You are on page 1of 16

Journal of Psychoactive Drugs

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/ujpd20

Psilocybin’s Potential Mechanisms in the


Treatment of Depression: A Systematic Review

Harrison J Lee, Vivian WL Tsang, Brandon S Chai, Michelle CQ Lin, Andrew


Howard, Christopher Uy & Julius O Elefante

To cite this article: Harrison J Lee, Vivian WL Tsang, Brandon S Chai, Michelle CQ Lin, Andrew
Howard, Christopher Uy & Julius O Elefante (2023): Psilocybin’s Potential Mechanisms in
the Treatment of Depression: A Systematic Review, Journal of Psychoactive Drugs, DOI:
10.1080/02791072.2023.2223195

To link to this article: https://doi.org/10.1080/02791072.2023.2223195

View supplementary material

Published online: 29 Jun 2023.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=ujpd20
JOURNAL OF PSYCHOACTIVE DRUGS
https://doi.org/10.1080/02791072.2023.2223195

Psilocybin’s Potential Mechanisms in the Treatment of Depression: A Systematic


Review
Harrison J Lee M.D. a, Vivian WL Tsang M.D., M.C.G., etc.a, Brandon S Chaib, Michelle CQ Linb,
Andrew Howard M.D., F.R.C.P.C.a, Christopher Uy M.D., F.R.C.P.C.c, and Julius O Elefante M.D., F.R.C.P.C.a
a
Department of Psychiatry, Faculty of Medicine, University of British Columbia, Vancouver, BC, Canada; bFaculty of Medicine, University of
British Columbia, Vancouver, BC, Canada; cDepartment of Neurology, Faculty of Medicine, University of British Columbia, Vancouver, BC,
Canada

ABSTRACT ARTICLE HISTORY


Evidence suggests that psilocybin has therapeutic benefit for treating depression. However, there Received 1 December 2022
is little consensus regarding the mechanism by which psilocybin elicits antidepressant effects. This Revised 05 March 2023
systematic review summarizes existing evidence. Ovid MEDLINE, EMBASE, psychINFO, and Web of Accepted 15 March 2023
Science were searched, for both human and animal studies, using a combination of MeSH Terms KEYWORDS
and free-text keywords in September 2021. No other mood disorders or psychiatric diagnoses were Psilocybin; Mechanism of
included. Original papers in English were included. The PRISMA framework was followed for the action; Treatment;
screening of papers. Two researchers screened the retrieved articles from the literature search, and Depression
a third researcher resolved any conflicts. Of 2,193 papers identified, 49 were selected for full-text
review. 14 articles were included in the qualitative synthesis. Six supported psilocybin’s mechanism
of antidepressant action via changes to serotonin or glutamate receptor activity and three papers
found an increase in synaptogenesis. Thirteen papers investigated changes in non-receptor or
pathway-specific brain activity. Five papers found changes in functional connectivity or neuro­
transmission, most commonly in the hippocampus or prefrontal cortex. Several neuroreceptors,
neurotransmitters, and brain areas are thought to be involved in psilocybin’s ability to mitigate
depressive symptoms. Psilocybin appears to alter cerebral blood flow to the amygdala and
prefrontal cortex, but the evidence on changes in functional connectivity and specific receptor
activity remains sparse. The lack of consensus between studies suggests that psilocybin’s mechan­
ism of action may involve a variety of pathways, demonstrating the need for more studies on
psilocybin’s mechanism of action as an antidepressant.

Introduction
impacts will increase as the prevalence of depression
Depression is a leading cause of world-wide disease rises globally (Moreno-Agostino et al. 2021).
burden, as quantified by disability-adjusted life years The monoamine theory of depression hypothesizes
(DALYs) (Lépine and Briley 2011). This is due to its that reductions in the monoamines serotonin, norepi­
high lifetime prevalence of up to 21%, in addition to its nephrine, and/or dopamine in specific areas of the
potentially devastating impacts on quality of life brain are associated with clinical depression and has
(Gutiérrez-Rojas et al. 2020). Compared to healthy con­ led to the development of many antidepressant med­
trols, those with depressive disorders are at risk of ications (Mulinari 2012). Antidepressants such as ser­
a higher mortality rate, tend to be of lower socioeco­ otonin selective reuptake inhibitors (SSRIs) are widely
nomic status, have poorer physical health outcomes, recognized as first-line pharmacological treatment for
and have higher prevalence of comorbid psychiatric depression. Unfortunately, these and other first-line
conditions such as generalized anxiety disorder agents for major depressive disorder require several
(Gutiérrez-Rojas et al. 2020; Walker, McGee, and weeks to achieve therapeutic effect and often cause
Druss 2015). The burden of depression strains health­ unwanted side effects; up to 43% of patients discon­
care systems because of its high impact on economic tinue antidepressants due to adverse effects (Carvalho
productivity in all age groups (König, König, and et al. 2016). Although several clinical guidelines exist,
Konnopka 2019). Consequently, financial and health there is a lack of agreement over the appropriate

CONTACT Harrison J Lee hjefflee@gmail.com Department of Psychiatry, Faculty of Medicine, University of British Columbia, 2255 Wesbrook Mall V6T
2A1, Vancouver, BC, Canada
*First authorship is shared between these two individuals.
Supplemental data for this article can be accessed online at https://doi.org/10.1080/02791072.2023.2223195.
© 2023 Taylor & Francis Group, LLC
2 H. J. LEE ET AL.

treatment for those with treatment-resistant depres­ aim of better understanding its therapeutic role in the
sion, defined as inadequate response to two or more treatment of depression.
antidepressants (Gabriel et al. 2020; Kennedy et al.
2016; McIntyre et al. 2014; Taylor, Barnes, and
Methods
Young 2018). Furthermore, only 40–60% of patients
will respond to initial pharmacological treatment and The Preferred Reporting Items for Systematic Review
only 30% will attain remission (Gabriel et al. 2020; and Meta-Analysis (PRISMA) guidelines were used to
MacQueen et al. 2017). ensure clarity and transparency in the methodology.
Recently, psychedelics have reemerged as a potential A protocol for this review has been registered with
therapeutic option for treatment-resistant depression. PROSPERO (registration number: 282710).
There is mounting evidence that psychedelics are effec­ The databases Ovid MEDLINE (1946–
tive at reducing symptoms of depression and are gen­ September 2021), EMBASE (1952-September 2021),
erally well tolerated (Kuypers 2020; Muttoni, psychINFO (1959–September 2021) and Web of
Ardissino, and John 2019). Psilocybin is of particular Science (1900-September 2021) were searched using
interest, with initial studies demonstrating both rapid a combination of MeSH Terms, non-Mesh keywords
and sustained antidepressant activity (Carhart-Harris and free-text keywords (see supplementary material).
et al. 2018; Griffiths et al. 2016). In their randomized, Search items were adapted with database-specific filters.
double-blind trial, Griffiths et al. (2016) found that References of all included papers were hand-mined, and
about 80% of participants sustained clinically signifi­ any additional documents were added from gray litera­
cant reductions in depressed mood six months follow­ ture such as from theses, dissertations, and Google
ing treatment. Similarly, Carhart-Harris et al. (2018) Scholar. The last search was completed on
found psilocybin treatment to significantly improve September 28, 2021.
depressive symptoms within one week, and these The search included human and animal studies
effects remained significant at six months with focusing on the mechanism of psilocybin’s antidepres­
a favorable safety profile compared to other psychede­ sant effects. Original papers, including randomized con­
lics (Hendricks, Johnson, and Griffiths 2015). Castro trolled trials, observational, case control, and cohort
Santos and Gama Marques (2021) also found no sig­ studies were included. Review articles, posters, abstracts,
nificant differences in reducing depressive symptoms conference proceedings, dissertations, commentaries,
when comparing psilocybin to escitalopram in letters, editorials, and toxicology or coroner’s reports
a 6-week trial of patients with depression. As a result, were excluded. Only articles available in English or
psilocybin is one of the most prominent psychedelics those that could be translated into English were
being studied for depression (Castro Santos and Gama included. Papers were excluded if they: 1) did not dis­
Marques 2021; Li et al. 2022; Thomas, Malcolm, and cuss specific mechanism(s) of action, 2) did not specifi­
Lastra 2017; Vargas et al. 2020). cally mention psilocybin, depression or antidepressant
Psilocybin’s antidepressant pharmacology has been effects, or 3) focused on other psychedelic drugs or
proposed to result from its binding to 5-HT2A and psychiatric illnesses unrelated to psilocybin and
5-HT1A receptors with high and lower affinity, respec­ depression.
tively (Lowe et al. 2021). However, the exact mechanism Articles were reviewed systematically using the
of action of its antidepressant effect is still uncertain PRISMA flow diagram (Figure 1). Two of four reviewers
(Inserra et al. 2021). Previous studies have explored (BC, MCQL) independently screened each title and
the neurobiological mechanisms behind psilocybin abstract, strictly adhering to the inclusion and exclusion
causing the “psychedelic experience” itself, character­ criteria to minimize bias. Conflicts were resolved by
ized by alteration in perception, space, and self a third reviewer (VWLT or HL). At least two of the
(Tagliazucchi et al. 2014), while others have focused four independent reviewers evaluated each full-text
on specific receptor and neurotransmitter activity manuscript and inconsistencies were resolved by dis­
(Vollenweider et al., 1999). Although many studies sup­ cussion and consensus. Data on demographics, meth­
port the therapeutic benefit of psilocybin, there is little ods, and outcomes were compiled using
agreement on its antidepressant mechanism of action a predetermined, standardized table by all reviewers.
(Castro Santos and Gama Marques 2021; Li et al. 2022; Quality synthesis and evaluation of bias for article
Thomas, Malcolm, and Lastra 2017; Vargas et al. 2020). inclusion was completed in alignment with the
The current systematic review consolidates the litera­ Modified Newcastle-Ottawa Quality Assessment
ture on the evidence for various proposed mechanism(s) scale for cohort studies (Shamsrizi et al. 2020).
of action of psilocybin antidepressant effect, with the Randomized controlled trials were evaluated with
JOURNAL OF PSYCHOACTIVE DRUGS 3

Figure 1. PRIMSA Flow Diagram.

the Jaded Scales (Halpern & Douglas 2005). Animal studies and nine studies with human participants.
studies were analyzed with SYRCLE risk of bias tools These results will be discussed separately in the fol­
for animal studies (Hooijmans et al. 2014). For case lowing section (Tables 2, 3).
series, the JBI Checklist was used (Munn et al. 2020).
Meta-analyses and reviews were not included in this
study. Functional imaging observations
The imaging modality most used in our included arti­
cles was fMRI, including blood oxygenation level depen­
Results
dent (BOLD) and arterial spin labeling (ASL) imaging
Of the 14 included articles, six supported psilocybin’s techniques. In studies that employed these modalities,
mechanism of antidepressant action via its changes to the amygdala and prefrontal cortex were areas of speci­
serotonin or glutamate receptor activity and three fic focus.
found an increase in synaptogenesis in regions such Carhart-Harris et al. used both BOLD/ASL fMRI in
as the medial frontal cortex and hippocampus healthy human participants and found contrasting pre-
(Table 1). Thirteen papers investigated changes in post changes between placebo and psilocybin, with psi­
gross, non-receptor or pathway-specific brain activity locybin inducing relative decreased activity in the med­
resulting from the administration of psilocybin. Of ial prefrontal cortex (mPFC), ventral posterior cingulate
these, four articles proposed reduced blood flow to cortex (vPCC), putamen, and subthalamic nuclei. In
the amygdala, two found altered blood flow to the addition, decreased brain blood flow and venous oxyge­
prefrontal cortex and one found a reduction in delta nation, reduced cerebral blood flow and BOLD signal in
power during sleep. Five papers found changes in the anterior cingulate cortex (ACC) and mPFC, and
functional connectivity or neurotransmission, most decreased positive coupling between mPFC and PCC
commonly in the hippocampus or prefrontal cortex, were observed. There were no differences seen in BOLD
but with little consensus. There were five animal signal during breath hold for patients on psilocybin
4 H. J. LEE ET AL.

Table 1. Summary of findings and study quality.


Title Summary of findings Study quality
Randomized Controlled Trials
Bernasconi et al. Increased top-down control of the prefrontal cortex and reduced neural responses to neutral Jadad Scales for RCT 3/5
(2014) and emotional faces was found after psilocybin administration
Carhart-Harris Psilocybin-induced decrease in activity in mPFC, vPCC, putamen, and subthalamus as shown 3/5
et al. (2012) on BOLD/ASL fMRI relative to placebo. Reduced cerebral blood flow and BOLD signal to ACC
and mPFC. Decreased positive coupling between mPFC and PCC.
Dudysová et al. Decreased absolute delta power during slow-wave sleep during the first sleep cycle was found 2/5
(2020) post-psilocybin administration. Increased REM sleep onset latency and a trend to reduced
REM sleep duration was also noticed.
Kometer et al. Patients taking psilocybin had enhanced positive mood states, decreased recognition of 4/5
(2012) negative facial expression, and increased behavior toward positive cues in comparison to
negative cues. Psilocybin-induced emotional bias is possibly due to stimulation of 5-HT1A
or 5-HT2C receptors rather than 5-HT2A receptors.
Kraehenmann Reduced right amygdala activation to negative and neutral pictures post-psilocybin treatment 3/5
et al. (2015) as measured by BOLD fMRI. Decreased left amygdala activation to negative pictures.
Reduced visual cortex activation.
Mason, et al., Higher relative glutamate concentration levels in the mPFC and lower relative glutamate 4/5
2020 concentration levels in the hippocampus were found in patients receiving psilocybin.
Animal Studies
Shao et al. Psilocybin showed long-lasting increases in spine density and spine head width in mouse SYRCLE risk of bias tool for 8/10
(2021) medial frontal cortex, elevating the formation rate of dendritic spines in vivo. Pretreatment animal studies
with ketanserin did not block structural plasticity from psilocybin. Elevated excitatory
neurotransmission was also noticed post-psilocybin.
Hesselgrave Psilocybin reversed anhedonic responses in mice after being subjected to a chronic 7/10
et al. (2021) multimodal stress paradigm. Excitatory synapses in the hippocampus were strengthened.
Pretreatment with ketanserin did not prevent psilocybin effectivity.
Raval et al. Psilocybin increased synaptogenesis, with higher synaptic vesicle protein 2A density in both 6/10
(2021) the hippocampus and PFC.
Grandjean et al. Reduced focal connectivity in the ventral striatum, along with increased focal connectivity 5/10
(2021) between the serotonin system and various striatal, cortical, and thalamic regions were
observed following psilocybin administration
Mahmoudi et al. Reduced total distance traveled in open field tests for the high-dose psilocybin group as 5/10
(2018) compared to placebo. No decrease in total distance moved between the low dose ketamine
or psilocybin group and the placebo group. However, co-administration of low dose
ketamine and psilocybin resulted in decreases in movement time similar to fluoxetine.
Case Series
Mertens et al. Reduced functional connectivity between the amygdala and vmPFC using BOLD fMRI. JBI checklist for case series Good
(2020) Increased functional connectivity between the vmPFC and the parietal and occipital lobe quality
during face processing. Enhanced connectivity in the precuneus, right angular gyrus, lingual
gyrus, supracalcarine cortex, intracalcarine cortex and lateral occipital cortex. Increased
functional connectivity between vmpFC with right angular gyrus and lateral occipital cortex
when viewing fearful faces.
Roseman et al. Increased amygdala activity using BOLD/ASL fMRI data post-psilocybin treatment and Good
(2018) increased functional connectivity between amygdala and visual areas in response to quality
emotional faces.
Cohort Study
Carhart-Harris Reduced cerebral blood flow as measured by BOLD/ASL fMRI in the temporal cortex including Modified Newcastle Ottawa 5
et al. (2017) the left amygdala which are associated with reductions in depressive symptoms. Increased Quality Assessment Scale
resting state functional connectivity (RSFC) within the default mode network. Increased
ventromedial-prefrontal cortex-bilateral inferior lateral parietal cortex RSFC in patients
responding to psilocybin treatment. Decreases in cerebral blood flow in Heschl’s gyrus, left
precentral gyrus, left planum temporale, left superior temporal gyrus, left amygdala, right
supramarginal nucleus and right parietal operculum.

versus placebo indicating the lack of effect of psilocybin the default mode network. Increased ventromedial pre­
on vasculature directly (Carhart-Harris et al. 2012). frontal cortex-bilateral inferior lateral parietal cortex
The same group subsequently studied BOLD/ASL RSFC and decreased parahippocampal-prefrontal cortex
fMRI in individuals with treatment-resistant depression RSFC was observed in patients who responded to psilo­
given oral psilocybin across two sessions (10 mg, 25 mg). cybin treatment (Carhart-Harris et al. 2017). Specific
They observed reduced cerebral blood flow in the tem­ areas demonstrating statistically significant decreases in
poral cortex, including the left amygdala, with CBF included left Heschl’s gyrus, left precentral gyrus, left
a statistically significant relationship found when com­ planum temporale, left superior temporal gyrus, left
pared to reductions in depressive symptoms (Carhart- amygdala, right supramarginal gyrus and right parietal
Harris et al. 2017). There was also increased resting operculum. Using data from the same study, Roseman
state functional connectivity (RSFC) observed within et al. found increased amygdala activity after psilocybin
JOURNAL OF PSYCHOACTIVE DRUGS 5

Table 2. Results Table Condensed.


Author last
name, year of Number of participants (n), patient Neurobiological changes and proposed mechanisms of
publication type, type of study Method of imaging (if any) action
Bernasconi et al. 30, healthy humans, double-blind, Continuous EEG Increased top-down control of the PFC, resulting in reduced
(2014) placebo-controlled RCT initial neural response to both negative and neutral faces
Carhart-Harris et al. 15, healthy human controls, subject- ASL (arterial spin labeling) and Decreased brain blood flow and venous oxygenation.
(2012) controlled RCT fMRI/BOLD Reduced CBF in ACC/mPFC;
Decreased positive coupling between mpFC and PCC;
Decreased mPFC activity via 5-HT2A receptor stimulation
Carhart-Harris et al. 19, patients suffering from treatment- ASL and BOLD RSFC (resting Decreased cerebral blood flow in whole brain;
(2017) resistant depression, open-label state functional Increased default mode network integrity;
clinical trial connectivity) analysis Reduced amygdala CBF and RSFC between bilateral
parahippocampus and PFC
Dudysová et al. 20, healthy human participants, Whole-night polysomnography Decreased absolute delta power during SWS in first sleep
(2020) double-blind placebo-controlled (EEG spectral analysis) cycle (similar to other antidepressants)
crossover design
Grandjean et al. 50, mice, placebo-controlled trial fMRI/BOLD – Resting state Reduced FC within ventral striatum;
(2021) The RSN predominantly affected by psilocybin was enriched
in DA receptors
Hesselgrave et al. 8, mice, exposed to a chronic Strengthened excitatory synapses in the hippocampus and
(2021) multimodal stress paradigm increased AMPA/NMDA ratios, unaffected by ketanserin
(5-HT2A blocker)
Kometer et al. 17, healthy human controls, double EEG recordings, with additional Higher P300 amplitudes for positive and negative faces
(2012) blinded RCT electrodes attached near the compared with neutral stimuli, unaffected by ketanserin
eyes (5-HT2A blocker)
Kraehenmann et al. 25, healthy human controls, double- fMRI/BOLD Reduced amygdala activation to negative pictures;
(2015) blinded, placebo-controlled RCT Decreased activation in visual cortex
Mahmoudi et al. 8 groups (8 mice each), placebo- NMDA antagonist and psilocybin co-administration led to
(2018) controlled behavioral trial most significant antidepressant behavioral change
Mason, et al., 2020 60, healthy human controls, double fMRI/BOLD and structural MRI Higher relative glutamate concentration levels in the mPFC,
blind RCT and lower relative glutamate concentration levels in the
hippocampus
Mertens et al. 20, humans with moderate to severe fMRI/BOLD – Anatomical scan Increased FC between amygdala and visual areas;
(2020) MDD, case series study Increased FC between vmPFC and other regions
Raval et al. (2021) 24, pigs, placebo-controlled trial In vitro autoradiography Transient decrease in 5-HT2AR density in hippocampus and
PFC; Higher SV2A density in hippocampus
Roseman et al. 20, humans with moderate to severe fMRI/BOLD – Anatomical scan Increased BOLD signal in right amygdala (and visual areas)
(2018) MDD, case series study
Shao et al. (2021) 12, mice, placebo-controlled trial Two-photon microscopy Increased neurotransmission and spine density in medial
frontal cortex, unaffected by ketanserin (5-HT2A blocker)

administration, as well as increased functional connectiv­ but not neutral pictures. There was also reduced activa­
ity between amygdala and the visual areas in response to tion in the visual cortex (Kraehenmann et al. 2015).
emotional faces (Roseman et al. 2018). Further examina­ Mason et al. explored structural MRI/MRS in healthy
tion of the sample group by Mertens et al. (2020), found human participants to analyze relative glutamate levels
reduced functional connectivity between the amygdala in the brain. They found that oral psilocybin (0.17 mg/
and the ventromedial prefrontal cortex (vmPFC). The kg) administration resulted in higher glutamate concen­
authors also reported increased functional connectivity trations in the mPFC and lower concentrations in the
between the vmPFC and the parietal lobe and occipital hippocampus. The authors also reported an increase in
lobe during face processing. Specific regions with between-network functional connectivity after psilocy­
enhanced connectivity included the precuneus and the bin, but not placebo, in all networks other than the
right angular gyrus of the parietal lobe and the lingual lateral motor network. Within their respective net­
gyrus, supracalcarine cortex, intracalcarine cortex and works, there was a decrease in co-activation in visual
lateral occipital cortex. The right angular gyrus and lateral networks 1 and 2 and both subcomponents of the
occipital cortex also demonstrated increased functional default mode network (DMN) post-psilocybin in com­
connectivity with the vmPFC, but only when viewing parison to post-placebo (Mason et al. 2020).
fearful faces (Mertens et al. 2020).
In a study with healthy human participants using Electrophysiological observations
BOLD fMRI, Kraehenmann et al. found reduced right
amygdala activation to both negative and neutral pictures EEG and other electrophysiological data were used as
after oral psilocybin (0.16 mg/kg) administration, as well investigative tools for psilocybin’s effects in three
as decreased activation in the left amygdala to negative human studies. Bernasconi et al. performed analyses
Table 3. Results table.
6

Proposed mechanism of
action – i.e. receptor
Number of activity/expression
Author last participants (n), Method of psilocybin (5-HT2a, dopamine,
name, year of patient type, type of administration (route, glutamate) and or Mode of neurological Resulting brain activity/Functional
pub study dose, timing) synaptogenesis investigation connectivity/[Enhanced neurotransmission] Clinical outcomes
Bernasconi 30, healthy humans, Oral, 170ug/kg Continuous EEG (electrical Increases top-down control of the PFC,
et al. (2014) double-blind, neuroimaging analyses resulting in reduced neural response to
H. J. LEE ET AL.

(3) placebo- on visual evoked both neutral and emotional faces


controlled RCT potentials [VEPs])
Carhart-Harris 15, healthy human Intravenous, 2mmg in 10 Decreased mPFC activity ASL (arterial spin labeling) Psilocybin decreased brain blood flow and Subjective ratings on nineteen items
et al. (2012) controls, subject- mL saline over 60s via 5-HT2A receptor and fMRI/BOLD brain venous oxygenation. immediately after the ASL scan showed an
(4) controlled RCT stimulation scans Reduced CBF in ACC/mPFC; decreased increase in items such as seeing the
positive coupling between mpFC and PCC surroundings change, seeing geometric
patterns, the experience having
a dreamlike quality, etc.
Carhart-Harris 19, patients Two doses of psilocybin, Decreasing cerebral blood ASL and BOLD RSFC Increased default mode network integrity; Mean depression score measured by the
et al. (2017) suffering from 10 mg followed by 25 flow in whole brain (resting state functional reduced amygdala CBF QIDS-SR16 was lower in all patients one
(5) treatment- mg one week apart connectivity) analysis Reduced RSF between bilateral week post-treatment, and all but one
resistant parahippocampus and PFC patient showed some decrease in QIDS-
depression, SR16 score at week 5.
open-label
clinical trial
Dudysová 20, healthy human Capsules adjusted N/A Whole-night Decreased absolute delta power during SWS Psilocybin increases REM sleep onset latency
et al. (2020) participants, according to body polysomnography (EEG in first sleep cycle (similar to other and shows a trend to reduced REM sleep
(6) double-blind weight, ~0.26 mg/kg, spectral analysis) with antidepressants) duration.
placebo- around 9am a gel cap with 19
controlled electrodes
crossover design
Grandjean 50, mice, placebo- IV, 1 mg/kg, 2 mg/kg,or Resting-state, BOLD fMRI − Reduced FC within ventral striatum
et al. (2021) controlled trial vehicle only, 20 min − The RSN predominantly affected by
(7) prior to imaging psilocybin is enriched in DA receptors
Hesselgrave 8, 8-week-old male Intraperitoneal injection, 1 [Independent of 5-HT2A N/A Strengthened excitatory synapses in the Psilocybin causes a rapid beneficial action in
et al. (2021) mice exposed to mg/kg, an hour after a i. (elicited effects despite hippocampus models of chronic stress and depression-
(8) a chronic p injection of ketanserin ketanserin)] [enhanced neurotransmission] relevant hedonic behaviours
multimodal (2 mg/kg)
stress paradigm,
[animal study?]
Kometer et al. 17, healthy human [route?], Placebo or 5-HT1A or 5-HT2C Electroencephalogram Contrary to the crucial contribution of 5-HT2A Psilocybin enhanced positive mood states;,
(2012) (10) controls, double ketanserin (50 mg) pathways rather than recordings were taken receptors in modulating mood states and decreased recognition of negative facial
blinded RCT (pretreatment) followed 5-HT2A pathway. using 64 scalp emotional face recognition, the psilocybin- expression; and, increased behavior toward
after 1 hour by placebo electrodes, with induced relative bias toward the positive relative to negative cues. Validated
or psyilocybin (215 g/ additional electrodes processing of positive emotions in the go/ scales used include: 5-Dimensions Altered
kg) (treatment). attached near the eyes nogo task were not blocked by States of Consciousness (5D-ASC) the
pretreatment with ketanserin. Thus, the German version of the Positive and
psilocybin-induced emotional bias might Negative Affect Schedule (PANAS) and the
rather be due to a stimulation of 5-HT1A or State-Trait Anxiety Inventory (STAI)
5-HT2C receptors than 5-HT2A receptors.
However, the strong valence-independent
reduction of the P300 seen after psilocybin
administration was partially reversed by
ketanserin, indicating an involvement of
5-HT2A receptors in the valence-
independent attentional performance.
(Continued)
Table 3. (Continued).
Proposed mechanism of
action – i.e. receptor
Number of activity/expression
Author last participants (n), Method of psilocybin (5-HT2a, dopamine,
name, year of patient type, type of administration (route, glutamate) and or Mode of neurological Resulting brain activity/Functional
pub study dose, timing) synaptogenesis investigation connectivity/[Enhanced neurotransmission] Clinical outcomes
Kraehenmann 25, healthy human Oral, 0.16 mg/kg, two Inhibiting glutamate? fMRI/BOLD Reduced right amygdala activation; Positive and Negative Affect Schedule
et al. (2015) controls, Double- sessions (14 days apart) (NMDA receptor decreased activation in visual cortex (PANAS): Psilocybin significantly increased
(11) blinded, antagonists) positive affect (Bonferroni- corrected p
randomized = .001, Figure 1) but not negative affect
placebo- (Bonferroni-corrected p 5 .87) or state
controlled trial. anxiety (Bonferroni-corrected)
Mahmoudi 8 mice per group, Intraperitoneal injections, Co-administration of low N/A Associated with a rapid beneficial result in
et al. (2018) male NMRI mice 10, 40, or 100 mg/kg dose ketamine and mouse model of chronic-stress induced
(12) weighing 20-25 PCE, 30 minutes before psilocybin caused deficits in depression-relevant hedonic
g, [RCT animal each test a significant decrease in behaviours
study?] movement time (similar
to fluoxetine)
Mason, et al., 60, healthy human [route?], 0.17 mg/kg Glutamate was higher in MRI – structural and fMRI/ Higher relative glutamate concentration Psilocybin was associated with increased
2020 (13) controls, double mPFC; Glutamate was BOLD levels in the mPFC, and lower relative ratings in dimensions of 5 Dimensions of
blind RCT lower in hippocampus glutamate concentration levels in the Altered States of Consciousness (5D-ASC)
hippocampus. and Ego Dissolution Inventory (EDI)
Mertens et al. 20, Human Oral, low and 25 mg dose, N/A Anatomical scan using Increased FC between amygdala and visual N/A?
(2020) (14) participants with low dose on trial day BOLD fMR areas; Increased between vmPFC and other
moderate to and 25 mg dose one regions
severe MDD, case week later
series study
Raval et al. 24, pigs, placebo- IV, 0.08 mg/kg, 1 day prior Transient decrease in In vitro autoradiography
(2021) (15) controlled trial to euthanization 5-HT2AR density in
hippocampus and PFC
Higher SV2A in
hippocampus
(synaptogenesis)
Roseman et al. 20, Human subjects Oral, low and 25 mg dose, N/A Anatomical scan using Increase BOLD in right amygdala (and visual Beck’s Depression Inventory (BDI): Change
(2018) (16) with moderate to low dose on trial day BOLD fMRI areas) (10.2 ± 5.3), response (63.2%) and
severe MDD, and 25 mg dose one remission (57.9%) in BDI scores at 1-week
Case series study week later Quick Inventory of Depressive Symptomology
(QIDS): Response at 1-day (68.4%), 1-week
(63.2%) and 3-weeks (63.2%)
13 participants were responders 1-day after
therapy and 9 maintained response 5
weeks after therapy.
Shao et al. 12, mice, placebo- IV, 1 mg/kg, 24 h before − Increase in spine density Two-photon microscopy Elevates neurotransmission in medial frontal
(2021) (17) controlled trial euthanization in medial frontal cortex cortex
- Increase in density [enhanced neurotransmission]
unaffected by
ketansarin (5-HT2A
JOURNAL OF PSYCHOACTIVE DRUGS

blocker)
7
8 H. J. LEE ET AL.

on visual evoked potentials (VEPs) in response to serotonin system and various striatal, cortical, and tha­
facial expressions (fearful, happy, and neutral) after lamic regions (Grandjean et al. 2021).
placebo administration or oral psilocybin(0.17 mg/kg). Using two-photon microscopy to image layer 5 pyr­
They found a significant global field power (GFP) amidal neurons in the mouse medial frontal cortex,
interaction between treatment and face valence, as Shao et al. found that a single dose of intraperitoneal
well as within the visual areas, limbic areas, tempor­ psilocybin (1 mg/kg) increased the rate of dendritic
oparietal areas, and the prefrontal areas. In these spine formation, leading to significant increases in
areas, reductions of brain activity were observed for spine protrusion lengths, width and density. In particu­
both neutral and fearful faces, but not for happy faces lar, Cg1/M2 (homologous to the mid-anterior cingulate
(Bernasconi et al. 2014). cortex/medial prefrontal cortex in humans) was an area
Kometer et al. used EEG to analyze emotional go/no- of interest which sustained greater spine density and
go tasks in response to face recognition, using oral protrusion length; however, increases in spine density
psilocybin (0.215 mg/kg), placebo, and pre-treatment were greater in female mice compared to males.
with ketanserin (a 5-HT2A receptor antagonist). Dendritic spine formation rate peaked shortly after psi­
Psilocybin enhanced positive mood and attenuated locybin administration dose, then returned to baseline
recognition of negative facial expression, as well as to match the unchanged elimination rate in both
increased goal-directed behavior toward positive com­ females and males. Long-term changes in psilocybin-
pared with negative cues. Psilocybin administration also induced dendritic modeling were found, as a portion of
facilitated positive and inhibited negative emotional psilocybin-induced spines were retained after 1 month.
effects, and valence-dependently attenuated P300 After ketanserin pretreatment, they found that head-
amplitude. Ketansarin alone had no effects but blocked twitch responses induced by psilocybin were eliminated,
psilocybin-induced mood enhancement, and decreased but not psilocybin-induced dendritic remodeling (indi­
recognition of negative facial expression (Kometer et al. cating that this effect was likely not mediated by sero­
2012). tonergic neurotransmission). Consistent with this,
Dudysova et al. conducted a placebo-controlled sleep excitatory neurotransmission in the medial frontal cor­
study that found a small but significant increase in REM tex was enhanced by psilocybin, where most dendritic
latency the night after oral psilocybin (0.26 mg/kg) spines have functional glutamatergic synapses (Shao
administration. Absolute delta power (which is repre­ et al. 2021).
sentative of deeper stage 3/4 sleep) during slow wave Raval et al. used in vitro autoradiography to study the
sleep (SWS) in the first sleep cycle was found to be synaptic vesicle protein 2A (SV2A; a membrane protein
significantly lower after psilocybin administration in used to measure pre-synaptic density) and 5-HT2AR
comparison to placebo, and locally at averaged frontal, densities in the pig hippocampus and PFC. They
central, parietal, temporal, and occipital derivations. found increased SV2A protein and reduced 5-HT2AR
Sleep latency, total sleep time, sleep efficiency, and the density in the hippocampus one-day post-treatment
number of sleep cycles, however, were not significantly with intravenous psilocybin (0.08 mg/kg), with only
different in placebo and psilocybin conditions the SV2A density remaining significantly increased
(Dudysová et al. 2020). after seven days. In the PFC, SV2A density remained
unchanged and 5-HT2AR density was reduced one day
after treatment; however, SV2A density was increased,
Observations from imaging studies in animals
with no change in 5-HT2AR from baseline after seven
Imaging studies using psilocybin in animals used days (Raval et al. 2021).
a variety of modalities including BOLD fMRI, micro­
scopy, and autoradiography. Grandjean et al. used rest­
Observations from behavioral studies in animals
ing-state BOLD fMRI in mice and observed a reduction
in focal connectivity within the ventral striatum post- There were two animal studies that focused on beha­
intravenous psilocybin administration (1 or 2 mg/kg). vioral analyses. Mahmoudi et al. (2018) used open field
Psilocybin was also shown to affect resting state striatal tests, tail suspension tests, and forced swimming tests in
networks enriched in dopamine receptors to the greatest mice after various doses of intraperitoneal psilocybin
degree. When testing for the effects of psilocybin on the (10, 40, 100 mg/kg), ketamine (NMDA receptor antago­
serotonin and dopamine associated networks, it was nist), fluoxetine, and placebo administration. They
found that while focal connectivity decreased between found a significant difference in the total distance
the dopamine system and striatal regions as mentioned, moved between the applied group in open fields at
there was an increase in focal connectivity between the high, but not low doses of psilocybin and ketamine
JOURNAL OF PSYCHOACTIVE DRUGS 9

compared to control. However, co-administration of atrophy and impaired neuroplasticity at the molecular
low-dose ketamine and psilocybin did cause level within these areas (Duman and Aghajanian 2012;
a significant decrease in movement time (similar to Duman et al. 2016). Chronic stress may induce activa­
fluoxetine) (Mahmoudi et al. 2018). tion of localized immune cells in the brain (microglia),
Hesselgrave et al. exposed a group of mice to which engulf nearby pyramidal neurons, contributing to
a chronic multi-modal stress paradigm (CMMS) under neuronal atrophy (Wohleb et al. 2018). Psilocybin
psilocybin and placebo conditions, with or without pre- potentially reverses these pathogenic mechanisms.
treatment with ketanserin (5-HT2A receptor antago­ Depression is also associated with altered functional
nist). They reported significantly increased sucrose and connectivity across the PFC and limbic areas, including
female urine preferences following stress in ketanserin- the amygdala and hippocampus (Siegle et al. 2007), with
pretreated mice receiving psilocybin, whereas ketan­ alterations in affective processing, and establishment of
serin pretreatment alone had no significant effect on negative biases and appraisal of self and the environ­
either behavior. After completion of the behavioral ment (Gotlib and Joormann 2010). Our review illus­
assays, hippocampal brain slices were taken from the trates evidence of alteration in functional connectivity
animals for extracellular recordings. They found that between these key structures after psilocybin adminis­
AMPA/NMDA receptor ratios in slices taken from psi­ tration, as well as in specific neurotransmitter pathways.
locybin-injected CMMS-susceptible mice were signifi­
cantly greater than those in slices taken from CMMS
Amygdala
susceptible animals injected with vehicle or ketanserin
alone, and pretreatment with ketanserin did not impair The amygdala plays a key role in affective regulation,
the ability of psilocybin to restore AMPA/NMDA recep­ including evocation of negative emotions and response
tor ratios (Hesselgrave et al. 2021). to fearful stimuli. Morphological changes in the amyg­
dala, as part of an altered limbic-thalamic-cortical cir­
cuit, have been implicated in the pathogenesis of
Study quality
depression (Bellani, Baiano, and Brambilla 2010; Zou
Of the 14 papers included in this review, one was et al. 2010). Studies have also shown increased amygdala
a cohort study, six were randomized control trials, five CBF and functional connectivity in negative affect states
were animal studies, and two were case series. The (Abercrombie et al. 1998; Coombs et al. 2014). Stress
cohort study was of moderate quality (a score of 5) also increases excitability in the amygdala in response to
due to the lack of a control cohort and short follow-up stress, resulting in increased reactivity and decreased
duration. Of the six randomized controlled trials, two cognitive processing (Dean and Keshavan 2017). In the
were of good quality, ranking 4 out of 5 only losing articles reviewed, variation in amygdala activity and
points for lack of long-term follow-up. Four studies functional connectivity were found following psilocybin
ranked 2 or 3 out of 5, losing points for the lack appro­ administration.
priate randomization and long-term follow-up. Of the Carhart-Harris et al. (2017), reported decreased CBF,
five animal studies, two scored 5 out of 10, one scored 6 and therefore reduced activity in the left amygdala, in
out of 10, one scored 7 out of 10 and one scored 8 out of a population with treatment-resistant depression
10 most commonly losing points for lack of blinding, one day after psilocybin administration, which corre­
inadequate allocation sequence application, and pre­ lated with reduced depressive symptoms. There is
sence of confounders. This systematic review included uncertainty as to whether or not psilocybin may
two case series which were of overall good quality, increase whole-brain CBF (Lewis et al. 2017) as acute
scoring low only for the lack of consecutive inclusion increases in CBF have been reported with other psyche­
of participants and lack of clear reporting of the pre­ delics such as LSD (Carhart-Harris et al. 2012, 2016).
senting site’s demographic information (Table 4). This post-acute reduction in amygdala CBF reported by
Carhart-Harris (in addition to other structures) could
represent a “reset mechanism,” and re-integration in
Discussion
a lower activity state.
Upon review of the literature, it appears that various In the same population as (Carhart-Harris et al.
brain regions and neurotransmitter pathways are poten­ 2017), Roseman et al. and Mertens et al. observed
tially involved in the potential antidepressant mechan­ increased amygdala activity and decreased functional
ism of action of psilocybin. Anatomical areas of interest connectivity between the amygdala and the vmPFC
include the amygdala, prefrontal cortex, and hippocam­ one day after psilocybin administration, in response to
pus. Depression has been characterized by neuronal negative and neutral facial affective recognition tasks.
10 H. J. LEE ET AL.

Table 4. Study Quality.


Scale Used and Score

JBI
Author Last Modified Newcastle Ottawa Jadad Checklist SYRCLE Risk of
Name, Year of Quality Assessment Scale Scales for Case Bias Tool for
Publication Study Methodology Data Analysis (COHORT STUDIES) for RCT Series Animal Studies
Bernasconi et al. Randomized, double-blind, 3/5
(2014) (3) placebo-controlled trial
Carhart-Harris Two trials, each of which Seed-based pharmaco- 3/5
et al. (2012) (4) had their own placebo physiological interaction
functional connectivity
analysis
Carhart-Harris Open-label clinical trial ASL and BOLD RSFC, QIDS-SR16 5
et al. (2017) (5)
Dudysová et al. Double-blind RCT 2/5
(2020) (6)
Grandjean et al. Placebo-controlled trial 5/10
(2021) (7) (animal)
Hesselgrave et al. Placebo-controlled trial 7/10
(2021) (8) (animal)
Kometer et al. Double-blinded RCT More complicated version of 4/5
(2012) (10) ANOVA
Kraehenmann Randomized, double-blind, 3/5
et al. (2015) placebo controlled,
(11) crossover design
Mahmoudi et al. Animal study 5/10
(2018) (12)
Mason et al. Double-blind, placebo Nonparametric Mann-Whitney 4/5
(2020) (13) controlled, parallel U tests
group design
Mertens et al. Follow-up to an open-label Psychophysiological interaction 7
(2020) (!4) trial analyses on fMRI imaging data
Raval et al. (2021) Placebo controlled trial 2/5
(15) (animal)
Roseman et al. Open-label trial 7
(2018) (16)
Shao et al. (2021) Randomized placebo 8/10
(17) controlled trial (animal)

They also found increased functional connectivity Mertens found in the post-acute phase. This may high­
between the amygdala and visual areas after administra­ light the difference in the acute/post-acute effects of
tion of psilocybin. This increase in emotional reactivity psilocybin, but are also confounded by methodological
in the post-acute phase may represent a temporary differences (lower dosing, healthy patient population).
increase in emotional salience, opening an avenue for A future study looking at resting state, acute, post-acute,
psychotherapeutic intervention during this period, prior and long-term changes to amygdala reactivity/func­
to the “reset.” One recognized mechanism of action of tional connectivity may help reconcile these findings.
traditional antidepressants, specifically SSRIs, is of nor­
malizing overactivity in the amygdala (Godlewska et al.
Prefrontal cortex
2012), given prior fMRI findings of increased amygdala
activation in clinically depressed individuals (Yang et al. According to the neuroplasticity models of depression,
2010). It may be that these traditional antidepressants the PFC and the hippocampus suffer from neuronal
have a general effect on reducing responsiveness to all loss, reduced synaptic density, and functionality result­
stimuli regardless of valence (Price, Cole, and Goodwin ing from prolonged reductions in brain-derived neuro­
2009), which is seen clinically with overall emotional trophic factor (BDNF) and chronic stress (Price and
blunting. Further studies looking at longer term changes Duman 2020). Additionally, reduced glutamatergic
after psilocybin administration would help determine if metabolites in the medial frontal cortex have been
there is a similar decrease in reactivity beyond the acute/ observed in patients with depression compared to
post-acute phase. healthy controls (Moriguchi et al. 2019). The changes
Kraehenmann et al. found decreased amygdala acti­ seen in the PFC are thought to lead to impaired func­
vation in response to negative emotional stimuli 210 tional connectivity within key areas of the limbic sys­
min after psilocybin administration in healthy partici­ tem, leading to dysfunctional top-down control of the
pants, which is in opposition to what Roseman and amygdala and hippocampus (Arco and Mora 2009;
JOURNAL OF PSYCHOACTIVE DRUGS 11

Zhao et al. 2020). Clinically, this manifests as dysfunc­ in over 1700 patients with MDD demonstrating robust
tional stress coping, environmental adaptability and reductions in volume (Schmaal et al. 2016).
learning (Price and Duman 2020). Experimentally, this Studies reviewed here varied in terms of hippocampal
may be recognized by modulation of attentional findings with the majority observing greater synaptic
resource allocation, and biased facial emotion proces­ plasticity and neurogenesis. Similar to their results in
sing (Stuhrmann, Suslow, and Dannlowski 2011), with the PFC, Raval et al. (2021) found changes to SV2A
the PFC playing a key role in generating top-down protein and 5-HT2AR densities in pigs after psilocybin
control in attention switching (Rossi et al. 2009). treatment, indicating plasticity in this brain region.
Stuhrmann, Suslow, and Dannlowski (2011) reported Supporting this finding, Hesselgrave et al (2021)
mood-congruent face processing bias in major depres­ observed greater AMPA/NMDA receptor ratios in the
sive disorder patients, with hyperactivation to negative hippocampus of psilocybin-injected mice, despite pre-
and hypoactivation to positive stimuli in a variety of treatment with ketansarin, a potent 5-HT2AR antago­
brain areas, however with inconsistent findings in the nist. This finding correlated with significant reduction
PFC. Bernasconi et al. (2014) found reduced PFC and in observed stress behaviors. Studies of traditional
limbic activity in their EEG study, with associated initial monoamine antidepressants have also found increased
reduction in response to neutral and fearful faces in AMPA/NMDA receptor ratios (Barbon et al. 2011;
comparison to happy faces, which may represent psilo­ Skolnick et al. 1996), which may represent
cybin’s modulation of this pathway. a concurrent pathway of psilocybin’s antidepressant
Psilocybin appears to increase functional connectiv­ effect independent of the 5-HT2A receptor, as well as
ity and neuroplasticity in the prefrontal cortex which reflecting an increase in neuroplasticity.
could explain its antidepressant effects. Mason et al. These findings suggest there may be a role for psilo­
(2020) found higher levels of glutamate in the mPFC cybin in alleviating the effects of hypercortisolemia from
post-psilocybin administration compared to baseline in hypothalamic-pituitary-adrenal axis dysfunction on the
healthy patients. Although this study was not conducted hippocampus in patients with depression (Dean and
on clinically depressed patients, psilocybin’s mechanism Keshavan 2017). Mason et al. (2020), however, found
of action for treating depression may involve increasing reduced glutamate activity in the hippocampus after
glutamate neurotransmission in the mPFC, and its sub­ immediate psilocybin administration and it remains
sequent effects on local BDNF production (Mattson unclear if this reduction persists or reverses and what
2008). In the animal imaging studies reviewed, Shao effect this might have in the longer term on AMPA/
et al. (2021) and Raval et al. (2021) both found increased NMDA receptor ratios, neuronal plasticity via BDNF
PFC synaptic plasticity initially after psilocybin admin­ production in this region.
istration. Maintenance of PFC plasticity was also
observed by Shao et al. (2021) one month after treat­
Observations of changes in clinically depressed
ment. This suggests sustained improvement in PFC
states
functionality and supports clinical studies showing
long-term reduction in depressive symptoms after psi­ While clinical outcomes were not the focus of our review,
locybin administration (Carhart-Harris et al. 2018; six of the included studies reported clinical measures such
Griffiths et al. 2016). These studies support the neuro­ as the Quick Inventory of Depressive Symptomology
trophic hypothesis of depression, postulating that (QIDS)-SR16, Beck Depression Inventory (BDI), the
increasing neurogenesis and plasticity in brain regions Positive and Negative Affect Schedule (PANAS) and the
like the PFC are important for improving depressive State-Trait Anxiety Inventory (STAI). Carhart-Harris
symptoms (Bus and Molendijk 2016; Dean and et al. (2017) reported lower mean scores on the QIDS-
Keshavan 2017). SR16 in participants with depression, and Roseman et al.
(2018) observed clinical improvement on the QIDS and
BDI following psilocybin administration. However, these
Hippocampus
clinical measures were not quantitatively correlated with
Similar to the PFC, chronic stress leads to increased biological measures to decisively determine psilocybin’s
neuronal atrophy, and decreased plasticity, BDNF mechanism of action.
expression, and synaptic density in the hippocampus
(Price and Duman 2020). The hippocampus appears
Limitations
particularly susceptible to atrophy in a depressed popu­
lation and is supported by a 2016 meta-analysis of Our study was limited by several factors inherent in the
imaging studies looking at subcortical brain volumes methodology of the studies selected. We were only able
12 H. J. LEE ET AL.

to include a small number of studies after application of References


exclusion criteria. This resulted in a heterogeneous
Abercrombie, H. C., S. M. Schaefer, C. L. Larson, T. R. Oakes,
selection of animal and human studies, as well as both K. A. Lindgren, J. E. Holden, S. B. Perlman, P. A. Turski,
healthy and depressed patients. The studies reviewed D. D. Krahn, R. M. Benca, et al. 1998. Metabolic rate in the
were also varied in terms of the neuroimaging and right amygdala predicts negative affect in depressed
other investigative modalities used, with their integra­ patients. Neuroreport 9 (14):3301–07. doi:10.1097/
tion requiring reference to other studies positing path­ 00001756-199810050-00028.
ways yet to be firmly established in the literature. Arco, A. D., and F. Mora. 2009. Neurotransmitters and pre­
frontal cortex–limbic system interactions: Implications for
Dosing parameters of psilocybin also varied, both in
plasticity and psychiatric disorders. Journal of Neural
amount and route of administration (e.g. oral vs. intra­ Transmission 116 (8):941–52. doi:10.1007/s00702-009-
venous), which may have affected the consistency of 0243-8.
neurobiological response. Our conclusions should be Barbon, A., L. Caracciolo, C. Orlandi, L. Musazzi, A. Mallei,
viewed as an early interpretation of potential neurobio­ L. La via, D. Bonini, C. Mora, D. Tardito, M. Gennarelli,
logical pathways, hopefully encouraging further study to et al. 2011. Chronic antidepressant treatments induce a
confirm the validity of these findings. time-dependent up-regulation of AMPA receptor subunit
protein levels. Neurochemistry International
59 (6):896–905. doi:10.1016/j.neuint.2011.07.013.
Bellani, M., M. Baiano, and P. Brambilla. 2010. Brain anatomy
Conclusion of major depression I. Focus on hippocampus.
Epidemiologia E Psichiatria Sociale 19 (4):298–301. doi:10.
Psilocybin may exert antidepressant effects through 1017/S1121189X00000634.
a variety of mechanisms. These were investigated in Bernasconi, F., A. Schmidt, T. Pokorny, M. Kometer,
a selection of studies including functional imaging E. Seifritz, and F. X. Vollenweider. 2014. Spatiotemporal
and electrophysiological techniques in humans, and brain dynamics of emotional face processing modulations
observations from imaging and behavioral studies in induced by the serotonin 1A/2A receptor agonist
animals. Although there is limited consensus and few psilocybin. Cerebral Cortex 24 (12):3221–31. doi:10.1093/
cercor/bht178.
methodologically rigorous studies, it appears that
Bus, B. A. A., and M. L. Molendijk. 2016. The neurotrophic
psilocybin alters amygdala activity and glutamate hypothesis of depression. Tijdschrift Voor Psychiatrie
levels in the medial prefrontal cortex, and increases 58 (3):215–22.
functional connectivity and synaptic plasticity in the Carhart-Harris, R. L., M. Bolstridge, C. M. J. Day, J. Rucker,
hippocampus. While psilocybin’s impact on seroto­ R. Watts, D. E. Erritzoe, M. Kaelen, B. Giribaldi,
nergic receptor systems has been described, this M. Bloomfield, S. Pilling, et al. 2018. Psilocybin with psy­
review suggests that its psychotropic therapeutic chological support for treatment-resistant depression:
Six-month follow-up. Psychopharmacology
effects may not be restricted to this neurotransmitter 235 (2):399–408. doi:10.1007/s00213-017-4771-x.
system. Further research, in particular pre-, intra-, Carhart-Harris, R. L., D. Erritzoe, T. Williams, J. M. Stone,
and post-treatment studies with molecular and ima­ L. J. Reed, A. Colasanti, R. J. Tyacke, R. Leech, A. L. Malizia,
ging biomarkers will hopefully shed light on a more K. Murphy, et al. 2012. Neural correlates of the psychedelic
unified understanding of how psilocybin may lead to state as determined by fMRI studies with psilocybin.
sustained mood improvement in individuals suffer­ Proceedings of the National Academy of Sciences
109 (6):2138–43. doi:10.1073/pnas.1119598109.
ing from depression.
Carhart-Harris, R. L., S. Muthukumaraswamy, L. Roseman,
M. Kaelen, W. Droog, K. Murphy, E. Tagliazucchi,
E. E. Schenberg, T. Nest, C. Orban, et al. 2016. Neural
Disclosure statement correlates of the LSD experience revealed by multimodal
neuroimaging. Proceedings of the National Academy of
No potential conflict of interest was reported by the authors. Sciences 113 (17):4853–58. doi:10.1073/pnas.1518377113.
Carhart-Harris, R. L., L. Roseman, M. Bolstridge,
L. Demetriou, J. N. Pannekoek, M. B. Wall, M. Tanner,
Funding M. Kaelen, J. McGonigle, K. Murphy, et al. 2017. Psilocybin
for treatment-resistant depression: FMRI-measured brain
The author(s) reported there is no funding associated with the mechanisms. Scientific Reports 7 (1): Article 1. doi:10.1038/
work featured in this article. s41598-017-13282-7.
Carvalho, A. F., M. S. Sharma, A. R. Brunoni, E. Vieta, and
G. A. Fava. 2016. The safety, tolerability and risks asso­
ORCID ciated with the use of newer generation antidepressant
drugs: a critical review of the literature. Psychotherapy
Harrison J Lee M.D. http://orcid.org/0000-0002-1966-5104 and Psychosomatics 85 (5):270–88. doi:10.1159/000447034.
JOURNAL OF PSYCHOACTIVE DRUGS 13

Castro Santos, H., and J. Gama Marques. 2021. What is the based Obstetric Anesthesia, 237–38. John Wiley & Sons,
clinical evidence on psilocybin for the treatment of psy­ Ltd. doi: 10.1002/9780470988343.app1.
chiatric disorders? A systematic review. Porto Biomedical Hendricks, P. S., M. W. Johnson, and R. R. Griffiths. 2015.
Journal 6 (1):e128. doi:10.1097/j.pbj.0000000000000128. Psilocybin, psychological distress, and suicidality. Journal
Coombs, G., M. L. Loggia, D. N. Greve, D. J. Holt, and of Psychopharmacology (Oxford, England) 29 (9):1041–43.
C. Soriano-Mas. 2014. Amygdala perfusion is predicted by doi:10.1177/0269881115598338.
its functional connectivity with the ventromedial prefrontal Hesselgrave, N., T. A. Troppoli, A. B. Wulff, A. B. Cole, and
cortex and negative affect. PloS One 9 (5):e97466. doi:10. S. M. Thompson. 2021. Harnessing psilocybin:
1371/journal.pone.0097466. Antidepressant-like behavioral and synaptic actions of psi­
Dean, J., and M. Keshavan. 2017. The neurobiology of depres­ locybin are independent of 5-HT2R activation in mice.
sion: An integrated view. Asian Journal of Psychiatry Proceedings of the National Academy of Sciences 118 (17):
27:101–11. doi:10.1016/j.ajp.2017.01.025. e2022489118. doi:10.1073/pnas.2022489118.
Dudysová, D., K. Janků, M. Šmotek, E. Saifutdinova, Hooijmans, C. R., M. M. Rovers, R. B. de Vries, M. Leenaars,
J. Kopřivová, J. Bušková, B. A. Mander, M. Brunovský, M. Ritskes-Hoitinga, and M. W. Langendam. 2014.
P. Zach, J. Korčák, et al. 2020. The effects of daytime SYRCLE’s risk of bias tool for animal studies. BMC
psilocybin administration on sleep: implications for anti­ Medical Research Methodology 14 (1):43. doi:10.1186/
depressant action. Frontiers in Pharmacology 11:602590. 1471-2288-14-43.
doi:10.3389/fphar.2020.602590. Inserra, A., D. D. Gregorio, G. Gobbi, and M. Nader. 2021.
Duman, R. S., and G. K. Aghajanian. 2012. Synaptic dysfunc­ Psychedelics in Psychiatry: Neuroplastic, immunomodula­
tion in depression: Potential therapeutic targets. Science tory, and neurotransmitter mechanisms. Pharmacological
(New York, NY) 338 (6103):68–72. doi:10.1126/science. Reviews 73 (1):202–77. doi:10.1124/pharmrev.120.000056.
1222939. Kennedy, S. H., R. W. Lam, R. S. McIntyre, S. V. Tourjman,
Duman, R. S., G. K. Aghajanian, G. Sanacora, and V. Bhat, P. Blier, M. Hasnain, F. Jollant, A. J. Levitt,
J. H. Krystal. 2016. Synaptic plasticity and depression: G. M. MacQueen, et al. 2016. Canadian network for
New insights from stress and rapid-acting antidepressants. mood and anxiety treatments (CANMAT) 2016 clinical
Nature Medicine 22 (3):238–49. doi:10.1038/nm.4050. guidelines for the management of adults with major depres­
Gabriel, F. C., D. O. de Melo, R. Fráguas, N. C. Leite-Santos, sive disorder: section 3. pharmacological treatments. The
R. A. Mantovani da Silva, E. Ribeiro, and G. Fischer. 2020. Canadian Journal of Psychiatry. 61 (9): 540–60. doi:10.
Pharmacological treatment of depression: A systematic 1177/0706743716659417.
review comparing clinical practice guideline Kometer, M., A. Schmidt, R. Bachmann, E. Studerus,
recommendations. PloS One 15 (4):e0231700. doi:10.1371/ E. Seifritz, and F. X. Vollenweider. 2012. Psilocybin biases
journal.pone.0231700. facial recognition, goal-directed behavior, and mood state
Godlewska, B. R., R. Norbury, S. Selvaraj, P. J. Cowen, and toward positive relative to negative emotions through dif­
C. J. Harmer. 2012. Short-term SSRI treatment normalises ferent serotonergic subreceptors. Biological Psychiatry
amygdala hyperactivity in depressed patients. Psychological 72 (11):898–906. doi:10.1016/j.biopsych.2012.04.005.
Medicine 42 (12):2609–17. doi:10.1017/ König, H., H.-H. König, and A. Konnopka. 2019. The excess
S0033291712000591. costs of depression: A systematic review and meta-analysis.
Gotlib, I. H., and J. Joormann. 2010. Cognition and depres­ Epidemiology and Psychiatric Sciences 29: doi:10.1017/
sion: Current status and future directions. Annual Review S2045796019000180.
of Clinical Psychology 6 (1):285–312. doi:10.1146/annurev. Kraehenmann, R., K. H. Preller, M. Scheidegger, T. Pokorny,
clinpsy.121208.131305. O. G. Bosch, E. Seifritz, and F. X. Vollenweider. 2015.
Grandjean, J., D. Buehlmann, M. Buerge, H. Sigrist, E. Seifritz, Psilocybin-induced decrease in amygdala reactivity corre­
F. X. Vollenweider, C. R. Pryce, and M. Rudin. 2021. lates with enhanced positive mood in healthy volunteers.
Psilocybin exerts distinct effects on resting state networks Biological Psychiatry 78 (8):572–81. doi:10.1016/j.biopsych.
associated with serotonin and dopamine in mice. 2014.04.010.
NeuroImage 225:117456. doi:10.1016/j.neuroimage.2020. Kuypers, K. P. C. 2020. The therapeutic potential of micro­
117456. dosing psychedelics in depression. Therapeutic Advances in
Griffiths, R. R., M. W. Johnson, M. A. Carducci, A. Umbricht, Psychopharmacology 10:2045125320950567. doi:10.1177/
W. A. Richards, B. D. Richards, M. P. Cosimano, and 2045125320950567.
M. A. Klinedinst. 2016. Psilocybin produces substantial Lépine, J.-P., and M. Briley. 2011. The increasing burden of
and sustained decreases in depression and anxiety in depression. Neuropsychiatric Disease and Treatment
patients with life-threatening cancer: A randomized 7 (Supplement 1):3–7. doi:10.2147/NDT.S19617.
double-blind trial. Journal of Psychopharmacology Lewis, C. R., K. H. Preller, R. Kraehenmann, L. Michels,
(Oxford, England) 30 (12):1181–97. doi:10.1177/ P. Staempfli, and F. X. Vollenweider. 2017. Two dose
0269881116675513. investigation of the 5-HT-agonist psilocybin on relative
Gutiérrez-Rojas, L., A. Porras-Segovia, H. Dunne, and global cerebral blood flow. NeuroImage 159:70–78.
N. Andrade-González, and J. A. Cervilla. 2020. Prevalence doi:10.1016/j.neuroimage.2017.07.020.
and correlates of major depressive disorder: A systematic Li, N.-X., Y.-R. Hu, W.-N. Chen, and B. Zhang. 2022. Dose
review. Brazilian Journal of Psychiatry 42 (6):657–72. effect of psilocybin on primary and secondary depression:
doi:10.1590/1516-4446-2020-0650. A preliminary systematic review and meta-analysis.
S. H. Halpern, Douglas. ed. 2005. Appendix: Jadad Scale for Journal of Affective Disorders 296:26–34. doi:10.1016/j.
Reporting Randomized Controlled Trials. In Evidence- jad.2021.09.041.
14 H. J. LEE ET AL.

Lowe, H., N. Toyang, B. Steele, H. Valentine, J. Grant, A. Ali, Muttoni, S., M. Ardissino, and C. John. 2019. Classical psy­
W. Ngwa, and L. Gordon. 2021. The therapeutic potential chedelics for the treatment of depression and anxiety:
of psilocybin. Molecules 26 (10):2948. doi:10.3390/ A systematic review. Journal of Affective Disorders
molecules26102948. 258:11–24. doi:10.1016/j.jad.2019.07.076.
MacQueen, G., P. Santaguida, H. Keshavarz, N. Jaworska, Price, J., V. Cole, and G. M. Goodwin. 2009. Emotional
M. Levine, J. Beyene, and P. Raina. 2017. Systematic side-effects of selective serotonin reuptake inhibitors:
Review of Clinical Practice Guidelines for Failed Qualitative study. British Journal of Psychiatry
Antidepressant Treatment Response in Major Depressive 195 (3):211–17. doi:10.1192/bjp.bp.108.051110.
Disorder, Dysthymia, and Subthreshold Depression in Price, R. B., and R. Duman. 2020. Neuroplasticity in cognitive
Adults. Canadian Journal of Psychiatry Revue Canadienne and psychological mechanisms of depression: An integra­
De Psychiatrie 62 (1):11–23. doi:10.1177/ tive model. Molecular Psychiatry Article 3. 25 (3):530–43.
0706743716664885. doi:10.1038/s41380-019-0615-x.
Mahmoudi, E., M. Faizi, R. Hajiaghaee, and A. Razmi*. 2018. Raval, N. R., A. Johansen, L. L. Donovan, N. F. Ros,
Alteration of depressive-like behaviors by psilocybe cuben­ B. Ozenne, H. D. Hansen, and G. M. Knudsen. 2021.
sis alkaloid extract in mice: The role of glutamate pathway. A single dose of psilocybin increases synaptic density and
Research Journal of Pharmacognosy 5 (2):17–24. doi:10. decreases 5-HT2A receptor density in the pig brain.
22127/rjp.2018.58486. International Journal of Molecular Sciences 22 (2):E835.
Mason, N. L., K. P. C. Kuypers, F. Müller, J. Reckweg, doi:10.3390/ijms22020835.
D. H. Y. Tse, S. W. Toennes, N. R. P. W. Hutten, Roseman, L., L. Demetriou, M. B. Wall, D. J. Nutt, and
J. F. A. Jansen, P. Stiers, A. Feilding, et al. 2020. Me, myself, R. L. Carhart-Harris. 2018. Increased amygdala responses
bye: Regional alterations in glutamate and the experience of to emotional faces after psilocybin for treatment-resistant
ego dissolution with psilocybin. Neuropsychopharmacology depression. Neuropharmacology 142:263–69. doi:10.1016/j.
Article 12. 45 (12):2003–11. doi:10.1038/s41386-020-0718-8. neuropharm.2017.12.041.
Mattson, M. P. 2008. Glutamate and neurotrophic factors in Rossi, A. F., L. Pessoa, R. Desimone, and L. G. Ungerleider.
neuronal plasticity and disease. Annals of the New York 2009. The prefrontal cortex and the executive control of
Academy of Sciences 1144 (1):97–112. doi:10.1196/annals. attention. Experimental Brain Research Experimentelle
1418.005. Hirnforschung Experimentation Cerebrale 192 (3):489–97.
McIntyre, R. S., M.-J. Filteau, L. Martin, S. Patry, A. Carvalho, doi:10.1007/s00221-008-1642-z.
D. S. Cha, M. Barakat, and M. Miguelez. 2014. Treatment- Schmaal, L., D. J. Veltman, T. G. M. van Erp, P. G. Sämann,
resistant depression: Definitions, review of the evidence, T. Frodl, N. Jahanshad, E. Loehrer, H. Tiemeier, A. Hofman,
and algorithmic approach. Journal of Affective Disorders W. J. Niessen, et al. 2016. Subcortical brain alterations in
156:1–7. doi:10.1016/j.jad.2013.10.043. major depressive disorder: Findings from the ENIGMA
Mertens, L. J., M. B. Wall, L. Roseman, L. Demetriou, major depressive disorder working group. Molecular
D. J. Nutt, and R. L. Carhart-Harris. 2020. Therapeutic Psychiatry 21 (6):806–12. doi:10.1038/mp.2015.69.
mechanisms of psilocybin: Changes in amygdala and pre­ Shamsrizi, P., B. P. Gladstone, E. Carrara, D. Luise, A. Cona,
frontal functional connectivity during emotional proces­ C. Bovo, and E. Tacconelli. 2020. Variation of effect esti­
sing after psilocybin for treatment-resistant depression. mates in the analysis of mortality and length of hospital stay
Journal of Psychopharmacology (Oxford, England) in patients with infections caused by bacteria-producing
34 (2):167–80. doi:10.1177/0269881119895520. extended-spectrum beta-lactamases: A systematic review
Moreno-Agostino, D., Y.-T. Wu, C. Daskalopoulou, and meta-analysis. British Medical Journal Open 10 (1):
M. T. Hasan, M. Huisman, and M. Prina. 2021. Global e030266. doi:10.1136/bmjopen-2019-030266.
trends in the prevalence and incidence of depression: Shao, L.-X., C. Liao, I. Gregg, P. A. Davoudian, N. K. Savalia,
A systematic review and meta-analysis. Journal of K. Delagarza, and A. C. Kwan. 2021. Psilocybin induces
Affective Disorders 281:235–43. doi:10.1016/j.jad.2020. rapid and persistent growth of dendritic spines in frontal
12.035. cortex in vivo. Neuron 109 (16):2535–44.e4. doi:10.1016/j.
Moriguchi, S., A. Takamiya, Y. Noda, N. Horita, M. Wada, neuron.2021.06.008.
S. Tsugawa, E. Plitman, Y. Sano, R. Tarumi, M. ElSalhy, Siegle, G. J., W. Thompson, C. S. Carter, S. R. Steinhauer, and
et al. 2019. Glutamatergic neurometabolite levels in major M. E. Thase. 2007. Increased amygdala and decreased dorso­
depressive disorder: A systematic review and meta-analysis lateral prefrontal BOLD responses in unipolar depression:
of proton magnetic resonance spectroscopy studies. Related and independent features. Biological Psychiatry
Molecular Psychiatry 24 (7): 952–64. Article 7. doi:10. 61 (2):198–209. doi:10.1016/j.biopsych.2006.05.048.
1038/s41380-018-0252-9. Skolnick, P., R. T. Layer, P. Popik, G. Nowak, I. A. Paul, and
Mulinari, S. 2012. Monoamine theories of depression: histor­ R. Trullas. 1996. Adaptation of N-Methyl-D-Aspartate
ical impact on biomedical research. Journal of the History of (NMDA) receptors following antidepressant treatment:
the Neurosciences 21 (4):366–92. doi:10.1080/0964704X. implications for the pharmacotherapy of depression.
2011.623917. Pharmacopsychiatry 29 (1):23–26. doi:10.1055/s-2007-
Munn, Z., T. H. Barker, S. Moola, C. Tufanaru, C. Stern, 979537.
A. McArthur, M. Stephenson, and E. Aromataris. 2020. Stuhrmann, A., T. Suslow, and U. Dannlowski. 2011. Facial
Methodological quality of case series studies: An introduc­ emotion processing in major depression: A systematic
tion to the JBI critical appraisal tool. JBI Evidence Synthesis review of neuroimaging findings. Biology of Mood &
18 (10):2127. doi:10.11124/JBISRIR-D-19-00099. Anxiety Disorders 1 (1):10. doi:10.1186/2045-5380-1-10.
JOURNAL OF PSYCHOACTIVE DRUGS 15

Tagliazucchi, E., R. Carhart‐Harris, R. Leech, D. Nutt, and JAMA Psychiatry 72 (4):334–41. doi:10.1001/jamapsychia
D. R. Chialvo. 2014. Enhanced repertoire of brain dynami­ try.2014.2502.
cal states during the psychedelic experience. Human Brain Wohleb, E. S., R. Terwilliger, C. H. Duman, and R. S. Duman.
Mapping 35 (11):5442–56. doi:10.1002/hbm.22562. 2018. Stress-induced neuronal colony stimulating factor 1
Taylor, D., T. R. E. Barnes, and A. H. Young. 2018. Prescribing provokes microglia-mediated neuronal remodeling and
Guidelines in Psychiatry. 13th ed. ed. Wiley Blackwell. depressive-like behavior. Biological Psychiatry
https://dl.uswr.ac.ir/bitstream/Hannan/32636/1/ 83 (1):38–49. doi:10.1016/j.biopsych.2017.05.026.
9781119442608.pdf. Yang, T. T., A. N. Simmons, S. C. Matthews, S. F. Tapert,
Thomas, K., B. Malcolm, and D. Lastra. 2017. Psilocybin- G. K. Frank, J. E. Max, A. Bischoff-Grethe, A. E. Lansing,
assisted therapy: a review of a novel treatment for psychia­ G. Brown, I. A. Strigo, et al. 2010. Adolescents with major
tric disorders. Journal of Psychoactive Drugs 49 (5):446–55. depression demonstrate increased amygdala activation—
doi:10.1080/02791072.2017.1320734. ScienceDirect. Journal of the American Academy of Child
Vargas, A. S., Â. Luís, M. Barroso, E. Gallardo, and L. Pereira. & Adolescent Psychiatry 49 (1):42–51. doi:10.1016/j.jaac.
2020. Psilocybin as a new approach to treat depression and 2009.09.004.
anxiety in the context of life-threatening diseases—a systema­ Zhao, J.-L., W.-T. Jiang, X. Wang, Z.-D. Cai, Z.-H. Liu, and
tic review and meta-analysis of clinical trials. Biomedicines G.-R. Liu. 2020. Exercise, brain plasticity, and depression.
Article 9. 8 (9):331. doi:10.3390/biomedicines8090331. CNS Neuroscience & Therapeutics 26 (9):885–95. doi:10.
Vollenweider F. 1999. 5-HT Modulation of Dopamine Release 1111/cns.13385.
in Basal Ganglia in Psilocybin-Induced Psychosis in Man—A Zou, K., W. Deng, T. Li, B. Zhang, L. Jiang, C. Huang, X. Sun, and
PET Study with [11C]raclopride. Neuropsychopharmacology X. Sun. 2010. Changes of brain morphometry in first-episode,
20 (5):424–433. doi:10.1016/S0893-133X(98)00108-0. drug-naïve, non–late-life adult patients with major depression:
Walker, E. R., R. E. McGee, and B. G. Druss. 2015. Mortality an optimized voxel-based morphometry study. Biological
in Mental disorders and global disease burden implications. Psychiatry 67 (2):186–88. doi:10.1016/j.biopsych.2009.09.014.

You might also like