Introduction
Medicinal and recreational cannabis use has increased globally, and continuation of this trend is anticipated as its use becomes legalized internationally.1,2Cannabis sativa is composed of over 100 “cannabinoids,”3,4 but the psychoactive compound delta-9-tetrahydrocannabinol (Δ9-THC), isolated in 1964, and the nonpsychoactive compound cannabidiol (CBD), isolated in 1940,5 represent the most abundant components. Consumption of cannabis products occurs through diverse routes (inhaled smoke, vaping of liquid extracts, resins or waxes, lotions, edibles).6,7 Inhaled cannabinoids are rapidly absorbed in the lungs8 but less so by other routes (e.g., dermal, oral, rectal).9 Due to their highly lipophilic properties, they are stored in adipose tissue for weeks or months and are concentrated in the breast milk of rodents and humans.10,11 CBD products can have beneficial health effects and aid in various medical disorders (e.g., Parkinson’s disease, anxiety, and epilepsy).12,13 Accumulating evidence also indicates there are neurotoxic and reproductive effects from exposure.14–18
Due to increasing cannabis use, exposure to Δ9-THC presents concerning health risks because use will likely also increase in pregnant or breastfeeding women, affecting all stages of brain and neurodevelopment of their offspring.19–24 Along with increased legalization, social acceptance, and use, a change in the ratio of Δ9-THC to CBD in cannabis has also occurred, leading to a change in potency (the Δ9-THC:CBD ratio increased from 14:1 in 1995 to 80:1 in 2014).25 Ultimately, the extent of cannabis neurotoxicity26 is dependent on many variables, including the Δ9-THC exposure level, purity,25 route of administration,7,9,27 developmental age at exposure,23,28–30 health status,31,32 pregnancy status,21,33–36 lactational status,37,38 and others.39 Further, due to the lipophilic nature of these compounds, it has been shown that exposure at low, realistically achievable in-vivo concentrations causes specific molecular targets to be affected, resulting in behavioral or cognitive deficits in those with Δ9THC exposure,39–41 or potential benefits that greatly improve the health of those with neurodegenerative diseases.42–44
In this review, both the risks and benefits of exposure associated with Δ9THC and CBD were investigated. Notably, the risks from CBD exposure, which is usually considered to be safe, are associated with reproductive and developmental health effects.45 Recently, concerns have been raised about CBD use, since it is available in numerous over-the-counter products, with little data supporting its safety or efficacy.46 The side effects and adverse health effects, along with questions regarding the ingredients, are often unknown. On the other hand, Δ9THC exposure has been associated with adverse effects, depending on the dose, yet the benefits of this drug need to be emphasized. These phytocannabinoids were selected because they are the dominant compounds in cannabis, and they are often used as treatments for physical ailments as well as for recreational use. There is a vast amount of literature characterizing these compounds and their effects during development and throughout life in both animal and human studies, but it is important to present the risks as well as the benefits.
The endocannabinoid system (eCBS)
The eCBS was discovered in the 1990s while investigating the mode of action (MOA) of Δ9THC. It is innate and multifaceted, affecting metabolic pathways throughout the body [e.g., muscle, adipose tissue, gastrointestinal tract, liver, and central nervous system (CNS)].47 It helps to shape neuronal connectivity in the brain throughout development and into adulthood,48 affecting the gamma-aminobutyric acid (GABA)ergic, glutamatergic, opioid, and dopaminergic systems.49 Cell membrane-bound cannabinoid-1 receptors (CB1Rs) are the most abundant in the brain, while CB2Rs are mainly expressed on immune cells (T-cells, macrophages) in the periphery or glia/microglia in the brain.47,50 Some researchers have suggested that the transient receptor potential cation channel subfamily V member 1 (TRPV1 or vanilloid receptor 1) could be classified as CB3R, as it is activated by CBD.51 Each receptor type can act independently; however, depending on their location, CB1Rs and CB2Rs (possibly also CB3Rs) can act together, competitively, or in opposite directions, potentially through dimerization to regulate physiological effects.
Normally, neurotransmitters [e.g., glutamate, GABA, serotonin (5-HT), dopamine (DA), acetylcholine (ACh), or norepinephrine] in the CNS are released presynaptically via neuronal stimulation, or by G protein-coupled receptors and voltage-gated ion channel calcium (Ca+2) and potassium (K+) influx.50,52 However, the elevation in postsynaptic Ca+2 affected by neurotransmitters/receptors through the ion channels [e.g., ionotropic glutamate receptors, N-methyl-D-aspartate (NMDA), or GABA],53 stimulates endocannabinoid (eCB) postsynaptic biosynthesis.50,52,54
There are two principal eCB ligands [2-arachidonoylglycerol (2-AG) and anandamide (AEA)], which are synthesized postsynaptically from arachidonic acid by N-acyl phosphatidylethanolamine phospholipase D and diacylglycerol lipase alpha/beta (DAGLα/β), respectively.55–57 These eCBs are produced, as needed,47 postsynaptically by Ca+2-dependent transacyclase and other enzymes, then they migrate from postsynaptic neurons to the presynaptic CBR.53,58 Signaling then occurs as CBR couples to the guanosine-5′-triphosphate (Gi/o)/α-protein subunit dimer58,59 and binds adenyl cyclase to generate cyclic adenosine monophosphate. The cascade decreases presynaptic Ca+2 influx by blocking the activity of voltage-dependent N-, P/Q- and L-type Ca+2 channels60,61 and activation of some K+ channels.53,62 The retrograde eCB (AEA and 2-AG) transmitters in the brain presynaptically inhibit the release of the neurotransmitters GABA,63,64 glutamate,63,65,66 DA,65,67,68 norepinephrine,69 5-HT67,70 and ACh,71,72 thereby decreasing the probability of neurotransmitter release. eCBs are then degraded by the serine hydrolase monoacylglycerol lipase (MAGL) in the presynaptic cell and fatty acid amide hydrolase (FAAH) located in the postsynaptic cell.49,57,73
Figure 1 compares the lipophilic structures of the eCBs (2-AG and AEA) with cannabinoids (e.g., Δ9THC and CBD). Δ9THC and CBD toxicity or neuroprotection depends on factors such as potency, exposure, duration/frequency, vehicle, route of administration, and species-specific differences. Pharmacokinetic and pharmacodynamic parameters determine the extent of P450 (CYP1A, 3A4, 2C9, and 2C19) metabolic activation and glucuronidation elimination of Δ9THC and CBD.9,74 A tipping point leading to an adverse health effect would depend on an individual’s ability to handle various exposure loads based on age, genetic makeup, health status, and diet, among other influences.75,76 These risk factors are often difficult to characterize in humans, since hepatic metabolism studies are, by necessity, generally performed in vitro.75
Δ9THC-associated mechanisms and neurotoxicity
To understand the effects of Δ9THC on the brain, it is helpful to know which areas are affected. The eCBS/CBRs throughout the brain77 help to regulate glutamatergic (excitatory), GABAergic (inhibitory),78,79 dopaminergic, and serotonergic neurotransmitter release at presynaptic terminals.80,81 The interactions among these systems are complex, occurring via direct and indirect stimulation, which may or may not be overseen by the eCBS to regulate neuroplasticity and excitability toward locomotor activity, cognition (learning and memory), executive functions, reward, motivation, and neuroendocrine control, among other functions.78,80,82–86 The striatum in the basal ganglia contains inhibitory GABAergic medium spiny neurons that are affected by the glutamatergic (AMPAR) and dopaminergic (i.e., D1 and D2) receptor inputs from the ventral tegmental area (VTA), substantia nigra (SNc), and prefrontal cortex (PFC).87
Table 1 summarizes some of the main brain regions, pathways, and neurotransmitters involving the neuronal connections in the eCBS and affected by Δ9THC.53,78,80,81,84,85,87–98
Table 1Brain regions and pathways affected by endocannabinoids, Δ9-THC and/or CBD
Neurotransmitter/Pathway | Brain region associations | Behavior/processes involving eCBS | Reference |
---|
Dopamine: DA |
Mesolimbic | DA from ventral tegmental area (VTA; midbrain) → ventral striatum (amygdala, pyriform cortex, lateral septal nuclei, nucleus accumbens) | Reward-related cognition (e.g., incentive: wanting; pleasure: liking; positive reinforcement, associative learning) & emotion | 78,80,81,88–91 |
Mesocortical | DA from VTA (midbrain) → prefrontal cortex + hippocampus | Cognition: executive function (e.g., planning, attention, working memory, planning, self-control, etc.), emotion | |
Nigrostriatal | DA from substantia nigra (pars compacta; substantia nigra SNc: midbrain) → dorsal striatum (i.e., caudate nucleus + putamen) | Neuromotor function, reward-related cognition, associative learning | |
Tuberoinfundibular | DA from the hypothalamic arcuate (infundibular) + paraventricular nucleus → pituitary gland median eminence | Inhibits the release of prolactin. | |
Glutamate |
Glutamatergic | Hippocampus, neocortex and over 90% of synapses in human brain. | Excitatory effects on VTA & SNc neurons, memory, learning, neural communication | 53,90,92,93 |
ɣ-Aminobutyric Acid: GABA |
GABAergic | Hippocampus, thalamus, basal ganglia, hypothalamus, brainstema | Inhibitory effects on VTA and SNc neurons | 90,94–96 |
Serotonin: 5HT |
Serotonergic | Dorsal raphe nuclei, cortex, hippocampus | Modulator of receptors with effects depending on subtype (i.e., biphasic effect on VTA neurons) | 80,84,85,97,98 |
Cannabinoid signaling can be disrupted through agonistic activity of Δ9THC at the CB1Rs throughout areas of the brain. This process leads to inhibition of accumulation of 2-AG and AEA in the brain.73,99,100 While there are many other neuronal circuits associated with the eCBS, the ones mentioned above are most frequently associated with cannabis.
Δ9THC-associated neurotoxicity in rodent and nonhuman primate models
Δ9THC exposure throughout all life stages is associated with effects on behavior, cognition, locomotor activity, birth weight, learning, and other adverse effects.101–104 Cannabis smoke was listed as a reproductive toxicant on 3 January 2020, under California’s Proposition 65.104 However, to control for the dose intake and other technical issues, many neurodevelopmental studies performed in animals used intravenous (i.v.) Δ9THC administration. Although this is not a likely exposure scenario for humans, the immediate absorption by i.v. could be compared to pulmonary exposure by inhalation.105,106 Subcutaneous (s.c.), oral (i.e., gavage), and intraperitoneal (i.p.) administration are more slowly absorbed and are subject to local metabolic processes prior to entering the blood stream.107,108 Other considerations contributing to potential variabilities in evaluating the study results are as follows: 1) often only a single exposure dose was used, limiting potential observations of a dose–response relationship; 2) Δ9THC dosing vehicles varied among studies; 3) different species/strains of rodent were used; 4) different exposure scenarios were used; and 5) many different laboratories contributed to the list of studies.
Gestational exposure to Δ9THC
The eCBS is involved in the earliest developmental stages, including fertilization, implantation, and neuronal progenitors in the brain, leading to migration, morphogenesis, and axonal guidance.94,109,110 The effects of Δ9THC on these processes can be seen in rodents’ pulmonary exposure by inhalation.105,106 Administration via a s.c., oral, or i.p. route is more slowly absorbed and is subject to local metabolic processes prior to entering the blood stream.107,111Δ9THC has profound effects on CB1Rs in areas of the brain regulating GABA, 5-HT, glutamate neurotransmitters, and DA release, influencing, for example, the development of locomotor activity, cognition, learning, memory, and emotional regulation (Table 2).11,29,34,112–141 Notably, the lowest doses of Δ9THC (0.15 mg/kg/day) in the offspring of Long-Evans rats treated in utero affected preproenkephalin, an endogenous opioid precursor in the nucleus accumbens, amygdala, and striatum, in addition to showing evidence of decreased cognition and other behavioral effects.112–114 Treatment in utero or from paternal exposure during a full cycle of sperm development, even at low Δ9THC doses (0.15 mg/kg/day), resulted in developmental deficits and epigenetic transmission.112,113,115–117 Male Wistar adult rats treated throughout sperm development (gavage, 2.0 mg/kg/day) had offspring with affected locomotor activity, feeding behavior, and visual operant signaling.118 Moreover, epidemiological evidence supported findings that cannabis exposure during gestation or during male sperm development results in children with cognitive, motor, and behavioral (including severe psychoses) effects.33,142–145 Infants with gestational exposure to cannabis may show an exaggerated startle response or an inability to adapt to novel stimuli.146,147 Furthermore, women who used cannabis during pregnancy had an increase in fetal deaths, premature births, heart rhythm disorders, and fetal intrauterine growth restrictions.36
Table 2Neurotoxic and behavioral effects from Δ9THC treatment during development in animal studies
Animal strain/Sex/Duration/Dose/Vehicle
| Day tested
| Effects
| LOEL (mg/kg/day)
| Reference
|
---|
Δ9THC in animal studies |
---|
Gestational treatment |
Long-Evans Dam: GD 5-PND 2; F1 fostered PND 2–21. Dose: i.v. 0.15 mg/kg/day. Vehicle: Tween 80/saline | F1 M/F Pups: PND 2 or PND 62, Adult | NAc: ↓striatal DRD2 mRNA expression; ↓DR2 receptor & binding sites; epigenetic regulation of DRD2 mRNA expression disrupted; affected DA receptor gene regulation. Significance: Increase in sensitivity to opiate reward in adulthood | 0.15* | 112 |
Long-Evans Dam: GD 5-PND 2 fostered PND 2-21. Dose: i.v. 0.15 mg/kg/day. Vehicle: Tween 80/saline | F1 M Pups PND 55, Adult | ↓PENK mRNA expression NAc (pup), ↑PENK in NAc & amygdala (adults); ↑Self-administer heroin; ↓latency between active lever press; ↑active lever press; ↑responses on stress test; ↑total responses on active lever on 1st & last extinction days; ↓distance traveled during acquisition & maintenance. Significance: Increased opioid seeking behavior (motivation/reward) & stress response in adulthood | 0.15* | 114 |
Long-Evans Dam every 3rd day; PND 28–49; mated PND 64–68; F1 fostered. Dose: i.p. 1.5 mg/kg/day. Vehicle: saline/Tween 80 | F1 M/F Pups: PND 35 (Adolescence) or PND 62 Adult | Striatal dysregulation of CB1R gene expression, affecting striatal plasticity; ventral to dorsal striatum disruptions between adolescence & adulthood; F ↓novelty seeking. Significance: Supports relevance to age-dependent vulnerability for neuropsychiatric disorders | 1.5* | 130 |
Long-Evans Dam every 3rd day PND 28–49; mated PND 64–68; F1 fostered. Dose: i.p. 1.5 mg/kg/day. Vehicle: saline/Tween 80 | F1 M/F Pups: PND 35 (Adolescence) or PND 62, Adult | Epigenetic effects & altered CB1R mRNA expressions in NAc associated with glutamatergic system regulation; F ↓ locomotor activity. Significance: Cross-generational epigenetic vulnerability to drug abuse | 1.5* | 117 |
Wistar Dam: GD 15–PND 9. Dose: Gavage 3.0 mg/kg/day. Vehicle: sesame oil | F1 M Pup: PND 90, Adult | Disrupted hippocampal GABAergic system; ↓GABA outflow & uptake in hippocampus; ↓ CB1 binding; cognitive impairments. Significance: Long term cognitive deficiency & disrupted GABA neuronal development | 5.0* | 115 |
Wistar Dam: GD 5–14, 16, 18, 21 & PND 1 & 5. Dose: Gavage 5.0 mg/kg/day. Vehicle sesame oil | F1 M/F GD 14, 16, 18, 21 + PND 1 & 5 Neonate | Disrupted tyrosine hydroxylase gene activation (rate limiting in DA production); ↑ DOPACL DA metabolite forebrain. Significance: Tyrosine hydroxylase plays a large part in neurodevelopment through DA production | 5.0* | 131 |
Wistar Dam: GD 7–22. Dose: i.p. 3 mg/kg/day. Vehicle: Not stated | F1 M/F Behavior PND 70–100 | M: ↓Time on light side of test box (↑anxiety); ↑transition to light; ↓Time in open arm of EPM; ↑VTA spike activity; ↓DA & NMDAR2B PND 21; ↑GAD87 PND 21; F: ↑GAD67, vGLUT1-2; PPARα & PPARϒ1-2 & NMDAR2B in the mesolimbic system (VTA-NAc); M/F: ↑Altered fatty acid concentrations in the nucleus accumbens core & shell up to PND 120 (M) or PND 21 (F). Significance: Sex difference with M more affected than F; Fatty acid deficits disrupt the DA/GLUT/GABAergic neurotransmissions affecting neurodevelopment | 3.0* | 132 |
SD Dam: GD 5–PND 2 foster-nursed PND 2–21. Dose: i.v. 0.15 mg/kg/day. Vehicle: Tween80/saline | F1 M/F Pups: PND 22, 45 & 60 Weaning, adolescent, adult | Pup: ↓anxiety: ↓active place avoidance acquisition; ↑active place avoidance reversal phase entries; Adult: ↓attention (acquisition, reversal & distraction) & cognition. Significance: Decreased anxiety, attention & cognitive function | 0.15* | 113 |
SD Dam: Group 1: GD 5–20. Group 2: GD 5–20 + PND 15. Dose: s.c. 2.0 mg/kg/day; PND 15 2.5 mg/kg/day. Vehicle: Tween80/saline | F F1 Pups: Groups 1 & 2: PND 15–28 Juvenile | Group 1 & 2: Male behaviors affected: ↑distance traveled; ↓stretch-attend postures; Group 2: ↓ latency in passive avoidance training; ↑AMPA from DA cells; ↓stretch-attend postures; ↓DA 240 min postacute dose. Significance: Behavioral effects from mesolimbic (NAc) dopaminergic disruptions are greater in males & greater after Δ9THC challenge | 2.0* | 1331 |
SD Dam: GD 5–GD 20. Dose: s.c. 2.0 mg/kg/day. Vehicle Tween 80/saline | F1 M/F Pups: Tests done PND 24–28, Juvenile | VTA DA neuron effects: ↑ firing rate; ↓ cells/track; ↓ spikes/burst, burst rate; ↓after hyperpolarization period; ↑DRD2 sensitivity & acute stress vulnerability; ↑ activity, ↓ PPI average in acute restraint & forced swim test. Significance: Sensorimotor gating deficits leading to an increase in susceptivity to stimuli triggering psychotic-like behaviors | 2.0* | 134 |
SD M Adult 28 days; mated 2 days post dose. Dose: s.c. 2.0, 4.0 mg/kg/day. Vehicle: Tween 80/saline | F1 M Pups: PND 30, 60, 100 & 150 Adolescent, adult | ↓ACh activity; ↑ ChAT: ACh biomarker for number of ACh terminals in striatum; ↓ ChAT hippocampus; ↓HC3/ChAT (ACh activity index) n frontal/parietal cortex & striatum. Significance: Paternal Δ9THC leads to disruptions in developmental trajectory of ACh potentially affecting attention | 2.0 | 116 |
Wild-type Mouse Dam: GD 12.5–16.5. Dose: i.p. 3.0 mg/kg/day. Vehicle: saline/DMSO/Tween 80 | F1 M/F Pups: PND 20; 2 months; Juvenile, adult | CB1R →affected cortical neuron synaptic signaling development →affected connectivity in cortical GABAergic & glutamatergic systems → ↓fine motor skills; ↓ skilled motor function; 2 months: ↓ success in pellet retrieved in skilled steps test; ↑seizure. Significance: Disrupted CB1 signaling leading to disrupted glutamate & GABA signaling leads to increased susceptibility to seizures and cortico-spinal function.in adulthood | 3.0* | 135 |
C57Bl/6 Mouse Dam: GD 14.5–18.5. Dose: i.p. 3.0 mg/kg/day. Vehicle: DMSO | F1 M/F: GD 18.5; PND 10 & 120, Fetal, pup, adult | ↓ CB1R & misrouted hippocampal CB1R afferents, ↑CB1R density in striatum; Impaired LTD in pyramidal cell synapsis; ↓ synaptic plasticity in the cortical circuitry; Impaired cortical axonal development; ↓2-AG signaling, ↓ CB1R & ↑ MAGL expression, ↓DAGL; abnormal growth cones & cytoskeleton in axonal region. Significance: Abnormal axonal development in growth cone disrupts neuronal circuitry, memory encoding, cognition & executive skills | 3.0* | 136 |
Postnatal Treatment |
Wistar Dam: PND 1–10. Dose: s.c.2.0 mg/kg/day. Vehicle: DMSO/cremophore/saline | F1 M/F Pups: PND 10, 15, 20; 9–21 Preweaning, juvenile | ↓Bodyweight gain; GABA excitatory to inhibitory switch in PFC (eCB disruption); ↓upregulation & expression of KCC2 (K+ transporter), Vocalizations ↑in frequency (kHZ). Significance: Delayed development of GABA switch leads to sensorimotor gating deficits, potential autism, epilepsies, schizophrenia-like behavior. | 2* | 11 |
Wistar M Adult: 12 days mated to untreated F. Dose: Gavage 2.0 mg/kg/day. Vehicle: EtOH/TritonX100/saline | F1 M/F Pups: PND 28–140, Adolescent, adult | ↑Habituation of locomotor activity, Novelty suppressed feeding: ↓latency to begin eating; ↓Visual operant signal. Significance: Impaired operant attention into adulthood | 2* | 118 |
SD Juvenile M/F: PND 10–16. Dose: Gavage 10 mg/kg/day. Vehicle: corn oil | F1 M/F Pups: PND 29 & 38, Adolescent | ↓Bodyweight gain; High Illumination: ↑entries & time in open arm; Low Illumination: ↑stretch attend posture; ↑head dips; ↓ exploration, ↑frequency of nape attacks; ↑time & frequency play fighting. Significance: Altered social behavior in adolescence. | 10* | 137 |
C57BL/6J Mice M Pup: PND 5–16 & 5–35. Dose: s.c. 1.0, 5.0 mg/kg/day. Vehicle not stated | F1 M Pup: PND 16 or PND 35 Preweaning, adolescent | Hippocampal cell rearranged CB1R; changes key molecular constituents of mitochondrial respiratory chain; Thinning of pyramidal cell layer; Neurochemical deficits Significance: Developmental deficits from neuronal disorganization, misrouted differentiation & associated pathologies. | 1.0 | 119 |
Adolescent Treatment |
Long-Evans M PND 28 each 3rd day to PND 50. Dose: i.p. 1.5 mg/kg/day. Vehicle: saline/H2O/Tween80 | M: PND 50 or PND 63, Adolescent, adult | Adolescent: Disrupted development of dendritic arbors PFC (pyramidal neurons); Adult: prolonged atrophy in distal apical arbors of PFC neurons; Prematurely pruned dendritic spines attenuated neuroplasticity. Significance: Disrupted PFC neural networks lead to decreased cognitive & emotional dysregulation & affected decision making similar to pathology in human schizophrenia | 1.5* | 29 |
Long-Evans F PND 35–75. Dose: i.p. 5.6 mg/kg/day. Vehicle: saline | F: PND 75–160 & 159 to 200 Adult | Adult: ↑CB1R density; Persistent impairment of working memory & task performance. Significance: Long term effects on operant learning | 5.6 | 123 |
Long-Evans M/F “Puberty Onset” for 14 days. Dose: i.p. 5 mg/kg/day. Vehicle: EtOH/Cremophor/saline | M/F: Day 14 treatment | M/F combined: ↓Total attacks, total pins, percent defense & complete rotation. | 5.0 (only dose) | 138 |
Wistar M i.p. 1.0 mg/kg/day PND 28–30 →5.0 mg/kg/day alternate days PND 34–52 or PND 60–62 →5.0 mg/kg/day alternate days PND 66-84 or Acute: 5 mg/kg/day: PND 52. Vehicle: Tween80/saline | M: PND 52, 55, 67, 70, 71, 72, 84, 87, 99, 102, 103, 104, Adolescent, adult | Adolescent: ↑Latency to emerge; ↓time in open areas; ↓rearings; ↓novel object preference; ↑memory deficits; alterations in hippocampal structure/function remaining to adulthood. Significance: Hippocampal alterations lead to persistent memory deficits that developed in adolescence | 1.0 | 120 |
SD M/F PND 35–37; 5; 38–41; 10; 42–45. Dose: i.p. 2.5 mg/kg/day, twice/day. Vehicle: EtOH/cremophor/saline | M/F: PND 75: Adult | Adult: ↓ Bodyweight & food intake; ↓CB1R binding & stimulation (NAc, amygdala, VTA, hippocampus); ↓sucrose preference (anhedonia); ↓CREB activation in prefrontal cortex, NAc, hippocampus; ↑ dynorphin (indicates depression). Significance: Disruption of neural circuitry related to emotion and depression during adolescence | 5.0 | 34 |
SD M PND 35–37; 5; 38–41; 10; 42–45. Dose: i.p. 2.5 mg/kg/day, twice/day. Vehicle: EtOH/cremophor/saline | M: PND 75 Adult | Adult: ↓Radial maze learning; ↓dendritic length in hippocampal dentate gyrus; ↓spine density; ↓NMDA receptors & biomarkers indicating ↓neuroplasticity. Significance: Spatial memory & cognitive deficits | 5.0 | 122 |
CD1 Mice M PND 28–48; 69–89. Dose: i.p. 3.0 mg/kg/day. Vehicle EtOH/chermophor/saline | PND 49–53 & PND 90–94 Adolescent. PND 90–94 & PND 131–135 Adult | Adolescent: Impaired object recognition/working memory (novel object recognition & discrimination); repetitive/compulsive behaviors (↑percent shredded in nestlet; ↑marble burying); ↓delayed anxiety to move out of the dark; Adult: ↓novel object recognition performance; elevated plus maze ↓ anxiety to venture out. Significance: Behaviors common to those seen in animal schizophrenia models & humans | 3.0* | 121 |
Adult treatment |
Long-Evans M. Dose: Acute i.p. 1.0, 1.5, 2.0 mg/kg. Vehicle: detergent/EtOH/saline | ∼15 min time increments postdose | ↓Attention; ↓hippocampal functional cell types. Significance: Information not likely to be encoded correctly & unlikely to be accurately retrieved or recalled | 0.5 | 127 |
Long-Evans M. Dose: Acute i.p. 0.01, 1.0 mg/kg. Vehicle: Tween80/saline | 30 min postdose | ↑Trials to achieving reversal task between stimulus & reward; affects c-fos expression associated with negative behavioral effects (orbital limbic & striatal regions in brain). Significance: Effects in orbitofrontal cortex & striatum (potential inelasticity) leading to an inability to perform reversal discriminations | 1.0 | 124 |
Wistar M: 5 days. Dose: i.p. 2.0, 4.0 mg/kg/day. Vehicle: Tween 80/saline | 30 min postdose | ↓Short-term memory & discrimination affected by eCB increase at the CB1R. Significance: Disrupted CB1Rs is detrimental to memory & cognition | 2.0 | 139 |
Wistar M: i.p. 7 days per dose. Dose: i.p. 1.0, 3.0, 10 mg/kg/day. Vehicle: EtOH/Tween 80/saline | 20 min postdose | ↓ Body weight; Anxiety measures: ↓time spent in emergence test; ↑ hide time; ↓ open field time; ↓percent open arm time; ↓active time; ↓total social interaction time & distance traveled; Place conditioning: ↓preference for the conditioned side; ↓CB1 R binding in hippocampus, substantia nigra, caudate putamen, cigulate gyrus. Significance: Affected anxiety, learning, memory & social interaction due to disruptions in CB1R binding in critical brain regions | 1.0, 3.0, 10 | 129 |
SD M 2 times/day for 14 days. Dose: i.p. 5.0 mg/kg twice per day. Vehicle: Tween 80/saline | Post terminal dose | ↓Performance attention, executive functions, memory, cognition associated with ↓DA in PFC. Significance: Disruption of the cortical dopaminergic pathways lead to cognitive & attention dysfunction | 20 | 140 |
SD M: Acute (1 treatment). Dose: i.p. 5.0 mg/kg. Vehicle: OH-β-cyclodextrin/saline | 30 min postdose | ↑Working memory impairments; ↑DA turnover (DOPAC/DA); ↑NE turnover PFC. Significance: Cognitive impairment | 5.0* | 141 |
C57BL/6JArc mice M: 1 or 21 days. Acute & Chronic Doses: i.p. 0.3, 1.0, 3.0, 10 mg/kg. Vehicle: EtOH/Tween 80/saline | Acute & chronic 60 min postdose | Acute & chronic: ↑analgesia & catalepsy; ↓ thermic response & locomotor activity; Anxiety: ↓ distance traveled light/dark; ↓ frequency of entries in elevated + maze; ↓ vertical activity, rearing & head dipping; ↓startle response; ↓passive avoidance/anogenital sniffing, social interaction; ↑latency passive avoidance; ↑prepulse inhibition. Significance: Effect on neurotoxicity (anxiety) occurs after both acute & chronic exposure | 1.0 | 126 |
CD1 mice M. Dose: Acute i.p. 0.2, 0.4, 0.8, 1.6, 3.2, 6.4, 12, 48 mg/kg/day. Vehicle: EtOH/CremophorEL/saline | 30 min postdose | ↑Percent time in the open arm in the elevated plus maze; ↓anxiety; ↑percent swim time; ↑closed arm entries. Significance: ↓Anxiety & depression behaviors | 0.8 | 125 |
In support of the gestational exposure findings, a meta-analysis was performed on the behavioral effects in animal offspring exposed to Δ9THC during gestation and lactation.148 A compilation and meta-analysis of behavior in offspring from 15 selected studies in Long-Evans, Sprague-Dawley, or Wistar female and/or male rats exposed from mothers exposed via oral, i.v., or s.c. administration indicated significant effects on cognitive, locomotor, and emotional behavior.
Postnatal exposure to Δ9THC
Postnatal exposures to young C57BL/6J male mouse pups resulted in behavioral effects from Δ9THC treatment at 1.0 mg/kg/day administered s.c.119 This and other studies performed in male and female Wistar rats at 2.0 mg/kg/day (s.c.)11 or 10 mg/kg/day (gavage)149 included effects on anxiety and neurodevelopmental deficits similar to those seen in autism, epilepsy, and schizophrenia.11,119,149 Perinatal exposure in children would likely be from nursing, secondhand smoke, or accidental ingestion, causing long-term effects.37,150 The transfer of cannabis in the milk to nursing babies was shown to affect DA receptors, resulting in hyperactivity, poor coordination, and cognitive function and leading to an increased risk of future drug abuse.37,150 For example, GABA is primarily excitatory in early development and then it switches to inhibitory postnatally. Disruption of this process in humans may result in neurodevelopmental patterns affecting chronic pain, neuroplasticity, and psychiatric diseases (e.g., autism, epilepsy, and schizophrenia).11,151 Data indicate that perinatal cannabis exposure increases the risk of future drug use.152
Adolescent exposure to Δ9THC
Exposure to Δ9THC i.p. in Long-Evans and Wistar male rats throughout adolescence at low doses (1 or 1.5 mg/kg/day) showed disrupted neural development in the PFC and hippocampus resulting in effects on neuroplasticity, cognition, social interactions, memory and others.29,120 Similar effects (i.e., increased: CB1R density, anxiety, learning deficits, anhedonia) were observed at higher doses (2.5–10 mg/kg/day) in Long-Evans females, male and female Sprague-Dawley rats and male CD1 mice receiving various dosing regimens (Table 2).120–123,153Δ9THC treatment in adolescents disrupted development of brain areas (e.g., PFC) associated with adverse behaviors like schizophrenia in humans, which often occurs in adolescence.155 Adolescence is a stage of peak eCB (2AG and AEA) and CB1R expression.156 The brain is still developing and is at heightened risk for disruption of normal neurodevelopmental processes.157,158 Where pre- or postnatal exposures may be involuntary in developing young, adolescence is where preteens and teens may begin to experiment with cannabis on their own.159 Vaping cannabis has become one of the most preferred methods of consumption, that will not only increase the concentration of Δ9THC but also potentially increase exposure to residues of pesticides used on cannabis crops.99,160–162 Cannabis use in adolescents greatly increases the risk of psychosis by 3–4-fold and has been shown to lower the age of schizophrenia onset.163,164 Further, adolescent cannabis use will increase the probability of future drug use,165 as shown by evidence from animal and epidemiological studies.152,166,167
Adult exposure to Δ9THC
Acute adult effects in Long-Evans male rats as well as C57BL/6Arc and CD1 male mice showed behavioral effects (attention and learning, decreased anxiety and locomotor activity) at low Δ9THC doses (i.p.: 0.25, 0.8, or 1.0 mg/kg/day; Table 2).124–127 Notably, these studies used 2–8 treatment levels and could therefore establish a dose–response relationship. C57BL/6J male mice treated at 10 mg/kg/day also experienced a decreased thermic response and increased catalepsy and analgesia.126 This study demonstrated the “cannabinoid tetrad”: increased catalepsy, hypomobility, hypothermia, and antinociception.128 At the low acute doses, the animals showed decreased anxiety; but at higher doses graduating from 1 to 3 to 10 mg/kg/day at 7-day intervals, the animals had increased anxiety measures with both acute and chronic exposures in male Wistar rats.129 Human studies also showed that cannabis use versus nonuse was associated with an earlier onset of psychoses, death by suicide, depression, mania, anhedonia, cognitive deficits, and anxiety/paranoia as well as brain effects (decreases in glutamine, affected DA, and decreased hippocampal volume (systematic review).168 This review also reported associated harmful effects of exposure on driving, stroke, pulmonary function, vision, and negative drug-drug interactions. With cannabis legalization, it is likely that there will be more health-related deficits and an increased need for public and clinical policy changes. Table 2 lists the lowest-observed-effect levels (LOELs) reported from each in-vivo study (mg/kg/day).
Nonhuman primate exposure to Δ9THC
Studies in nonhuman primates have been performed in pregnant animals. Rhesus macaques were fed Δ9THC in a cookie at 2.5 g/7 kg at gestation day (GD) 0–155.169 There were decreases in the amniotic fluid volume throughout pregnancy and decreased placental perfusion (oxygen availability decreases) accompanied by increased placental microinfarctions. In addition, there were significant changes in the RNA signature sequences in the placental transcriptome. These data indicate that disruptions in vascular development and angiogenesis affect the offspring through decreased testes weights and relative heart weights. Adult male rhesus macaques were treated with Δ9THC in a cookie at 0.5 mg/7 kg/day (1–70 days), 1.0 mg/7 kg/day (71–140 days), and 2.5 mg/7 kg/day (141–210 days). At 210 days, there were dose-related decreases in testicular and epididymal weights.170 Follicle-stimulating hormone, luteinizing hormone, and prolactin were increased, and total testosterone and estradiol were decreased. These effects indicate potential disruption of the hypothalamus–pituitary–gonadotropin axis, impacting testicular function.171 In another study, adult female rhesus macaques were treated with Δ9THC in a cookie at 0.5 mg/7 kg/day (1–3 weeks), 1.0 mg/7 kg/day (4–6 weeks), 2.0 mg/7 kg/day (7–9 weeks), and 2.5 mg/7 kg/day (10–12 weeks). At 12 weeks, the animals showed increases in menstrual cycle length and increased follicle-stimulating hormone concentrations, another indication of hypothalamus–pituitary–gonadotropin axis disruption.171 The disruptions in hormonal balance, menstrual cycle, and ovulatory function would likely affect fecundity.172
Δ9THC-associated effects in humans
A review by Frau and Melis173 provides evidence showing that in utero, transplacental Δ9THC exposure deregulates the mesolimbic dopaminergic system in males, potentially predisposing them to schizophrenia. Prenatal exposure in humans can act to prime the sensorimotor gating development in the brain, primarily in the VTA region associated with the dopaminergic system. Subsequent environmental exposures such as Δ9THC or other stressors can lower the threshold to initiation of psychotic-like effects.134 In addition, Δ9THC exposure to infants during breastfeeding can continue more than 6 weeks after the last maternal consumption, potentially affecting brain development.9,38,142,174 Monfort, Ferreira, Leclair, and Lodygensky22 have described the pharmacokinetics of cannabinoid exposures during pregnancy, in infants, and during breastfeeding. While consumption may be due to depression, anxiety, nausea, or pain, data indicate that there are significant irreversible risks to neuronal development in fetuses, neonates, and the developing young.22 Data also support the increased risks of dysregulated glucose-insulin measurements as well as obesity in children after maternal use of cannabis during pregnancy.175
Although Δ9THC (cannabis) is not federally legal in the United States, acute and repeated human exposure to Δ9THC is regulated by the European Food Safety Authority.176 Human data were used by this agency to establish a lowest-observed-adverse-effect level (LOAEL) for an administered Δ9THC exposure of 2.5 mg/kg/day (corresponding to an internal dose of 0.036 mg/kg/day). Applying an uncertainty factor of 3 to extrapolate from a LOAEL to a no-observed-adverse-effect level (NOAEL) and 10 for intraspecies differences produced 1 µg/kg/day (acute reference dose: ARfD = [0.036 mg/kg/day ÷ 30] = 1 µg/kg/day). However, it is evident from gestational treatment in Table 2 that offspring experienced neurodevelopmental effects related to motivation/reward, stress response, and increased sensitivity to opiate reward in adulthood at 0.15 mg/kg/day.112,114 Establishing an ARfD would require the same uncertainty factors in addition to an interspecies default of 10 ([LOAEL 0.15 mg/kg/day ÷ 3 = NOAEL 0.05 mg/kg/day] ÷ [10 interspecies × 10 intraspecies]) = 0.5 µg/kg/day.177–179 Gestational exposure to Δ9THC may need a different ARfD than that of adults, since effects occur at very low doses. This is especially critical to re-evaluate because the low-dose animal studies used only one dose, and there were no doses below 0.15 mg/kg/day in which effects might also be seen in developing fetuses.
CBD-associated mechanisms
While it can make up as much as 40% of cannabis extract,180 CBD has been purified in products for use by people and even their pets. CBD is one of the most actively studied therapies for a broad spectrum of neurological, inflammatory, and mental diseases because of its efficacy, low toxicity, and availability (e.g., over the counter and online order). The exact mechanism for the therapeutic effects is still under investigation,42,181 but the proposed MOA for CBD indicates several targets associated with neuroprotection (Fig. 2182 and Table 3).86,183,184 Like Δ9THC, CBD has effects on many interacting targets, and there is evidence for direct and indirect CBD actions on inflammatory and neurological parameters.180
Table 3In-vivo and in-vitro examples of neuroprotective effects of CBD in different neurological diseases183
Model | CBD dose | Treatment | Biological/pharmacological effect | Neurological disease |
---|
Neuroprotection through activation of A2ARs |
SJL/J mice: F | 5.0 mg/kg, i.p. | Days 1–7 post infection | Microglia activation attenuated, downregulating the expression of VCAM1, CCL2 and CCL5 & proinflammatory cytokine IL1β. CBD improved motor deficits in the chronic phase of the disease | Multiple sclerosis |
Newborn C57BL6 mice: M/F | 0.1–1,000 µM | 15 min pre-incubation | ↓Acute brain damage & apoptosis; ↓glutamate concentration, IL6 & expression of TNFα, COX2, and iNOS | Hypoxic-ischemic brain damage |
Primary rat microglial & N13 microglial cells & C57Bl/6 mice: M/F | 20 mg/kg, i.v. | 1/day for 7 days; 3 days/week for 2 weeks | Inhibited ATP-induced intracellular Ca2+ increase in cultured N13 & primary microglial cells and A2A receptors may be involved in this mechanism. In vivo: ↓gene expression of proinflammatory cytokine IL6 & prevented cognitive impairment induced by βA | Alzheimer’s disease |
Sabra mice: F | 5.0 mg/kg, i.p. | 28 days | ↓Hippocampal TNFα-R 1 gene expression but ↑expression of the BDNF gene. Indirect activation of A2AR, ↑cognitive & motor function in rats with hepatic encephalopathy. | Hepatic encephalopathy |
Neuroprotection through the activation of the 5-HT1A |
MCA occlusion mice: M | 3.0 or 10 mg/kg, i.p. | Before & 3 h after damage | CBD significantly ↓infarct volume induced by MCA occlusion through 5-HT1A receptor | Cerebral ischemia |
Swiss mice: M | 5.0, 15, 30, or 60 mg/kg, i.p. | 30 min before receiving drugs to induce catalepsy | CBD pretreatment ↓catalepsy in a dose-dependent manner, through the 5-HT1AR | Striatal disorders |
Swiss mice: M | 15–60 mg/kg or 60 nmol, i.p. | 30 min before or 2.5 h after receiving the drugs to induce catalepsy | CBD pretreatment ↓catalepsy in a dose-dependent manner, through the 5-HT1AR | Striatal disorders |
Wistar Kyoto rats: M | 100 mg/kg | 60 min before seizure induction | CBD significantly mitigated PTZ-induced seizure | Seizure disorders |
Adult Wistar rats: M | 0.1–1.0 mg/kg & 5.0 mg/kg, i.p. | Acute treatment + cumulative injections every 5 min & repeated at 5 mg/kg/day for 7 days | CBD protected nerve injury-induced deficits in dorsal raphe nucleus 5-HT neuronal activity & exerted antiallodynic effects by TRPV1 activation & anxiolytic properties through 5-HT1A receptor activation | Allodynia & anxiety |
Sabra mice: F | 5.0 mg/kg, i.p. | 28 days | CBD, by 5-HT1AR activation, ↑cognition & motor function, impaired by bile-duct ligation. CBD ↓neuroinflammation, ↑BDNF gene expression & ↓TNFαR 1 gene expression in hepatic encephalopathy model | Hepatic Encephalopathy |
Sabra mice: F | 5.0 mg/kg, i.p. | Single acute dose | CBD ameliorated cognitive deficits & locomotor activity; restored brain 5-HT levels & improved liver function | Hepatic Encephalopathy |
C57BL/6J mice: M | 30 mg/kg/day, i.p. | 7 days | ↑Time spent interacting; ↓psychotic-like behaviors acting through 5-HT 1A receptors | Schizophrenia |
Neuroprotection by antagonistic activation of GPR55 |
Scn1a mutant mice (DS model): M/F | 10, 20, 100, or 200 mg/kg/day | Twice/day for 7 days | Acute CBD ↓thermally induced seizures & ↓spontaneous seizure rate. Low doses ameliorated autism-type social interaction deficits in genetically induced DS model, ↑GABA inhibitory transmission impaired in DS mediated by GPR55 | DS |
Adult C57BL/6 mice: M | 5.0 mg/kg | 5 days/week, 5 weeks | ↓Density of microglial cells in the cell body. In the haloperidol-induced catalepsy model, through GPR55-activation. | Parkinson’s disease |
C67BL/6 mice M/F | 5.0–10 & 50 mg/kg | Increasing doses from 5.0 to 10 mg/kg 3 times/week, or daily, at 50 mg/kg, for 23 days | EAE disease ameliorated (all doses), ↓encephalitogenic cell vitality, ↓levels of IL6, production of ROS, ↓apoptosis & GPR55R in CNS | EAE disease |
Neuroprotection through activation of the TRPV receptors |
Wistar rat: M | 10 mg/kg, i.p. | 2 h after the induction of model | CBD inhibited carrageenan-induced hyperalgesia by desensitization of TRPV1R | Hyperalgesia |
hPBMECs & hCMEC/D3 Cellsa | 0.1, 0.3, 1.0, 3.0, 10, 15 µM | 7 or 24 h of incubation | Dose-related ↑in intracellular Ca2+ through activation of TRPV2 enhanced cell proliferation, cell migration & tubulogenesis in human brain endothelial cells. | — |
U87MG Human glioblastoma cell line | 10 µM | Cells treated with different CBD doses 1 day or co-treated with CBD 10 µM & chemo drugs 6 h | TRPV2 activation & ↑Ca2+ improved chemotherapy drug action by enhancing absorption & ameliorating cytotoxic activity in human glioma cells | — |
Human gingival mesenchymal stem cells | 5 µM | 24-h incubation | TRPV1 desensitization promoted the PI3K/Akt pathwayb signaling, which can reduce Alzheimer hallmarks | Alzheimer’s disease |
Neuroprotection through the activation of the PPARɣ |
SH-SY5YAPP+ cells | 10−9–10−6 M | 24 h | ↓Expression of amyloid precursor protein & its ubiquitination, leading to ↓Aβ & neuronal apoptosis. Effects mediated by PPARɣ activation | Alzheimer’s disease |
Primary rat astrocytes & SD rat: M | 10−9–10−7 M: in vitro; 10 mg/kg: in vivo, i.p. | 15 days | In vitro: Dose response ↓ in Aβ mediated through inhibition of NF-кB; Aβ-induced neuronal damage led to ↓gliosis & glial fibrillary acidic protein. Effects exerted through PPARɣ activation | Alzheimer’s disease |
Hippocampal slices from C57Bl/6 mice | 10 µM | 30 min before addition of Aβ | Improved synaptic transmission & long-term potentiation in the hippocampus slice of C57BL/6 mice, protecting it from cognitive deficits induced by Aβ 1–42. CBD effects exerted through interaction with PPARɣ | Alzheimer’s disease |
Newborn C57/BL6 & Swiss mice primary microglial cultures: M/F | 60 mg/kg: in vivo, i.p.; 10 µM: in vitro | 2 injections/day 30 min prior to haloperidol: 21 days | Dyskinesia prevented after induction haloperidol. ↓Oxidative stress in corpus striatum, ↓activation of microglial, inflammatory cytokine (e.g., IL1β and TNFα), ↑anti-inflammatory cytokine IL10. CBD affects PPARɣ actions on lipopolysaccharide-stimulated microglial cells | Tardive dyskinesia |
Adult C57/BL6 mice: M | 15, 30, or 60 mg/kg, i.p. | 15 min before L-DOPA administration for 3 days | CBD did not prevent L-DOPA-induced dyskinesia. Cotreatment of CBD + capsazepine, acting through CB1R & PPARɣ, ameliorated dyskinesia. | Parkinson’s disease |
Human brain microvascular endothelial cell/human astrocyte co-cultures | 100 nM, 1.0 & 10 µM | Before or directly after induction of ischemic damage | 10 µM prevented enhanced BBB permeability after ischemic damage induced by oxygen-glucose deprivation, through by activating PPARɣ & 5HT1AR. | Ischemic stroke |
Neuroprotection through positive allosteric modulation of GABAA receptors |
Surgical human DS & TSC cortical tissue in Xenopus oocytes | 5.0 µM | Pre-incubation of cells 10 s before co-application of GABA & CBD | Positive modulation of GABAAR, ↑amplitude of GABA-evoked current in brain tissues of patients with DS &TSC | DS & TSC |
Scn1a+/- mice (M/F) & Xenopus oocytes expressing GABAA receptors | in vivo 12 or 100 mg/kg, i.p.; in vitro 10 µM | In vivo: CBD administered i.p. 45 min before CLB. In vitro: CBD (10 µM) co-applied with GABA, for 60 s | ↑CLB concentration & active metabolite N-CLB in plasma & brain. Co-administration ↑anticonvulsant effect by enhancing the activity of the GABAA receptor | DS |
Glial cells
CNS connective tissue (e.g., macroglia: astrocytes and microglia) consists of nonneuronal cells that link neuronal cells to the blood supply (blood–brain barrier), regulate blood flow to the brain, and regulate neurotransmission (macroglia) or serve as macrophages to mount immune responses in the brain (microglia).185 When neuronal injury occurs, astrocytes can signal microglia to initiate an immune response; however, when the immune response becomes unbalanced, neuronal injury will occur.181 CBD can decrease the microglial immune response to injured dopaminergic neurons in diseases like Parkinson’s disease, and it increases the recruitment of astrocytes to promote neuronal regeneration through brain-derived neurotrophic factor (BDNF) (Table 3).
Adenosine receptor 2A (A2AR)
Adenosine acts at a G protein-coupled receptor (A2AR) on neuronal membranes to suppress immune responses due to inflammation or cell stress. CBD serves as an agonist to decrease adenosine reuptake, thereby increasing adenosine signaling and decreasing neuroinflammation.186,187 CBD exposure decreases proinflammatory cytokine interleukin (IL)1β, microglial activity, tumor necrosis factor-alpha (TNFα), cyclooxygenase-2 (COX2), and inducible nitric oxide synthase (iNOS) activity in the brain (Table 3). These pathways have been shown to improve the effects of multiple sclerosis, hypoxic-ischemic brain damage, Alzheimer’s disease, and hepatic encephalopathy.183
5-HT receptors
The dorsal raphe nucleus (DRN) is the primary serotonergic center (5-HT) in the brain where GCPR 5-HT1A receptors are expressed. Receptor stimulation inhibits voltage-gated Ca2+ channels, activates K+ channels, and inhibits neurotransmission in the DRN.44,188 CBD has an anxiolytic effect by acting through 5-HT1A receptor in male Wistar rats, previously stressed by foot shocks or restraint, but it can also induce anxiogenic behaviors in rats experiencing contextual fear conditioning,182–190 perhaps by serving as an agonist at the 5-HT1A receptor.188 Acting on the serotonergic system, CBD is associated with improved locomotor activity (after striatal damage), cognition, cerebral ischemia, seizure disorders, and hepatic encephalopathy (Table 3).183 Through the 5-HT1A receptor, CBD is associated with antiepileptic, anticataleptic, neuroprotective, antiemetic, anxiolytic, antidepressant, antipsychotic, and analgesic effects.86,191–195 Others have also indicated that CBD acts via a negative allosteric mechanism in DRN somatodendritic 5-HT1A receptors that does not require CB1, 5-HT2A, or GABAA receptors.86,186
CB1Rs and CB2Rs
CBD at the CB1Rs regulate excitotoxicity by inhibiting glutamate release to the NMDA receptors and normalizing glutamatergic activity. CBD acts to increase the blood supply to areas after ischemic incidents by decreasing endothelial-derived endothelin-1 or nitric oxide to increase vasoconstriction.197 Neurodegeneration occurs with activation of microglial cells (immune cells in the brain); however, CB1R activation by CBD leads to a decrease of TNFα and IL12 and an increase of IL10. Activation of CB2 then decreases the proliferation and migration of microglial cells while decreasing TNFα by inhibiting nuclear factor kappa-light-chain-enhancer of activated B cells (NFкB; Table 3).198,199 The anti-inflammatory action of CBD has been shown to improve neuronal damage from ischemic stroke, Tardive dyskinesia, and Parkinson’s disease.
FAAH
CBD can act indirectly at the CB1R through inhibition of FAAH and the AEA transporter, leading to increased AEA and activation of CB1R.200,201 Increased CB1R agonism leads to decreased eCB degradation and transport (Table 3).
TRPV1
TRPV consists of a vanilloid channel on the plasma membrane, considered by some to be a CB3R,51 that induces neuropeptide release associated with pain perception, neuroinflammation, and body temperature regulation.200 CBD at TRPV-1 channels leads to increased Ca2+ levels, resulting in desensitization and subsequent decreased pain. TRPV1 binding decreases microglial activation and migration as well as oxidative stress (Table 3). In addition, CBD can increase AEA levels by inhibition of FAAH.202 However, AEA and CBD are both TRPV1 channel agonists. TRPV1 channel activation by CBD presynaptically increases glutamate release in the brain, which may serve to counteract/antagonize the inhibitory action of CB1R binding by CBD on colocalized glutamatergic neurons. TRPV1 activation by CBD agonism can increase the PI3K/Akt pathway signaling to decrease the incidence of hallmarks of Alzheimer’s disease.
G-coupled protein receptor 55 (GPR55)
GPR55 binding protects against excitotoxicity potentially through GABAA receptor. CBD, as an antagonist, decreases GPR55 activation in the CNS to regulate such processes as neuropathic pain and antiepileptic activity.203 CBD has a high affinity for GPR55, resulting in a decreased glutamate release in the hippocampus, thus causing anti-convulsive effects, also seen in human subjects.180 Moreover, the use of CBD has been shown to result in improved Parkinson’s disease and Dravet syndrome (DS) symptoms (Table 3).183,204
Peroxisome proliferator-activated receptor gamma (PPARɣ) receptors
CBD is an agonist of PPARɣ, a nuclear receptor and ligand-inducible transcription factor that produces anti-inflammatory and antioxidative effects.199 PPARɣ modulates inflammation by inducing ubiquitin-proteasomal degradation of p65, resulting in inhibition of proinflammatory gene expression of cyclooxygenase (COX2) and proinflammatory mediators (e.g., TNFα, IL1β, and IL6) in addition to inhibition of NFкB-mediated inflammatory signaling. CBD agonist activity with PPARɣ also contributes to the inhibition of TNFα, IL1β, and IL6 transcription to prevent NFкB signaling, and it also produces antioxidant properties.198,199 It increases eCBs by antagonist activity at CB2Rs, and the eCBs then act as PPARɣ agonists to promote anti-inflammatory and antioxidant actions. Furthermore, Alzheimer’s disease has been demonstrated to be improved via the PPARɣ-mediated protective effects of CBD (Table 3).
GABAA receptors
As the main inhibitory neurotransmitter in the CNS, GABA disruption is associated with neurological diseases, including cognitive deficits, drug addiction, chronic stress and anxiety, epileptic disorders, and Huntington’s disease.180,205 CBD stimulates GABAergic neurotransmission, meaning that the inhibitory neurotransmission and frequency are increased.206 Seizure frequency, duration, and severity were reduced in addition to increased social behaviors in a mouse model of DS and other diseases after CBD treatment. In addition, overexcitation in the dentate gyrus of the hippocampus was decreased through CBD effects on GABAA receptors.206 Therefore, with CBD bound to the GABAA receptor, anticonvulsant and anxiolytic actions are seen in the CNS. Moreover, since CBD does not bind competitively with the benzodiazepine receptor, it is potentially useful in patients resistant to benzodiazepines, which is the standard antiseizure treatment (Table 3).207
CBD-associated neuroprotection in animal studies
CBD has shown neuroprotective effects in animal models with several neural-associated disease states (Table 3).86,181,183,208 The areas studied have focused mainly on neuroprotection and treatment of brain-related diseases (e.g., multiple sclerosis, Alzheimer’s disease, and schizophrenia), rather than effects on other areas of the body (e.g., local pain). CBD at doses from 5.0 mg/kg/day in rodents has many beneficial effects (Table 3). Note that doses administered in vivo were by i.p.; therefore, CBD is more slowly absorbed and subject to local metabolic processes prior to entering the blood stream, as would occur with oral exposure.107,111Table 3 indicates pathways specifically shown to be associated with CBD exposure.
CBD neuroprotection in human studies
The neuroprotective effects of CBD observed in animal studies are supported by observations in human subjects. CBD is well tolerated in children and adults and has a broad spectrum of therapeutic benefits to help with significant neurological disease states,209 including neurological damage and disorders, brain tumors, Parkinson’s disease, Huntington’s disease, Alzheimer’s disease, multiple sclerosis, neuropathic pain, and childhood seizures (e.g., Lennox-Gastaut syndrome and DS).180,210 Additionally, synthetic forms of CBD have been used to treat drug-resistant epilepsies in children (age ≥2 and older) (Lennox-Gastaut syndrome or DS).210 Epidiolex/Epidyolex (>99% CBD) is approved by the United States Food and Drug Administration and the European Medicines Agency to treat these diseases.211 The benefits of CBD also have been shown in human subjects to treat anxiety, depression, post-traumatic stress disorder, and obsessive-compulsive disorders;212,213 furthermore, it has demonstrated antipsychotic properties in those with schizophrenia.214 A few examples of CBD affecting neurological diseases are listed in Table 4 (review).42
Table 4Neuroprotection for Parkinson’s disease initiated with cannabidiol treatment
CBD target | Biological effect |
---|
CBD neuroprotection in Parkinson’s disease (review)42 |
CB1 activation | ↓Microglial activation and microglial NADPH oxidase expression; ↓Production of proinflammatory agents (IL1β, TNFα, iNOs, COX2); ↓Dopaminergic neuronal damage; ↓Excitotoxicity (↓glutamate release); ↓ROS and lipid peroxidation |
CB1 antagonism | ↑Astrocyte activation in substantia nigra pars compacta |
CB2 activation | ↓Microglia number and production of proinflammatory agents (IL1β, TNFα, iNOs, nitric oxide); ↓Dopamine depletion; ↓Myeloperoxidase-positive astrocytes; ↑Antioxidant enzyme activity and antioxidant agents |
MAGL inhibition | ↓Microglia and astrocyte number; ↑CB2 activation; ↑GDNF |
FAAH inhibition | ↑Motor activity; prevents excitotoxicity by inhibiting glutamate release due to neuroinflammation; ↓Protein carbonylation; ↓ROS and lipid peroxidation |
PPARɣ activation | ↓ROS |
CBD neuroprotection in Huntington’s disease (review)42 |
CB1 activation | ↓Excitotoxicity (↓glutamate release) |
CB2 activation | ↓Reactive microglial cell number; ↓Production of proinflammatory agents (TNFα); ↓ROS and nitric oxide; ↑Production of neurotrophins & anti-inflammatory mediators (IL10, IL1 antagonist) |
Phytocannabinoid structure | ↓ROS (phenolic structure acts as an ROS scavenger) |
PPARɣ activation | Interference with the NFκB signaling pathway; Induction of antioxidant enzymes |
CBD neuroprotection in Alzheimer’s disease (review)42 |
PPARɣ activation | ↓Apoptosis during neurodegeneration; ↓Astrocyte activation; ↓Expression of proinflammatory cytokine IL1β and iNOS (↓neuroinflammation); ↓Amyloid plaque and inflammation |
CB1 activation | ↓Amyloid β-induced memory impairment |
CB2 activation | ↓Proinflammatory mediators from microglial cells and astrocytes; ↓Neuroinflammation |
Parkinson’s disease
The hallmark of Parkinson’s disease is the accumulation of α-synuclein and the degeneration of dopaminergic neurons in the SNa in addition to motor alterations (bradykinesia, resting tremors, rigidity, and postural instability), depression, and dementia (review).42 Improvement in the disease by CBD occurs via numerous pathways acting through the eCBS (e.g., CB1Rs, CB2Rs, FAAH, and MAGL) to modulate excitotoxicity, dopaminergic neuronal degeneration through inflammation, and microglial inhibition (Table 4).43,202,215–217 Importantly, CBD has been used to improve the effects of Parkinson’s disease in human subjects (review).218
Huntington’s disease
Huntington’s disease is an autosomal-dominant neurodegenerative disease that is progressive, leading to degeneration of striatal GABA and dopaminergic neuronal destruction in the globus pallidus.43 CB1R activation by CBD in the striatum can inhibit glutamatergic transmission to protect damaged neurons and serve as an antioxidant (Table 4).43,217,219,220
Alzheimer’s disease
CBD has been shown to decrease or block hyperphosphorylation of tau protein, acetylcholinesterase activity, oxidative stress, apoptosis, neuroinflammation, gliosis, and deposition and expression of beta-amyloid (βA).210 The mechanism is associated with selective activation of PPARɣ, resulting in increased clearance of βA peptides through autophagy in the hippocampus, ubiquitination of amyloid precursor proteins, and decreased βA deposition (Table 4).43,210
CBD-associated toxicity
Since it is not considered to be intoxicating, compared to Δ9THC, CBD has been widely used for medicinal purposes and is of great interest to medical communities.17 While CBD use has increased in humans for a plethora of conditions, little is known about the potential for risks from consumption during pregnancy or in children using CBD to treat epilepsy.17,221 The effects of CBD on brain development in utero are not well understood; however, C57BL6/J dams treated with 3.0 mg/kg s.c. GD 5-18 had pups with sex-specific behavioral effects (Table 5).15–17,23,24,183,222,223,228, The male pups showed higher body weights, and there were effects on ultrasonic vocalizations (both sexes), homing behavior, and decreased motor and discriminatory abilities (females). These findings indicate that CBD has effects on psychopathology after in-utero exposure at 3.0 mg/kg/day and may not be as safe as previously considered when consumed during pregnancy.
Table 5Neurotoxic, behavioral, and reproductive effects from CBD treatment during development in animal studies
Animal strain sex/duration/dose/vehicle | Day tested | Effects | LOEL (mg/kg/day) | Reference |
---|
Gestational treatment |
In-vitro C57Bl/6J mouse whole embryos. 6 somite embryos for 24–30 h of culture. Dose: 0, 15, 30 µM CBD. Vehicle: EtOH | 24–30 h | No effects on embryo growth. ↓cranial neural tube closure 15, 30 µM. Significance: Adverse effects on brain development in vitro | 15 µM | 23 |
C57Bl/6J mouse M/F: GD 5–18 S.C. Dose: 0, 3 mg/kg/day. Vehicle: Cremophor EL, EtOH, saline | PND 10 and 13 | 10d: Mean USV duration ↓ (M) and frequency ↑ (F); PND 10, 13, 16, 19, 22: ↓body weight (M); Syllabic repertoire of sound communication sex specific; ↓Homing behavior: Distance moved, velocity, movement distance moved from nest (F). Significance: Adverse neuronal development in vivo | 3 mg/kg/day (only dose tested) | 24 |
Postnatal treatment |
In Vitro Wistar primary neonatal (PND 2) rat cerebral cortices (astrocytes + neurons) 1–24 h. Dose: 0, 0.5, 1, 5 µM CBD. Vehicle: EtOH | 24 h | Neuron: All doses tested: Viability ↓ LDH ↑ at ≥0.1 µM; Only 0.1 µM tested: Change in mitochondrial membrane potential, ↑ATP depletion & caspase 4/7 activation, ↑apoptosis & chromatin condensation; ↓dendrite length; Astrocytes: All doses tested: Viability ↓ LDH ↑ at ≥0.5 µM; Only 0.5 µM tested: dysregulated mitochondrial membrane potential, ↑ATP depletion & caspase 8, 9, 4/7 activation, ↑apoptosis & necrosis. Significance: Cytotoxic to neurons and astrocytes in vitro | Neurons: 0.1 µM; Astrocytes: 0.5 µM | 17 |
In vitro 18-week-old human M: Sertoli cells mouse sertoli cell line. Dose: Human: 7, 8, 9, 10 µM. Mouse: 10, 12.5, 15, 17.5, 20 µM. Vehicle: DMSO | 24 h | Human & Mouse: ↑Cytotoxicity & cell senescence; ↓DNA replication & DNA repair; disruptions in cell-cycle related genes; ↓Cell viability; inhibition of G1/S phase cell cycle transition; ↓mRNA for Wilms’ tumor 1 biomarker. Significance: Adverse effects on human Sertoli cells in vitro | Human: 7.0 µM; Mouse: 10 µM | 228 |
Adolescent treatment |
Swiss mice M: PND 21–55 (4 spermatogenic cycles), gavage. Dose: 0, 15 & 30 mg/kg/day. Vehicle: Sunflower oil | PND 90 | ↓Testosterone (30 mg/kg/day); ↓spermatogenesis (≥15 mg/kg/day); ↑sperm with head abnormalities & cytoplasmic droplets (≥15 mg/kg/day); affected seminiferous tubule morphology (≥15 mg/kg/day). Significance: Disrupted sperm development likely affected fertility | 15 mg/kg/day | 223 |
Swiss Mice M: PND 21–55 (4 spermatogenic cycles), gavage. Dose: 0, 15 & 30 mg/kg/day. Vehicle: Sunflower oil | PND 90 | Germinal epithelium stages disrupted & seminiferous tubule dysmorphology during spermatogenesis (≥15 mg/kg/day); ↑malonaldehyde & ↓sperm motility, super oxide dismutase & catalase at 30 mg/kg/day; ↑abnormal acrosome reaction & sperm velocity (≥15 mg/kg/day). Significance: Potentially affected fertility & ↑oxidative stress | 15 mg/kg/day | 15 |
Adult CBD treatment |
Wistar rat: 1 treatment (M/F) or 4 days (F). Dose: 0, 0.3, 3, 30 mg/kg. Vehicle: Not stated | F: pro- and late diestrus 1 h/4 d; M: 1 h | Acute: Late diestrus ↑entries into & time spent in open arms EPM (0.3 mg/kg/day F; 3.0 mg/kg/day M); 4-day F: Late diestrus ↑entries & time spent into open arms EPM. Significance: Disrupted behavior, indicating neuronal damage in both sexes. | F: 0.3 mg/kg/day; M; 3.0 mg/kg/day | 16 |
In adults, aspects of CBD neurotoxicity are related to sex and strain in rodent studies.208 For example, male and female Swiss and C57BL/6 mice were treated with a single dose of CBD at 0 (saline/Tween 80), 10, and 20 mg/kg/day, and Flanders-sensitive line rats and Flanders- resistant line rats were treated with CBD at 0, 10, 30, and 60 mg/kg/day i.p. The mice were tested in the elevated plus maze, which measures anxiety behavior, and in the tail suspension test, which measures immobility and antidepressant behavior) 30 min after treatment. There were no effects from treatment with either strain of females in the tests, but male Swiss mice showed increased immobility in the tail suspension test at all doses (antidepressant). In the elevated plus maze test, the female Swiss mice showed decreased entries into the enclosed arm, indicating decreased exploratory behavior (antidepressant-like effect). Meanwhile, male and female C57BL/6 mice did not show effects in the elevated plus maze test. Rats were also tested 50 min after treatment in the forced swim and open field tests. The Flanders-sensitive line rats showed decreases at all doses in the forced swim test (measure of immobility), with no effects on distance traveled in the open field test and no effects in these tests with Flanders-resistant line rats. When the interval between treatment and testing was increased to 2 h, there was a slight increase in immobility in the Flanders-sensitive line rats at 30 mg/kg CBD. Therefore, it is significant to note that the exposure time, sex, strain, and species differences with CBD treatment were related to anxiety/depressive behaviors. The doses used in this study and those shown in Table 5 are within the range of those showing neuroprotection in Table 3, also administered i.p. In-vitro studies with mouse embryos also support the toxic effects of CBD during development.16
Animal studies have shown that doses of CBD that are neuroprotective (Table 3), can be toxic to the male reproductive tract.14,15,223–225 CBD treatment at 15 mg/kg/day (gavage) for three sperm development cycles in mice can lead to disrupted sperm development, abnormal seminiferous epithelium, decreased testes weights, and other effects that would impact fertility.14 Studies also have demonstrated reduced testosterone, inhibition of sperm maturation, and thinning, atrophied cells, pyknosis in seminiferous tubules, and other pathologies.14 The presumptive MOA involves CBD inhibition of 17α-hydroxylase in Leydig cells, leading to decreased testosterone production. However, in humans, the effects on sperm and other reproductive parameters in males have been mainly attributed to the Δ9THC content in cannabis, rather than CBD.226,227 But based on animal studies, CBD in cannabis could contribute to the negative effects in males; hence, this area needs more research. In-vitro studies performed on human and mouse Sertoli cells obtained postnatally support the toxic effects of CBD observed in animal studies.18 Dose exposure, route, species, sex, frequency of consumption, and susceptibility to the effects from exposure contribute to health outcomes.
Future directions
With increasing use of cannabis with higher concentrations of Δ9THC, there are concomitant risks to safety in the general population from intoxication while driving or in the workplace. Methods have been developed to measure impairment from cannabis in a timely manner on site (e.g., in a car or workplace) through brain imaging to provide assessments of intoxication.39 Functional near-infrared spectroscopy provides a measurable signature of neural impairment of the PFC, and the results are supported by blood and urine assessments to indicate whether participants were exposed but not impaired or exposed and impaired. Such measures acknowledge the growing need for detection and mitigating safety measures due to cognitive impairment from cannabis use.
Neurotoxicity of CBD is also in need of more study. For example, CBD injured neonatal rat cortical neurons and astrocytes in vitro at low therapeutic levels that could affect patients treated with CBD.17,221 CBD is known to be neuroprotective in Parkinson’s disease, where dopaminergic neurons of the substantia nigra pars compacta are shown to degenerate.78,194,229 Conversely, in animal models, dopaminergic pathways are attenuated by CBD, resulting in decreased motor functions.24 While data indicate that for some, the benefits of CBD may outweigh the risks, it is clearly necessary to continue researching optimal treatment levels related to disease improvement. Persons exposed to higher doses of CBD for severe illnesses, such as DS to control seizures (Epidiolex®, Epidyolex® in Europe), may need to weigh the risk versus benefit and exert caution for use in pregnant women and children.
Finally, one of the biggest challenges in characterizing the effects of cannabis during developmental life stages is knowing the exposure and individual health risk factors. In laboratory experiments, the exact dose, purity of cannabinoids, animal strain/sex/pregnancy status, duration of exposure, and other parameters are controlled; however, with human subjects, it is difficult to characterize exposure. Nevertheless, knowledge of the dose and product components being consumed as well as the life stage of exposure, route of exposure (i.e., inhalation, s.c., i.v., oral, or i.p.), body fat composition, age, health status, frequency of use, and other factors will determine the absorption, distribution, and metabolism of cannabinoids entering the blood stream.107,111 Many of these parameters are not consistent among studies performed in animals (e.g., different animal species/strains, dosing regimens, vehicles), and data may be difficult to obtain in epidemiological studies with human subjects. Thus, there is a need for further study to protect fetuses, infants, and children from harmful exposures during development. There is also a need for further research related to risks for male reproductive toxicity.
Conclusions
This review focused on neurotoxicity and neuroprotection of the most thoroughly characterized phytocannabinoids in cannabis—Δ9THC and CBD. Most cannabis exposure is not a pure form of either compound, but it contains a combination of those and over 100 others. Due to the increasing use of cannabis or CBD, not just recreationally, but for the treatment of diseases (e.g., depression, anxiety, inflammation, pain, and seizures) and a plethora of other conditions, it is critical for the industry to thoroughly characterize expected exposures. The extent of the risk versus beneficial effects of compounds in cannabis is dependent on many factors, but, as indicated by studies with Δ9THC, there is a high risk for long-lasting neurodevelopmental effects from exposure to fetuses, infants, children, and adolescents, including severe mental dysfunction (e.g., depression, anxiety, and schizophrenia), decreased cognition, drug dependency tendencies, and decreased motor function. Adolescent use can present unique challenges because adolescence is a developmental stage of increased independence and potential for experimentation with cannabis. In addition, brain development as well as major dynamic changes in the eCBS continue for the first 25, or more, years of life; hence, cannabis exposure during adolescence can still attenuate brain development. Adolescent exposure has been shown to lead to persistent adverse neurodevelopmental changes, increasing the risks for major depressive disorder, drug addiction, and severe psychotic disorders.
On the other hand, CBD is nonpsychotropic and has positive therapeutic applications to treat childhood epilepsy, multiple sclerosis, stroke, Alzheimer’s disease, Parkinson’s disease, and other severe disorders. The focus has been mainly on the health benefits; however, the reported developmental effects from exposure in utero, effects on male reproduction, and associations with human genotoxicity have not been well studied, and a significant data gap remains.
Abbreviations
- ACh:
acetylcholine
- 2-AG:
2-arachidonoylglycerol
- 5-HT:
serotonin
- AEA:
anandamide
- ARfD:
acute reference dose
- βA:
beta-amyloid
- BDNF:
brain-derived neurotropic factor
- Ca+2:
calcium
- CB1R:
cannabinoid 1 receptor
- CBD:
cannabidiol
- CNS:
central nervous system
- COX2:
cyclooxygenase-2
- D1 or D2:
dopamine receptors
- DA:
dopamine
- DAGL:
diacylglycerol lipase
- DRN:
dorsal raphe nucleus
- DS:
Dravet syndrome
- eCB:
endocannabinoid
- eCBS:
endocannabinoid system
- FAAH:
fatty acid amide hydrolase
- GABA:
gamma-aminobutyric acid
- GD:
gestation day
- GPR55:
G-coupled protein receptor 55
- i.p.:
intraperitoneal
- i.v.:
intravenous
- IL:
interleukin
- iNOS:
inducible nitric oxide synthase
- K+:
potassium
- LOAEL:
lowest-observed-adverse-effect level
- LOEL:
lowest-observed-effect level
- MAGL:
monoacylglycerol lipase
- MOA:
mode of action
- NAc:
nucleus accumbens
- NFкB:
nuclear factor kappa-light-chain-enhancer of activated B cells
- NMDA:
N-methyl-D-aspartate
- NOAEL:
no-observed-adverse-effect level
- PFC:
prefrontal cortex
- PPARɣ:
peroxisome proliferator-activated receptor gamma
- ROS:
reactive oxygen species
- s.c.:
subcutaneous
- SNc:
substantia nigra
- TNF:
tumor necrosis factor
- TRPV1:
transient receptor potential cation channel subfamily V member 1, or vanilloid receptor 1
- VTA:
ventral tegmental area
- Δ9THC:
delta-9-tetrahydrocannabinol
Declarations
Acknowledgement
I would like to thank Dr. Poorni Iyer, DVM, PhD for helpful discussions and for her work with cannabis and developmental effects.
Funding
This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.
Conflict of interest
I have no competing interests (financial/personal) to declare.
Authors’ contributions
MHS is the sole author of this work.