version]
  77. 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] [PubMed] [Green Version]
  78. Stella, N.; Piomelli, D. Receptor-dependent formation of endogenous cannabinoids in cortical neurons. Eur. J. Pharmacol. 2001, 425, 189–196. [Google Scholar] [CrossRef] [Green Version]
  79. Franklin, A.; Parmentier-Batteur, S.; Walter, L.; Greenberg, D.A.; Stella, N. Palmitoylethanolamide increases after focal cerebral ischemia and potentiates microglial cell motility. J. Neurosci. 2003, 23, 7767–7775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Franklin, A.; Stella, N. Arachidonylcyclopropylamide increases microglial cell migration through cannabinoid CB2 and ab-normal-cannabidiol-sensitive receptors. Eur. J. Pharmacol. 2003, 474, 195–198. [Google Scholar] [CrossRef]
  81. Kreutz, S.; Koch, M.; Böttger, C.; Ghadban, C.; Korf, H.-W.; Dehghani, F. 2-Arachidonoylglycerol elicits neuroprotective effects on excitotoxically lesioned dentate gyrus granule cells via abnormal-cannabidiol-sensitive receptors on microglial cells. Glia 2009, 57, 286–294. [Google Scholar] [CrossRef] [PubMed]
  82. Wake, H.; Moorhouse, A.J.; Jinno, S.; Kohsaka, S.; Nabekura, J. Resting Microglia Directly Monitor the Functional State of Synapses In Vivo and Determine the Fate of Ischemic Terminals. J. Neurosci. 2009, 29, 3974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Weinhard, L.; di Bartolomei, G.; Bolasco, G.; Machado, P.; Schieber, N.L.; Neniskyte, U.; Exiga, M.; Vadisiute, A.; Raggioli, A.; Schertel, A.; et al. Microglia remodel synapses by presynaptic trogocytosis and spine head filopodia in-duction. Nat. Commun. 2018, 9, 1228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Joly, E.; Hudrisier, D. What is trogocytosis and what is its purpose? Nat. Immunol. 2003, 4, 815. [Google Scholar] [CrossRef]
  85. Jafari, M.; Schumacher, A.-M.; Snaidero, N.; Gavilanes, E.M.U.; Neziraj, T.; Kocsis-Jutka, V.; Engels, D.; Jürgens, T.; Wagner, I.; Weidinger, J.D.F.; et al. Phagocyte-mediated synapse removal in cortical neuroinflammation is promoted by local calcium accumulation. Nat. Neurosci. 2021, 24, 355–367. [Google Scholar] [CrossRef]
  86. Combes, V.; Guillemin, G.J.; Chan-Ling, T.; Hunt, N.H.; Grau, G.E. The crossroads of neuroinflammation in infectious diseases: Endothelial cells and astrocytes. Trends Parasitol. 2012, 28, 311–319. [Google Scholar] [CrossRef]
  87. Hume, R.; Dingledine, R.; Heinemann, S.F. Identification of a site in glutamate receptor subunits that controls calcium permeability. Science 1991, 253, 1028–1031. [Google Scholar] [CrossRef]
  88. Malenka, R.C.; Kauer, J.A.; Perkel, D.J.; Mauk, M.D.; Kelly, P.T.; Nicoll, R.A.; Waxham, M.N. An essential role for postsynaptic calmodulin and protein kinase activity in long-term potentiation. Nat. Cell Biol. 1989, 340, 554–557. [Google Scholar] [CrossRef] [PubMed]
  89. Malinow, R.; Schulman, H.; Tsien, R.W. Inhibition of postsynaptic PKC or CaMKII blocks induction but not expression of LTP. Science 1989, 245, 862–866. [Google Scholar] [CrossRef]
  90. Murakoshi, H.; Wang, H.; Yasuda, R. Local, persistent activation of Rho GTPases during plasticity of single dendritic spines. Nat. Cell Biol. 2011, 472, 100–104. [Google Scholar] [CrossRef]
  91. Shi, S.H.; Hayashi, Y.; Petralia, R.S.; Zaman, S.H.; Wenthold, R.J.; Svoboda, K.; Malinow, R. Rapid spine delivery and redistribution of AMPA receptors after synaptic NMDA receptor activation. Science 1999, 284, 1811–1816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Park, M.; Penick, E.C.; Edwards, J.G.; Kauer, J.A.; Ehlers, M.D. Recycling Endosomes Supply AMPA Receptors for LTP. Science 2004, 305, 1972–1975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Derkach, V.; Barria, A.; Soderling, T.R. Ca2+/calmodulin-kinase II enhances channel conductance of al-pha-amino-3-hydroxy-5-methyl-4-isoxazolepropionate type glutamate receptors. Proc. Natl. Acad. Sci. USA 1999, 96, 3269–3274. [Google Scholar] [CrossRef] [Green Version]
  94. Fortin, D.A.; Davare, M.A.; Srivastava, T.; Brady, J.D.; Nygaard, S.; Derkach, V.A.; Soderling, T.R. Long-term potentiation-dependent spine enlargement requires synaptic Ca2+-permeable AMPA receptors recruited by CaM-kinase I. J. Neurosci. 2010, 30, 11565–11575. [Google Scholar] [CrossRef] [Green Version]
  95. Mulkey, R.M.; Endo, S.; Shenolikar, S.; Malenka, R.C. Involvement of a calcineurin/ inhibitor-1 phosphatase cascade in hippocampal long-term depression. Nat. Cell Biol. 1994, 369, 486–488. [Google Scholar] [CrossRef] [PubMed]
  96. Lee, H.-K.; Kameyama, K.; Huganir, R.L.; Bear, M.F. NMDA Induces Long-Term Synaptic Depression and Dephosphorylation of the GluR1 Subunit of AMPA Receptors in Hippocampus. Neuron 1998, 21, 1151–1162. [Google Scholar] [CrossRef] [Green Version]
  97. Ehlers, M.D. Reinsertion or Degradation of AMPA Receptors Determined by Activity-Dependent Endocytic Sorting. Neuron 2000, 28, 511–525. [Google Scholar] [CrossRef] [Green Version]
  98. Mauna, J.C.; Miyamae, T.; Pulli, B.; Thiels, E. Protein phosphatases 1 and 2A are both required for long-term depression and associated dephosphorylation of cAMP response element binding protein in hippocampal area CA1 in vivo. Hippocampus 2011, 21, 1093–1104. [Google Scholar] [CrossRef] [Green Version]
  99. Rick, J.T.; Milgram, N.W. Frequency dependence of long-term potentiation and depression in the dentate gyrus of the freely moving rat. Hippocampus 1996, 6, 118–124. [Google Scholar] [CrossRef]
  100. Cui, Y.; Paillé, V.; Xu, H.; Genet, S.; Delord, B.; Fino, E.; Berry, H.; Venance, L. Endocannabinoids mediate bidirectional striatal spike-timing-dependent plasticity. J. Physiol. 2015, 593, 2833–2849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Varma, N.; Carlson, G.C.; Ledent, C.; Alger, B.E. Metabotropic Glutamate Receptors Drive the Endocannabinoid System in Hippocampus. J. Neurosci. 2001, 21, RC188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Kasatkina, L.A.; Gumenyuk, V.P.; Sturm, E.M.; Heinemann, A.; Bernas, T.; Trikash, I.O. Modulation of neurosecretion and approaches for its multistep analysis. Biochim. Biophys. Acta BBA Gen. Subj. 2018, 1862, 2701–2713. [Google Scholar] [CrossRef]
  103. Singla, S.; Kreitzer, A.C.; Malenka, R.C. Mechanisms for Synapse Specificity during Striatal Long-Term Depression. J. Neurosci. 2007, 27, 5260–5264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Sjöström, P.J.; Turrigiano, G.G.; Nelson, S.B. Neocortical LTD via Coincident Activation of Presynaptic NMDA and Can-nabinoid Receptors. Neuron 2003, 39, 641–654. [Google Scholar] [CrossRef] [Green Version]
  105. Straiker, A.; Mackie, K. Depolarization-induced suppression of excitation in murine autaptic hippocampal neurones. J. Physiol. 2005, 569, 501–517. [Google Scholar] [CrossRef]
  106. Colmers, P.L.; Bains, J.S. Presynaptic mGluRs Control the Duration of Endocannabinoid-Mediated DSI. J. Neurosci. 2018, 38, 10444–10453. [Google Scholar] [CrossRef] [Green Version]
  107. Maejima, T.; Hashimoto, K.; Yoshida, T.; Aiba, A.; Kano, M. Presynaptic Inhibition Caused by Retrograde Signal from Metabotropic Glutamate to Cannabinoid Receptors. Neuron 2001, 31, 463–475. [Google Scholar] [CrossRef] [Green Version]
  108. Chávez, A.E.; Chiu, C.Q.; Castillo, P.E. TRPV1 activation by endogenous anandamide triggers postsynaptic long-term depression in dentate gyrus. Nat. Neurosci. 2010, 13, 1511–1518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Grueter, B.A.; Brasnjo, G.; Malenka, R.C. Postsynaptic TRPV1 triggers cell type–specific long-term depression in the nucleus accumbens. Nat. Neurosci. 2010, 13, 1519–1525. [Google Scholar] [CrossRef]
  110. Gibson, H.E.; Edwards, J.G.; Page, R.S.; Van Hook, M.J.; Kauer, J.A. TRPV1 Channels Mediate Long-Term Depression at Synapses on Hippocampal Interneurons. Neuron 2008, 57, 746–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Puente, N.; Cui, Y.; Lassalle, O.; Lafourcade, M.; Georges, F.; Venance, L.; Grandes, P.; Manzoni, O.J. Polymodal activation of the endocannabinoid system in the extended amygdala. Nat. Neurosci. 2011, 14, 1542–1547. [Google Scholar] [CrossRef] [PubMed]
  112. Mechoulam, R.; Fride, E.; Hanu, L.; Sheskin, T.; Bisogno, T.; Di Marzo, V.; Bayewitch, M.; Vogel, Z. Anandamide may mediate sleep induction. Nat. Cell Biol. 1997, 389, 25–26. [Google Scholar] [CrossRef]
  113. Nayak, D.; Roth, T.L.; McGavern, D.B. Microglia Development and Function. Annu. Rev. Immunol. 2014, 32, 367–402. [Google Scholar] [CrossRef] [Green Version]
  114. Ling, E.; Ng, Y.; Wu, C.; Kaur, C. Microglia: Its development and role as a neuropathology sensor. Progress Brain Res. 2001, 132, 61–79. [Google Scholar] [CrossRef]
  115. Colton, C.A.; Gilbert, D.L. Production of superoxide anions by a CNS macrophage, the microglia. FEBS Lett. 1987, 223, 284–288. [Google Scholar] [CrossRef] [Green Version]
  116. Takahashi, K.; Rochford, C.D.; Neumann, H. Clearance of apoptotic neurons without inflammation by microglial triggering receptor expressed on myeloid cells-2. J. Exp. Med. 2005, 201, 647–657. [Google Scholar] [CrossRef] [Green Version]
  117. Levi, G.; Minghetti, L.; Aloisi, F. Regulation of prostanoid synthesis in microglial cells and effects of prostaglandin E2 on microglial functions. Biochimie 1998, 80, 899–904. [Google Scholar] [CrossRef]
  118. Nakamura, Y. Regulating Factors for Microglial Activation. Biol. Pharm. Bull. 2002, 25, 945–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Zhou, H.; Lapointe, B.M.; Clark, S.R.; Zbytnuik, L.; Kubes, P. A Requirement for Microglial TLR4 in Leukocyte Recruitment into Brain in Response to Lipopolysaccharide. J. Immunol. 2006, 177, 8103–8110. [Google Scholar] [CrossRef] [Green Version]
  120. Russo, M.V.; McGavern, D.B. Immune Surveillance of the CNS following Infection and Injury. Trends Immunol. 2015, 36, 637–650. [Google Scholar] [CrossRef] [Green Version]
  121. Davalos, D.; Grutzendler, J.; Yang, G.; Kim, J.V.; Zuo, Y.; Jung, S.; Littman, D.R.; Dustin, M.L.; Gan, W.-B. ATP mediates rapid microglial response to local brain injury in vivo. Nat. Neurosci. 2005, 8, 752–758. [Google Scholar] [CrossRef] [PubMed]
  122. Banati, R.B.; Gehrmann, J.; Schubert, P.; Kreutzberg, G.W. Cytotoxicity of microglia. Glia 1993, 7, 111–118. [Google Scholar] [CrossRef] [PubMed]
  123. Tang, Y.; Le, W. Differential Roles of M1 and M2 Microglia in Neurodegenerative Diseases. Mol. Neurobiol. 2016, 53, 1181–1194. [Google Scholar] [CrossRef]
  124. Block, M.L.; Zecca, L.; Hong, J.-S. Microglia-mediated neurotoxicity: Uncovering the molecular mechanisms. Nat. Rev. Neurosci. 2007, 8, 57–69. [Google Scholar] [CrossRef]
  125. Banati, R.B. Visualising microglial activation in vivo. Glia 2002, 40, 206–217. [Google Scholar] [CrossRef] [PubMed]
  126. Banati, R.B. Neuropathological imaging: In vivo detection of glial activation as a measure of disease and adaptive change in the brain. Br. Med. Bull. 2003, 65, 121–131. [Google Scholar] [CrossRef] [Green Version]
  127. Lehnardt, S.; Massillon, L.; Follett, P.; Jensen, F.E.; Ratan, R.; Rosenberg, P.; Volpe, J.J.; Vartanian, T. Activation of innate immunity in the CNS triggers neurodegeneration through a Toll-like receptor 4-dependent pathway. Proc. Natl. Acad. Sci. USA 2003, 100, 8514–8519. [Google Scholar] [CrossRef] [Green Version]
  128. Heneka, M.T.; Sastre, M.; Dumitrescu-Ozimek, L.; Dewachter, I.; Walter, J.; Klockgether, T.; Van Leuven, F. Focal glial acti-vation coincides with increased BACE1 activation and precedes amyloid plaque deposition in APP[V717I] transgenic mice. J. Neuroinflamm. 2005, 2, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Song, M.; Jin, J.; Lim, J.E.; Kou, J.; Pattanayak, A.; Rehman, J.A.; Kim, H.D.; Tahara, K.; Lalonde, R.; Fukuchi, K. TLR4 mutation reduces microglial activation, increases Abeta deposits and exacerbates cognitive deficits in a mouse model of Alzheimer’s disease. J. Neuroinflamm. 2011, 8, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Caso, J.R.; Pradillo, J.M.; Hurtado, O.; Lorenzo, P.; Moro, M.A.; Lizasoain, I. Toll-Like Receptor 4 Is Involved in Brain Damage and Inflammation After Experimental Stroke. Circulation 2007, 115, 1599–1608. [Google Scholar] [CrossRef]
  131. Maresz, K.; Carrier, E.J.; Ponomarev, E.D.; Hillard, C.J.; Dittel, B.N. Modulation of the cannabinoid CB2 receptor in mi-croglial cells in response to inflammatory stimuli. J. Neurochem. 2005, 95, 437–445. [Google Scholar] [CrossRef]
  132. Hohmann, U.; Pelzer, M.; Kleine, J.; Hohmann, T.; Ghadban, C.; Dehghani, F. Opposite Effects of Neuroprotective Cannabinoids, Palmitoylethanolamide, and 2-Arachidonoylglycerol on Function and Morphology of Microglia. Front. Neurosci. 2019, 13, 1180. [Google Scholar] [CrossRef] [PubMed]
  133. 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]
  134. 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 CB2Receptor-Dependent Mechanism. Mol. Pharmacol. 2004, 65, 999–1007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Araque, A.; Parpura, V.; Sanzgiri, R.P.; Haydon, P.G. Tripartite synapses: Glia, the unacknowledged partner. Trends Neurosci. 1999, 22, 208–215. [Google Scholar] [CrossRef]
  136. Han, J.; Kesner, P.; Metna-Laurent, M.; Duan, T.; Xu, L.; Georges, F.; Koehl, M.; Abrous, 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] [Green Version]
  137. Rossi, D.; Volterra, A. Astrocytic dysfunction: Insights on the role in neurodegeneration. Brain Res. Bull. 2009, 80, 224–232. [Google Scholar] [CrossRef]
  138. Liddelow, S.A.; Barres, B.A. Reactive Astrocytes: Production, Function, and Therapeutic Potential. Immunity 2017, 46, 957–967. [Google Scholar] [CrossRef] [Green Version]
  139. Lian, H.; Yang, L.; Cole, A.; Sun, L.; Chiang, A.C.; Fowler, S.W.; Shim, D.J.; Rodriguez-Rivera, J.; Taglialatela, G.; Jankowsky, J.L.; et al. NFkappaB-activated astroglial release of complement C3 compromises neuronal morphology and function associated with Alzheimer’s disease. Neuron 2015, 85, 101–115. [Google Scholar] [CrossRef] [Green Version]
  140. Phatnani, H.P.; Guarnieri, P.; Friedman, B.A.; Carrasco, M.A.; Muratet, M.; O’Keeffe, S.; Nwakeze, C.; Behn, F.P.; Newberry, K.M.; Meadows, S.K.; et al. Intricate interplay between astrocytes and motor neurons in ALS. Proc. Natl. Acad. Sci. USA 2013, 110, E756–E765. [Google Scholar] [CrossRef] [Green Version]
  141. Aebischer, J.; Cassina, P.; Otsmane, B.; Moumen, A.; Seilhean, D.; Meininger, V.; Barbeito, L.; Pettmann, B.; Raoul, C. IFNgamma triggers a LIGHT-dependent selective death of motoneurons contributing to the non-cell-autonomous effects of mutant SOD1. Cell Death Differ. 2011, 18, 754–768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.-S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef]
  143. Okada, S.; Nakamura, M.; Katoh, H.; Miyao, T.; Shimazaki, T.; Ishii, K.; Yamane, J.; Yoshimura, A.; Iwamoto, Y.; Toyama, Y.; et al. Conditional ablation of Stat3 or Socs3 discloses a dual role for reactive astrocytes after spinal cord injury. Nat. Med. 2006, 12, 829–834. [Google Scholar] [CrossRef]
  144. Christopherson, K.S.; Ullian, E.M.; Stokes, C.C.A.; Mullowney, C.E.; Hell, J.W.; Agah, A.; Lawler, J.; Mosher, D.F.; Bornstein, P.; Barres, B.A. Thrombospondins Are Astrocyte-Secreted Proteins that Promote CNS Synaptogenesis. Cell 2005, 120, 421–433. [Google Scholar] [CrossRef] [Green Version]
  145. Wang, L.; Lin, F.; Wang, J.; Wu, J.; Han, R.; Zhu, L.; Zhang, G.; DiFiglia, M.; Qin, Z. Truncated N-terminal huntingtin fragment with expanded-polyglutamine (htt552-100Q) suppresses brain-derived neurotrophic factor transcription in astrocytes. Acta Biochim. Biophys. Sin. 2012, 44, 249–258. [Google Scholar] [CrossRef] [Green Version]
  146. Arregui Jorge, L.; Benítez, J.A.; Razgado Paula, L.F.; Vergara, P.; Segovia, J. Adenoviral Astrocyte-Specific Expression of BDNF in the Striata of Mice Transgenic for Huntington’s Disease Delays the Onset of the Motor Phenotype. Cell. Mol. Neurobiol. 2011, 31, 1229–1243. [Google Scholar] [CrossRef] [PubMed]
  147. Walter, L.; Dinh, T.; Stella, N. ATP induces a rapid and pronounced increase in 2-arachidonoylglycerol production by astro-cytes, a response limited by monoacylglycerol lipase. J. Neurosci. 2004, 24, 8068–8074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Walter, L.; Stella, N. Endothelin-1 increases 2-arachidonoyl glycerol (2-AG) production in astrocytes. Glia 2003, 44, 85–90. [Google Scholar] [CrossRef]
  149. Hegyi, Z.; Olah, T.; Koszeghy, A.; Piscitelli, F.; Hollo, K.; Pal, B.; Csernoch, L.; Di Marzo, V.; Antal, M. CB1 receptor activation induces intracellular Ca(2+) mobilization and 2-arachidonoylglycerol release in rodent spinal cord astrocytes. Sci. Rep. 2018, 8, 10562. [Google Scholar] [CrossRef] [PubMed]
  150. Del Pulgar, T.G.; de Ceballos, M.L.; Guzmán, M.; Velasco, G. Cannabinoids Protect Astrocytes from Ceramide-induced Apoptosis through the Phosphatidylinositol 3-Kinase/Protein Kinase B Pathway. J. Biol. Chem. 2002, 277, 36527–36533. [Google Scholar] [CrossRef] [Green Version]
  151. Carracedo, A.; Geelen, M.J.H.; Diez, M.; Hanada, K.; Guzman, M.; Velasco, G. Ceramide sensitizes astrocytes to oxidative stress: Protective role of cannabinoids. Biochem. J. 2004, 380, 435–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Molina-Holgado, F.; Molina-Holgado, E.; Guaza, C.; Rothwell, N.J. Role of CB1 and CB2 receptors in the inhibitory effects of cannabinoids on lipopolysaccharide-induced nitric oxide release in astrocyte cultures. J. Neurosci. Res. 2002, 67, 829–836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Wang, S.; Zhang, H.; Geng, B.; Xie, Q.; Li, W.; Deng, Y.; Shi, W.; Pan, Y.; Kang, X.; Wang, J. 2-arachidonyl glycerol modulates astrocytic glutamine synthetase via p38 and ERK1/2 pathways. J. Neuroinflamm. 2018, 15, 220. [Google Scholar] [CrossRef]
  154. Feliu, A.; Mestre, L.; Carrillo-Salinas, F.J.; Yong, V.W.; Mecha, M.; Guaza, C. 2-arachidonoylglycerol reduces chondroitin sulphate proteoglycan production by astrocytes and enhances oligodendrocyte differentiation under inhibitory conditions. Glia 2020, 68, 1255–1273. [Google Scholar] [CrossRef] [PubMed]
  155. Wang, F.; Li, M.; Li, X.; Kinden, R.; Zhou, H.; Guo, F.; Wang, Q.; Xiong, L. 2-Arachidonylglycerol Protects Primary Astrocytes Exposed to Oxygen-Glucose Deprivation Through a Blockade of NDRG2 Signaling and STAT3 Phosphorylation. Rejuvenation Res. 2016, 19, 215–222. [Google Scholar] [CrossRef]
  156. Scuderi, C.; Esposito, G.; Blasio, A.; Valenza, M.; Arietti, P.; Steardo, L., Jr.; Carnuccio, R.; De Filippis, D.; Petrosino, S.; Iuvone, T.; et al. Palmitoylethanolamide counteracts reactive astrogliosis induced by beta-amyloid peptide. J. Cell Mol. Med. 2011, 15, 2664–2674. [Google Scholar] [CrossRef] [Green Version]
  157. Gajardo-Gomez, R.; Labra, V.C.; Maturana, C.J.; Shoji, K.F.; Santibanez, C.A.; Saez, J.C.; Giaume, C.; Orellana, J.A. Cannabinoids prevent the amyloid beta-induced activation of astroglial hemichannels: A neuroprotective mechanism. Glia 2017, 65, 122–137. [Google Scholar] [CrossRef]
  158. Beggiato, S.; Borelli, A.C.; Ferraro, L.; Tanganelli, S.; Antonelli, T.; Tomasini, M.C. Palmitoylethanolamide Blunts Amy-loid-beta42-Induced Astrocyte Activation and Improves Neuronal Survival in Primary Mouse Cortical Astrocyte-Neuron Co-Cultures. J. Alzheimers Dis. 2018, 61, 389–399. [Google Scholar] [CrossRef]
  159. Beggiato, S.; Cassano, T.; Ferraro, L.; Tomasini, M.C. Astrocytic palmitoylethanolamide pre-exposure exerts neuroprotective effects in astrocyte-neuron co-cultures from a triple transgenic mouse model of Alzheimer’s disease. Life Sci. 2020, 257, 118037. [Google Scholar] [CrossRef]
  160. Bronzuoli, M.R.; Facchinetti, R.; Steardo, L.; Romano, A.; Stecca, C.; Passarella, S.; Cassano, T.; Scuderi, C. Palmitoylethanolamide Dampens Reactive Astrogliosis and Improves Neuronal Trophic Support in a Triple Transgenic Model of Alzheimer’s Disease: In Vitro and In Vivo Evidence. Oxid. Med. Cell. Longev. 2018, 2018, 4720532. [Google Scholar] [CrossRef] [Green Version]
  161. Luo, D.; Zhang, Y.; Yuan, X.; Pan, Y.; Yang, L.; Zhao, Y.; Zhuo, R.; Chen, C.; Peng, L.; Li, W.; et al. Oleoylethan-olamide inhibits glial activation via moudulating PPARalpha and promotes motor function recovery after brain ischemia. Pharmacol. Res. 2019, 141, 530–540. [Google Scholar] [CrossRef]
  162. Scuderi, C.; Valenza, M.; Stecca, C.; Esposito, G.; Carratù, M.R.; Steardo, L. Palmitoylethanolamide exerts neuroprotective effects in mixed neuroglial cultures and organotypic hippocampal slices via peroxisome proliferator-activated receptor-α. J. Neuroinflamm. 2012, 9, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Zhang, L.Y.; Zhou, Y.Q.; Yu, Z.P.; Zhang, X.Q.; Shi, J.; Shen, H.W. Restoring glutamate homeostasis in the nucleus accumbens via endocannabinoid-mimetic drug prevents relapse to cocaine seeking behavior in rats. Neuropsychopharmacology 2021, 46, 970–981. [Google Scholar] [CrossRef] [PubMed]
  164. Moreno-García, Álvaro; Bernal-Chico, A.; Colomer, T.; Rodríguez-Antigüedad, A.; Matute, C.; Mato, S. Gene Expression Analysis of Astrocyte and Microglia Endocannabinoid Signaling during Autoimmune Demyelination. Biomolecules 2020, 10, 1228. [Google Scholar] [CrossRef] [PubMed]
  165. Viader, A.; Blankman, J.L.; Zhong, P.; Liu, X.; Schlosburg, J.E.; Joslyn, C.M.; Liu, Q.-S.; Tomarchio, A.J.; Lichtman, A.H.; Selley, D.E.; et al. Metabolic Interplay between Astrocytes and Neurons Regulates Endocannabinoid Action. Cell Rep. 2015, 12, 798–808. [Google Scholar] [CrossRef] [Green Version]
  166. Chen, X.; Hu, Y.; Cao, Z.; Liu, Q.; Cheng, Y. Cerebrospinal Fluid Inflammatory Cytokine Aberrations in Alzheimer’s Disease, Parkinson’s Disease and Amyotrophic Lateral Sclerosis: A Systematic Review and Meta-Analysis. Front. Immunol. 2018, 9, 2122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Almutairi, M.M.A.; Gong, C.; Xu, Y.G.; Chang, Y.; Shi, H. Factors controlling permeability of the blood–brain barrier. Cell. Mol. Life Sci. 2016, 73, 57–77. [Google Scholar] [CrossRef]
  168. Rothwell, N.J.; Luheshi, G.N. Interleukin 1 in the brain: Biology, pathology and therapeutic target. Trends Neurosci. 2000, 23, 618–625. [Google Scholar] [CrossRef]
  169. Garcia, E.; Aguilar-Cevallos, J.; Silva-Garcia, R.; Ibarra, A. Cytokine and Growth Factor Activation In Vivo and In Vitro after Spinal Cord Injury. Mediat. Inflamm. 2016, 2016, 1–21. [Google Scholar] [CrossRef] [Green Version]
  170. Barnes, C.A. Long-term potentiation and the ageing brain. Philos. Trans. R. Soc. B Biol. Sci. 2003, 358, 765–772. [Google Scholar] [CrossRef] [Green Version]
  171. Zhu, H.; Wang, Z.; Yu, J.; Yang, X.; He, F.; Liu, Z.; Che, F.; Chen, X.; Ren, H.; Hong, M.; et al. Role and mechanisms of cytokines in the secondary brain injury after intracerebral hemorrhage. Prog. Neurobiol. 2019, 178, 101610. [Google Scholar] [CrossRef]
  172. Li, B.; Chen, M.; Guo, L.; Yun, Y.; Li, G.; Sang, N. Endocannabinoid 2-arachidonoylglycerol protects inflammatory insults from sulfur dioxide inhalation via cannabinoid receptors in the brain. J. Environ. Sci. 2017, 51, 265–274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Grabner, G.F.; Eichmann, T.O.; Wagner, B.; Gao, Y.; Farzi, A.; Taschler, U.; Radner, F.P.; Schweiger, M.; Lass, A.; Holzer, P.; et al. Deletion of Monoglyceride Lipase in Astrocytes Attenuates Lipopolysaccharide-induced Neuroinflammation. J. Biol. Chem. 2016, 291, 913–923. [Google Scholar] [CrossRef] [Green Version]
  174. Wen, J.; Jones, M.; Tanaka, M.; Selvaraj, P.; Symes, A.J.; Cox, B.; Zhang, Y. WWL70 protects against chronic constriction injury-induced neuropathic pain in mice by cannabinoid receptor-independent mechanisms. J. Neuroinflamm. 2018, 15, 9. [Google Scholar] [CrossRef] [PubMed]
  175. Piro, J.R.; Suidan, G.L.; Quan, J.; Pi, Y.; O’Neill, S.M.; Ilardi, M.; Pozdnyakov, N.; Lanz, T.A.; Xi, H.; Bell, R.D.; et al. Inhibition of 2-AG hydrolysis differentially regulates blood brain barrier permeability after injury. J. Neuroinflamm. 2018, 15, 142. [Google Scholar] [CrossRef]
  176. 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 hippo-campus of aged rats and restores an age-related deficit in long-term potentiation. J. Neuroinflamm. 2012, 9, 79. [Google Scholar] [CrossRef] [Green Version]
  177. Rivera, P.; Fernández-Arjona, M.D.M.; Silva-Peña, D.; Blanco, E.; Vargas, A.; López-Ávalos, M.D.; Grondona, J.M.; Serrano, A.; Pavón, F.J.; De Fonseca, F.R.; et al. Pharmacological blockade of fatty acid amide hydrolase (FAAH) by URB597 improves memory and changes the phenotype of hippocampal microglia despite ethanol exposure. Biochem. Pharmacol. 2018, 157, 244–257. [Google Scholar] [CrossRef] [PubMed]
  178. Henry, R.J.; Kerr, D.M.; Flannery, L.E.; Killilea, M.; Hughes, E.M.; Corcoran, L.; Finn, D.P.; Roche, M. Pharmacological inhibition of FAAH modulates TLR-induced neuroinflammation, but not sickness behaviour: An effect partially mediated by central TRPV1. Brain Behav. Immun. 2017, 62, 318–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Paterniti, I.; Impellizzeri, D.; Crupi, R.; Morabito, R.; Campolo, M.; Esposito, E.; Cuzzocrea, S. Molecular evidence for the involvement of PPAR-delta and PPAR-gamma in anti-inflammatory and neuroprotective activities of palmitoylethanolamide after spinal cord trauma. J. Neuroinflamm. 2013, 10, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Impellizzeri, D.; Cordaro, M.; Bruschetta, G.; Siracusa, R.; Crupi, R.; Esposito, E.; Cuzzocrea, S. N-Palmitoylethanolamine-Oxazoline as a New Therapeutic Strategy to Control Neuroinflammation: Neuroprotective Effects in Experimental Models of Spinal Cord and Brain Injury. J. Neurotrauma 2017, 34, 2609–2623. [Google Scholar] [CrossRef]
  181. Cordaro, M.; Impellizzeri, D.; Paterniti, I.; Bruschetta, G.; Siracusa, R.; De Stefano, D.; Cuzzocrea, S.; Esposito, E. Neuropro-tective Effects of Co-UltraPEALut on Secondary Inflammatory Process and Autophagy Involved in Traumatic Brain Injury. J. Neurotrauma 2016, 33, 132–146. [Google Scholar] [CrossRef]
  182. Gugliandolo, E.; D’Amico, R.; Cordaro, M.; Fusco, R.; Siracusa, R.; Crupi, R.; Impellizzeri, D.; Cuzzocrea, S.; Di Paola, R. Effect of PEA-OXA on neuropathic pain and functional recovery after sciatic nerve crush. J. Neuroinflamm. 2018, 15, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Siracusa, R.; Paterniti, I.; Impellizzeri, D.; Cordaro, M.; Crupi, R.; Navarra, M.; Cuzzocrea, S.; Esposito, E. The Association of Palmitoylethanolamide with Luteolin Decreases Neuroinflammation and Stimulates Autophagy in Parkinson’s Disease Model. CNS Neurol. Disord. Drug Targets 2015, 14, 1350–1366. [Google Scholar] [CrossRef]
  184. Crupi, R.; Impellizzeri, D.; Cordaro, M.; Siracusa, R.; Casili, G.; Evangelista, M.; Cuzzocrea, S. N-palmitoylethanolamide Prevents Parkinsonian Phenotypes in Aged Mice. Mol. Neurobiol. 2018, 55, 8455–8472. [Google Scholar] [CrossRef] [PubMed]
  185. Orefice, N.S.; Alhouayek, M.; Carotenuto, A.; Montella, S.; Barbato, F.; Comelli, A.; Calignano, A.; Muccioli, G.G.; Orefice, G. Oral Palmitoylethanolamide Treatment Is Associated with Reduced Cutaneous Adverse Effects of Interferon-beta1a and Circulating Proinflammatory Cytokines in Relapsing-Remitting Multiple Sclerosis. Neurotherapeutics 2016, 13, 428–438. [Google Scholar] [CrossRef]
  186. Sayd, A.; Antón, M.; Alén, F.; Caso, J.R.; Pavón, J.; Leza, J.C.; De Fonseca, F.R.; García-Bueno, B.; Orio, L. Systemic Administration of Oleoylethanolamide Protects from Neuroinflammation and Anhedonia Induced by LPS in Rats. Int. J. Neuropsychopharmacol. 2015, 18, pyu111. [Google Scholar] [CrossRef] [Green Version]
  187. Sahu, P.; Mudgal, J.; Arora, D.; Kinra, M.; Mallik, S.B.; Rao, C.M.; Pai, K.S.R.; Nampoothiri, M. Cannabinoid receptor 2 activation mitigates lipopolysaccharide-induced neuroinflammation and sickness behavior in mice. Psychopharmacology 2019, 236, 1829–1838. [Google Scholar] [CrossRef]
  188. Li, L.; Yun, D.; Zhang, Y.; Tao, Y.; Tan, Q.; Qiao, F.; Luo, B.; Liu, Y.; Fan, R.; Xian, J.; et al. A cannabinoid receptor 2 agonist reduces blood–brain barrier damage via induction of MKP-1 after intracerebral hemorrhage in rats. Brain Res. 2018, 1697, 113–123. [Google Scholar] [CrossRef]
  189. Sun, L.; Dong, R.; Xu, X.; Yang, X.; Peng, M. Activation of cannabinoid receptor type 2 attenuates surgery-induced cognitive impairment in mice through anti-inflammatory activity. J. Neuroinflamm. 2017, 14, 138. [Google Scholar] [CrossRef]
  190. Javed, H.; Azimullah, S.; Haque, M.E.; Ojha, S.K. Cannabinoid Type 2 (CB2) Receptors Activation Protects against Oxidative Stress and Neuroinflammation Associated Dopaminergic Neurodegeneration in Rotenone Model of Parkinson’s Disease. Front. Neurosci. 2016, 10, 321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Roche, M.; Diamond, M.; Kelly, J.P.; Finn, D.P. In vivo modulation of LPS-induced alterations in brain and peripheral cyto-kines and HPA axis activity by cannabinoids. J. Neuroimmunol. 2006, 181, 57–67. [Google Scholar] [CrossRef] [PubMed]
  192. Lou, Z.-Y.; Zhao, C.-B.; Xiao, B.-G. Immunoregulation of experimental autoimmune encephalomyelitis by the selective CB1 receptor antagonist. J. Neurosci. Res. 2012, 90, 84–95. [Google Scholar] [CrossRef]
  193. Chen, Y.; Luo, X.; Liu, S.; Shen, Y. Neuroprotective effect of cannabinoid receptor 1 antagonist in the MNU-induced retinal degeneration model. Exp. Eye Res. 2018, 167, 145–151. [Google Scholar] [CrossRef]
  194. Reichenbach, Z.W.; Li, H.; Ward, S.J.; Tuma, R.F. The CB1 antagonist, SR141716A, is protective in permanent photo-thrombotic cerebral ischemia. Neurosci. Lett. 2016, 630, 9–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Holmes, C.; Cunningham, C.; Zotova, E.; Woolford, J.; Dean, C.; Kerr, S.; Culliford, D.; Perry, V.H. Systemic inflammation and disease progression in Alzheimer disease. Neurology 2009, 73, 768–774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Rodriguez-Oroz, M.C.; Jahanshahi, M.; Krack, P.; Litvan, I.; Macias, R.; Bezard, E.; Obeso, J.A. Initial clinical manifestations of Parkinson’s disease: Features and pathophysiological mechanisms. Lancet Neurol. 2009, 8, 1128–1139. [Google Scholar] [CrossRef] [Green Version]
  197. Raymond, L.A.; Andre, V.M.; Cepeda, C.; Gladding, C.M.; Milnerwood, A.J.; Levine, M.S. Pathophysiology of Huntington’s disease: Time-dependent alterations in synaptic and receptor function. Neuroscience 2011, 198, 252–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Crotti, A.; Glass, C.K. The choreography of neuroinflammation in Huntington’s disease. Trends Immunol. 2015, 36, 364–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Robberecht, W.; Philips, T. The changing scene of amyotrophic lateral sclerosis. Nat. Rev. Neurosci. 2013, 14, 248–264. [Google Scholar] [CrossRef] [PubMed]
  200. Philips, T.; Robberecht, W. Neuroinflammation in amyotrophic lateral sclerosis: Role of glial activation in motor neuron disease. Lancet Neurol. 2011, 10, 253–263. [Google Scholar] [CrossRef]
  201. Ransohoff, R.M. How neuroinflammation contributes to neurodegeneration. Science 2016, 353, 777–783. [Google Scholar] [CrossRef]
  202. Wang, Q.; Peng, Y.; Chen, S.; Gou, X.; Hu, B.; Du, J.; Lu, Y.; Xiong, L. Pretreatment with Electroacupuncture Induces Rapid Tolerance to Focal Cerebral Ischemia Through Regulation of Endocannabinoid System. Stroke 2009, 40, 2157–2164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Martinez-Vargas, M.; Morales-Gomez, J.; Gonzalez-Rivera, R.; Hernandez-Enriquez, C.; Perez-Arredondo, A.; Estrada-Rojo, F.; Navarro, L. Does the Neuroprotective Role of Anandamide Display Diurnal Variations? Int. J. Mol. Sci. 2013, 14, 23341–23355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Boccella, S.; Marabese, I.; Iannotta, M.; Belardo, C.; Neugebauer, V.; Mazzitelli, M.; Pieretti, G.; Maione, S.; Palazzo, E. Metabotropic Glutamate Receptor 5 and 8 Modulate the Ameliorative Effect of Ultramicronized Palmitoylethanolamide on Cognitive Decline Associated with Neuropathic Pain. Int. J. Mol. Sci. 2019, 20, 1757. [Google Scholar] [CrossRef] [Green Version]
  205. Scuderi, C.; Bronzuoli, M.R.; Facchinetti, R.; Pace, L.; Ferraro, L.; Broad, K.D.; Serviddio, G.; Bellanti, F.; Palombelli, G.; Carpinelli, G.; et al. Ultramicronized palmitoylethanolamide rescues learning and memory impairments in a triple transgenic mouse model of Alzheimer’s disease by exerting anti-inflammatory and neuroprotective effects. Transl. Psychiatry 2018, 8, 32. [Google Scholar] [CrossRef]
  206. Fusco, R.; Gugliandolo, E.; Campolo, M.; Evangelista, M.; Di Paola, R.; Cuzzocrea, S. Effect of a new formulation of micronized and ultramicronized N-palmitoylethanolamine in a tibia fracture mouse model of complex regional pain syndrome. PLoS ONE 2017, 12, e0178553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Schweiger, V.; Martini, A.; Bellamoli, P.; Donadello, K.; Schievano, C.; Balzo, G.D.; Sarzi-Puttini, P.; Parolini, M.; Polati, E. Ultramicronized Palmitoylethanolamide (um-PEA) as Add-on Treatment in Fibromyalgia Syndrome (FMS): Retrospective Observational Study on 407 Patients. CNS Neurol. Disord. Drug Targets 2019, 18, 326–333. [Google Scholar] [CrossRef]
  208. Caltagirone, C.; Cisari, C.; Schievano, C.; Di Paola, R.; Cordaro, M.; Bruschetta, G.; Esposito, E.; Cuzzocrea, S.; Stroke Study Group. Co-ultramicronized Palmitoylethanolamide/Luteolin in the Treatment of Cerebral Ischemia: From Rodent to Man. Transl. Stroke Res. 2016, 7, 54–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Aso, E.; Ferrer, I. CB2 Cannabinoid Receptor as Potential Target against Alzheimer’s Disease. Front. Neurosci. 2016, 10, 243. [Google Scholar] [CrossRef] [Green Version]
  210. Han, Q.W.; Yuan, Y.H.; Chen, N.H. The therapeutic role of cannabinoid receptors and its agonists or antagonists in Parkinson’s disease. Prog. Neuropsychopharmacol. Biol. Psychiatry 2020, 96, 109745. [Google Scholar] [CrossRef]
  211. González, S.; Scorticati, C.; García-Arencibia, M.; de Miguel, R.; Ramos, J.A.; Fernández-Ruiz, J. Effects of rimonabant, a selective cannabinoid CB1 receptor antagonist, in a rat model of Parkinson’s disease. Brain Res. 2006, 1073–1074, 209–219. [Google Scholar] [CrossRef]
  212. Di Marzo, V.; Hill, M.P.; Bisogno, T.; Crossman, A.R.; Brotchie, J.M. Enhanced levels of endogenous cannabinoids in the globus pallidus are associated with a reduction in movement in an animal model of Parkinson’s disease. FASEB J. 2000, 14, 1432–1438. [Google Scholar] [CrossRef]
  213. Mesnage, V.; Houeto, J.L.; Bonnet, A.M.; Clavier, I.; Arnulf, I.; Cattelin, F.; Le Fur, G.; Damier, P.; Welter, M.L.; Agid, Y. Neurokinin B, Neurotensin, and Cannabinoid Receptor Antagonists and Parkinson Disease. Clin. Neuropharmacol. 2004, 27, 108–110. [Google Scholar] [CrossRef] [PubMed]
  214. Benito, C.; Nunez, E.; Tolon, R.M.; Carrier, E.J.; Rabano, A.; Hillard, C.J.; Romero, J. Cannabinoid CB2 receptors and fatty acid amide hydrolase are selectively overexpressed in neuritic plaque-associated glia in Alzheimer’s disease brains. J. Neurosci. 2003, 23, 11136–11141. [Google Scholar] [CrossRef] [Green Version]
  215. Núñez, E.; Benito, C.; Pazos, M.R.; Barbachano, A.; Fajardo, O.; González, S.; Tolón, R.M.; Romero, J. Cannabinoid CB2receptors are expressed by perivascular microglial cells in the human brain: An immunohistochemical study. Synapse 2004, 53, 208–213. [Google Scholar] [CrossRef]
  216. Ren, S.-Y.; Wang, Z.-Z.; Zhang, Y.; Chen, N.-H. Potential application of endocannabinoid system agents in neuropsychiatric and neurodegenerative diseases—focusing on FAAH/MAGL inhibitors. Acta Pharmacol. Sin. 2020, 41, 1263–1271. [Google Scholar] [CrossRef]
  217. Deng, H.; Li, W. Monoacylglycerol lipase inhibitors: Modulators for lipid metabolism in cancer malignancy, neurological and metabolic disorders. Acta Pharm. Sin. B 2020, 10, 582–602. [Google Scholar] [CrossRef] [PubMed]
  218. Montanari, S.; Scalvini, L.; Bartolini, M.; Belluti, F.; Gobbi, S.; Andrisano, V.; Ligresti, A.; Di Marzo, V.; Rivara, S.; Mor, M.; et al. Fatty Acid Amide Hydrolase (FAAH), Acetylcholinesterase (AChE), and Butyrylcholinesterase (BuChE): Networked Targets for the Development of Carbamates as Potential Anti-Alzheimer’s Disease Agents. J. Med. Chem. 2016, 59, 6387–6406. [Google Scholar] [CrossRef]
  219. Bottemanne, P.; Muccioli, G.G.; Alhouayek, M. N-acylethanolamine hydrolyzing acid amidase inhibition: Tools and potential therapeutic opportunities. Drug Discov. Today 2018, 23, 1520–1529. [Google Scholar] [CrossRef] [PubMed]
  220. Piomelli, D.; Scalvini, L.; Fotio, Y.; Lodola, A.; Spadoni, G.; Tarzia, G.; Mor, M. N-Acylethanolamine Acid Amidase (NAAA): Structure, Function, and Inhibition. J. Med. Chem. 2020, 63, 7475–7490. [Google Scholar] [CrossRef] [PubMed]
  221. Malamas, M.S.; Farah, S.I.; Lamani, M.; Pelekoudas, D.N.; Perry, N.T.; Rajarshi, G.; Miyabe, C.Y.; Chandrashekhar, H.; West, J.; Pavlopoulos, S.; et al. Design and synthesis of cyanamides as potent and selective N-acylethanolamine acid amidase inhibitors. Bioorg. Med. Chem. 2020, 28, 115195. [Google Scholar] [CrossRef] [PubMed]
  222. Kerbrat, A.; Ferré, J.-C.; Fillatre, P.; Ronzière, T.; Vannier, S.; Carsin-Nicol, B.; Lavoué, S.; Vérin, M.; Gauvrit, J.-Y.; Le Tulzo, Y.; et al. Acute Neurologic Disorder from an Inhibitor of Fatty Acid Amide Hydrolase. N. Engl. J. Med. 2016, 375, 1717–1725. [Google Scholar] [CrossRef] [Green Version]
  223. Pasquarelli, N.; Porazik, C.; Bayer, H.; Buck, E.; Schildknecht, S.; Weydt, P.; Witting, A.; Ferger, B. Contrasting effects of selective MAGL and FAAH inhibition on dopamine depletion and GDNF expression in a chronic MPTP mouse model of Parkinson’s disease. Neurochem. Int. 2017, 110, 14–24. [Google Scholar] [CrossRef]
  224. Tchantchou, F.; Tucker, L.B.; Fu, A.H.; Bluett, R.J.; McCabe, J.T.; Patel, S.; Zhang, Y. The fatty acid amide hydrolase inhibitor PF-3845 promotes neuronal survival, attenuates inflammation and improves functional recovery in mice with traumatic brain injury. Neuropharmacology 2014, 85, 427–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Johnston, T.H.; Huot, P.; Fox, S.H.; Wakefield, J.D.; Sykes, K.A.; Bartolini, W.P.; Milne, G.T.; Pearson, J.P.; Brotchie, J.M. Fatty acid amide hydrolase (FAAH) inhibition reduces L-3,4-dihydroxyphenylalanine-induced hyperactivity in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-lesioned non-human primate model of Parkinson’s disease. J. Pharmacol. Exp. Ther. 2011, 336, 423–430. [Google Scholar] [CrossRef] [Green Version]
  226. Viveros-Paredes, J.; Gonzalez-Castañeda, R.; Escalante-Castañeda, A.; Tejeda-Martínez, A.; Castañeda-Achutiguí, F.; Flores-Soto, M. Effect of inhibition of fatty acid amide hydrolase on MPTP-induced dopaminergic neuronal damage. Neurología 2019, 34, 143–152. [Google Scholar] [CrossRef] [PubMed]
  227. Celorrio, M.; Fernández-Suárez, D.; Rojo-Bustamante, E.; Echeverry-Alzate, V.; Ramírez, M.J.; Hillard, C.J.; López-Moreno, J.A.; Maldonado, R.; Oyarzábal, J.; Franco, R.; et al. Fatty acid amide hydrolase inhibition for the symptomatic relief of Parkinson’s disease. Brain Behav. Immun. 2016, 57, 94–105. [Google Scholar] [CrossRef] [Green Version]
  228. Pryce, G.; Cabranes, A.; Fernández-Ruiz, J.; Bisogno, T.; Di Marzo, V.; Long, J.Z.; Cravatt, B.F.; Giovannoni, G.; Baker, D. Control of experimental spasticity by targeting the degradation of endocannabinoids using selective fatty acid amide hydrolase inhibitors. Mult. Scler. J. 2013, 19, 1896–1904. [Google Scholar] [CrossRef]
  229. Kerr, D.M.; Harhen, B.; Okine, B.N.; Egan, L.J.; Finn, D.P.; Roche, M. The monoacylglycerol lipase inhibitor JZL184 atten-uates LPS-induced increases in cytokine expression in the rat frontal cortex and plasma: Differential mechanisms of action. Br. J. Pharmacol. 2013, 169, 808–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Zhang, J.; Chen, C. Alleviation of Neuropathology by Inhibition of Monoacylglycerol Lipase in APP Transgenic Mice Lacking CB2 Receptors. Mol. Neurobiol. 2017, 55, 4802–4810. [Google Scholar] [CrossRef]
  231. Pihlaja, R.; Takkinen, J.; Eskola, O.; Vasara, J.; López-Picón, F.R.; Haaparanta-Solin, M.; Rinne, J.O. Monoacylglycerol lipase inhibitor JZL184 reduces neuroinflammatory response in APdE9 mice and in adult mouse glial cells. J. Neuroinflamm. 2015, 12, 81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Mayeux, J.; Katz, P.; Edwards, S.; Middleton, J.W.; Molina, P.E. Inhibition of Endocannabinoid Degradation Improves Outcomes from Mild Traumatic Brain Injury: A Mechanistic Role for Synaptic Hyperexcitability. J. Neurotrauma 2017, 34, 436–443. [Google Scholar] [CrossRef] [PubMed]
  233. Darmani, N.A.; Izzo, A.; Degenhardt, B.; Valenti, M.; Scaglione, G.; Capasso, R.; Sorrentini, I.; Di Marzo, V. Involvement of the cannabimimetic compound, N-palmitoyl-ethanolamine, in inflammatory and neuropathic conditions: Review of the available pre-clinical data, and first human studies. Neuropharmacology 2005, 48, 1154–1163. [Google Scholar] [CrossRef]
  234. Cisar, J.S.; Weber, O.D.; Clapper, J.R.; Blankman, J.L.; Henry, C.L.; Simon, G.M.; Alexander, J.P.; Jones, T.K.; Ezekowitz, R.A.B.; O’Neill, G.P.; et al. Identification of ABX-1431, a Selective Inhibitor of Monoacylglycerol Lipase and Clinical Candidate for Treatment of Neurological Disorders. J. Med. Chem. 2018, 61, 9062–9084. [Google Scholar] [CrossRef] [Green Version]
  235. Fucich, E.A.; Stielper, Z.F.; Cancienne, H.L.; Edwards, S.; Gilpin, N.W.; Molina, P.E.; Middleton, J.W. Endocannabinoid degradation inhibitors ameliorate neuronal and synaptic alterations following traumatic brain injury. J. Neurophysiol. 2020, 123, 707–717. [Google Scholar] [CrossRef] [PubMed]
  236. Toma, W.; Caillaud, M.; Patel, N.H.; Tran, T.H.; Donvito, G.; Roberts, J.; Bagdas, D.; Jackson, A.; Lichtman, A.; Gewirtz, D.A.; et al. N-acylethanolamine-hydrolysing acid amidase: A new potential target to treat paclitaxel-induced neuropathy. Eur. J. Pain 2021. [Google Scholar] [CrossRef]
  237. Migliore, M.; Pontis, S.; Fuentes de Arriba, A.L.; Realini, N.; Torrente, E.; Armirotti, A.; Romeo, E.; Di Martino, S.; Russo, D.; Pizzirani, D.; et al. Second-Generation Non-Covalent NAAA Inhibitors are Protective in a Model of Multiple Sclerosis. Angew. Chem. Int. Ed. 2016, 55, 11193–11197. [Google Scholar] [CrossRef]
  238. Yang, L.; Li, L.; Chen, L.; Li, Y.; Chen, H.; Li, Y.; Ji, G.; Lin, D.; Liu, Z.; Qiu, Y. Potential analgesic effects of a novel N-acylethanolamine acid amidase inhibitor F96 through PPAR-α. Sci. Rep. 2015, 5, srep13565. [Google Scholar] [CrossRef] [Green Version]
  239. Petrosino, S.; Ahmad, A.; Marcolongo, G.; Esposito, E.; Allarà, M.; Verde, R.; Cuzzocrea, S.; Di Marzo, V. Diacerein is a potent and selective inhibitor of palmitoylethanolamide inactivation with analgesic activity in a rat model of acute inflammatory pain. Pharmacol. Res. 2015, 91, 9–14. [Google Scholar] [CrossRef]
  240. Sasso, O.; Moreno-Sanz, G.; Martucci, C.; Realini, N.; Dionisi, M.; Mengatto, L.; Duranti, A.; Tarozzo, G.; Tarzia, G.; Mor, M.; et al. Antinociceptive effects of the N-acylethanolamine acid amidase inhibitor ARN077 in rodent pain models. Pain 2013, 154, 350–360. [Google Scholar] [CrossRef] [Green Version]
  241. Banks, W. Blood-Brain Barrier Transport of Cytokines: A Mechanism for Neuropathology. Curr. Pharm. Des. 2005, 11, 973–984. [Google Scholar] [CrossRef]
  242. Varatharaj, A.; Galea, I. The blood-brain barrier in systemic inflammation. Brain Behav. Immun. 2017, 60, 1–12. [Google Scholar] [CrossRef] [Green Version]
  243. Panikashvili, D.; Shein, N.A.; Mechoulam, R.; Trembovler, V.; Kohen, R.; Alexandrovich, A.; Shohami, E. The endocannabinoid 2-AG protects the blood–brain barrier after closed head injury and inhibits mRNA expression of proinflammatory cytokines. Neurobiol. Dis. 2006, 22, 257–264. [Google Scholar] [CrossRef] [PubMed]
  244. Katz, P.S.; Sulzer, J.K.; Impastato, R.A.; Teng, S.X.; Rogers, E.K.; Molina, P.E. Endocannabinoid Degradation Inhibition Improves Neurobehavioral Function, Blood–Brain Barrier Integrity, and Neuroinflammation following Mild Traumatic Brain Injury. J. Neurotrauma 2015, 32, 297–306. [Google Scholar] [CrossRef] [Green Version]
  245. Tchantchou, F.; Zhang, Y. Selective Inhibition of Alpha/Beta-Hydrolase Domain 6 Attenuates Neurodegeneration, Alleviates Blood Brain Barrier Breakdown, and Improves Functional Recovery in a Mouse Model of Traumatic Brain Injury. J. Neurotrauma 2013, 30, 565–579. [Google Scholar] [CrossRef] [Green Version]
  246. Wang, Z.; Li, Y.; Cai, S.; Li, R.; Cao, G. Cannabinoid receptor 2 agonist attenuates blood-brain barrier damage in a rat model of intracerebral hemorrhage by activating the Rac1 pathway. Int. J. Mol. Med. 2018, 42, 2914–2922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Chung, Y.C.; Shin, W.-H.; Baek, J.Y.; Cho, E.J.; Baik, H.H.; Kim, S.R.; Won, S.-Y.; Jin, B.K. CB2 receptor activation prevents glial-derived neurotoxic mediator production, BBB leakage and peripheral immune cell infiltration and rescues dopamine neurons in the MPTP model of Parkinson’s disease. Exp. Mol. Med. 2016, 48, e205. [Google Scholar] [CrossRef] [PubMed]
  248. Yang, M.-C.; Zhang, H.-Z.; Wang, Z.; You, F.-L.; Wang, Y.-F. The molecular mechanism and effect of cannabinoid-2 receptor agonist on the blood–spinal cord barrier permeability induced by ischemia-reperfusion injury. Brain Res. 2016, 1636, 81–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Li, L.; Tao, Y.; Tang, J.; Chen, Q.; Yang, Y.; Feng, Z.; Chen, Y.; Yang, L.; Yang, Y.; Zhu, G.; et al. A Cannabinoid Receptor 2 Agonist Prevents Thrombin-Induced Blood–Brain Barrier Damage via the Inhibition of Microglial Activation and Matrix Metalloproteinase Expression in Rats. Transl. Stroke Res. 2015, 6, 467–477. [Google Scholar] [CrossRef] [PubMed]
  250. Amenta, P.S.; Jallo, J.I.; Tuma, R.F.; Hooper, D.C.; Elliott, M.B. Cannabinoid receptor type-2 stimulation, blockade, and deletion alter the vascular inflammatory responses to traumatic brain injury. J. Neuroinflamm. 2014, 11, 191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Vendel, E.; de Lange, E.C. Functions of the CB1 and CB 2 receptors in neuroprotection at the level of the blood-brain barrier. Neuromol. Med. 2014, 16, 620–642. [Google Scholar] [CrossRef] [PubMed]
  252. Fujii, M.; Sherchan, P.; Krafft, P.R.; Rolland, W.B.; Soejima, Y.; Zhang, J.H. Cannabinoid type 2 receptor stimulation attenuates brain edema by reducing cerebral leukocyte infiltration following subarachnoid hemorrhage in rats. J. Neurol. Sci. 2014, 342, 101–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Kong, W.; Li, H.; Tuma, R.F.; Ganea, D. Selective CB2 receptor activation ameliorates EAE by reducing Th17 differentiation and immune cell accumulation in the CNS. Cell. Immunol. 2014, 287, 1–17. [Google Scholar] [CrossRef] [Green Version]
  254. Rom, S.; Zuluaga-Ramirez, V.; Dykstra, H.; Reichenbach, N.L.; Pacher, P.; Persidsky, Y. Selective Activation of Cannabinoid Receptor 2 in Leukocytes Suppresses Their Engagement of the Brain Endothelium and Protects the Blood-Brain Barrier. Am. J. Pathol. 2013, 183, 1548–1558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Amenta, P.S.; Jallo, J.I.; Tuma, R.F.; Elliott, M.B. A cannabinoid type 2 receptor agonist attenuates blood-brain barrier damage and neurodegeneration in a murine model of traumatic brain injury. J. Neurosci. Res. 2012, 90, 2293–2305. [Google Scholar] [CrossRef] [PubMed]
  256. Ramirez, S.H.; Haskó, 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]
  257. Zhang, M.; Adler, M.W.; Abood, M.E.; Ganea, D.; Jallo, J.; Tuma, R.F. CB2 receptor activation attenuates microcirculatory dysfunction during cerebral ischemic/reperfusion injury. Microvasc. Res. 2009, 78, 86–94. [Google Scholar] [CrossRef] [Green Version]
  258. Hind, W.H.; Tufarelli, C.; Neophytou, M.; Anderson, S.I.; England, T.J.; O’Sullivan, S.E. Endocannabinoids modulate human blood-brain barrier permeability in vitro. Br. J. Pharmacol. 2015, 172, 3015–3027. [Google Scholar] [CrossRef] [Green Version]
  259. Zhou, Y.; Yang, L.; Ma, A.; Zhang, X.; Li, W.; Yang, W.; Chen, C.; Jin, X. Orally administered oleoylethanolamide protects mice from focal cerebral ischemic injury by activating peroxisome proliferator-activated receptor α. Neuropharmacology 2012, 63, 242–249. [Google Scholar] [CrossRef] [PubMed]
  260. Li, Y.; Xu, L.; Zeng, K.; Xu, Z.; Suo, D.; Peng, L.; Ren, T.; Sun, Z.; Yang, W.; Jin, X.; et al. Propane-2-sulfonic acid octadec-9-enyl-amide, a novel PPARα/γ dual agonist, protects against ischemia-induced brain damage in mice by inhibiting inflammatory responses. Brain Behav. Immun. 2017, 66, 289–301. [Google Scholar] [CrossRef]
  261. Mestre, L.; Inigo, P.M.; Mecha, M.; Correa, F.G.; Hernangomez-Herrero, M.; Loria, F.; Docagne, F.; Borrell, J.; Guaza, C. Anandamide inhibits Theiler’s virus induced VCAM-1 in brain endothelial cells and reduces leukocyte transmigration in a model of blood brain barrier by activation of CB(1) receptors. J. Neuroinflamm. 2011, 8, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Kivisäkk, P.; Mahad, D.J.; Callahan, M.K.; Trebst, C.; Tucky, B.; Wei, T.; Wu, L.; Baekkevold, E.S.; Lassmann, H.; Staugaitis, S.M.; et al. Human cerebrospinal fluid central memory CD4+T cells: Evidence for trafficking through choroid plexus and meninges via P-selectin. Proc. Natl. Acad. Sci. USA 2003, 100, 8389–8394. [Google Scholar] [CrossRef] [Green Version]
  263. Wilson, E.H.; Weninger, W.; Hunter, C.A. Trafficking of immune cells in the central nervous system. J. Clin. Investig. 2010, 120, 1368–1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Abadier, M.; Boscacci, R.; Vestweber, D.; Barnum, S.; Deutsch, U.; Engelhardt, B.; Lyck, R.; Cardoso-Alves, L.; Jahromi, N.H. Cell surface levels of endothelial ICAM-1 influence the transcellular or paracellular T-cell diapedesis across the blood-brain barrier. Eur. J. Immunol. 2015, 45, 1043–1058. [Google Scholar] [CrossRef]
  265. Banks, W.A.; Niehoff, M.L.; Ponzio, N.M.; Erickson, M.A.; Zalcman, S.S. Pharmacokinetics and modeling of immune cell trafficking: Quantifying differential influences of target tissues versus lymphocytes in SJL and lipopolysaccharide-treated mice. J. Neuroinflamm. 2012, 9, 231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Bohatschek, M.; Werner, A.; Raivich, G. Systemic LPS Injection Leads to Granulocyte Influx into Normal and Injured Brain: Effects of ICAM-1 Deficiency. Exp. Neurol. 2001, 172, 137–152. [Google Scholar] [CrossRef]
  267. Wang, H.; Sun, J.; Goldstein, H. Human Immunodeficiency Virus Type 1 Infection Increases the In Vivo Capacity of Peripheral Monocytes to Cross the Blood-Brain Barrier into the Brain and the In Vivo Sensitivity of the Blood-Brain Barrier to Disruption by Lipopolysaccharide. J. Virol. 2008, 82, 7591–7600. [Google Scholar] [CrossRef] [Green Version]
  268. Graham, E.; Angel, C.; Schwarcz, L.; Dunbar, P.; Glass, M. Detailed Characterisation of CB2 Receptor Protein Expression in Peripheral Blood Immune Cells from Healthy Human Volunteers Using Flow Cytometry. Int. J. Immunopathol. Pharmacol. 2010, 23, 25–34. [Google Scholar] [CrossRef] [Green Version]
  269. Jean-Gilles, L.; Braitch, M.; Latif, M.L.; Aram, J.; Fahey, A.J.; Edwards, L.J.; Robins, R.A.; Tanasescu, R.; Tighe, P.J.; Gran, B.; et al. Effects of pro-inflammatory cytokines on cannabinoid CB1and CB2receptors in immune cells. Acta Physiol. 2015, 214, 63–74. [Google Scholar] [CrossRef] [PubMed]
  270. Carlisle, S.; Marciano-Cabral, F.; Staab, A.; Ludwick, C.; Cabral, G. 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]
  271. Han, K.H.; Lim, S.; Ryu, J.; Lee, C.-W.; Kim, Y.; Kang, J.-H.; Kang, S.-S.; Ahn, Y.K.; Park, C.-S.; Kim, J.J. CB1 and CB2 cannabinoid receptors differentially regulate the production of reactive oxygen species by macrophages. Cardiovasc. Res. 2009, 84, 378–386. [Google Scholar] [CrossRef] [PubMed]
  272. Kurihara, R.; Tohyama, Y.; Matsusaka, S.; Naruse, H.; Kinoshita, E.; Tsujioka, T.; Katsumata, Y.; Yamamura, H. Effects of Peripheral Cannabinoid Receptor Ligands on Motility and Polarization in Neutrophil-like HL60 Cells and Human Neutrophils. J. Biol. Chem. 2006, 281, 12908–12918. [Google Scholar] [CrossRef] [Green Version]
  273. Balenga, N.A.; Aflaki, E.; Kargl, J.; Platzer, W.; Schroder, R.; Blattermann, S.; Kostenis, E.; Brown, A.J.; Heinemann, A.; Waldhoer, M. GPR55 regulates cannabinoid 2 receptor-mediated responses in human neutrophils. Cell Res. 2011, 21, 1452–1469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Murikinati, S.; Jüttler, E.; Keinert, T.; Ridder, D.A.; Muhammad, S.; Waibler, Z.; Ledent, C.; Zimmer, A.; Kalinke, U.; Schwaninger, M. Activation of cannabinoid 2 receptors protects against cerebral ischemia by inhibiting neutrophil recruitment. FASEB J. 2009, 24, 788–798. [Google Scholar] [CrossRef] [Green Version]
  275. Joseph, J.; Niggemann, B.; Zaenker, K.S.; Entschladen, F. Anandamide is an endogenous inhibitor for the migration of tumor cells and T lymphocytes. Cancer Immunol. Immunother. 2004, 53, 723–728. [Google Scholar] [CrossRef] [PubMed]
  276. Coopman, K.; Smith, L.D.; Wright, K.L.; Ward, S.G. Temporal variation in CB2R levels following T lymphocyte activation: Evidence that cannabinoids modulate CXCL12-induced chemotaxis. Int. Immunopharmacol. 2007, 7, 360–371. [Google Scholar] [CrossRef] [PubMed]
  277. Chouinard, F.; Lefebvre, J.S.; Navarro, P.; Bouchard, L.; Ferland, C.; Lalancette-Hébert, M.; Marsolais, D.; LaViolette, M.; Flamand, N. The Endocannabinoid 2-Arachidonoyl-Glycerol Activates Human Neutrophils: Critical Role of Its Hydrolysis and De Novo Leukotriene B4 Biosynthesis. J. Immunol. 2011, 186, 3188–3196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Chouinard, F.; Turcotte, C.; Guan, X.; LaRose, M.-C.; Poirier, S.; Bouchard, L.; Provost, V.; Flamand, L.; Grandvaux, N.; Flamand, N. 2-Arachidonoyl-glycerol- and arachidonic acid-stimulated neutrophils release antimicrobial effectors against E. coli, S. aureus, HSV-1, and RSV. J. Leukoc. Biol. 2012, 93, 267–276. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Main pathways of endocannabinoid (eCB) generation and degradation. Monoacylglycerols, such as 2—arachidonylglycerol (2-AG) and N-acylethanolamines such as arachidonoylethanolamide (AEA) are generated from membrane phospholipids and provide precursors used in the synthesis of lipoaminoacids and N-acylneurotransmitters. Production of eCBs is associated with the generation of lysophospholipids, like lysophosphatidylinositol (LPI) and lysophosphatidic acid (LPA). ABHD6/12, alpha/beta-hydrolase domain containing 6/12; ADH, alcohol dehydrogenase; 2-AG, 2-arachidonylglycerol; ALDH, aldehyde dehydrogenase; COX, cyclooxygenase; DAG, diacylglycerol; DAGL, diacylglycerol lipase; FAAH, fatty acid amide hydrolase; HPETE-Gs, glycerol esters of 12- or 15-hydroperoxyeicosatetraenoic acid; IP3R, inositol trisphosphate receptor; LOXs, lipoxygenases; MAGL, monoacylglycerol lipase; NAAA, NAE-hydrolyzing acid amidase; NAE, N-acylethanolamine, NAPE, N-acyl-phosphatidylethanolamine; NAPE-AT, NAPE forming N-acyltransferase; PA, phosphatidic acid; PAM, peptidyl-glycine alpha-amidating monooxygenase; PGE2-G, prostaglandin E2–glycerol ester; PGE2S, prostaglandin E2 synthase; PI, phosphatidylinositol; PIP, phosphatidylinositol 3-phosphate; PIP2, phosphatidylinositol 4,5-bisphosphate; PL, phospholipase; PTPN22, protein tyrosine phosphatase non-receptor type 22; SHIP1, SH-2 containing inositol 5′ polyphosphatase 1.
Figure 1. Main pathways of endocannabinoid (eCB) generation and degradation. Monoacylglycerols, such as 2—arachidonylglycerol (2-AG) and N-acylethanolamines such as arachidonoylethanolamide (AEA) are generated from membrane phospholipids and provide precursors used in the synthesis of lipoaminoacids and N-acylneurotransmitters. Production of eCBs is associated with the generation of lysophospholipids, like lysophosphatidylinositol (LPI) and lysophosphatidic acid (LPA). ABHD6/12, alpha/beta-hydrolase domain containing 6/12; ADH, alcohol dehydrogenase; 2-AG, 2-arachidonylglycerol; ALDH, aldehyde dehydrogenase; COX, cyclooxygenase; DAG, diacylglycerol; DAGL, diacylglycerol lipase; FAAH, fatty acid amide hydrolase; HPETE-Gs, glycerol esters of 12- or 15-hydroperoxyeicosatetraenoic acid; IP3R, inositol trisphosphate receptor; LOXs, lipoxygenases; MAGL, monoacylglycerol lipase; NAAA, NAE-hydrolyzing acid amidase; NAE, N-acylethanolamine, NAPE, N-acyl-phosphatidylethanolamine; NAPE-AT, NAPE forming N-acyltransferase; PA, phosphatidic acid; PAM, peptidyl-glycine alpha-amidating monooxygenase; PGE2-G, prostaglandin E2–glycerol ester; PGE2S, prostaglandin E2 synthase; PI, phosphatidylinositol; PIP, phosphatidylinositol 3-phosphate; PIP2, phosphatidylinositol 4,5-bisphosphate; PL, phospholipase; PTPN22, protein tyrosine phosphatase non-receptor type 22; SHIP1, SH-2 containing inositol 5′ polyphosphatase 1.
Ijms 22 05431 g001
Figure 2. The endocannabinoids(eCBs), related compounds and their molecular targets. Major eCBs are monoacylglycerols (such as 2-arachidonylglycerol), N-acylethanolamines (such as arachidonoylethanolamide), lipoaminoacids, N-acylneurotransmitters or primary fatty acid amides, and have different affinity for cannabinoid receptors (CB)1/2, non-cannabinoid G-protein coupled receptors (GPRs), and transient receptor potential cation channel subfamily V member 1 and 4 (TRPV1 and TRPV4). 2-AGE, 2-arachidonyl glyeryl ether (nolandin ether); DEA, docosatetraenoylethanolamide; DHEA, docosahexaenoylethanolamide; EPEA, eicosapentaenoylethanolamide; HEA, homo-γ-linolenylethanolamide; LEA, linoleylethanolamide; 2-LG, 2-linoleoylglycerol; LPA5, lysophosphatidic acid receptor 5; OEA, oleoylethanolamide; 2-OG, 2-oleoylglycerol, PEA, palmitoylethanolamide; SEA, stearoylethanolamide; TRPM8, transient receptor potential cation channel subfamily M (melastatin) member 8.
Figure 2. The endocannabinoids(eCBs), related compounds and their molecular targets. Major eCBs are monoacylglycerols (such as 2-arachidonylglycerol), N-acylethanolamines (such as arachidonoylethanolamide), lipoaminoacids, N-acylneurotransmitters or primary fatty acid amides, and have different affinity for cannabinoid receptors (CB)1/2, non-cannabinoid G-protein coupled receptors (GPRs), and transient receptor potential cation channel subfamily V member 1 and 4 (TRPV1 and TRPV4). 2-AGE, 2-arachidonyl glyeryl ether (nolandin ether); DEA, docosatetraenoylethanolamide; DHEA, docosahexaenoylethanolamide; EPEA, eicosapentaenoylethanolamide; HEA, homo-γ-linolenylethanolamide; LEA, linoleylethanolamide; 2-LG, 2-linoleoylglycerol; LPA5, lysophosphatidic acid receptor 5; OEA, oleoylethanolamide; 2-OG, 2-oleoylglycerol, PEA, palmitoylethanolamide; SEA, stearoylethanolamide; TRPM8, transient receptor potential cation channel subfamily M (melastatin) member 8.
Ijms 22 05431 g002
Figure 3. Signaling pathway and anti-inflammatory actions of NAEs. Arachidonoylethanolamide (AEA) signals via the cannabinoid receptors CB1/2, the non-cannabinoid G protein coupled receptor (GPR)55, the transient receptor potential cation channel subfamily V member 1 (TRPV1), and the peroxisome proliferator-activated receptor (PPAR)-α and γ. Palmitoylethanolamide (PEA) has been shown to inhibit the fatty acid amide hydrolase (FAAH) and to signal via GPR55, TRPV1, and PPAR-α, while CB1/2binding is still controversial. Oleoylethanolamide (OEA) activates TRPV1 and PPAR-α, while stearoylethanolamide (SEA) inhibits FAAH and activates PPAR-α. Linoleylethanolamide (LEA) was shown to activate TRPV1, GPR119, and to inhibit FAAH, while docosahexaenoylethanolamide (DHEA) signaling involves the activation of GPR110.
Figure 3. Signaling pathway and anti-inflammatory actions of NAEs. Arachidonoylethanolamide (AEA) signals via the cannabinoid receptors CB1/2, the non-cannabinoid G protein coupled receptor (GPR)55, the transient receptor potential cation channel subfamily V member 1 (TRPV1), and the peroxisome proliferator-activated receptor (PPAR)-α and γ. Palmitoylethanolamide (PEA) has been shown to inhibit the fatty acid amide hydrolase (FAAH) and to signal via GPR55, TRPV1, and PPAR-α, while CB1/2binding is still controversial. Oleoylethanolamide (OEA) activates TRPV1 and PPAR-α, while stearoylethanolamide (SEA) inhibits FAAH and activates PPAR-α. Linoleylethanolamide (LEA) was shown to activate TRPV1, GPR119, and to inhibit FAAH, while docosahexaenoylethanolamide (DHEA) signaling involves the activation of GPR110.
Ijms 22 05431 g003
Figure 4. The development of synaptic dysfunction and the involvement of the ECS in neuroprotective responses. Endocannabinoid produced in the glutamatergic synapse under excitotoxic conditions attract microglial cells into close proximity to spines. By adopting a pro-inflammatory or pro-survival phenotype microglia largely define the fate of the injured cells and spines. eCB signaling decrease the presynaptic neurotransmitter release, support the glutamate–glutamine cycle, and balanced glutamate/GABAergic transmission. AMPA, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; BDNF, brain-derived neurotrophic factor; CB, cannabinoid receptor; EAATs, excitatory amino acid transporters; eCBs, endocannabinoids; ER, endoplasmic reticulum; GDNF, glial cell line-derived neurotrophic factor; Gln, glutamine; Glu, glutamic acid; IFNγ, interferon gamma; IL, interleukin; mGluRs, metabotropic glutamate receptors; NGF, nerve growth factor; NMDA, N-methyl-D-aspartate receptor; NO, nitrogen monoxide; PGD2, prostaglandin D2; QUIN, quinolinic acid; ROS, reactive oxygen species; TGFβ, transforming growth factor-beta; TNFα, tumor necrosis factor alpha; VGCC, voltage-gated calcium channels.
Figure 4. The development of synaptic dysfunction and the involvement of the ECS in neuroprotective responses. Endocannabinoid produced in the glutamatergic synapse under excitotoxic conditions attract microglial cells into close proximity to spines. By adopting a pro-inflammatory or pro-survival phenotype microglia largely define the fate of the injured cells and spines. eCB signaling decrease the presynaptic neurotransmitter release, support the glutamate–glutamine cycle, and balanced glutamate/GABAergic transmission. AMPA, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; BDNF, brain-derived neurotrophic factor; CB, cannabinoid receptor; EAATs, excitatory amino acid transporters; eCBs, endocannabinoids; ER, endoplasmic reticulum; GDNF, glial cell line-derived neurotrophic factor; Gln, glutamine; Glu, glutamic acid; IFNγ, interferon gamma; IL, interleukin; mGluRs, metabotropic glutamate receptors; NGF, nerve growth factor; NMDA, N-methyl-D-aspartate receptor; NO, nitrogen monoxide; PGD2, prostaglandin D2; QUIN, quinolinic acid; ROS, reactive oxygen species; TGFβ, transforming growth factor-beta; TNFα, tumor necrosis factor alpha; VGCC, voltage-gated calcium channels.
Ijms 22 05431 g004
Figure 5. The stimuli involved in the establishment of synaptic plasticity and retrograde eCB signaling in the glutamatergic synapse. Glutamate release from the presynaptic terminal activates ionotropic (NMDA, AMPA, and kainate) and metabotropic glutamate receptors (mGluRs), that together with the Ca2+ influx through VGCC are potent triggers for eCB production (indicated in green). Ca2+ influx trough NMDA receptors is involved in the regulation of AMPA receptor trafficking, spine enlargement, and plastic changes in the strength of the synaptic transmission (indicated in purple). AC, adenylate cyclase; AMPA, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; CaM, calmodulin; CaMKII, calcium-calmodulin dependent protein kinase II; Cdc42, cell division control protein 42 homolog; DAGL, diacylglycerol lipase; EAATs, excitatory amino acid transporters; eCBs, endocannabinoids; Glu, glutamic acid; IP3, inositol 3-phosphate; Kir, inward-rectifier potassium channel; MAPK, mitogen-activated protein kinase; mGluRs, metabotropic glutamate receptors; NAT, N-acyltransferase; NMDA, N-methyl-D-aspartate receptor; PI3K, phosphatidylinositol 3-kinase; PL, phospholipase; RhoA, Ras homolog family member A; VGCC, voltage-gated calcium channels.
Figure 5. The stimuli involved in the establishment of synaptic plasticity and retrograde eCB signaling in the glutamatergic synapse. Glutamate release from the presynaptic terminal activates ionotropic (NMDA, AMPA, and kainate) and metabotropic glutamate receptors (mGluRs), that together with the Ca2+ influx through VGCC are potent triggers for eCB production (indicated in green). Ca2+ influx trough NMDA receptors is involved in the regulation of AMPA receptor trafficking, spine enlargement, and plastic changes in the strength of the synaptic transmission (indicated in purple). AC, adenylate cyclase; AMPA, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; CaM, calmodulin; CaMKII, calcium-calmodulin dependent protein kinase II; Cdc42, cell division control protein 42 homolog; DAGL, diacylglycerol lipase; EAATs, excitatory amino acid transporters; eCBs, endocannabinoids; Glu, glutamic acid; IP3, inositol 3-phosphate; Kir, inward-rectifier potassium channel; MAPK, mitogen-activated protein kinase; mGluRs, metabotropic glutamate receptors; NAT, N-acyltransferase; NMDA, N-methyl-D-aspartate receptor; PI3K, phosphatidylinositol 3-kinase; PL, phospholipase; RhoA, Ras homolog family member A; VGCC, voltage-gated calcium channels.
Ijms 22 05431 g005
Table 1. Schematic representation of functional contribution of CB1/2 in neurons, astrocytes and microglia.
Table 1. Schematic representation of functional contribution of CB1/2 in neurons, astrocytes and microglia.
ReceptorLocation/
Cell Type
Process RegulatedIntracellular
Pathway Involved
Physiological Relevance
CB1NeuronsPresynaptic vesicular neurotransmitter releaseMultiple, including inhibition of AC and VGCC;
resulting in regulation of [Ca2+]i
Synaptic
plasticity
[Ca2+] influxNeuronal survival
AstrocytesGlutamate–glutamine cycleUpregulation of glutamine-synthasePerisynaptic
glutamate scavenging, prevention of
excitoxicity
GliotransmissionCa2+ influxModulation of synaptic strength;
role in synaptic
plasticity
Endothelial cellsGrowth and proliferationCoupled to the MAP
kinase cascade
Maintenance of BBB integrity
and selectivity
CB2Ventral tegmental area dopamine neuronsNeuronal excitabilityModulation of K+
channels [30]
Long-term neuronal hyperpolarization; synaptic plasticity
Hippocampal CA3/CA2
pyramidal
neurons
Activation of the Na+/Bicarbonate co-transporter [34]
MicrogliaMigrationExtracellular signal-regulated kinase (ERK) 1/2 signal transduction pathwayShift between the pro-inflammatory and pro-resolving phenotype
Table 2. The profiles and activity of eCBs and their congeners in pathological conditions of the CNS.
Table 2. The profiles and activity of eCBs and their congeners in pathological conditions of the CNS.
CompoundOrganismDisease/ModelFindingReferences
2-AGMouseClosed head injuryIncreased 2-AG brain levels; neuroprotection[56]
RatInflammatory model of painAntinociceptive effect[57]
Stress-induced analgesiaIncreased 2-AG and AEA mid-brain levels;
analgesia
[58]
AEAMouseExperimental autoimmune encephalomyelitisIncreased AEA brain levels[59]
HumanMultiple sclerosisIncreased AEA levels in cerebrospinal fluid[59]
RatFocal cerebral ischemia modelIncreased AEA brain levels[60]
NADAMouseAcute systemic inflammationReduces inflammation in vivo;
increases survival in endotoxemic mice
[61]
2-AG, AEA, PEAMouseTransgenic model of Huntington’s diseaseChanged levels of 2-AG, AEA, and PEA in a disease phase- and brain region-specific way[62]
PEARatGranulomatous inflammationReduced granuloma-induced hyperalgesia[63]
MousePaw model of hyperalgesiaReduced mechanical hyperalgesia[39]
Model of neuropathic painAnti-allodynic and anti-hyperalgesic effects[64]
Alzheimer’s diseaseRescued cognitive deficit and reduced
neuroinflammation and oxidative stress
[40]
Traumatic brain injury modelReduced edema and brain infractions; blocked
infiltration of astrocytes
[65]
SEAMouseLPS-induced
neuroinflammation
Decreased activation of resident microglia and leukocyte trafficking into the brain; increased CB1/2
expression and 2-AG brain levels
[47]
Table 3. FAAH, MAGL, and NAAA inhibitors and their effects in neuroinflammation and neurodegenerative diseases.
Table 3. FAAH, MAGL, and NAAA inhibitors and their effects in neuroinflammation and neurodegenerative diseases.
InhibitorCompoundDisease/ModelFindingReferences
FAAH
inhibitors
PF-3845Chronic mouse model of PDdecreased CB2 expression[223]
TBI mouse modelEnhanced AEA brain levels; reduced neurodegeneration in the dentate gyrus; suppressed production of amyloid precursor protein; suppressed expression of iNOS and COX-2[224]
URB597Mouse model of PDIncreased AEA, PEA, and OEA levels; reduced L-DOPA-induced side effects
Inhibition of dopaminergic neuronal death; decreased microglial immunoreactivity; improved motor capacity
[225]
[226]
URB597
JNJ1661010
TCF2
Chronic mouse model of PDAnti-cataleptic effects[227]
URB597
CAY100400
CAY100402
Experimental autoimmune encephalomyelitis (EAE) in miceAnti-spastic effects[228]
MAGL
inhibitors
JZL184LPS-induced neuroinflammation in ratsDecreased peripheral and central cytokine production [229]
EAE in miceAnti-spastic effects[228]
Mouse model of ADImprovements in spatial learning and memory
decreased proinflammatory reactions of microglia and reduced amyloid β burden
[230]
[231]
TBI rat modelImproved neurobehavioral recovery;
reduce astrocyte activation;
less glutamate dyshomeostasis
[232]
CPD-4645LPS-induced
neuroinflammation and focal photothrombotic ischemic insult in mice
Restored functional homeostasis
of the brain vasculature
[175]
KML29chronic MPTP/probenecid
mouse model of PD
Attenuated striatal dopamine depletion;
increase in Gdnf expression
[223]
Ulcerative colitis in humans
back pain in humans
Increased PEA levels in colon
enhanced PEA blood levels
after therapeutic manipulation
[233]
ABX-1431 Formalin pain model in ratsIncreased brain 2-AG concentrations;
suppressed pain behavior
[234]
dual MAGL
FAAH
inhibitors
JZL195Traumatic brain injury in ratsAttenuated neuronal dysfunction[235]
NAAA
inhibitors
AM9053Neuropathy in miceReversed and prevented peripheral neuropathy via PPAR-α[236]
Compound 8EAE in miceDelayed disease onset; attenuated symptom intensity; normalized body weight; reduced leukocyte infiltration and microglia activation[237]
oxazolidin-2-one (F96)Ear edema model in micePrevented allodynia via PPAR-α[238]
EPT4900Carrageenan-induced pain in ratsInhibited inflammation as well as hyperalgesia[239]
ARN077Rodent models of hyperalgesia and allodyniaPrevented heat hyperalgesia and mechanical allodynia via PPAR-α[240]
Table 4. Effects of CB1/CB2 receptor activation on leukocyte function.
Table 4. Effects of CB1/CB2 receptor activation on leukocyte function.
CB receptorCompoundCell TypeEffectReferences
CB1AEA
ACEA
Monocytes
Macrophages
Promotes ROS and cytokine production[271]
CB2JWH-015MacrophagesInhibits ROS production[271]
2-AGNeutrophilsInhibits migration[272]
Promotes migration[273]
JWH-133NeutrophilsInhibits adhesion to
cerebral endothelial cells
[274]
T cellsInhibits migration[275,276]
2-AGT cellsInhibits migration[276]
CB1/CB2
independent
2-AGNeutrophilsStimulates myeloperoxidase and leukotriene release, kinase activation, and calcium mobilization[277]
Release antimicrobial effectors[278]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kasatkina, L.A.; Rittchen, S.; Sturm, E.M. Neuroprotective and Immunomodulatory Action of the Endocannabinoid System under Neuroinflammation. Int. J. Mol. Sci. 2021, 22, 5431. https://doi.org/10.3390/ijms22115431

AMA Style

Kasatkina LA, Rittchen S, Sturm EM. Neuroprotective and Immunomodulatory Action of the Endocannabinoid System under Neuroinflammation. International Journal of Molecular Sciences. 2021; 22(11):5431. https://doi.org/10.3390/ijms22115431

Chicago/Turabian Style

Kasatkina, Ludmila A., Sonja Rittchen, and Eva M. Sturm. 2021. "Neuroprotective and Immunomodulatory Action of the Endocannabinoid System under Neuroinflammation" International Journal of Molecular Sciences 22, no. 11: 5431. https://doi.org/10.3390/ijms22115431

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