Cannabinoid receptor
Cannabinoid receptors are G protein-coupled receptors (GPCRs) that bind endocannabinoids produced endogenously in vertebrates, as well as phytocannabinoids such as Δ9-tetrahydrocannabinol (THC) from Cannabis sativa.[1] The primary subtypes, CB1 and CB2, couple mainly to inhibitory Gi/o proteins, modulating adenylyl cyclase activity, ion channels, and intracellular signaling pathways to regulate diverse physiological processes.[1] These receptors form a core component of the endocannabinoid system, which maintains homeostasis through retrograde neurotransmission and peripheral signaling.[1] CB1 receptors were first cloned from rat brain tissue in 1990, with the CB2 subtype identified in human promyelocytic leukemia cells in 1993.[2] Structurally, both feature seven transmembrane domains typical of class A GPCRs, sharing approximately 44% amino acid sequence identity, though CB1 exhibits greater structural plasticity for ligand binding.[2] Crystal structures of CB1 resolved in 2016 and CB2 in 2019 have elucidated activation mechanisms involving toggle switches in the binding pocket, facilitating agonist-induced conformational changes.[2] CB1 receptors predominate in the central nervous system, particularly at presynaptic terminals of GABAergic and glutamatergic neurons in regions like the hippocampus, cortex, and basal ganglia, where they inhibit neurotransmitter release and modulate synaptic plasticity essential for learning and memory.[3] In contrast, CB2 receptors are chiefly expressed on immune cells such as microglia and macrophages, exerting anti-inflammatory effects by suppressing cytokine production and immune cell migration, with emerging evidence of neuronal expression under pathological conditions like neurodegeneration.[1] Dysregulation of these receptors has been implicated in disorders including chronic pain, epilepsy, and neuroinflammatory diseases, positioning them as targets for therapeutic modulation, though CB1 agonists carry risks of psychoactive side effects and tolerance.[3]Discovery and Historical Context
Initial Identification and Early Research
The hypothesis of specific receptors mediating cannabinoid effects gained traction in the 1970s through structure-activity relationship studies of THC analogs, which suggested receptor involvement due to stereoselectivity and potency differences, though direct binding evidence was lacking owing to THC's low affinity for labeling.[4] Breakthroughs occurred in Allyn Howlett's laboratory at Saint Louis University in the mid-1980s, where experiments first demonstrated stereospecific, high-affinity binding of potent synthetic cannabinoids to rat brain membrane preparations and their inhibition of forskolin-stimulated adenylate cyclase via pertussis toxin-sensitive G-proteins, indicating a G-protein-coupled receptor mechanism.[4] In November 1988, William A. Devane, Floyd A. Dysarz III, and colleagues published the definitive characterization of a cannabinoid receptor in rat brain synaptic plasma membranes, using the tritiated synthetic agonist [³H]CP-55,940 as a radioligand.[5] This binding site displayed high affinity (K_d ≈ 1-5 nM), saturability (B_max ≈ 100-150 fmol/mg protein), reversibility, and pronounced stereoselectivity, with the levorotatory enantiomer of CP-55,940 binding potently while the dextrorotatory form did not; regional distribution varied, with highest densities in globus pallidus and substantia nigra.[5][4] Subsequent early research in the late 1980s and early 1990s employed autoradiography to map receptor localization, revealing enrichment in cerebral cortex, hippocampus, basal ganglia, and cerebellum, consistent with behavioral effects of cannabinoids like hypolocomotion and memory impairment.[4] Functional studies confirmed Gi/o protein coupling, leading to decreased cAMP, increased potassium conductance, and suppressed calcium influx in neurons, establishing the receptor's role in modulating neurotransmitter release.[4] These findings, reliant on synthetic ligands due to endogenous THC's limitations, laid the groundwork for molecular cloning efforts.[5]Key Milestones in Characterization
In 1988, researchers identified high-affinity, stereoselective binding sites for cannabinoids in rat brain synaptic membranes using the radioligand [³H]CP 55,940, providing the first direct evidence for a specific cannabinoid receptor, later designated CB₁.[4] Concurrent functional studies demonstrated that Δ⁹-tetrahydrocannabinol (THC) and synthetic analogs inhibited forskolin-stimulated adenylate cyclase activity in N18TG₂ neuroblastoma cells, indicating mediation by a pertussis toxin-sensitive G protein-coupled mechanism.[4] Autoradiographic mapping further localized these sites predominantly to brain regions involved in cognition, memory, and motor control, such as the hippocampus, cerebellum, and basal ganglia.[6] Molecular cloning marked a pivotal advance in receptor characterization. In 1990, the CB₁ receptor gene was isolated from rat cerebral cortex cDNA libraries through expression cloning, yielding a 472-amino-acid protein with seven transmembrane domains and features typical of the G protein-coupled receptor (GPCR) superfamily, including conserved aspartate residues in transmembrane helices II and III for ligand binding.[4] The human CB₁ ortholog was cloned independently the same year, sharing 97% amino acid identity with the rat sequence and confirming its neuronal enrichment via in situ hybridization.[4] These efforts revealed coupling to Gᵢ/Gₒ proteins, inhibiting adenylyl cyclase and modulating ion channels such as voltage-gated calcium channels.[7] The discovery of a second receptor type followed in 1993, when the CB₂ gene was cloned from human promyelocytic leukemia HL60 cell cDNA, encoding a 360-amino-acid protein with 68% sequence homology to CB₁ but lacking significant central nervous system expression.[4] Pharmacological profiling distinguished CB₂ by its prominence in spleen, leukocytes, and tonsils, with binding affinities for THC analogs mirroring CB₁ but enabling tissue-specific functional assays.[8] Early characterization confirmed CB₂'s Gi/o-mediated signaling, including cAMP suppression and mitogen-activated protein kinase activation in immune cells, broadening understanding of cannabinoid effects beyond the brain.[7] The development of selective antagonists, such as SR141716A for CB₁ in 1994, further delineated receptor subtypes through blockade of THC-induced behaviors and immune modulation.[4]Classification and Molecular Structure
CB1 Receptor Structure and Genetics
The CB1 receptor, also known as cannabinoid receptor type 1, is encoded by the CNR1 gene located on the long arm of human chromosome 6 at position 6q15.[9] [10] The gene spans 26.1 kilobases and comprises four exons, with the primary protein-coding region residing in exon 4, which is the largest exon and most commonly expressed.[11] [12] The CNR1 open reading frame encodes a 472-amino-acid polypeptide chain in humans, exhibiting 97–99% sequence identity with rodent orthologs.[13] Genetic variations in CNR1 include single-nucleotide polymorphisms (SNPs) such as rs1049353 (G1359A in exon 3, also known as the 1359 G/A variant) and microsatellite repeats like (AAT)n in the 3' untranslated region, which can influence mRNA stability and receptor expression levels.[14] [15] Rare coding variants in CNR1 have been identified, potentially altering receptor function, though their prevalence is low (e.g., heterozygous variants associated with altered pain sensitivity in some cohorts).[16] As a class A rhodopsin-like G protein-coupled receptor (GPCR), the CB1 protein features a characteristic architecture with seven α-helical transmembrane domains (TMDs), an extracellular N-terminal domain of 116 amino acids lacking a canonical signal peptide, three extracellular loops, three intracellular loops, and an intracellular C-terminal tail.[17] [18] [19] The ligand-binding pocket is primarily buried within the TMD bundle, with key residues in helices 2–7 contributing to agonist and antagonist interactions, as elucidated by cryo-electron microscopy and X-ray crystallography structures resolved to 2.6–3.3 Å resolution.[20] [21] These structural insights highlight a toggle switch mechanism involving Trp6.48 (Ballesteros-Weinstein numbering) for receptor activation and allosteric modulation sites distinct from the orthosteric pocket.[20]CB2 Receptor Structure and Genetics
The CNR2 gene encodes the CB2 cannabinoid receptor and is located on the short arm of human chromosome 1 at cytogenetic band 1p36.11, oriented on the minus strand.[22][23] The gene spans roughly 90 kilobases and contains six exons, with alternative splicing yielding at least two mRNA isoforms: the full-length CNR2A and a shorter CNR2B variant lacking exon 2, which results in a frameshift and truncated protein.[24][25] The predominant isoform, CNR2A, produces a 347-amino-acid protein, shorter than the CB1 receptor due to a more compact extracellular N-terminus and intracellular C-terminus.[25] As a class A G protein-coupled receptor (GPCR), CB2 features seven transmembrane α-helices connected by three intracellular and three extracellular loops, with an extracellular N-terminal domain and intracellular C-terminal tail involved in signaling and trafficking.[26] The ligand-binding pocket lies within the transmembrane bundle, accommodating diverse agonists like endocannabinoids and synthetic cannabinoids through hydrophobic interactions and hydrogen bonding, as revealed by structural analyses.[27] High-resolution structures have elucidated CB2's conformational dynamics. A 2.8 Å crystal structure of inactive CB2 bound to antagonist AM10257, stabilized by a fusion protein and lipidic cubic phase, shows a closed orthosteric site and distinct helix orientations compared to CB1.[26] Subsequent cryo-electron microscopy (cryo-EM) studies, including a 3.2 Å resolution complex with agonist AM10233 and Gi protein, depict an active conformation with inward TM6 displacement and outward TM7 movement, enabling G protein coupling.[28] Additional cryo-EM structures with agonists like WIN 55,212-2 confirm conserved GPCR activation mechanisms while highlighting CB2-specific residues for ligand selectivity, such as Phe117 in TM3 and Trp258 in TM6.[27][29] Genetic variations in CNR2 include single nucleotide polymorphisms (SNPs) like rs2501431 (Q63R) in exon 1, which alters receptor trafficking and signaling efficiency, and rs35761398 in the promoter region affecting expression levels.[30] These polymorphisms exhibit population-specific frequencies and have been linked to immune modulation, though functional impacts vary; for instance, the Q63R variant reduces β-arrestin recruitment without abolishing G protein signaling.[30] Interspecies differences in CNR2 structure, including exon-intron organization and isoform prevalence, underscore challenges in translating rodent models to human pharmacology.[25]Evidence for Non-Classical Receptors
Research has identified several orphan G-protein-coupled receptors (GPCRs) as potential mediators of cannabinoid signaling independent of classical CB1 and CB2 receptors, based on binding affinities for endocannabinoids like anandamide and synthetic ligands such as CP55940, coupled with functional responses in cellular assays.[31] These candidates include GPR55 and GPR18, which exhibit partial overlap in pharmacology but distinct signaling profiles, such as GPR55-mediated increases in intracellular calcium via Gq/11 proteins rather than the Gi/o inhibition typical of CB1/CB2.[32] However, their classification as bona fide "cannabinoid receptors" remains debated due to inconsistencies in ligand selectivity, species differences in expression, and lack of comprehensive genetic validation akin to CB1/CB2 knockouts.[33] GPR55, first proposed as a cannabinoid-sensitive receptor in 2007, binds endocannabinoids including anandamide and virodhamine with micromolar affinity and is activated by lysophosphatidylinositol (LPI) as an endogenous agonist, leading to RhoA activation and cytoskeletal remodeling in cells like dorsal root ganglion neurons.[31] Functional evidence includes GPR55 knockout mice showing reduced mechanical hyperalgesia in inflammatory and neuropathic pain models, implicating it in pain sensitization without the psychoactive effects of CB1.[34] In bone physiology, GPR55 antagonism inhibits osteoclast function and prevents bone loss in ovariectomized mice, suggesting a role in regulating bone mass via cannabinoid ligands like cannabidiol (CBD).[35] Despite these findings, GPR55 does not respond uniformly to all classical cannabinoids—e.g., Δ9-tetrahydrocannabinol (THC) shows weak activation—and its signaling diverges in human versus rodent orthologs, complicating therapeutic translation.[36] A 2024 study in glutamate neurons found GPR55 expression but questioned its direct cannabinoid mediation due to absent colocalization with endocannabinoid machinery in key brain regions.[33] GPR18, another orphan GPCR, responds to atypical cannabinoids like abnormal cannabidiol (Abn-CBD) and N-arachidonyl glycine, promoting biased agonism that favors β-arrestin recruitment over G-protein signaling, as evidenced by migration assays in microglia and endometrial cells.[37][38] In BV-2 microglia, siRNA knockdown of GPR18 abolishes Abn-CBD-induced apoptosis suppression under pro-inflammatory conditions, confirming receptor-specific effects on survival pathways.[39] Cardiovascular studies demonstrate GPR18 activation in the rostral ventrolateral medulla lowers blood pressure via enhanced neuronal adiponectin and nitric oxide production while reducing reactive oxygen species, effects blocked by GPR18 antagonists.[40] GPR18 shares only ~13% sequence homology with CB1 and ~8% with CB2, and its expression in immune and reproductive tissues supports roles in inflammation resolution and cellular migration, though ligand promiscuity with non-cannabinoid fatty acid amides tempers claims of specificity.[41][42] Other proposed non-classical targets, such as GPR119, show weaker evidence, with limited binding to cannabinoids and primary activation by lipid-derived agonists in metabolic contexts rather than canonical endocannabinoid signaling.[43] Transient receptor potential (TRP) channels like TRPV1 also mediate some hyperalgesic effects of anandamide but function as ion channels rather than GPCRs, distinguishing them from receptor paradigms.[44] Overall, while pharmacological and genetic data support modulatory roles for these receptors in cannabinoid-evoked responses—particularly in pain, inflammation, and metabolism—definitive proof requires orthogonal validation, as many effects may arise from off-target interactions or downstream crosstalk rather than dedicated receptor activation.[45] Reviews emphasize that non-CB1/CB2 mechanisms account for only a subset of observed cannabinoid bioactivity, underscoring the need for selective ligands to disentangle contributions.Tissue Distribution and Expression Patterns
CB1 Distribution in the Central Nervous System
The CB1 receptor exhibits one of the highest expression levels among G protein-coupled receptors in the central nervous system, with dense localization primarily on presynaptic terminals of neurons.[46] This distribution enables modulation of neurotransmitter release, particularly of GABA and glutamate, across multiple brain regions.[3] In the cerebral cortex, CB1 receptors show high density, including in the neocortex, frontal cortex, cingulate gyrus, and motor cortex, where they are enriched on GABAergic interneurons.[46][47] The hippocampus displays prominent CB1 expression, predominantly in the CA1-CA3 regions and dentate gyrus, again favoring presynaptic sites on cholecystokinin-positive GABAergic basket cells over principal glutamatergic neurons.[48][49] Basal ganglia structures, such as the striatum, globus pallidus, and substantia nigra, contain some of the highest CB1 densities in the brain, supporting roles in motor control and reward processing through inhibition of GABAergic and dopaminergic transmission.[50][51] The cerebellum exhibits intense CB1 immunoreactivity, particularly in the molecular layer on parallel fiber-Purkinje cell synapses, contributing to fine-tuned motor coordination.[47][51] Additional regions with notable CB1 presence include the amygdala, especially the lateral and basal nuclei, where expression on GABAergic terminals influences emotional processing and fear responses.[49] Lower but detectable levels occur in areas like the thalamus and hypothalamus, with overall CNS distribution reflecting evolutionary conservation across mammals, as confirmed by immunocytochemical and autoradiographic studies in primates and rodents.[51][47] Subcellular organization often features CB1 receptors in periodic hotspots along axons, enhancing signaling efficiency upon endocannabinoid activation.[52]CB2 Predominance in Peripheral Tissues
The CB2 receptor, also known as the peripheral cannabinoid receptor, demonstrates markedly higher expression in non-neuronal peripheral tissues compared to the central nervous system (CNS), where its levels are minimal under basal conditions. This distribution pattern was established through early cloning and transcript analysis, revealing CB2 mRNA abundance in immune organs such as the spleen and thymus, with transcript levels approximately 10- to 100-fold greater than those of CB1 in these sites.[53][54] In contrast, CB2 transcripts are virtually absent from brain tissue, underscoring its role primarily outside the CNS.[53] Within peripheral tissues, CB2 expression is concentrated in cells of hematopoietic origin, including macrophages, B lymphocytes, T lymphocytes, natural killer cells, and polymorphonuclear neutrophils, as confirmed by in situ hybridization and flow cytometry studies on human leukocytes.[55][1] Notably, the receptor was first cloned from the human promyelocytic leukemia cell line HL-60, highlighting its prominence in myeloid-derived immune cells.[55] High CB2 density is also observed in secondary lymphoid organs like tonsils and the marginal zone of the spleen, where it modulates immune responses.[53] This peripheral bias positions CB2 as a key regulator of inflammation and immune function, with expression upregulated in activated immune states such as during infection or tissue injury.[56] Although some studies have detected low-level CB2 expression in peripheral neurons or under pathological CNS conditions (e.g., neuroinflammation via microglia), these findings do not alter the receptor's overall predominance in extraneural tissues, where it comprises up to 70-80% of total cannabinoid receptor binding in spleen homogenates.[57][58] This tissue-specific profile supports CB2's therapeutic potential in peripheral disorders like autoimmune diseases and chronic pain without prominent psychoactive effects associated with CB1 agonism.[59]Regulation of Expression
The expression of CNR1 (encoding CB1) and CNR2 (encoding CB2) genes is controlled at multiple levels, including transcriptional initiation via promoter elements, epigenetic modifications such as DNA methylation and histone acetylation, and post-transcriptional regulation by microRNAs (miRNAs). These mechanisms allow dynamic adaptation to physiological states, developmental stages, and pathological conditions like inflammation or neurodegeneration. Epigenetic alterations, in particular, enable heritable changes in receptor density without altering the DNA sequence, influencing endocannabinoid signaling efficacy.[60][13] For CB1 receptors, CNR1 promoter DNA methylation negatively correlates with mRNA levels in human postmortem brain tissues, including the prefrontal cortex, hippocampus, and caudate nucleus, with higher methylation associated with reduced expression in regions vulnerable to psychiatric disorders.[61] In schizophrenia patients, CNR1 mRNA is downregulated in subcortical brain areas such as the thalamus and putamen, potentially contributing to altered endocannabinoid tone, while upregulated in peripheral blood leukocytes, suggesting tissue-specific regulatory divergence.[62] Developmental and stress-related epigenetic reprogramming can further modulate CNR1, as seen in rodent models of early-life adversity combined with adult stressors, where histone modifications lead to transient upregulation followed by normalization.[63] CB2 receptor expression exhibits greater inducibility, particularly in immune and glial cells, where inflammatory cytokines and lipopolysaccharide (LPS) challenge upregulate CNR2 mRNA via transcription factor activation and reduced promoter methylation.[64] In peripheral sensory neurons, nerve injury promotes CNR2 upregulation through bivalent histone modifications (H3K4me3 and H3K27me3), shifting chromatin from a poised to an active state and enhancing anti-nociceptive signaling.[65] Post-transcriptionally, miR-187-3p binds the CNR2 3' untranslated region to suppress translation, limiting CB2 levels in non-stimulated states, while stress-induced epigenetic silencing in microglia maintains low basal expression until neuroinflammatory triggers demethylate the locus.[13][64] Exogenous cannabinoids can reciprocally influence receptor expression epigenetically; for instance, Δ9-tetrahydrocannabinol (THC) alters DNA methylation and miRNA profiles in immune cells, potentially downregulating CB1/2 to prevent overstimulation, though chronic exposure risks tolerance via sustained histone deacetylation.[66] These regulatory patterns underscore CB1's relatively stable neuronal expression versus CB2's adaptive role in peripheral immunity, with dysregulation implicated in conditions like chronic pain and autoimmune disorders.[67]Signaling Pathways and Mechanisms
Canonical G-Protein Mediated Signaling
The canonical signaling pathway of cannabinoid receptors CB1 and CB2 involves coupling to pertussis toxin-sensitive heterotrimeric G proteins of the Gi/o family upon activation by agonists such as endocannabinoids (e.g., anandamide or 2-arachidonoylglycerol).[68][69] This coupling promotes GDP-to-GTP exchange on the Gα subunit, leading to dissociation of the Gαi/o subunit from the Gβγ dimer, which then modulates downstream effectors.[70] Both receptors exhibit this primary Gi/o-mediated mechanism, though CB1 demonstrates higher coupling efficiency in neuronal contexts, while CB2 predominates in immune and peripheral cells.[71] The Gαi/o subunit primarily inhibits adenylyl cyclase isoforms (AC1, AC3, AC5, AC6, and others), reducing cyclic AMP (cAMP) production and subsequent protein kinase A (PKA) activation, which alters gene transcription and cellular excitability.[72][73] This cAMP suppression is a core feature, measurable via assays showing dose-dependent decreases in forskolin-stimulated cAMP levels in cells expressing CB1 or CB2 transfectants, with EC50 values around 1-10 nM for potent agonists like CP55,940.[74] Additionally, Gαi/o can activate phospholipase Cβ (PLCβ) in some systems, generating inositol trisphosphate (IP3) and diacylglycerol (DAG), though this is less dominant than AC inhibition.[69] The free Gβγ subunits exert direct effects on ion channels and kinases, particularly for CB1 in presynaptic terminals: they inhibit voltage-gated N-, P-, and Q-type calcium channels (reducing Ca²⁺ influx and neurotransmitter release) and activate G protein inwardly rectifying potassium (GIRK) channels (hyperpolarizing membranes).[70][71] These βγ-mediated actions occur independently of cAMP changes and are pertussis toxin-sensitive, as demonstrated in patch-clamp studies on hippocampal neurons where CB1 agonists like WIN55,212-2 suppress Ca²⁺ currents by 20-50% within seconds.[3] For CB2, βγ signaling similarly modulates immune cell effectors, including reduced chemotaxis via channel modulation, though with less emphasis on neuronal ion channels.[56] Gi/o coupling also engages mitogen-activated protein kinase (MAPK) cascades, such as ERK1/2 phosphorylation via Ras-Raf-MEK pathways, contributing to long-term cellular adaptations like synaptic plasticity for CB1.[73] This pathway persists even after β-arrestin blockade, underscoring its G-protein dependence.[75] While both receptors share these mechanisms, ligand bias can modulate Gi/o efficacy; for instance, synthetic agonists like HU-308 show preferential CB2 Gi/o activation without strong β-arrestin recruitment.[76] Pertussis toxin pretreatment abolishes these effects, confirming Gi/o specificity across studies.[77]Non-Canonical Pathways and Crosstalk
Cannabinoid receptors CB1 and CB2, classically coupled to Gi/o proteins to inhibit adenylyl cyclase and modulate ion channels, also engage non-canonical pathways that diversify their signaling outputs. These include coupling to stimulatory G proteins such as Gs and Gq/11, recruitment of β-arrestins for scaffold-dependent kinase activation, and downstream effectors like MAPK/ERK1/2 and PI3K/Akt, often context-dependent on cell type, ligand, and receptor localization.[78][68] For CB1, non-canonical Gs coupling elevates cAMP levels, as demonstrated in HEK293 cells and hippocampal neurons with agonists like WIN55,212-2, contrasting the canonical inhibitory effect.[68] Gq/11 activation via CB1 stimulates phospholipase C (PLC), increasing intracellular calcium and IP3 in astrocytes and neurons, contributing to glutamate release modulation.[78] β-Arrestin recruitment represents a key non-canonical mechanism for both receptors, facilitating receptor desensitization, internalization, and biased signaling from endosomal compartments. In CB1-transfected cells and neurons, β-arrestin-1 and -2 mediate late-phase ERK1/2 phosphorylation via scaffolds involving Src and PKCε, independent of G-protein activation, as shown in studies using phosphorylation-deficient mutants.[46][78] For CB2, β-arrestin pathways regulate sustained signaling post-internalization in immune cells, with agonists like CP55,940 exhibiting full recruitment efficacy, influencing anti-inflammatory cytokine profiles.[68] Ligand bias amplifies these effects; for instance, Δ9-tetrahydrocannabinol (THC) biases CB1 toward β-arrestin-1 over G-protein paths in neuronal cultures, while JWH133 favors β-arrestin at CB2 in microglia, altering nitric oxide release.[68] Such bias, quantified via operational models in HEK293 assays, underscores how synthetic and endogenous ligands like 2-arachidonoylglycerol (2-AG) selectively engage effectors, impacting proliferation and survival.[68] Crosstalk arises through heterodimerization of CB1 or CB2 with other GPCRs, altering canonical signaling and enabling emergent pathways. CB1-D2 dopamine receptor heteromers in striatal neurons switch coupling to Gs upon co-activation, elevating cAMP and ERK1/2 contrary to individual Gi/o inhibition, as evidenced in co-transfected cells and rat striatum.[78][46] Similarly, CB1-A2A adenosine receptor interactions attenuate motor effects in behavioral models, while CB1-CB2 heteromers in pineal gland and nucleus accumbens reduce Akt phosphorylation, modulating neuroprotection.[78] For CB2, crosstalk with CXCR4 in cancer cells diminishes migration via altered ERK signaling, confirmed in heteromer-expressing lines.[78] These interactions, detected via co-immunoprecipitation and FRET in primary cells, highlight causal roles in fine-tuning responses like inflammation and synaptic plasticity, with implications for therapeutic selectivity.[78]Allosteric Sites and Recent Structural Insights
Allosteric sites on cannabinoid receptors CB1 and CB2 enable modulation of receptor activity without direct competition at the orthosteric binding pocket occupied by endogenous ligands like anandamide or exogenous agonists such as Δ9-tetrahydrocannabinol (THC). These sites, located extracellularly or within the transmembrane helices interfacing with the lipid bilayer, facilitate positive allosteric modulators (PAMs) that enhance orthosteric ligand efficacy or negative allosteric modulators (NAMs) that reduce it, offering potential for biased signaling and subtype selectivity.[79][80] The first high-resolution structure of an allosteric modulator bound to CB1, determined by X-ray crystallography in 2019, revealed that the NAM ORG27569 occupies an extrahelical pocket in the inner leaflet of the membrane, overlapping with a conserved cholesterol interaction site between transmembrane helices 2, 3, and 4 (TM2–TM3–TM4). This binding stabilizes an intermediate conformation that impairs G-protein coupling while allowing orthosteric agonist binding, demonstrating non-competitive antagonism.[81][82] Subsequently, cryo-EM structures of CB1 with the PAM ZCZ011, reported in 2024, showed binding to a distinct yet overlapping site on the TM2–TM3–TM4 surface, where it promotes Gi-mediated signaling over β-arrestin pathways by altering intracellular loop dynamics and G-protein engagement. Unlike ORG27569, ZCZ011 induces subtle shifts that favor active-state conformations for productive transducer interactions, highlighting how allosteric ligands can bias efficacy.[83][84] For CB2, structural insights into allosteric sites remain less resolved, but mutagenesis-guided mapping in 2025 identified a pocket involving residues in TM3 and TM5, distinct from the orthosteric site, amenable to small-molecule modulation for immune-related applications. Computational docking and functional assays confirmed that ligands targeting this site enhance orthosteric agonist potency without psychoactivity associated with CB1 activation. Recent cryo-EM structures of agonist-bound CB2-Gi complexes, including those from 2023 onward, have indirectly informed allosteric modeling by revealing helix rearrangements that expose these sites upon activation.[85][84][80] These structural advances, combining X-ray crystallography with cryo-EM resolutions below 3 Å, underscore conserved yet receptor-specific allosteric mechanisms across CB1 and CB2, including lipid-cholesterol dependencies and helix 8 involvement, paving the way for designing modulators that selectively target pathological signaling without broad orthosteric disruption. For instance, allosteric sites enable decoupling of therapeutic anti-inflammatory effects from CB1-mediated cognitive side effects. Ongoing efforts integrate these insights with structure-activity relationship studies to optimize modulator selectivity and pharmacokinetics.[81][83][86]Ligands and Pharmacology
Endogenous Cannabinoids
Endogenous cannabinoids, or endocannabinoids, are lipid-derived signaling molecules produced on demand within the body that primarily activate the G-protein-coupled receptors CB1 and CB2, modulating various physiological processes including neurotransmission and inflammation.[87] The two principal endocannabinoids are N-arachidonoylethanolamine (anandamide, AEA) and 2-arachidonoylglycerol (2-AG), both derived from arachidonic acid and exhibiting agonist activity at cannabinoid receptors, though with differing potencies and tissue distributions.[88] AEA was the first identified in 1992 from porcine brain tissue, where it competitively inhibited binding of radiolabeled cannabinoid probes to synaptosomal membranes, demonstrating partial agonist effects at CB1 similar to Δ⁹-tetrahydrocannabinol (THC).[89] In contrast, 2-AG, discovered shortly thereafter, is more abundant in tissues and acts as a full agonist at both CB1 and CB2 receptors, with higher efficacy in stimulating downstream signaling pathways like inhibition of adenylyl cyclase.[90] AEA biosynthesis occurs via enzymatic hydrolysis of N-arachidonoyl-phosphatidylethanolamine (NArPE) by N-acyl phosphatidylethanolamine phospholipase D (NAPE-PLD), yielding AEA alongside phosphatidic acid and ethanolamine; this pathway is calcium-dependent and activated by neuronal activity.[91] 2-AG is synthesized from diacylglycerol (DAG) through sequential actions of phospholipase C (PLC) and sn-1-diacylglycerol lipase (DAGL), predominantly in postsynaptic neurons, facilitating its role as a retrograde messenger.[87] Degradation of AEA is mainly mediated by fatty acid amide hydrolase (FAAH), an integral membrane enzyme that hydrolyzes it to arachidonic acid and ethanolamine, while 2-AG is primarily broken down by monoacylglycerol lipase (MAGL), producing arachidonic acid and glycerol; these enzymes regulate endocannabinoid tone and prevent overstimulation of receptors.[92] Additional putative endocannabinoids include virodhamine (arachidonoyl glycine ether), which acts as an endogenous antagonist or partial agonist at CB1 while agonizing CB2, noladin ether (2-arachidonyl glyceryl ether), a CB1-selective agonist with lower potency, and N-arachidonoyldopamine (NADA), which binds CB1 and the transient receptor potential vanilloid 1 (TRPV1) channel.[93] However, AEA and 2-AG remain the most studied and functionally dominant, with 2-AG levels typically 100- to 1000-fold higher than AEA in brain tissue, underscoring their complementary roles in endocannabinoid signaling.[94] Structural variations, such as acyl chain length or saturation, influence receptor affinity; for instance, AEA's ethanolamide head group confers partial agonism at CB1, whereas 2-AG's glycerol ester enables broader efficacy across receptor subtypes.[95]Exogenous Ligands: Phytocannabinoids and Synthetics
Exogenous ligands for cannabinoid receptors encompass plant-derived phytocannabinoids and laboratory-synthesized compounds that interact with CB1 and CB2 receptors. These ligands mimic or block the effects of endogenous cannabinoids, influencing receptor signaling through agonism, partial agonism, antagonism, or inverse agonism. Phytocannabinoids, primarily isolated from Cannabis sativa, were among the first identified modulators, with Δ⁹-tetrahydrocannabinol (THC) discovered in 1964 by Gaoni and Mechoulam as the main psychoactive constituent binding to CB1 with moderate affinity (Ki ≈ 40 nM) and acting as a partial agonist, while exhibiting lower affinity for CB2 (Ki ≈ 360 nM).[2] Cannabidiol (CBD), another major phytocannabinoid, displays negligible direct binding affinity for both CB1 and CB2 (Ki > 10,000 nM), though it modulates receptor activity indirectly or allosterically, particularly antagonizing CB1 in the presence of THC.[96] Other phytocannabinoids include cannabigerol (CBG), cannabichromene (CBC), and cannabinol (CBN), which generally show lower affinities than THC. CBG acts as a partial agonist at both receptors with Ki values in the micromolar range for CB1 and moderate affinity for CB2, while CBN demonstrates higher selectivity for CB2 (Ki ≈ 100 nM) compared to CB1 (Ki ≈ 200 nM). CBC and CBDV exhibit weak binding to both receptors, often with Ki > 1 μM, limiting their direct pharmacological potency at cannabinoid sites.[97][98] These compounds' interactions are supported by radioligand binding assays in recombinant expression systems, revealing structure-activity relationships where variations in alkyl side chains and cyclization affect receptor selectivity.[99]| Phytocannabinoid | CB1 Ki (nM) | CB2 Ki (nM) | Activity Notes |
|---|---|---|---|
| Δ⁹-THC | ~40 | ~360 | Partial agonist at CB1; weaker at CB2[2] |
| CBD | >10,000 | >10,000 | Negligible affinity; indirect modulation[96] |
| CBG | >1,000 | ~1,000 | Partial agonist; low potency[98] |
| CBN | ~200 | ~100 | Partial agonist; CB2 preferential[100] |
Physiological Roles and Functions
Neuromodulation and Synaptic Plasticity
Cannabinoid receptor type 1 (CB1) receptors, densely expressed on presynaptic terminals throughout the central nervous system, exert neuromodulatory effects by inhibiting the release of neurotransmitters such as GABA, glutamate, dopamine, and acetylcholine.[104] This inhibition occurs primarily through Gi/o-protein coupling, which suppresses adenylyl cyclase activity, reduces cyclic AMP levels, closes voltage-gated calcium channels, and activates inwardly rectifying potassium channels, thereby decreasing presynaptic calcium influx and vesicular release probability.[104] Endocannabinoids like 2-arachidonoylglycerol (2-AG) and N-arachidonoylethanolamine (anandamide) serve as retrograde messengers, synthesized postsynaptically in response to calcium elevation or Gq-coupled receptor activation, diffusing across the synapse to bind presynaptic CB1 receptors.[105] A hallmark of this neuromodulation is depolarization-induced suppression of inhibition (DSI), where postsynaptic depolarization triggers endocannabinoid release that transiently suppresses GABAergic transmission via CB1 activation on inhibitory terminals, as demonstrated in hippocampal and cerebellar neurons.[106] Similarly, depolarization-induced suppression of excitation (DSE) reduces glutamatergic release at excitatory synapses, with both processes lasting seconds to minutes and requiring CB1 integrity, as evidenced by their abolition in CB1 knockout mice.[107][108] These short-term plasticities fine-tune network excitability, preventing overexcitation or excessive inhibition during high-activity states.[107] In synaptic plasticity, CB1-mediated endocannabinoid signaling drives specific forms of long-term depression (LTD), particularly endocannabinoid-dependent LTD (eCB-LTD) at excitatory and inhibitory synapses in regions including the hippocampus, striatum, amygdala, and prefrontal cortex.[105] For example, in hippocampal CA1 pyramidal neurons, repetitive low-frequency stimulation induces eCB-LTD via postsynaptic group I metabotropic glutamate receptor activation, leading to 2-AG synthesis and sustained presynaptic CB1 inhibition that persists for hours, blocked by CB1 antagonists such as rimonabant (SR141716).[105] Cerebellar parallel fiber-Purkinje cell synapses exhibit CB1-dependent LTD following conjunctive stimulation, essential for motor coordination refinement.[105] Conversely, endocannabinoids can bidirectionally modulate plasticity, sometimes facilitating LTP in dopamine-rich circuits like the nucleus accumbens under low-release conditions.[109] CB2 receptors contribute minimally to central neuromodulation and synaptic plasticity, with expression confined largely to microglia and immune cells rather than neurons, though peripheral or neuroinflammatory contexts may involve indirect modulation.[46] Pharmacological and genetic evidence underscores CB1's dominance: CB1-/- mice show deficits in DSI, DSE, and multiple LTD variants, linking these processes to behaviors like fear extinction and habit formation.[106][105] Dysregulation of this system, as seen with chronic cannabinoid exposure, impairs plasticity induction, potentially underlying cognitive deficits observed in heavy users.[110]Immune Modulation and Inflammation
Cannabinoid receptor 2 (CB2) predominates in immune modulation, with high expression on hematopoietic cells including macrophages, B and T lymphocytes, natural killer cells, dendritic cells, and microglia, particularly under inflammatory conditions.[56] This distribution positions CB2 to regulate innate and adaptive immune responses, primarily exerting immunosuppressive effects that curb excessive inflammation. Endogenous cannabinoids like 2-arachidonoylglycerol (2-AG) and anandamide bind CB2 to inhibit adenylyl cyclase via G_i/o proteins, lowering cyclic AMP levels and thereby dampening pro-inflammatory signaling.[111] CB1 receptors contribute peripherally but to a lesser extent, as their expression on immune cells is minimal compared to neuronal tissues.[112] CB2 activation suppresses key inflammatory mediators, reducing production of cytokines such as tumor necrosis factor-α (TNF-α), interleukin-1β (IL-1β), IL-6, and IL-12 in macrophages and microglia, while elevating anti-inflammatory IL-10.[111] These effects involve modulation of mitogen-activated protein kinase (MAPK) pathways, including ERK1/2 and p38, and inhibition of NF-κB translocation, which collectively impair immune cell activation.[56] In T cells and B cells, CB2 signaling inhibits proliferation and induces apoptosis through caspase activation and reactive oxygen species generation, particularly in hyperactivated states, thus preventing cytokine storms.[112] Leukocyte migration and chemotaxis are also attenuated, as CB2 ligands disrupt adhesion molecule expression and cytoskeletal remodeling in neutrophils and monocytes.[111] In vivo evidence underscores these roles: CB2 knockout mice display exacerbated inflammation, with increased neutrophil recruitment and cytokine levels in models like dinitrofluorobenzene-induced dermatitis.[111] Selective CB2 agonists, such as JWH-133 (K_i = 3.4 nM), reduce TNF-α by up to 50% in macrophage cultures and attenuate joint destruction in rheumatoid arthritis models by limiting synovial inflammation.[111] Similarly, in experimental autoimmune encephalomyelitis—a multiple sclerosis analog—CB2 activation decreases T-cell infiltration and prolongs survival by 56% in amyotrophic lateral sclerosis mice through reduced microglial activation.[56] These findings indicate CB2 physiologically maintains immune homeostasis by countering pathological hyperinflammation, though chronic activation risks impairing anti-microbial or anti-tumor defenses via excessive suppression.[112]Metabolic and Cardiovascular Effects
Cannabinoid receptor 1 (CB1) is expressed in peripheral tissues including adipose, liver, and pancreas, where it modulates lipid and glucose homeostasis. Activation of peripheral CB1 promotes lipogenesis in hepatocytes and adipocytes, contributing to hepatic steatosis and visceral fat accumulation observed in obesity. [113] Endocannabinoid levels, such as anandamide and 2-arachidonoylglycerol, are elevated in obese states, correlating with insulin resistance and dysregulated beta-cell function in pancreatic islets. [114] Selective blockade of peripheral CB1, without central nervous system penetration, reduces body weight, enhances insulin sensitivity, and improves glucose tolerance in rodent models of diet-induced obesity, independent of food intake suppression. [115] Clinical trials of peripherally restricted CB1 antagonists have shown promise in reversing hepatic steatosis and cardiometabolic risk factors in humans with obesity. [116] CB1 signaling also influences appetite and energy expenditure centrally and peripherally. Stimulation of hypothalamic CB1 enhances orexigenic neuropeptides like neuropeptide Y and ghrelin, driving hyperphagia, while antagonism induces sustained weight loss and normalized leptin sensitivity. [117] In diabetes models, CB1 overactivation exacerbates hyperglycemia via impaired incretin secretion and peripheral insulin signaling; genetic knockout or pharmacological inhibition mitigates these effects, reducing fasting glucose and HbA1c equivalents. [118] CB2 receptors play a lesser role in metabolism but may attenuate inflammation-driven insulin resistance in adipose tissue under high-fat diet conditions. [113] In the cardiovascular system, acute activation of CB1 by exogenous cannabinoids like Δ9-tetrahydrocannabinol induces tachycardia, with heart rate increases of 20-50% lasting up to three hours, alongside mild vasodilation and orthostatic hypotension. [119] Endogenous endocannabinoids exert tonic suppression of cardiac contractility and blood pressure via CB1 in brainstem nuclei and vascular endothelium, particularly in hypertensive states, where enhanced CB1 signaling provides cardiodepressor effects. [120] Observational data link frequent cannabis use to elevated risks of myocardial infarction (odds ratio 1.25-4.8 for daily users) and arrhythmias, potentially through prothrombotic shifts and sympathetic activation. [121] [122] CB2 receptors, abundant in immune cells infiltrating vascular and cardiac tissues, mediate anti-inflammatory and cardioprotective responses. Activation of CB2 reduces infarct size in myocardial ischemia-reperfusion models by limiting neutrophil infiltration and oxidative stress, with preclinical evidence supporting its role in attenuating atherosclerosis progression. [123] However, chronic CB1 agonism in cardiomyocytes promotes apoptosis and fibrosis, exacerbating heart failure, while CB2 agonism counters these via prosurvival pathways. [124] Meta-analyses of cannabis exposure yield mixed results on acute coronary events, with some finding no significant association after adjusting for confounders like tobacco co-use, though heavier consumption correlates with adverse outcomes in population studies. [125] The endocannabinoid system thus balances homeostatic regulation against potential risks from dysregulation or exogenous overload. [126]Therapeutic Applications
FDA-Approved Cannabinoid-Based Treatments
The U.S. Food and Drug Administration (FDA) has approved four cannabinoid-based medications, consisting of one purified cannabis-derived compound and three synthetic cannabinoids or analogs that primarily target CB1 and CB2 receptors to exert therapeutic effects. These approvals, dating from 1985 onward, focus on specific indications such as chemotherapy-induced nausea and vomiting, anorexia-associated weight loss in AIDS patients, and seizures in rare epilepsy syndromes, reflecting rigorous clinical evidence of efficacy and safety in controlled trials rather than broader cannabis legalization contexts.[127][128] Epidiolex (cannabidiol oral solution), the only FDA-approved cannabis-derived product, received initial approval on June 25, 2018, for treating seizures associated with Lennox-Gastaut syndrome or Dravet syndrome in patients aged 2 years and older. Its label was expanded on July 31, 2020, to include seizures linked to tuberous sclerosis complex in patients aged 1 year and older, based on randomized, placebo-controlled trials demonstrating reductions in seizure frequency. Cannabidiol modulates cannabinoid receptors indirectly while also interacting with other targets like serotonin and GABA systems, distinguishing it from direct agonists like THC.[129][130] Dronabinol, a synthetic delta-9-tetrahydrocannabinol (THC), is available as Marinol capsules, approved May 31, 1985, for nausea and vomiting from cancer chemotherapy refractory to conventional antiemetics, with a 1992 expansion for anorexia and weight loss in AIDS patients. The liquid formulation Syndros gained approval in July 2016 for AIDS-related anorexia and in March 2017 for chemotherapy-induced nausea, offering bioavailability advantages over capsules in some patients. As a partial agonist at CB1 (primarily in the central nervous system) and CB2 receptors, dronabinol alleviates symptoms through endocannabinoid system activation, though it carries risks of psychoactive effects and dependence.[131][128] Nabilone, marketed as Cesamet capsules, was approved December 27, 1985, for chemotherapy-induced nausea and vomiting unresponsive to standard treatments, supported by trials showing superior antiemetic control compared to placebo. This synthetic THC analog functions as a full CB1 agonist with minimal CB2 activity, providing symptom relief via central mechanisms but with potential for dysphoria and hallucinations at higher doses.[132][127]| Drug | Active Ingredient | Initial FDA Approval | Primary Indications |
|---|---|---|---|
| Epidiolex | Cannabidiol | 2018 | Seizures in Lennox-Gastaut syndrome, Dravet syndrome (expanded 2020 to tuberous sclerosis complex)[129] |
| Marinol | Dronabinol (synthetic THC) | 1985 | Chemotherapy-induced nausea/vomiting; AIDS-related anorexia/weight loss[131] |
| Syndros | Dronabinol (synthetic THC, oral solution) | 2016 (AIDS); 2017 (chemotherapy) | Same as Marinol[128] |
| Cesamet | Nabilone (synthetic THC analog) | 1985 | Chemotherapy-induced nausea/vomiting refractory to conventional antiemetics[132] |