Psychedelic-Induced Neural Plasticity: A Comprehensive Review and a Discussion of Clinical Implications
Abstract
:1. Introduction
2. Methods
3. Results
3.1. Preclinical Evidence
3.1.1. Psilocybin/Psilocin
References | Methods | Results | Comments |
---|---|---|---|
Raval et al., 2021 [52] doi:10.3390/ijms22020835 | Single dose of 0.08 mg/kg psilocybin in pigs In vitro autoradiography | Psilocybin increases SV2A expression in the hippocampus one and seven days post-injection. | Psilocybin significantly increases the expression of a marker of synaptic plasticity (synaptic vesicle protein 2A, SV2A) in the hippocampus. |
Jefsen et al., 2021 [54] doi:10.1177/0269881120959614 | Single dose of psilocybin (0.5–20 mg/kg) in rats Real-time qPCR | Prefrontal cortex: Cebpb↑, cFos↑, Dusp1↑, Fosb↑, Junb↑, Iκβ↑, Nr4a1↑, P11↑, Psd95↑, Sgk1↑, Clk1↓ Hippocampus: Arrdc2↑, Dusp1↑, Iκβ-α↑, Sgk1↑ Arc↓, Clk1↓, Egr2↓, Ptgs2↓ | Psilocybin acutely and dose-dependently induces the expression of genes associated with neuroplasticity in the prefrontal cortex and the hippocampus. |
Fadahunsi et al., 2022 [55] doi:10.1038/s41398-022-02103-9 | Mice were randomized in four groups (intra-peritoneal injection of 0.3 mg/kg psilocybin, 1 mg/kg psilocybin, 3 mg/kg psilocybin, or isotonic saline) Real-time qPCR | Prefrontal cortex: Nptn↑, BDNF↑, Negr1↑ | Psilocybin acutely and dose-dependently induces the expression of genes associated with neuroplasticity in the prefrontal cortex. |
Liu et al., 2023 [59] doi:10.3389/fnins.2023.1168911 | Single dose of 2.0 mg/kg psilocybin hydrochloride Immunofluorescence | EGR1↑ | Psilocybin acutely increases the expression of EGR1, a gene involved in neuronal plasticity, across the brain. |
Du et al., 2023 doi:10.1097/CM9.0000000000002647 [60] | Vehicle (sterile 0.9% saline) or psilocybin (0.1 mg/kg, 0.5 mg/kg, or 2.5 mg/kg) administered 30 min prior to Fear conditioning (FC) extinction training in mice Western blot | BDNF↑, mTOR↑ | Psilocybin reversed FC-associated BDNF and mTOR reduction 7 days after administration. |
Davoudian et al., 2023 doi:10.1021/acschemneuro.2c00637 [61] | Mice received either saline (10 mL/kg), ketamine (10 mg/kg) or psilocybin (1 mg/kg). c-FosGFP transgenic mice Whole-brain serial two-photon microscopy Light sheet microscopy | c-Fos↑ in anterior cingulate cortex, locus coeruleus, primary visual cortex, central amygdala, basolateral amygdala, hippocampus, medial habenula, lateral habenula, claustrum and anterior and midline thalamic nuclei, after both psilocybin and ketamine. c-Fos↑ in reticular nucleus of the thalamus, caudoputamen, periaqueductal gray and c-Fos↓ in dorsal raphe and insular cortex after psilocybin. | Psilocybin acutely modulates expression of c-Fos across different areas of the brain similarly to ketamine, but with some difference in dorsal raphe, insular cortex and hippocampus. |
Rijsketic et al., 2023 doi:10.1038/s41386-023-01613-4 [62] | Mice were administered saline or psilocybin (2 mg/kg, i.p.), before placement into their home cage or an enriched environment. After 2 weeks, mice were administered saline or psilocybin and confined to either environment for 2 h. Light sheet fluorescence microscopy | c-Fos↑ in subregions of the neocortex, caudoputamen tail, central amygdala. c-Fos↓ in the hypothalamus, cortical amygdala, striatum and pallidum. | Both psilocybin and environmental context acutely modulate expression of c-Fos across different areas of the brain, in a mainly additive manner. |
Funk et al., 2024 [64] doi:10.1016/j.neuroscience.2024.01.001 | Male rats were randomized into four groups (vehicle or subcutaneous psilocybin 0.1, 0.5, 3 mg/kg at 1 mL/kg). Immunohistochemistry | c-Fos↑ in neurons and oligodendrocytes in the frontal cortex, nucleus accumbens, central amygdala, basolateral amygdala, locus coeruleus. | Psilocybin acutely increases the expression of c-Fos both in neurons and oligodendrocytes across several brain areas. |
Shahar et al., 2024 [65] doi:10.1038/s41380-024-02477-w | Adult male mice were randomized into three groups (vehicle, psilocybin 4.4 mg/kg intraperitoneally or psychedelic mushroom extract (PME) containing the same amount of psilocybin). Western blot | Hippocampus: GAP43↑, synaptophysin↑, SV2A↑ (only in PME group). Amygdala: synaptophysin↑, SV2A↑ (only in psilocybin group). Frontal cortex: GAP43↑ Striatum: synaptophysin↑ (only in PME group). | Psilocybin and psychedelic mushroom extract both induce gene expression associated with neuroplasticity 11 days after administration. |
Lerer et al., 2024 [66] doi:10.3389/fphar.2024.1391412 | Male mice were randomized into three groups (vehicle, psilocybin 4.4 mg/kg or PME with equal dose of psilocybin). Real-time qPCR | c-Fos↑, egr1↑ in somatosensory cortex. | Both psilocybin and psychedelic mushroom extract acutely increase immediate early genes expression in somato-sensory cortex. |
Zhao et al., 2024 [68] doi:10.1177/02698811241249436 | Male mice, previously exposed to glucocorticoids for 21 days (CORT-induced depression model), were randomly divided into four groups (vehicle, 0.1 mg/kg, 1.0 mg/kg and 2.5 mg/kg psilocybin intraperitoneal injection). The same randomization was conducted on a population not exposed to glucocorticoids. Western blot Immunofluorescence | BDNF↑, TrkB↑, mTOR↑, Doublecortin↑ in the prefrontal cortex and the hippocampus. | Psilocybin reversed the reduction in neuro-plasticity biomarkers induced by chronic glucocorticoid exposure. |
3.1.2. DMT-Related Compounds
References | Methods | Results | Comments |
---|---|---|---|
Dakic et al., 2017 [73] doi:10.1038/s41598-017-12779-5 | Human embryonic stem cells were induced towards neural differentiation, expanded and grown into cerebral organoids; organoids were separated into a group treated with 5-methoxy-N,N-dimethyltryptamine (5-MeO-DMT) and a vehicle group. qPCR Immunohistochemistry | 360 downregulated and 574 upregulated proteins, predicting dendritic spine and cellular protrusion formation, microtubule and cytoskeletal organization, mild T lymphocyte differentiation and inhibiting neurodegeneration, cell death and brain lesion NMDAR↑, CaMK2↑, CREB↑, ERK1/2↑ NF-κB↓, mGluR5↓, PKC↓, PLC↓, CaM↓, AC1/8↓, IP3R↓, EPAC1↓, PKA↓ | 5-MeO-DMT caused molecular alterations in human cerebral organoids associated with neuroplasticity and anti-inflammatory mechanisms. |
Ly et al., 2018 [39] doi:10.1016/j.celrep.2018.05.022 | Cortical neurons were randomized into ANA-12-treated group, rapamycin, ketanserin, control, then each group was randomized into 2,5-Dimethoxy-4-iodoamphetamine, N,N-Dimethyltryptamine or LSD for 24 h. Droplet digital PCR. ELISA. | Psychedelics caused a non-significant increase in BDNF expression, which was not seen in neurons pre-treated with ANA-12. Treatment with rapamycin or ketanserin blocked psychedelic-induced neuritogenesis. | Inhibiting TrkB, mTOR, 5-HT2A receptor blocked psychedelic-induced increase in neuritogenesis or BDNF expression, suggesting a role of these macro-molecular structures in the effects of psychedelics. |
Nardai et al., 2020 [77] doi:10.1016/j.expneurol.2020.113245 | Transient middle cerebral artery occlusion (MCAO) was induced in male rats before being randomized into 2 groups: i.p. bolus of vehicle or 1 mg/kg DMT followed by 2 mg/Kg in 24 h. qPCR. ELISA. | APAF1↑, BNDF↑, IL-10↑, TNF-α↓, IL1-β↓, IL-6↓ | DMT reduced ischemic brain lesion volume and had a neurotrophic and anti-inflammatory effect, mediated by its modulation of cytokines and other neurotrophins. |
Almeida et al., 2022 [78] doi:10.1016/j.bbr.2021.113546 | Mice received 2.2 g/kg ethanol or saline i.p. injections every other day for nine days and then randomized into two groups (daily administration of a dose of ayahuasca corresponding to 1.76 mg/kg of N,N-dimethyltryptamine, DMT, or water). Western blot | Ayahuasca dampened the increase of 5-HT1A receptor levels in the hippocampus after induction by ethanol. Ethanol decreased the dynorphin/prodynorphin ratio in the striatum, and treatment with ayahuasca partially reverted this reduction. Ayahuasca prevented the increase of prodynorphin in the hippocampus induced by ethanol. | Ayahuasca attenuated anxiety-like behaviour, proposedly preventing ethanol-induced increases in 5-HT1A receptor and prodynorphin levels in the hippocampus. |
Kelley et al., 2022 [80] doi:10.1021/acschemneuro.1c00660 | A population of male rats was randomized into placebo or a 30-day stress regiment followed by administration of DMT 2 mg/kg i.p., harmaline 1.5 mg/kg i.p., DMT 2 mg/kg and harmaline 1.5 mg/kg i.p. combined (pharmahuasca), or vehicle every other day for 5 days. Electron paramagnetic resonance spectroscopy Real-time PCR | DMT + Harmaline mitochondrial transcripts↓, CYBA↓, CYBB↓, p47phox↓, Rap1↓, Il1r1↓, Il1α↓, Tlr4↓, Tlr6↓, Tlr7↓, Ifngr1↓, NF-κβ2↓, Nrf2↓, Keap1↑, IPA Oxidative Phosphorylation signaling pathway↓ DMT Mitochondrial transcripts↓, Rap1↓, Il1r1↓, Il1α↓, Tlr4↓, Tlr6↓, Tlr7↓, Ifngr1↓, Lrp8↑, Ntrk2↑, Ntrk3↑, mTORC1↑, mTORC2↑, Crebbp↑, Nrf2↓, Keap1↑ Harmaline Mitochondrial transcripts↓↑, Rap1↓, CYBB↓, Il1r1↓, Il1α↓, Tlr4↓, Tlr6↓, Tlr7↓, Ifngr1↓, IPA synaptogenesis pathway↑, Sst↑, Sstr1↑, Sstr2↑, Sstr3↑, Sstr4↑, Nrf2↓, Keap1↑, IPA Oxidative Phosphorylation signaling pathway↓ | All treatments reversed some differentially expressed genes observed in PTSD, each with differences. All treatments reduced expression of ROS production-associated pathways, increased expression of anti-inflammatory and synaptogenesis-associated pathways. |
3.1.3. LSD
References | Methods | Results | Comments |
---|---|---|---|
Nichols & Sanders-Bush, 2002 [85] doi:10.1016/S0893-133X(01)00405-5 | Rats were randomized into 1 mg/kg of LSD i.p. or control. DNA microarrays RNase protection assays | Prefrontal cortex: c-fos↑, arc↑, krox-20↑, NOR1↑, Ikβ-α↑, sgk↑, ania3↑ Midbrain: c-fos↑, NOR1↑, Ikβ-α↑, sgk↑, ania3↑ Hippocampus: c-fos↑, krox-20↑, Ikβ-α↑, sgk↑ | LSD increased expression of seven genes involved in synaptic plasticity, glutamatergic signaling and cytoskeletal architecture, in a region-specific manner. |
Nichols et al., 2003 [86] doi:10.1016/s0169-328x(03)00029-9 | Main experiment: rats were randomized into 1 mg/kg of LSD i.p. or saline i.p. Secondary experiments: rats were randomized into WAY100635 1 mg/kg+ LSD 1 mg/kg i.p, or WAY100635 1 mg/kg i.p alone, MDL100907 1 mg/kg + LSD 1 mg/kg i.p, or MDL100907 1 mg/kg i.p alone. DNA microarrays RNase protection assays and qPCR | c-fos↑, arc↑, krox-20↑, NOR1↑, Ikβ-α↑, sgk↑, ania3↑, homer1a↑ WAY100635 did not inhibit LSD-induced gene expression. MDL100907 inhibited LSD-induced gene expression. | LSD-induced expression of neurotrophism-associated genes in prefrontal cortex, and it appears to be largely mediated by 5-HT2A receptors. |
Nichols & Sanders-Bush, 2004 [87] doi:10.1111/j.1471-4159.2004.02515.x | Rats were randomized into 1 mg/kg of LSD i.p. or saline i.p. In another experiment rats were randomized into WAY100635 1 mg/kg+ LSD 1 mg/kg i.p, or WAY100635 1 mg/kg i.p alone, MDL100907 1 mg/kg + LSD 1 mg/kg i.p, or MDL100907 1 mg/kg i.p alone. DNA microarrays RNase protection assays and qPCR | Prefrontal cortex: C/EBP β↑, MKP-1↑, ILAD-1↑ WAY100635 did not inhibit LSD-induced gene expression. MDL100907 inhibited LSD-induced gene expression. Hippocampus: MKP-1↑, ILAD-1↑ Midbrain: C/EBP β↑, MKP-1↑ | LSD-induced expression of neurotrophism-associated genes in prefrontal cortex, and it appears to be largely mediated by 5-HT2A receptors. |
Martin et al., 2014 [89] doi:10.1016/j.neuropharm.2014.03.013 | Rats were randomized into saline or 0.16 mg/kg LSD i.p., every other day, for 90 days. qPCR | 28 days after the final treatment, 283 transcripts exhibited differential expression (>25%) (p < 0.05), including: Gabrb1↑, Gabrb2↑, Gabrg3↑, Slc6a13↓, NR2a↑, NR2b↑, Bdnf↑, Krox20↑, Drd1↓, Drd2↓, Ndufs8↓, Ndufb2↓, Ndufb7↓, Nudufa1↓, Atp6v0b↓, Atp6v0e2↓, Atp6v0c↓, Atp5d↓, Cox7a2↓, Cox8a↓, Cox4nb↓, Gstt2↓, Gstp2↓ | Chronic LSD administration causes long term changes in expression of genes implicated with a large variety of functions, including synaptic transmission, synaptic plasticity, cell projection organization, cell-cell signaling, cytoskeleton. |
Ly et al., 2018 [39] doi:10.1016/j.celrep.2018.05.022 | Cortical neurons were randomized into ANA-12-treated group, rapamycin, ketanserin, control, then each group was randomized into 2,5-Dimethoxy-4-iodoamphetamine, N,N-Dimethyltryptamine or LSD for 24 h. Droplet digital PCR ELISA | Psychedelics caused a non-significant increase in BDNF expression, which was not seen in neurons pre-treated with ANA-12 Treatment with rapamycin or ketanserin blocked psychedelic-induced neuritogenesis | Inhibiting TrkB, mTOR, 5-HT2A receptor blocked psychedelic-induced increase in neuritogenesis or BDNF expression, suggesting a role of these macro-molecular structures in the effects of psychedelics. |
Savino & Nichols, 2022 [91] doi:10.1111/jnc.15534 | Rats were randomized into saline or 0.16 mg/kg LSD i.p., every other day, for 90 days. RNA sequencing | Enriched gene ontology (GO) categories: “dendrite development”, “circadian rhythms”, “covalent chromatin modification/histone modification”. Weighted gene co-expression network analysis (WGCNA): six clusters showed differential activity between LSD and controls: GTP binding↓, serine hydrolase↓, ribosome↓, chromatin organization↑, vesicle mediated transport in synapse↑, cell-cell adhesion↑. Increased signaling entropy and reduced between sample entropy | LSD induced differential expression in genes involved in neural function and remodelling. LSD increased plasticity-like transcriptional entropy (signaling entropy) and reduced ageing-like transcriptional entropy (between sample entropy). |
Ornelas et al., 2022 [92] doi:10.1016/j.expneurol.2022.114148 | Brain organoids originating from induced pluripotent human stem cells (iPSCs) were randomized into LSD 10 nM (3.23 ng/mL) in 24 h or vehicle. Liquid chromatography–mass spectrometry (LC-MS) based proteomics | Out of 3448 identified proteins, 234 had significant LSD-induced modifications in expression (p < 0.05). Enrichment analysis for predicted affected pathways and biological processes (Metascape) found several processes potentially affected: DNA replication, axon guidance, synaptic vesicle cycle, mTOR pathway, long-term depression (LTD) and dopamine neurotransmitter release cycle. | LSD induced differential expression in genes involved in neuroplasticity. In particular, proteins involved in mTOR signaling pathway was significantly (p = 0.034) over-represented in the LSD-treated set (2.15%) compared to the control set; (0.79%). |
Inserra et al., 2022 [93] doi:10.1016/j.pnpbp.2022.110594 | Male rats were randomized into LSD 30 µg/kg/day for 7 days or vehicle. Whole-genome bisulfite sequencing Proteomic profiling | Modulation of 635 CpG site methylation, especially in autosomes Modulation of expression of 178 proteins involved in signaling pathways for the development of the nervous system, axon guidance, synaptic plasticity, quantity and cell viability of neurons and protein translation. | LSD modulated methylation of 635 CpG sites and induced differential expression in genes involved in neuroplasticity. |
3.1.4. Mescaline and Related Compounds
3.1.5. Synthetic 5-HT2A Agonists (Such as DOI and 25C-NBOMe)
References | Methods | Results | Comments |
---|---|---|---|
Pei et al., 2000 [49] doi:10.1016/s0028-3908(99)00148-3 | Rats were randomized into DOI 0.2 mg/kg i.p., DOI 1 mg/kg, DOI 2 mg/kg, ketanserin 2 mg/kg + DOI 1 mg/kg. In situ hybridization histochemistry. | DOI: dose-dependent Arc↑ in orbital cortex, cingulate cortex, frontal cortex, parietal cortex, less significantly in striatum; ketanserin completely inhibited DOI’s effects. | DOI increased Arc expression in many brain areas. |
Benekareddy et al., 2013 [96] doi:10.1017/S1461145712000168 | Rats were randomized into DOI 8 mg/kg or saline i.p. injection. Induced BDNF KO mice and wild type mice were each randomized into immobilization stress or being left undisturbed. BDNF KO mice and wild type mice were each randomized into saline or DOI injection. In situ hybridization Immunohistochemistry | Immobilization stress and DOI: Arc↑ in the neocortex but not in the hippocampus BDNF: Arc↑ in the neocortex and the hippocampus Induced BDNF KO mice: Arc↓ in the neocortex but not in the hippocampus; reduced Arc↑ with acute immobilization or DOI. | DOI and immobilization stress increase Arc expression, which appears to be regulated by BDNF. |
de la Fuente Revenga et al., 2021 [99] doi:10.1016/j.celrep.2021.109836 | Mice (n = 6 for each condition) were injected (i.p.) with vehicle or DOI. Frontal cortex samples were collected at 24 h, 48 h, or 7 days after administration. High-resolution, cell-type-specific and low-input ChIP-sequencing; RNA-sequencing | DOI significantly induced differential modifications in enhancer activation (H3K27ac) and gene expression in several Gene ontology (GO) categories. Some of these modifications were long-lasting (persisting after seven days). | DOI induces both transcriptomic and epigenomic modifications compatible with an increased plasticity state, the latter being largely persistent. The findings of this study are in agreement with a central role of 5-HT2A receptors in psychedelic-induced plasticity. |
Desouza et al., 2021 [100] doi:10.3389/fnmol.2021.790213 | 5HT2A KO mice and WT mice were randomized into i.p. DOI 2 mg/kg or vehicle. Male rats, CREBαδ KO mice and WT mice were randomized into i.p. DOI 8 mg/kg or vehicle. Another group of rats was randomized into electroconvulsive seizure treatment or sham. Immunofluorescence Western blot qPCR Chromatin immunoprecipitation In situ hybridization | 2 mg DOI in mice: Arc↑, Bdnf1↑, Cebpb↑, cFos↑, Egr1↑ and Egr2↑ in neocortex These effects were inhibited by MDL100,907 or U73122 administration. KN-62 inhibited the effect on cFos and U0126 inhibited the effect on Egr1; both inhibited the effect on Arc, Bdnf1, Cebpb, Egr2. DOI effect on Bdnf1 and cFos were inhibited in 5HT2A KO mice, while the effect on Arc, Egr1 and Egr2 was reduced. pCREB↑ with DOI, inhibited by KN-62 and U0126 8 mg DOI in rats:Arc↑, Atf3↑, Atf4↑, Bdnf1↑, Cebpb↑, Cebpd↑, Egr1↑, Egr2↑, Egr3↑, Egr4↑, cFos↑, JunB↑, Nfkbia↑. Arc, Bdnf1, Cebpb and cFos, but not Egr1 and Egr2 expression, was associated with pCREB enrichment. Electroconvulsive seizure treatment increased expression of Arc, Bdnf1, Cebpb, cFos and Egr2 similarly to DOI. CREBαδ KO mice had no significant difference in 5-HT2A or 5-HT2C expression after DOI administration compared to WT mice. DOI-dependent increase in Arc expression was reduced in CREBαδ KO mice. | Neurotrophic effects of DOI appear to be mediated by 5-HT2AR and Gq-coupled PLC pathway leading, through MAPK and CaMKII signaling to an increased phosphorylation of CREB in the cortex. |
Study | Methods | Results | Comments |
---|---|---|---|
Jha et al., 2008 [88] doi:10.1016/j.neulet.2008.06.028 | DOI (8 mg/kg) or LSD (0.5 mg/kg) was administered through intraperitoneal (i.p.) injection either once (acute) or once daily for 7 days (chronic). Immunohistochemistry study, using anti-BrdU (5-bromo-2-deoxyuridine) antibodies to assess the proliferation of adult hippocampal progenitors | No observed effect of DOI or LSD on the number of proliferating cells in the adult rat hippocampus | DOI and LSD do not seem to exert mitogenic effects on hippocampal progenitors. |
Ly et al., 2018 [39] doi:10.1016/j.celrep.2018.05.022 | For in vitro studies, cells were treated with LSD, DOI and DMT at a concentration of 10 μM, 10 μM and 90 μM, respectively. For the ex vivo experiment, rats were administered DMT i.p. at a dose of 10 mg/kg. For in vitro assays, structured illumination microscopy (SIM); for ex vivo assays Golgi–Cox staining. Tests with inhibitors of 5-HT2A receptors (ketanserin), TrkB receptors (ANA-12) and mTOR complex (rapamycin) were performed. | DOI, LSD and DMT induce structural plasticity (neuritogenesis, spinogenesis and synaptogenesis). Among the psychedelics tested, LSD appeared to be the one with the greatest psychoplastogenic effects. Ketanserin, ANA-12 and rapamycin abolished these effects. | DOI, LSD and DMT seem to augment neuronal remodeling, and this finding is confirmed by different methods in vitro and ex vivo (only for DMT). Psychoplastogenic effects are largely dependent on 5-HT2A receptors, TrkB receptors and mTOR complex signaling. |
Lima da Cruz et al., 2018 [75] doi:10.3389/fnmol.2018.00312 | Adult mouse single intracerebroventricular iniection of 1 μL 5-MeO-DMT solution BrdU Labeling, BrdU immunohistochemistry, clustering Analysis of BrdU+ cells, confocal microscopy | Cellular proliferation and increased dendritic complexity in granule cells of the dentate gyrus | DMT seems to exert a mitogenic effects on hippocampal progenitors. |
de la Fuente Revenga et al., 2021 [99] doi:10.1016/j.celrep.2021.109836 | Single administration of DOI or vehicle in 5-HT2AR−/− mice as compared to 5-HT2AR+/+ controls. 3D automated method for quantitative structural spine analysis. | Increased density of transitional dendritic spines, but not mature mushrooms spines, in 5-HT2AR+/+, but not in 5-HT2AR−/− mice. A significantly greater magnitude of LTP was evident in DOI-treated compared to vehicle-treated mice over the same time course. | 5-HT2ARs seem to be necessary for the plasticity-boosting effects of DOI. These latter could partly explain the LTP facilitatory properties of this compound. |
Shao et al., 2021 [53] doi:10.1016/j.neuron.2021.06.008 | Mice treated with 1 mg/kg intraperitoneal psylocibin compared to placebo. Two-photon microscopy | Increased size and density of cortical dendritic spines that appear within the first 24 h and last at least for a month. These effects were not blocked by ketanserin administration. | Psychedelics directly induce neuronal structures remodeling and growth in mice, allegedly through a 5-HT2AR-independent mechanism. |
Ly et al., 2021 [90] doi:10.1021/acsptsci.0c00065 | Treatment of rat neuronal cell cultures with 10 μM LSD. Structured illumination microscopy (SIM). | Transient exposures (from 15 min to 6 h) is sufficient to induce neuroplastic changes (neuritogenesis, spinogenesis and synaptogenesis) comparable to that observed following more prolonged periods of exposure (72 h) applied in the previous experiment. | Neuroplastic changes would persist after short exposures to LSD and would require a stimulation phase, mediated by the activation of TrkB receptor and a growth phase, involving the sustained activation of mTOR and the AMPA receptor. |
Ko et al., 2023 [102] doi:10.1016/j.biopsych.2023.02.388 | Injection of 3 mg/kg of DOI in wake-behaving mice compared to a vehicle-injected group. Miniature microscope to study hippocampal cells from mice that expressed the calcium sensor under the control of the oligodendrocyte precursor cell (OPC) promoter. | DOI is capable of influencing myelin plasticity, increasing myelin density in rat hippocampus 24 h after administration. | Not only does DOI exert neuroplastic effects on neurons, but it seems also to induce similar effects on oligodendrocytes. |
Vargas et al., 2023 [81] doi:10.1126/science.adf0435 | 10 μM of 5-methoxy-N,N-dimethyltryptamine (5-MeO) to wild-type (WT) and 5-HT2AR knock- out (KO) mice. Golgi–Cox staining | Neuroplastic structural effects are observed, essentially in terms of dendritogenesis, largely mediated by intracellular 5-HT2A receptors. | Intracellular 5-HT2A receptors seem to have a crucial role in the neuroplastic effects of tryptamines. |
Jefferson et al., 2023 [82] doi:10.1038/s41386-023-01572-w | In vivo two-photon microscopy of mouse medial frontal cortex after i.p. 5-MeO-DMT. | Intensification of dendritic spine formation in the mouse medial frontal cortex, resulting in a long-lasting increase in spine density | Tryptamines are able to increase cortical dendritic spine density in vivo. |
Moliner et al., 2023 [94] doi:10.1038/s41593-023-01316-5 | Multiple in vitro and in vivo/ex vivo experiments comparing LSD, psilocin and vehicle. Various techniques were employed, including binding assays, microscale thermophoresis assay, nuclear magnetic resonance spectroscopy, split-luciferase complementation assay, fluorescence recovery after photobleaching, real time qPCR, single-molecule localization microscopy, ELISA, Western blotting | LSD and psilocin bound TrkB with high affinity (LSD Ki = 3.38 ± 1.39 nM; Psilocin Ki = 673 ± 3.05 nM), Y433F and TrkA.TM, but not S440A mutation, impaired LSD binding. LSD and PSI induced a fast and long-lasting dimerization of TrkB but not in Y433F heterodimers. Pretreatment with ketanserin or M100907 did not prevent this effect. LSD potentiated the effect of low doses of BDNF. | Contrary to what was previously observed, LSD and psilocin might exert at least part of their neuroplastic effects regardless of 5-HT2A-mediated signaling, through a direct TrkB-dependent mechanism. LSD and psilocin bind to the transmembrane domain of TrkB and act as positive allosteric modulators. |
Du et al., 2023 [60] doi:10.1097/CM9.0000000000002647 | A single dose of psilocybin (2.5 mg/kg, i.p.) in mice, 30 min before extinction training in a fear conditioning (FC) protocol. Golgi staining for the dendritic complexity and spine density | Reduced decrease of hippocampal dendritic complexity and spine density induced by FC. | Psilocybin might be able to dampen the anti-plastic effects of stress, facilitating FC extinction. |
Zhao et al., 2024 [68] doi:10.1177/02698811241249436 | Psilocybin (0.1, 0.5, 1.0, 2.5, 5.0 mg/kg) i.p. injection in mice. Western blot, immunofluorescence, dendrite analysis after Golgi staining | A single administration of psilocybin can reverse the detrimental effects of chronic glucocorticoid exposure on structural plasticity (density and number of dendritic spines and branches). | Psychedelics might reverse anti-plastic processes known to be associated with chronic stress and restore the adaptive capacity of neurons. |
Study | Methods | Results | Comments |
---|---|---|---|
Cameron et al., 2019 [76] doi:10.1021/acschemneuro.8b00692 | Male and female Sprague Dawley rats treated with DMT·fumarate (1 mg/kg) over two months. Behavioral assays included elevated plus maze, forced swim test and fear conditioning. | DMT improved mood-related behaviors and fear extinction, particularly in females. No significant changes in anxiety or cognition. DMT induced a reopening of plasticity without any influence on working/short-term memory or social interactions. | Chronic low-dose DMT shows promise for mood improvement without cognitive deficits. |
de la Fuente Revenga et al., 2021 [99] doi:10.1016/j.celrep.2021.109836 | Adult male mice treated with DOI (2 mg/kg) via intraperitoneal injection. Behavioral tests included forced swimming test (FST), dark–light test, novel object recognition and contextual fear extinction. Synaptic plasticity was analyzed using histological and electrophysiological techniques in 5-HT2AR+/+ and 5-HT2AR−/− mice. | DOI significantly reduced immobility time in the FST, indicating antidepressant-like effects. No effects were observed on anxiety or cognitive performance. DOI accelerated fear extinction in 5-HT2AR+/+ mice but not in 5-HT2AR−/− mice. Enhanced synaptic plasticity was noted in the frontal cortex, characterized by an increase in dendritic spines. Enhanced synaptic plasticity was noted in the frontal cortex, characterized by an increase in dendritic spines. | DOI exhibits potential as an antidepressant, particularly in reducing behavioral despair without affecting anxiety. Its effects on fear extinction and synaptic plasticity via 5-HT2A receptor activation support its therapeutic potential for anxiety disorders. |
Nardou et al., 2023 [64] doi:10.1038/s41586-023-06204-3 | Social conditioned place preference (sCPP) test in adult male mice treated with psilocybin (0.3 mg/kg), LSD (1 µg/kg), ketamine (3 mg/kg), ibogaine (40 mg/kg) and saline as a control. | Psychedelics enhanced social reward learning in mice. The effects of psilocybin and LSD lasted 2–3 weeks; ibogaine lasted for 4 weeks and ketamine lasted for 1 week. Cocaine had no effect. Psilocybin induced a reopening of the critical period, with a duration similar to MDMA (approximately two weeks), longer than ketamine (approximately 48 h), but shorter than LSD (approximately three weeks) and ibogaine (longer than four weeks). | This suggests that psychedelics may facilitate social reward learning, with potential therapeutic applications in autism, PTSD and depression. |
Werle et al., 2024 [84] doi:10.1111/bph.16315 | Contextual fear conditioning in Wistar rats treated with ayahuasca (0.3 mg/kg of DMT). Antagonists for 5-HT2A and 5-HT1A receptors were infused into the infralimbic cortex. | Ayahuasca enhanced fear extinction without affecting anxiety or exploratory behavior. It facilitated conditioned fear extinction in rats through a mechanism dependent on the activation of 5-HT2A and 5-HT1A receptors in the infralimbic cortex. Effects were blocked by 5-HT2A and 5-HT1A receptor antagonists. | Ayahuasca enhances fear extinction without affecting anxiety or exploratory behavior. Effects are blocked by 5-HT2A and 5-HT1A receptor antagonists. |
Šabanović et al., 2024 [104] doi:10.1038/s41380-024-02439-2 | Male C57BL/6J mice treated with DOI (2 mg/kg). Behavioral assays included a probabilistic reversal learning task. Ex vivo MRI was used to assess brain structural changes 24 h post-treatment. | DOI enhanced cognitive flexibility, with mice adapting faster to task changes and showing novel strategies in response to reward omissions. MRI revealed increased brain volume in sensory and association areas, indicating structural plasticity. | DOI’s enhancement of cognitive flexibility and ability to modify learning strategies underscores its potential utility in treating conditions characterized by rigid cognitive patterns, such as OCD or addiction. |
Kelly et al., 2024 [67] doi:10.1038/s41386-023-01744-8 | Fear conditioning and extinction training in C57BL/6J mice treated with the psilocin analogue 4-Hydroxy-NN-diisopropyltryptamine (4-OH-DiPT) (1 mg/kg or 3 mg/kg). Anxiety tests included open field, light–dark box, elevated plus maze and novelty-suppressed feeding tests. | 4-OH-DiPT promoted fear extinction and spontaneous inhibitory post-synaptic currents (sIPSC) in the basolateral amygdala. Females exhibited significant reductions in freezing and avoidance behaviors. No significant effects were observed in male mice. | Highlights the sex-dependent effects of 4-OH-DiPT on fear extinction and anxiety responses. |
Hinchcliffe et al., 2024 [69] doi:10.1126/scitranslmed.adi2403 | Affective bias test (ABT) in Lister hooded rats treated with psilocybin and ketamine. Additional substances, such as scopolamine, were also tested. | Psilocybin and ketamine altered affective bias, while high doses of ketamine induced cognitive impairment with slower decision-making. Psilocybin also demonstrated “re-learning effects”, not only reducing negative biases similarly to ketamine but also uniquely increasing positive biases. | Demonstrates dose-dependent effects of psilocybin on emotional processing and cognitive function. |
Sekssaoui et al., 2024 [70] doi:10.1038/s41386-024-01794-6 | WT mice subjected to chronic despair conditioning protocol were randomized into DOI 1 mg/kg, psilocybin 1 mg/kg, lisuride 1 mg/kg or vehicle i.p. injection. 5-HT2A knock-out mice subjected to chronic despair conditioning protocol were randomized into psilocybin 1 mg/kg i.p., WAY-100635 0.5 mg/kg + psilocybin 1 mg/kg i.p., SCH23390 0.03 mg/kg, s.c. + psilocybin 1 mg/kg i.p., eticlopride + psilocybin 1 mg/kg i.p., DOI 0.05 mg/kg/day for 6 days or psilocybin 0.05 mg/kg/day for 6 days. Behavioral assays included novelty-suppressed feeding, sucrose preference and forced swim test. | Psilocybin, DOI and lisuride improved depressive-like behaviors in WT mice. DOI and lisuride failed to induce these effects in 5-HT2A receptor knockout mice. Psilocybin induced antidepressant effects in 5-HT2A knock-out mice and in mice pretreated with WAY-100635, SCH23390 or eticlopride. | Findings suggests that the antidepressant effects of psilocybin, oppositely to DOI and lisuride, might be partly independent of 5-HT2A receptor activation. |
Takaba et al., 2024 [71] doi:10.1007/s00210-023-02778-x | Male C57BL/6J mice (7–9 weeks old) were randomized into i.p. administration of psilocin, DOI, TCB-2 and vehicle. FST, TST and novelty-suppressed feeding test (NSFT) were performed 24 h after administration. Each test was repeated every week for a month after administration. | Treatment with psilocin, DOI and TCB-2 significantly reduced immobility times in the FST and TST, compared to vehicle.Volinanserin, a 5-HT2A antagonist, abolished these effects. The effects induced by psilocin endured for at least three weeks. Psilocin also reduced the latency to feed in the NSFT. Pretreatment with volinanserin did not diminish this efect. In contrast, DOI and TCB did not show this property. | Results show that the antidepressant and anxiolytic properties of psychedelics might involve different mechanisms. Only psilocin exerted anxiolytic effects in the NSFT, and these were not affected by volinanserin, suggesting a 5-HT2A-independent mechanism. Moreover, only psilocin produced sustained antidepressant effects in the FST and TST for three weeks. |
3.2. Clinical Evidence
3.2.1. Neuroimaging
3.2.2. Electrophysiology
3.2.3. Biohumoral Markers
3.2.4. Behavioral Studies
3.2.5. Case Reports
4. Discussion
Author Contributions
Funding
Conflicts of Interest
References
- Sulser, F.; Vetulani, J.; Mobley, P.L. Mode of Action of Antidepressant Drugs. Biochem. Pharmacol. 1978, 27, 257–261. [Google Scholar] [CrossRef] [PubMed]
- Charney, D.S.; Menkes, D.B.; Heninger, G.R. Receptor Sensitivity and the Mechanism of Action of Antidepressant Treatment. Implications for the Etiology and Therapy of Depression. Arch. Gen. Psychiatry 1981, 38, 1160–1180. [Google Scholar] [CrossRef] [PubMed]
- Stahl, S.M. Mechanism of Action of Serotonin Selective Reuptake Inhibitors. Serotonin Receptors and Pathways Mediate Therapeutic Effects and Side Effects. J. Affect. Disord. 1998, 51, 215–235. [Google Scholar] [CrossRef]
- Hyman, S.E.; Nestler, E.J. Initiation and Adaptation: A Paradigm for Understanding Psychotropic Drug Action. Am. J. Psychiatry 1996, 153, 151–162. [Google Scholar] [CrossRef] [PubMed]
- Nibuya, M.; Morinobu, S.; Duman, R.S. Regulation of BDNF and TrkB MRNA in Rat Brain by Chronic Electroconvulsive Seizure and Antidepressant Drug Treatments. J. Neurosci. Off. J. Soc. Neurosci. 1995, 15, 7539–7547. [Google Scholar] [CrossRef] [PubMed]
- Russo-Neustadt, A.A.; Beard, R.C.; Huang, Y.M.; Cotman, C.W. Physical Activity and Antidepressant Treatment Potentiate the Expression of Specific Brain-Derived Neurotrophic Factor Transcripts in the Rat Hippocampus. Neuroscience 2000, 101, 305–312. [Google Scholar] [CrossRef]
- Skaper, S.D. The Neurotrophin Family of Neurotrophic Factors: An Overview BT—Neurotrophic Factors: Methods and Protocols; Skaper, S.D., Ed.; Humana Press: Totowa, NJ, USA, 2012; pp. 1–12. ISBN 978-1-61779-536-7. [Google Scholar]
- Wang, C.S.; Kavalali, E.T.; Monteggia, L.M. BDNF Signaling in Context: From Synaptic Regulation to Psychiatric Disorders. Cell 2022, 185, 62–76. [Google Scholar] [CrossRef]
- Karege, F.; Perret, G.; Bondolfi, G.; Schwald, M.; Bertschy, G.; Aubry, J.-M. Decreased Serum Brain-Derived Neurotrophic Factor Levels in Major Depressed Patients. Psychiatry Res. 2002, 109, 143–148. [Google Scholar] [CrossRef] [PubMed]
- Karege, F.; Bondolfi, G.; Gervasoni, N.; Schwald, M.; Aubry, J.-M.; Bertschy, G. Low Brain-Derived Neurotrophic Factor (BDNF) Levels in Serum of Depressed Patients Probably Results from Lowered Platelet BDNF Release Unrelated to Platelet Reactivity. Biol. Psychiatry 2005, 57, 1068–1072. [Google Scholar] [CrossRef] [PubMed]
- Gonul, A.S.; Akdeniz, F.; Taneli, F.; Donat, O.; Eker, C.; Vahip, S. Effect of Treatment on Serum Brain-Derived Neurotrophic Factor Levels in Depressed Patients. Eur. Arch. Psychiatry Clin. Neurosci. 2005, 255, 381–386. [Google Scholar] [CrossRef]
- Kim, Y.-K.; Lee, H.-P.; Won, S.-D.; Park, E.-Y.; Lee, H.-Y.; Lee, B.-H.; Lee, S.-W.; Yoon, D.; Han, C.; Kim, D.-J.; et al. Low Plasma BDNF Is Associated with Suicidal Behavior in Major Depression. Prog. Neuropsychopharmacol. Biol. Psychiatry 2007, 31, 78–85. [Google Scholar] [CrossRef] [PubMed]
- Hauser, P.; Altshuler, L.L.; Berrettini, W.; Dauphinais, I.D.; Gelernter, J.; Post, R.M. Temporal Lobe Measurement in Primary Affective Disorder by Magnetic Resonance Imaging. J. Neuropsychiatry Clin. Neurosci. 1989, 1, 128–134. [Google Scholar] [CrossRef] [PubMed]
- Sheline, Y.I.; Wang, P.W.; Gado, M.H.; Csernansky, J.G.; Vannier, M.W. Hippocampal Atrophy in Recurrent Major Depression. Proc. Natl. Acad. Sci. USA 1996, 93, 3908–3913. [Google Scholar] [CrossRef] [PubMed]
- Bremner, J.D.; Narayan, M.; Anderson, E.R.; Staib, L.H.; Miller, H.L.; Charney, D.S. Hippocampal Volume Reduction in Major Depression. Am. J. Psychiatry 2000, 157, 115–118. [Google Scholar] [CrossRef] [PubMed]
- Mervaala, E.; Föhr, J.; Könönen, M.; Valkonen-Korhonen, M.; Vainio, P.; Partanen, K.; Partanen, J.; Tiihonen, J.; Viinamäki, H.; Karjalainen, A.K.; et al. Quantitative MRI of the Hippocampus and Amygdala in Severe Depression. Psychol. Med. 2000, 30, 117–125. [Google Scholar] [CrossRef]
- Kempermann, G. Regulation of Adult Hippocampal Neurogenesis—Implications for Novel Theories of Major Depression. Bipolar Disord. 2002, 4, 17–33. [Google Scholar] [CrossRef]
- Castrén, E. Is Mood Chemistry? Nat. Rev. Neurosci. 2005, 6, 241–246. [Google Scholar] [CrossRef]
- Benusková, L. Antidepressants and Synaptic Plasticity: A Hypothesis. Med. Hypotheses 1991, 35, 17–22. [Google Scholar] [CrossRef] [PubMed]
- Duman, R.S.; Malberg, J.; Thome, J. Neural Plasticity to Stress and Antidepressant Treatment. Biol. Psychiatry 1999, 46, 1181–1191. [Google Scholar] [CrossRef]
- D’Sa, C.; Duman, R.S. Antidepressants and Neuroplasticity. Bipolar Disord. 2002, 4, 183–194. [Google Scholar] [CrossRef]
- Castrén, E. Neurotrophic Effects of Antidepressant Drugs. Curr. Opin. Pharmacol. 2004, 4, 58–64. [Google Scholar] [CrossRef] [PubMed]
- Maya Vetencourt, J.F.; Sale, A.; Viegi, A.; Baroncelli, L.; De Pasquale, R.; O’Leary, O.F.; Castrén, E.; Maffei, L. The Antidepressant Fluoxetine Restores Plasticity in the Adult Visual Cortex. Science 2008, 320, 385–388. [Google Scholar] [CrossRef]
- Maya Vetencourt, J.F.; Tiraboschi, E.; Spolidoro, M.; Castrén, E.; Maffei, L. Serotonin Triggers a Transient Epigenetic Mechanism That Reinstates Adult Visual Cortex Plasticity in Rats. Eur. J. Neurosci. 2011, 33, 49–57. [Google Scholar] [CrossRef] [PubMed]
- O’Leary, O.F.; Wu, X.; Castren, E. Chronic Fluoxetine Treatment Increases Expression of Synaptic Proteins in the Hippocampus of the Ovariectomized Rat: Role of BDNF Signalling. Psychoneuroendocrinology 2009, 34, 367–381. [Google Scholar] [CrossRef]
- Castrén, E. Neuronal Network Plasticity and Recovery from Depression. JAMA Psychiatry 2013, 70, 983–989. [Google Scholar] [CrossRef] [PubMed]
- Castrén, E.; Antila, H. Neuronal Plasticity and Neurotrophic Factors in Drug Responses. Mol. Psychiatry 2017, 22, 1085–1095. [Google Scholar] [CrossRef]
- Umemori, J.; Winkel, F.; Didio, G.; Llach Pou, M.; Castrén, E. IPlasticity: Induced Juvenile-like Plasticity in the Adult Brain as a Mechanism of Antidepressants. Psychiatry Clin. Neurosci. 2018, 72, 633–653. [Google Scholar] [CrossRef] [PubMed]
- Berman, R.M.; Cappiello, A.; Anand, A.; Oren, D.A.; Heninger, G.R.; Charney, D.S.; Krystal, J.H. Antidepressant Effects of Ketamine in Depressed Patients. Biol. Psychiatry 2000, 47, 351–354. [Google Scholar] [CrossRef]
- Zarate Jr, C.A.; Singh, J.B.; Carlson, P.J.; Brutsche, N.E.; Ameli, R.; Luckenbaugh, D.A.; Charney, D.S.; Manji, H.K. A Randomized Trial of an N-Methyl-D-Aspartate Antagonist in Treatment-Resistant Major Depression. Arch. Gen. Psychiatry 2006, 63, 856–864. [Google Scholar] [CrossRef] [PubMed]
- Witkin, J.M.; Martin, A.E.; Golani, L.K.; Xu, N.Z.; Smith, J.L. Rapid-Acting Antidepressants. Adv. Pharmacol. 2019, 86, 47–96. [Google Scholar] [CrossRef] [PubMed]
- Autry, A.E.; Adachi, M.; Nosyreva, E.; Na, E.S.; Los, M.F.; Cheng, P.; Kavalali, E.T.; Monteggia, L.M. NMDA Receptor Blockade at Rest Triggers Rapid Behavioural Antidepressant Responses. Nature 2011, 475, 91–95. [Google Scholar] [CrossRef] [PubMed]
- Yang, C.; Hu, Y.-M.; Zhou, Z.-Q.; Zhang, G.-F.; Yang, J.-J. Acute Administration of Ketamine in Rats Increases Hippocampal BDNF and MTOR Levels during Forced Swimming Test. Ups. J. Med. Sci. 2013, 118, 3–8. [Google Scholar] [CrossRef] [PubMed]
- Zhou, W.; Wang, N.; Yang, C.; Li, X.-M.; Zhou, Z.-Q.; Yang, J.-J. Ketamine-Induced Antidepressant Effects Are Associated with AMPA Receptors-Mediated Upregulation of MTOR and BDNF in Rat Hippocampus and Prefrontal Cortex. Eur. Psychiatry 2014, 29, 419–423. [Google Scholar] [CrossRef]
- Yang, C.; Shirayama, Y.; Zhang, J.; Ren, Q.; Yao, W.; Ma, M.; Dong, C.; Hashimoto, K. R-Ketamine: A Rapid-Onset and Sustained Antidepressant without Psychotomimetic Side Effects. Transl. Psychiatry 2015, 5, e632. [Google Scholar] [CrossRef] [PubMed]
- Jaffe, R.J.; Novakovic, V.; Peselow, E.D. Scopolamine as an Antidepressant: A Systematic Review. Clin. Neuropharmacol. 2013, 36, 24–26. [Google Scholar] [CrossRef] [PubMed]
- Gründer, G.; Brand, M.; Kärtner, L.; Scharf, D.; Schmitz, C.; Spangemacher, M.; Mertens, L.J. Sind Psychedelika Schnell Wirksame Antidepressiva? Nervenarzt 2022, 93, 254–262. [Google Scholar] [CrossRef]
- Ghosal, S.; Bang, E.; Yue, W.; Hare, B.D.; Lepack, A.E.; Girgenti, M.J.; Duman, R.S. Activity-Dependent Brain-Derived Neurotrophic Factor Release Is Required for the Rapid Antidepressant Actions of Scopolamine. Biol. Psychiatry 2018, 83, 29–37. [Google Scholar] [CrossRef]
- Ly, C.; Greb, A.C.; Cameron, L.P.; Wong, J.M.; Barragan, E.V.; Wilson, P.C.; Burbach, K.F.; Soltanzadeh Zarandi, S.; Sood, A.; Paddy, M.R.; et al. Psychedelics Promote Structural and Functional Neural Plasticity. Cell Rep. 2018, 23, 3170–3182. [Google Scholar] [CrossRef]
- Olson, D.E. Psychoplastogens: A Promising Class of Plasticity-Promoting Neurotherapeutics. J. Exp. Neurosci. 2018, 12, 1179069518800508. [Google Scholar] [CrossRef]
- Mertens, L.J.; Koslowski, M.; Betzler, F.; Evens, R.; Gilles, M.; Jungaberle, A.; Jungaberle, H.; Majić, T.; Ströhle, A.; Wolff, M.; et al. Methodological Challenges in Psychedelic Drug Trials: Efficacy and Safety of Psilocybin in Treatment-Resistant Major Depression (EPIsoDE)—Rationale and Study Design. Neurosci. Appl. 2022, 1, 100104. [Google Scholar] [CrossRef]
- van Elk, M.; Fried, E.I. History Repeating: Guidelines to Address Common Problems in Psychedelic Science. Ther. Adv. Psychopharmacol. 2023, 13, 20451253231198464. [Google Scholar] [CrossRef]
- Evans, J.; Robinson, O.C.; Argyri, E.K.; Suseelan, S.; Murphy-Beiner, A.; McAlpine, R.; Luke, D.; Michelle, K.; Prideaux, E. Extended Difficulties Following the Use of Psychedelic Drugs: A Mixed Methods Study. PLoS ONE 2023, 18, e0293349. [Google Scholar] [CrossRef]
- Olson, D.E. Biochemical Mechanisms Underlying Psychedelic-Induced Neuroplasticity. Biochemistry 2022, 61, 127–136. [Google Scholar] [CrossRef]
- Ouzzani, M.; Hammady, H.; Fedorowicz, Z.; Elmagarmid, A. Rayyan—A Web and Mobile App for Systematic Reviews. Syst. Rev. 2016, 5, 210. [Google Scholar] [CrossRef]
- Gonzales, S.A.B.; Alexopoulos, C.; Arkfeld, D.G. Potential Benefits of Psilocybin for Lupus Pain: A Case Report. Curr. Rheumatol. Rev. 2024, 20, 97–99. [Google Scholar] [CrossRef] [PubMed]
- Zheng, Z.-H.; Lin, X.-C.; Lu, Y.; Cao, S.-R.; Liu, X.-K.; Lin, D.; Yang, F.-H.; Zhang, Y.-B.; Tu, J.-L.; Pan, B.-X.; et al. Harmine Exerts Anxiolytic Effects by Regulating Neuroinflammation and Neuronal Plasticity in the Basolateral Amygdala. Int. Immunopharmacol. 2023, 119, 110208. [Google Scholar] [CrossRef]
- Morales-García, J.A.; de la Fuente Revenga, M.; Alonso-Gil, S.; Rodríguez-Franco, M.I.; Feilding, A.; Perez-Castillo, A.; Riba, J. The Alkaloids of Banisteriopsis Caapi, the Plant Source of the Amazonian Hallucinogen Ayahuasca, Stimulate Adult Neurogenesis in Vitro. Sci. Rep. 2017, 7, 5309. [Google Scholar] [CrossRef]
- Pei, Q.; Lewis, L.; Sprakes, M.E.; Jones, E.J.; Grahame-Smith, D.G.; Zetterström, T.S. Serotonergic Regulation of MRNA Expression of Arc, an Immediate Early Gene Selectively Localized at Neuronal Dendrites. Neuropharmacology 2000, 39, 463–470. [Google Scholar] [CrossRef] [PubMed]
- Alkadhi, K.A.; Salgado-Commissariat, D.; Hogan, Y.H.; Akpaudo, S.B. Induction and Maintenance of Ganglionic Long-Term Potentiation Require Activation of 5-Hydroxytryptamine (5-HT3) Receptors. J. Physiol. 1996, 496 Pt 2, 479–489. [Google Scholar] [CrossRef] [PubMed]
- Edagawa, Y.; Saito, H.; Abe, K. The Serotonin 5-HT2 Receptor-Phospholipase C System Inhibits the Induction of Long-Term Potentiation in the Rat Visual Cortex. Eur. J. Neurosci. 2000, 12, 1391–1396. [Google Scholar] [CrossRef]
- Raval, N.R.; Johansen, A.; Donovan, L.L.; Ros, N.F.; Ozenne, B.; Hansen, H.D.; Knudsen, G.M. A Single Dose of Psilocybin Increases Synaptic Density and Decreases 5-HT(2A) Receptor Density in the Pig Brain. Int. J. Mol. Sci. 2021, 22, 835. [Google Scholar] [CrossRef] [PubMed]
- Shao, L.-X.; Liao, C.; Gregg, I.; Davoudian, P.A.; Savalia, N.K.; Delagarza, K.; Kwan, A.C. Psilocybin Induces Rapid and Persistent Growth of Dendritic Spines in Frontal Cortex in Vivo. Neuron 2021, 109, 2535–2544.e4. [Google Scholar] [CrossRef] [PubMed]
- Jefsen, O.H.; Elfving, B.; Wegener, G.; Müller, H.K. Transcriptional Regulation in the Rat Prefrontal Cortex and Hippocampus after a Single Administration of Psilocybin. J. Psychopharmacol. 2021, 35, 483–493. [Google Scholar] [CrossRef]
- Fadahunsi, N.; Lund, J.; Breum, A.W.; Mathiesen, C.V.; Larsen, I.B.; Knudsen, G.M.; Klein, A.B.; Clemmensen, C. Acute and Long-Term Effects of Psilocybin on Energy Balance and Feeding Behavior in Mice. Transl. Psychiatry 2022, 12, 330. [Google Scholar] [CrossRef]
- Thomas, C.W.; Blanco-Duque, C.; Bréant, B.J.; Goodwin, G.M.; Sharp, T.; Bannerman, D.M.; Vyazovskiy, V. V Psilocin Acutely Alters Sleep-Wake Architecture and Cortical Brain Activity in Laboratory Mice. Transl. Psychiatry 2022, 12, 77. [Google Scholar] [CrossRef]
- Avvenuti, G.; Bernardi, G. Chapter 3—Local Sleep: A New Concept in Brain Plasticity. In Handbook of Clinical Neurology; Quartarone, A., Ghilardi, M.F., Boller, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; Volume 184, pp. 35–52. ISBN 0072-9752. [Google Scholar]
- Heck, D.H.; Kozma, R.; Kay, L.M. The Rhythm of Memory: How Breathing Shapes Memory Function. J. Neurophysiol. 2019, 122, 563–571. [Google Scholar] [CrossRef]
- Liu, J.; Wang, Y.; Xia, K.; Wu, J.; Zheng, D.; Cai, A.; Yan, H.; Su, R. Acute Psilocybin Increased Cortical Activities in Rats. Front. Neurosci. 2023, 17, 1168911. [Google Scholar] [CrossRef] [PubMed]
- Du, Y.; Li, Y.; Zhao, X.; Yao, Y.; Wang, B.; Zhang, L.; Wang, G. Psilocybin Facilitates Fear Extinction in Mice by Promoting Hippocampal Neuroplasticity. Chin. Med. J. 2023, 136, 2983–2992. [Google Scholar] [CrossRef] [PubMed]
- Davoudian, P.A.; Shao, L.-X.; Kwan, A.C. Shared and Distinct Brain Regions Targeted for Immediate Early Gene Expression by Ketamine and Psilocybin. ACS Chem. Neurosci. 2023, 14, 468–480. [Google Scholar] [CrossRef] [PubMed]
- Rijsketic, D.R.; Casey, A.B.; Barbosa, D.A.N.; Zhang, X.; Hietamies, T.M.; Ramirez-Ovalle, G.; Pomrenze, M.B.; Halpern, C.H.; Williams, L.M.; Malenka, R.C.; et al. UNRAVELing the Synergistic Effects of Psilocybin and Environment on Brain-Wide Immediate Early Gene Expression in Mice. Neuropsychopharmacology 2023, 48, 1798–1807. [Google Scholar] [CrossRef] [PubMed]
- Nardou, R.; Sawyer, E.; Song, Y.J.; Wilkinson, M.; Padovan-Hernandez, Y.; de Deus, J.L.; Wright, N.; Lama, C.; Faltin, S.; Goff, L.A.; et al. Psychedelics Reopen the Social Reward Learning Critical Period. Nature 2023, 618, 790–798. [Google Scholar] [CrossRef] [PubMed]
- Funk, D.; Araujo, J.; Slassi, M.; Lanthier, J.; Atkinson, J.; Feng, D.; Lau, W.; Lê, A.; Higgins, G.A. Effect of a Single Psilocybin Treatment on Fos Protein Expression in Male Rat Brain. Neuroscience 2024, 539, 1–11. [Google Scholar] [CrossRef] [PubMed]
- Shahar, O.; Botvinnik, A.; Shwartz, A.; Lerer, E.; Golding, P.; Buko, A.; Hamid, E.; Kahn, D.; Guralnick, M.; Blakolmer, K.; et al. Effect of Chemically Synthesized Psilocybin and Psychedelic Mushroom Extract on Molecular and Metabolic Profiles in Mouse Brain. Mol. Psychiatry 2024, 29, 2059–2073. [Google Scholar] [CrossRef] [PubMed]
- Lerer, E.; Botvinnik, A.; Shahar, O.; Grad, M.; Blakolmer, K.; Shomron, N.; Lotan, A.; Lerer, B.; Lifschytz, T. Effects of Psilocybin, Psychedelic Mushroom Extract and 5-Hydroxytryptophan on Brain Immediate Early Gene Expression: Interaction with Serotonergic Receptor Modulators. Front. Pharmacol. 2024, 15, 1391412. [Google Scholar] [CrossRef]
- Kelly, T.J.; Bonniwell, E.M.; Mu, L.; Liu, X.; Hu, Y.; Friedman, V.; Yu, H.; Su, W.; McCorvy, J.D.; Liu, Q.-S. Psilocybin Analog 4-OH-DiPT Enhances Fear Extinction and GABAergic Inhibition of Principal Neurons in the Basolateral Amygdala. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2024, 49, 854–863. [Google Scholar] [CrossRef]
- Zhao, X.; Du, Y.; Yao, Y.; Dai, W.; Yin, Y.; Wang, G.; Li, Y.; Zhang, L. Psilocybin Promotes Neuroplasticity and Induces Rapid and Sustained Antidepressant-like Effects in Mice. J. Psychopharmacol. 2024, 38, 489–499. [Google Scholar] [CrossRef] [PubMed]
- Hinchcliffe, J.K.; Stuart, S.A.; Wood, C.M.; Bartlett, J.; Kamenish, K.; Arban, R.; Thomas, C.W.; Selimbeyoglu, A.; Hurley, S.; Hengerer, B.; et al. Rapid-Acting Antidepressant Drugs Modulate Affective Bias in Rats. Sci. Transl. Med. 2024, 16, eadi2403. [Google Scholar] [CrossRef] [PubMed]
- Sekssaoui, M.; Bockaert, J.; Marin, P.; Bécamel, C. Antidepressant-like Effects of Psychedelics in a Chronic Despair Mouse Model: Is the 5-HT(2A) Receptor the Unique Player? Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2024, 49, 747–756. [Google Scholar] [CrossRef] [PubMed]
- Takaba, R.; Ibi, D.; Yoshida, K.; Hosomi, E.; Kawase, R.; Kitagawa, H.; Goto, H.; Achiwa, M.; Mizutani, K.; Maeda, K.; et al. Ethopharmacological Evaluation of Antidepressant-like Effect of Serotonergic Psychedelics in C57BL/6J Male Mice. Naunyn. Schmiedebergs. Arch. Pharmacol. 2024, 397, 3019–3035. [Google Scholar] [CrossRef]
- Aboharb, F.; Davoudian, P.A.; Shao, L.-X.; Liao, C.; Rzepka, G.N.; Wojtasiewicz, C.; Dibbs, M.; Rondeau, J.; Sherwood, A.M.; Kaye, A.P.; et al. Classification of Psychedelic Drugs Based on Brain-Wide Imaging of Cellular c-Fos Expression. bioRxiv Prepr. Serv. Biol. 2024. [Google Scholar] [CrossRef]
- Dakic, V.; Minardi Nascimento, J.; Costa Sartore, R.; Maciel, R.d.M.; de Araujo, D.B.; Ribeiro, S.; Martins-de-Souza, D.; Rehen, S.K. Short Term Changes in the Proteome of Human Cerebral Organoids Induced by 5-MeO-DMT. Sci. Rep. 2017, 7, 12863. [Google Scholar] [CrossRef] [PubMed]
- Vargas-Perez, H.; Grieder, T.E.; Ting-A-Kee, R.; Maal-Bared, G.; Chwalek, M.; van der Kooy, D. A Single Administration of the Hallucinogen, 4-Acetoxy-Dimethyltryptamine, Prevents the Shift to a Drug-Dependent State and the Expression of Withdrawal Aversions in Rodents. Eur. J. Neurosci. 2017, 45, 1410–1417. [Google Scholar] [CrossRef]
- Lima da Cruz, R.V.; Moulin, T.C.; Petiz, L.L.; Leão, R.N. A Single Dose of 5-MeO-DMT Stimulates Cell Proliferation, Neuronal Survivability, Morphological and Functional Changes in Adult Mice Ventral Dentate Gyrus. Front. Mol. Neurosci. 2018, 11, 312. [Google Scholar] [CrossRef]
- Cameron, L.P.; Benson, C.J.; DeFelice, B.C.; Fiehn, O.; Olson, D.E. Chronic, Intermittent Microdoses of the Psychedelic N,N-Dimethyltryptamine (DMT) Produce Positive Effects on Mood and Anxiety in Rodents. ACS Chem. Neurosci. 2019, 10, 3261–3270. [Google Scholar] [CrossRef]
- Nardai, S.; László, M.; Szabó, A.; Alpár, A.; Hanics, J.; Zahola, P.; Merkely, B.; Frecska, E.; Nagy, Z. N,N-Dimethyltryptamine Reduces Infarct Size and Improves Functional Recovery Following Transient Focal Brain Ischemia in Rats. Exp. Neurol. 2020, 327, 113245. [Google Scholar] [CrossRef]
- Almeida, C.A.F.; Pereira-Junior, A.A.; Rangel, J.G.; Pereira, B.P.; Costa, K.C.M.; Bruno, V.; Silveira, G.O.; Ceron, C.S.; Yonamine, M.; Camarini, R.; et al. Ayahuasca, a Psychedelic Beverage, Modulates Neuroplasticity Induced by Ethanol in Mice. Behav. Brain Res. 2022, 416, 113546. [Google Scholar] [CrossRef] [PubMed]
- Daneluz, D.M.; Sohn, J.M.B.; Silveira, G.O.; Yonamine, M.; Stern, C.A. Evidence on the Impairing Effects of Ayahuasca on Fear Memory Reconsolidation. Psychopharmacology 2022, 239, 3325–3336. [Google Scholar] [CrossRef] [PubMed]
- Kelley, D.P.; Venable, K.; Destouni, A.; Billac, G.; Ebenezer, P.; Stadler, K.; Nichols, C.; Barker, S.; Francis, J. Pharmahuasca and DMT Rescue ROS Production and Differentially Expressed Genes Observed after Predator and Psychosocial Stress: Relevance to Human PTSD. ACS Chem. Neurosci. 2022, 13, 257–274. [Google Scholar] [CrossRef] [PubMed]
- Vargas, M.V.; Dunlap, L.E.; Dong, C.; Carter, S.J.; Tombari, R.J.; Jami, S.A.; Cameron, L.P.; Patel, S.D.; Hennessey, J.J.; Saeger, H.N.; et al. Psychedelics Promote Neuroplasticity through the Activation of Intracellular 5-HT2A Receptors. Science 2023, 379, 700–706. [Google Scholar] [CrossRef]
- Jefferson, S.J.; Gregg, I.; Dibbs, M.; Liao, C.; Wu, H.; Davoudian, P.A.; Woodburn, S.C.; Wehrle, P.H.; Sprouse, J.S.; Sherwood, A.M.; et al. 5-MeO-DMT Modifies Innate Behaviors and Promotes Structural Neural Plasticity in Mice. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2023, 48, 1257–1266. [Google Scholar] [CrossRef] [PubMed]
- Cameron, L.P.; Patel, S.D.; Vargas, M.V.; Barragan, E.V.; Saeger, H.N.; Warren, H.T.; Chow, W.L.; Gray, J.A.; Olson, D.E. 5-HT2ARs Mediate Therapeutic Behavioral Effects of Psychedelic Tryptamines. ACS Chem. Neurosci. 2023, 14, 351–358. [Google Scholar] [CrossRef]
- Werle, I.; Nascimento, L.M.M.; Dos Santos, A.L.A.; Soares, L.A.; Dos Santos, R.G.; Hallak, J.E.C.; Bertoglio, L.J. Ayahuasca-Enhanced Extinction of Fear Behaviour: Role of Infralimbic Cortex 5-HT(2A) and 5-HT(1A) Receptors. Br. J. Pharmacol. 2024, 181, 1671–1689. [Google Scholar] [CrossRef] [PubMed]
- Nichols, C.D.; Sanders-Bush, E. A Single Dose of Lysergic Acid Diethylamide Influences Gene Expression Patterns within the Mammalian Brain. Neuropsychopharmacology 2002, 26, 634–642. [Google Scholar] [CrossRef] [PubMed]
- Nichols, C.D.; Garcia, E.E.; Sanders-Bush, E. Dynamic Changes in Prefrontal Cortex Gene Expression Following Lysergic Acid Diethylamide Administration. Brain Res. Mol. Brain Res. 2003, 111, 182–188. [Google Scholar] [CrossRef] [PubMed]
- Nichols, C.D.; Sanders-Bush, E. Molecular Genetic Responses to Lysergic Acid Diethylamide Include Transcriptional Activation of MAP Kinase Phosphatase-1, C/EBP-Beta and ILAD-1, a Novel Gene with Homology to Arrestins. J. Neurochem. 2004, 90, 576–584. [Google Scholar] [CrossRef] [PubMed]
- Jha, S.; Rajendran, R.; Fernandes, K.A.; Vaidya, V.A. 5-HT2A/2C Receptor Blockade Regulates Progenitor Cell Proliferation in the Adult Rat Hippocampus. Neurosci. Lett. 2008, 441, 210–214. [Google Scholar] [CrossRef]
- Martin, D.A.; Marona-Lewicka, D.; Nichols, D.E.; Nichols, C.D. Chronic LSD Alters Gene Expression Profiles in the MPFC Relevant to Schizophrenia. Neuropharmacology 2014, 83, 1–8. [Google Scholar] [CrossRef] [PubMed]
- Ly, C.; Greb, A.C.; Vargas, M.V.; Duim, W.C.; Grodzki, A.C.G.; Lein, P.J.; Olson, D.E. Transient Stimulation with Psychoplastogens Is Sufficient to Initiate Neuronal Growth. ACS Pharmacol. Transl. Sci. 2021, 4, 452–460. [Google Scholar] [CrossRef]
- Savino, A.; Nichols, C.D. Lysergic Acid Diethylamide Induces Increased Signalling Entropy in Rats’ Prefrontal Cortex. J. Neurochem. 2022, 162, 9–23. [Google Scholar] [CrossRef]
- Ornelas, I.M.; Cini, F.A.; Wießner, I.; Marcos, E.; Araújo, D.B.; Goto-Silva, L.; Nascimento, J.; Silva, S.R.B.; Costa, M.N.; Falchi, M.; et al. Nootropic Effects of LSD: Behavioral, Molecular and Computational Evidence. Exp. Neurol. 2022, 356, 114148. [Google Scholar] [CrossRef]
- Inserra, A.; Campanale, A.; Cheishvili, D.; Dymov, S.; Wong, A.; Marcal, N.; Syme, R.A.; Taylor, L.; De Gregorio, D.; Kennedy, T.E.; et al. Modulation of DNA Methylation and Protein Expression in the Prefrontal Cortex by Repeated Administration of D-Lysergic Acid Diethylamide (LSD): Impact on Neurotropic, Neurotrophic, and Neuroplasticity Signaling. Prog. Neuropsychopharmacol. Biol. Psychiatry 2022, 119, 110594. [Google Scholar] [CrossRef] [PubMed]
- Moliner, R.; Girych, M.; Brunello, C.A.; Kovaleva, V.; Biojone, C.; Enkavi, G.; Antenucci, L.; Kot, E.F.; Goncharuk, S.A.; Kaurinkoski, K.; et al. Psychedelics Promote Plasticity by Directly Binding to BDNF Receptor TrkB. Nat. Neurosci. 2023, 26, 1032–1041. [Google Scholar] [CrossRef] [PubMed]
- Padró, J.; De Panis, D.N.; Luisi, P.; Dopazo, H.; Szajnman, S.; Hasson, E.; Soto, I.M. Ortholog Genes from Cactophilic Drosophila Provide Insight into Human Adaptation to Hallucinogenic Cacti. Sci. Rep. 2022, 12, 13180. [Google Scholar] [CrossRef]
- Benekareddy, M.; Nair, A.R.; Dias, B.G.; Suri, D.; Autry, A.E.; Monteggia, L.M.; Vaidya, V.A. Induction of the Plasticity-Associated Immediate Early Gene Arc by Stress and Hallucinogens: Role of Brain-Derived Neurotrophic Factor. Int. J. Neuropsychopharmacol. 2013, 16, 405–415. [Google Scholar] [CrossRef]
- Hu, L.; Liu, C.; Dang, M.; Luo, B.; Guo, Y.; Wang, H. Activation of 5-HT2A/2C Receptors Reduces the Excitability of Cultured Cortical Neurons. Neurosci. Lett. 2016, 632, 124–129. [Google Scholar] [CrossRef] [PubMed]
- Berthoux, C.; Barre, A.; Bockaert, J.; Marin, P.; Bécamel, C. Sustained Activation of Postsynaptic 5-HT2A Receptors Gates Plasticity at Prefrontal Cortex Synapses. Cereb. Cortex 2019, 29, 1659–1669. [Google Scholar] [CrossRef] [PubMed]
- de la Fuente Revenga, M.; Zhu, B.; Guevara, C.A.; Naler, L.B.; Saunders, J.M.; Zhou, Z.; Toneatti, R.; Sierra, S.; Wolstenholme, J.T.; Beardsley, P.M.; et al. Prolonged Epigenomic and Synaptic Plasticity Alterations Following Single Exposure to a Psychedelic in Mice. Cell Rep. 2021, 37, 109836. [Google Scholar] [CrossRef]
- Desouza, L.A.; Benekareddy, M.; Fanibunda, S.E.; Mohammad, F.; Janakiraman, B.; Ghai, U.; Gur, T.; Blendy, J.A.; Vaidya, V.A. The Hallucinogenic Serotonin(2A) Receptor Agonist, 2,5-Dimethoxy-4-Iodoamphetamine, Promotes CAMP Response Element Binding Protein-Dependent Gene Expression of Specific Plasticity-Associated Genes in the Rodent Neocortex. Front. Mol. Neurosci. 2021, 14, 790213. [Google Scholar] [CrossRef]
- Tang, Z.-H.; Yu, Z.-P.; Li, Q.; Zhang, X.-Q.; Muhetaer, K.; Wang, Z.-C.; Xu, P.; Shen, H.-W. The Effects of Serotonergic Psychedelics in Synaptic and Intrinsic Properties of Neurons in Layer II/III of the Orbitofrontal Cortex. Psychopharmacology 2023, 240, 1275–1285. [Google Scholar] [CrossRef] [PubMed]
- Ko, M.; Jaques, Y.; Khera, R.; Rorie, B.; Wang, C.; Jacobowitz, A.; Kaufer, D. 148. Glial Mechanisms Underlying the Effects of Psychedelics on Mood Behaviors of Mice. Biol. Psychiatry 2023, 93, S154. [Google Scholar] [CrossRef]
- Olson, R.J.; Bartlett, L.; Sonneborn, A.; Milton, R.; Bretton-Granatoor, Z.; Firdous, A.; Harris, A.Z.; Abbas, A.I. Decoupling of Cortical Activity from Behavioral State Following Administration of the Classic Psychedelic DOI. Neuropharmacology 2024, 257, 110030. [Google Scholar] [CrossRef]
- Šabanović, M.; Lazari, A.; Blanco-Pozo, M.; Tisca, C.; Tachrount, M.; Martins-Bach, A.B.; Lerch, J.P.; Walton, M.E.; Bannerman, D.M. Lasting Dynamic Effects of the Psychedelic 2,5-Dimethoxy-4-Iodoamphetamine ((±)-DOI) on Cognitive Flexibility. Mol. Psychiatry 2024, 29, 1810–1823. [Google Scholar] [CrossRef]
- Bouso, J.C.; Palhano-Fontes, F.; Rodríguez-Fornells, A.; Ribeiro, S.; Sanches, R.; Crippa, J.A.S.; Hallak, J.E.C.; de Araujo, D.B.; Riba, J. Long-Term Use of Psychedelic Drugs Is Associated with Differences in Brain Structure and Personality in Humans. Eur. Neuropsychopharmacol. J. Eur. Coll. Neuropsychopharmacol. 2015, 25, 483–492. [Google Scholar] [CrossRef]
- Lebedev, A.V.; Kaelen, M.; Lövdén, M.; Nilsson, J.; Feilding, A.; Nutt, D.J.; Carhart-Harris, R.L. LSD-Induced Entropic Brain Activity Predicts Subsequent Personality Change. Hum. Brain Mapp. 2016, 37, 3203–3213. [Google Scholar] [CrossRef] [PubMed]
- Barrett, F.S.; Doss, M.K.; Sepeda, N.D.; Pekar, J.J.; Griffiths, R.R. Emotions and Brain Function Are Altered up to One Month after a Single High Dose of Psilocybin. Sci. Rep. 2020, 10, 2214. [Google Scholar] [CrossRef] [PubMed]
- Mallaroni, P.; Mason, N.L.; Kloft, L.; Reckweg, J.T.; van Oorsouw, K.; Ramaekers, J.G. Cortical Structural Differences Following Repeated Ayahuasca Use Hold Molecular Signatures. Front. Neurosci. 2023, 17, 1217079. [Google Scholar] [CrossRef] [PubMed]
- Siegel, J.S.; Subramanian, S.; Perry, D.; Kay, B.P.; Gordon, E.M.; Laumann, T.O.; Reneau, T.R.; Metcalf, N.V.; Chacko, R.V.; Gratton, C.; et al. Psilocybin Desynchronizes the Human Brain. Nature 2024, 632, 131–138. [Google Scholar] [CrossRef] [PubMed]
- Timmermann, C.; Spriggs, M.J.; Kaelen, M.; Leech, R.; Nutt, D.J.; Moran, R.J.; Carhart-Harris, R.L.; Muthukumaraswamy, S.D. LSD Modulates Effective Connectivity and Neural Adaptation Mechanisms in an Auditory Oddball Paradigm. Neuropharmacology 2018, 142, 251–262. [Google Scholar] [CrossRef]
- Dudysová, D.; Janků, K.; Šmotek, M.; Saifutdinova, E.; Kopřivová, J.; Bušková, J.; Mander, B.A.; Brunovský, M.; Zach, P.; Korčák, J.; et al. The Effects of Daytime Psilocybin Administration on Sleep: Implications for Antidepressant Action. Front. Pharmacol. 2020, 11, 602590. [Google Scholar] [CrossRef] [PubMed]
- Skosnik, P.D.; Sloshower, J.; Safi-Aghdam, H.; Pathania, S.; Syed, S.; Pittman, B.; D’Souza, D.C. Sub-Acute Effects of Psilocybin on EEG Correlates of Neural Plasticity in Major Depression: Relationship to Symptoms. J. Psychopharmacol. 2023, 37, 687–697. [Google Scholar] [CrossRef] [PubMed]
- Murphy, R.J.; Godfrey, K.; Shaw, A.D.; Muthukumaraswamy, S.; Sumner, R.L. Modulation of Long-Term Potentiation Following Microdoses of LSD Captured by Thalamo-Cortical Modelling in a Randomised, Controlled Trial. BMC Neurosci. 2024, 25, 7. [Google Scholar] [CrossRef]
- Hutten, N.R.P.W.; Mason, N.L.; Dolder, P.C.; Theunissen, E.L.; Holze, F.; Liechti, M.E.; Varghese, N.; Eckert, A.; Feilding, A.; Ramaekers, J.G.; et al. Low Doses of LSD Acutely Increase BDNF Blood Plasma Levels in Healthy Volunteers. ACS Pharmacol. Transl. Sci. 2021, 4, 461–466. [Google Scholar] [CrossRef]
- Wießner, I.; Olivieri, R.; Falchi, M.; Palhano-Fontes, F.; Oliveira Maia, L.; Feilding, A.; Araujo, D.B.; Ribeiro, S.; Tófoli, L.F. LSD, Afterglow and Hangover: Increased Episodic Memory and Verbal Fluency, Decreased Cognitive Flexibility. Eur. Neuropsychopharmacol. J. Eur. Coll. Neuropsychopharmacol. 2022, 58, 7–19. [Google Scholar] [CrossRef] [PubMed]
- Kanen, J.W.; Luo, Q.; Rostami Kandroodi, M.; Cardinal, R.N.; Robbins, T.W.; Nutt, D.J.; Carhart-Harris, R.L.; den Ouden, H.E.M. Effect of Lysergic Acid Diethylamide (LSD) on Reinforcement Learning in Humans. Psychol. Med. 2022, 53, 6434–6445. [Google Scholar] [CrossRef]
- Ramachandran, V.; Chunharas, C.; Marcus, Z.; Furnish, T.; Lin, A. Relief from Intractable Phantom Pain by Combining Psilocybin and Mirror Visual-Feedback (MVF). Neurocase 2018, 24, 105–110. [Google Scholar] [CrossRef]
- Lyes, M.; Yang, K.H.; Castellanos, J.; Furnish, T. Microdosing Psilocybin for Chronic Pain: A Case Series. Pain 2023, 164, 698–702. [Google Scholar] [CrossRef]
- Kovacevich, A.; Weleff, J.; Claytor, B.; Barnett, B.S. Three Cases of Reported Improvement in Microsmia and Anosmia Following Naturalistic Use of Psilocybin and LSD. J. Psychoactive Drugs 2023, 55, 672–679. [Google Scholar] [CrossRef] [PubMed]
- Bahrami, S.; Drabløs, F. Gene Regulation in the Immediate-Early Response Process. Adv. Biol. Regul. 2016, 62, 37–49. [Google Scholar] [CrossRef] [PubMed]
- Kočárová, R.; Horáček, J.; Carhart-Harris, R. Does Psychedelic Therapy Have a Transdiagnostic Action and Prophylactic Potential? Front. Psychiatry 2021, 12, 661233. [Google Scholar] [CrossRef]
- Carhart-Harris, R.L.; Nutt, D.J. Serotonin and Brain Function: A Tale of Two Receptors. J. Psychopharmacol. 2017, 31, 1091–1120. [Google Scholar] [CrossRef] [PubMed]
- Vollenweider, F.X.; Kometer, M. The Neurobiology of Psychedelic Drugs: Implications for the Treatment of Mood Disorders. Nat. Rev. Neurosci. 2010, 11, 642–651. [Google Scholar] [CrossRef] [PubMed]
- Vollenweider, F.X.; Smallridge, J.W. Classic Psychedelic Drugs: Update on Biological Mechanisms. Pharmacopsychiatry 2022, 55, 121–138. [Google Scholar] [CrossRef] [PubMed]
- Stoliker, D.; Egan, G.F.; Friston, K.J.; Razi, A. Neural Mechanisms and Psychology of Psychedelic Ego Dissolution. Pharmacol. Rev. 2022, 74, 876–917. [Google Scholar] [CrossRef]
- Kyrios, M.; Nelson, B.; Ahern, C.; Fuchs, T.; Parnas, J. The Self in Psychopathology. Psychopathology 2015, 48, 275–277. [Google Scholar] [CrossRef] [PubMed]
- Gallagher, S. The Oxford Handbook of the Self; Oxford University Press: Oxford, UK, 2011; ISBN 9780199548019. [Google Scholar]
- Marrocu, A.; Kettner, H.; Weiss, B.; Zeifman, R.J.; Erritzoe, D.; Carhart-Harris, R.L. Psychiatric Risks for Worsened Mental Health after Psychedelic Use. J. Psychopharmacol. 2024, 38, 225–235. [Google Scholar] [CrossRef]
- Carhart-Harris, R.L.; Friston, K.J. REBUS and the Anarchic Brain: Toward a Unified Model of the Brain Action of Psychedelics. Pharmacol. Rev. 2019, 71, 316–344. [Google Scholar] [CrossRef]
- Carhart-Harris, R.L.; Chandaria, S.; Erritzoe, D.E.; Gazzaley, A.; Girn, M.; Kettner, H.; Mediano, P.A.M.; Nutt, D.J.; Rosas, F.E.; Roseman, L.; et al. Canalization and Plasticity in Psychopathology. Neuropharmacology 2023, 226, 109398. [Google Scholar] [CrossRef]
- Yaden, D.B.; Griffiths, R.R. The Subjective Effects of Psychedelics Are Necessary for Their Enduring Therapeutic Effects. ACS Pharmacol. Transl. Sci. 2021, 4, 568–572. [Google Scholar] [CrossRef]
- Olson, D.E. The Subjective Effects of Psychedelics May Not Be Necessary for Their Enduring Therapeutic Effects. ACS Pharmacol. Transl. Sci. 2021, 4, 563–567. [Google Scholar] [CrossRef]
- Cameron, L.P.; Tombari, R.J.; Lu, J.; Pell, A.J.; Hurley, Z.Q.; Ehinger, Y.; Vargas, M.V.; McCarroll, M.N.; Taylor, J.C.; Myers-Turnbull, D.; et al. A Non-Hallucinogenic Psychedelic Analogue with Therapeutic Potential. Nature 2021, 589, 474–479. [Google Scholar] [CrossRef] [PubMed]
- Luppi, A.I.; Girn, M.; Rosas, F.E.; Timmermann, C.; Roseman, L.; Erritzoe, D.; Nutt, D.J.; Stamatakis, E.A.; Spreng, R.N.; Xing, L.; et al. A Role for the Serotonin 2A Receptor in the Expansion and Functioning of Human Transmodal Cortex. Brain 2024, 147, 56–80. [Google Scholar] [CrossRef]
- Page, C.E.; Epperson, C.N.; Novick, A.M.; Duffy, K.A.; Thompson, S.M. Beyond the Serotonin Deficit Hypothesis: Communicating a Neuroplasticity Framework of Major Depressive Disorder. Mol. Psychiatry 2024, 29, 3802–3813. [Google Scholar] [CrossRef]
- Zagrebelsky, M.; Korte, M. Are TrkB Receptor Agonists the Right Tool to Fulfill the Promises for a Therapeutic Value of the Brain-Derived Neurotrophic Factor? Neural Regen. Res. 2024, 19, 29–34. [Google Scholar] [CrossRef] [PubMed]
- Rief, W.; Barsky, A.J.; Bingel, U.; Doering, B.K.; Schwarting, R.; Wöhr, M.; Schweiger, U. Rethinking Psychopharmacotherapy: The Role of Treatment Context and Brain Plasticity in Antidepressant and Antipsychotic Interventions. Neurosci. Biobehav. Rev. 2016, 60, 51–64. [Google Scholar] [CrossRef]
- Doidge, N. The Brain That Changes Itself: Stories of Personal Triumph from the Frontiers of Brain Science; Penguin Publishing Group: London, UK, 2007; ISBN 9781101147115. [Google Scholar]
- Nyberg, F. Structural Plasticity of the Brain to Psychostimulant Use. Neuropharmacology 2014, 87, 115–124. [Google Scholar] [CrossRef]
- Frei, K. Tardive Dyskinesia: Who Gets It and Why. Park. Relat. Disord. 2019, 59, 151–154. [Google Scholar] [CrossRef]
- Iasevoli, F.; Avagliano, C.; D’Ambrosio, L.; Barone, A.; Ciccarelli, M.; De Simone, G.; Mazza, B.; Vellucci, L.; de Bartolomeis, A. Dopamine Dynamics and Neurobiology of Non-Response to Antipsychotics, Relevance for Treatment Resistant Schizophrenia: A Systematic Review and Critical Appraisal. Biomedicines 2023, 11, 895. [Google Scholar] [CrossRef]
- Brouwer, A.; Carhart-Harris, R.L. Pivotal Mental States. J. Psychopharmacol. 2021, 35, 319–352. [Google Scholar] [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Weiss, F.; Magnesa, A.; Gambini, M.; Gurrieri, R.; Annuzzi, E.; Elefante, C.; Perugi, G.; Marazziti, D. Psychedelic-Induced Neural Plasticity: A Comprehensive Review and a Discussion of Clinical Implications. Brain Sci. 2025, 15, 117. https://doi.org/10.3390/brainsci15020117
Weiss F, Magnesa A, Gambini M, Gurrieri R, Annuzzi E, Elefante C, Perugi G, Marazziti D. Psychedelic-Induced Neural Plasticity: A Comprehensive Review and a Discussion of Clinical Implications. Brain Sciences. 2025; 15(2):117. https://doi.org/10.3390/brainsci15020117
Chicago/Turabian StyleWeiss, Francesco, Anna Magnesa, Matteo Gambini, Riccardo Gurrieri, Eric Annuzzi, Camilla Elefante, Giulio Perugi, and Donatella Marazziti. 2025. "Psychedelic-Induced Neural Plasticity: A Comprehensive Review and a Discussion of Clinical Implications" Brain Sciences 15, no. 2: 117. https://doi.org/10.3390/brainsci15020117
APA StyleWeiss, F., Magnesa, A., Gambini, M., Gurrieri, R., Annuzzi, E., Elefante, C., Perugi, G., & Marazziti, D. (2025). Psychedelic-Induced Neural Plasticity: A Comprehensive Review and a Discussion of Clinical Implications. Brain Sciences, 15(2), 117. https://doi.org/10.3390/brainsci15020117