Next Article in Journal
AXL-Receptor Targeted 14FN3 Based Single Domain Proteins (Pronectins™) from 3 Synthetic Human Libraries as Components for Exploring Novel Bispecific Constructs against Solid Tumors
Next Article in Special Issue
SLC6A3, HTR2C and HTR6 Gene Polymorphisms and the Risk of Haloperidol-Induced Parkinsonism
Previous Article in Journal
A Subset Screen of the Compounds Australia Scaffold Library Identifies 7-Acylaminodibenzoxazepinones as Potent and Selective Hits for Anti-Giardia Drug Discovery
Previous Article in Special Issue
Pharmacogenetics and Schizophrenia—Can Genomics Improve the Treatment with Second-Generation Antipsychotics?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Antipsychotics-Induced Changes in Synaptic Architecture and Functional Connectivity: Translational Implications for Treatment Response and Resistance

Section of Psychiatry, Laboratory of Translational and Molecular Psychiatry and Unit of Treatment-Resistant Psychosis, Department of Neuroscience, Reproductive Sciences and Odontostomatology, University Medical School of Naples “Federico II”, Via Pansini 5, 80131 Naples, Italy
*
Author to whom correspondence should be addressed.
Biomedicines 2022, 10(12), 3183; https://doi.org/10.3390/biomedicines10123183
Submission received: 27 October 2022 / Revised: 2 December 2022 / Accepted: 4 December 2022 / Published: 8 December 2022
(This article belongs to the Special Issue Antipsychotics: 70 Years)

Abstract

:
Schizophrenia is a severe mental illness characterized by alterations in processes that regulate both synaptic plasticity and functional connectivity between brain regions. Antipsychotics are the cornerstone of schizophrenia pharmacological treatment and, beyond occupying dopamine D2 receptors, can affect multiple molecular targets, pre- and postsynaptic sites, as well as intracellular effectors. Multiple lines of evidence point to the involvement of antipsychotics in sculpting synaptic architecture and remodeling the neuronal functional unit. Furthermore, there is an increasing awareness that antipsychotics with different receptor profiles could yield different interregional patterns of co-activation. In the present systematic review, we explored the fundamental changes that occur under antipsychotics’ administration, the molecular underpinning, and the consequences in both acute and chronic paradigms. In addition, we investigated the relationship between synaptic plasticity and functional connectivity and systematized evidence on different topographical patterns of activation induced by typical and atypical antipsychotics.

1. Introduction

Seventy years after their introduction in therapy, antipsychotics still represent the cornerstone of schizophrenia pharmacological treatment as well as a relevant therapy, along with mood stabilizers, in the treatment of bipolar disorders and a validated strategy for augmenting antidepressants’ action [1,2]. Most antipsychotics, albeit with individual differences, share the common feature of occupying the dopamine D2 receptor (D2R), which is considered the main mechanism responsible for their therapeutic effect [3]. Despite the relevance of dopamine-related action, atypical antipsychotics and novel compounds in experimental trials modulate other neurotransmitter pathways, including the serotonergic, glutamatergic, and adrenergic ones with a putative beneficial effect on psychotic symptoms [4,5,6]. Mounting evidence suggests that antipsychotics can lead to morphological and synaptic changes in the brain through the modulation of processes that regulate synaptic plasticity, dendritic spine architecture, and postsynaptic density [5,7]. In addition, antipsychotics may promote a functional reorganization of neural connections and modify the correlation pattern existing between pairs of brain regions [8,9].
The investigation of antipsychotics’ mechanism of action at a molecular and network level is particularly relevant for those diseases that are conceptualized as disorders of both synaptic plasticity and brain cortical-subcortical connectivity, such as schizophrenia [10,11,12]. Specifically, several studies have supported the neurodevelopmental hypothesis for schizophrenia by the direct detection of synaptic plasticity and postsynaptic density (PSD) alterations in patients with psychosis compared to controls [13,14]. The PSD, which is a major contributor to changes in synaptic architecture under multiple stimuli including antipsychotics’ administration, is a highly organized structure beneath the neuronal membrane mainly at glutamate synapses and comprises receptors, scaffolding, and adaptor proteins [15,16]. Moreover, in recent years, the psychopathological manifestations of schizophrenia have been attributed to the loss of neuronal synchrony ensuing in a global dysconnectivity between and within brain networks [17]. Thus, even though the anatomical alteration in schizophrenia is placed at a microscopic level, for instance at PSD, the resulting macroscale manifestation is the impairment in functional interactions among brain regions [18].
Here, we addressed the following questions:
  • How do changes in gene expression induced by antipsychotics impact brain synapses and connectivity?
  • Where do these changes occur in terms of brain region topography?
  • Are the brain regions involved coherent with the modeling of schizophrenia pathophysiology?
  • And finally, how could these changes translate in terms of implications for response and resistance to antipsychotic treatment?
To answer these questions, we described here the molecular mechanisms underlying the interrelationship between changes in synaptic plasticity, mainly PSD, and brain functional connectivity, assessing the impact of antipsychotics on synaptic architecture and neuronal activity, as well as highlighting differences in the topographical pattern of co-activation induced by different antipsychotics alone or in combination. Further, we analyzed the regional expression of immediate early genes (IEGs) after the administration of antipsychotics, which increases rapidly and transiently in response to neuronal activation and represents a valuable marker of postsynaptic enhancement [19]. In addition, we recapitulated the major findings from preclinical studies investigating the impact of molecules on brain functional connectivity and matched this information with putative beneficial antipsychotic effects. Lastly, we critically evaluated the methods used to explore connectivity in preclinical models of schizophrenia, focusing on the pros and cons of each procedure and pointing to novel future applications for responsiveness and resistance to antipsychotics.

2. Materials and Methods

We performed a systematic search of the literature available on Medline/PubMed and Embase in accordance with the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) criteria [20] by combining the following keywords in a search string: antipsychotic, chlorpromazine, haloperidol, paliperidone, risperidone, asenapine, olanzapine, clozapine, quetiapine, amisulpride, aripiprazole, brexpiprazole, cariprazine, lurasidone, lumateperone, xanomeline, Arc/Arg, BDNF, c-fos, fos, c-Jun, Jun, Egr1, Delta, Narp1, NPAS-4, Homer, Nor1, Nurr1, NGFI-B/Nur77, Nerve Growth Factor Inducible-B, ISH, fMRI, EEG, functional connectivity, connectome, network. We registered the review protocol on the International Platform of Registered Systematic Review and Meta-Analysis Protocols (INPLASY2022110107). The search was conducted on 10 September 2022, and updates were checked until 14 October 2022. Hand-searching and checking the reference lists of the included articles turned up further papers. The search strings are available in the Supplementary Material. The search returned 5899 articles, which resulted in 4455 after removing duplicates. We deemed eligible English-written preclinical studies published in peer-reviewed journals that were pertinent to the topic, i.e., investigating the antipsychotic effects on synaptic architecture, IEGs expression, and, on a larger scale, anatomical and functional brain connectivity, without time or design methodology constraints. Although the names of the most widely tested antipsychotics in preclinical studies were used as keywords, others, selected from the generic term “antipsychotic”, also fit the inclusion criteria. EndNote X9 was used to manage all records. Two independent reviewers independently carried out screening and full-text assessment procedures (AB and GDS). Overall, 54 studies were included in the qualitative synthesis. The PRISMA flow diagram in Figure 1 reports the specifics of the search and selection procedures.

3. The Close Relationship between Brain Functional Connectivity and Postsynaptic Density

In recent years, the hypothesis that cognition and behavior rely on the neuronal capability to process and integrate information within and between specific brain network communities has attracted growing interest [21]. Network modules or communities are clusters of nodes that exhibit a highly correlated activity to attend to the same specialized tasks and are interlinked by hub regions [22]. Nodes represent the basic elements of networks, and their interconnections are depicted as weighted or unweighted, directed or undirected edges [22]. Individual nodes can show a low or high number of connections and, in the latter case, are called hubs [22]. A graphical representation of network attributes is shown in Figure 2. In brain networks, nodes are often but not exclusively used to provide topographical information on interregional patterns of activity and generally are representative of cortical and subcortical portions [23,24]. Given the relevance of brain networks in the expression of cognitive and behavioral phenotypes, it is not surprising that functional connectivity, also defined as functional networks, could have a major role in the understanding of the neurobiological correlates of schizophrenia and treatment-resistant schizophrenia [25,26,27].
While structural connectivity aims to identify white matter tracts that physically bridge brain regions, functional connectivity is involved in describing statistical co-activation patterns between different areas [28]. However, differences in mesoscale properties, such as nodes clustering, have been detected between functional and structural networks probably for the existence of polysynaptic connections not mediated by white matter fibers [28]. Specifically, while the analysis of correlations between the spontaneous blood oxygenation level-dependent (BOLD) signals in different brain regions allows for the identification of intrinsic connectivity in large-scale networks such as the default mode network (DMN), white matter tracts connecting parts of the parietal cortex and precuneus involved in the DMN have not been found [28]. Therefore, functional connectivity is not necessarily reflected by white fiber-mediated anatomical connections, emphasizing the importance of functional versus structural information in the study of psychotic disorders [28].
Accurate task-specific patterns of functional connectivity between large-scale brain circuits rest on well-organized processes of signal transmission and depend on synaptic integrity at the neuronal level [29]. In neurons, the information flow is conducted as an electrical signal along the axon, from dendrites to synapses, where a translation into a chemical message takes over [30]. An integral and functional synapsis is then required to spread the information flow over other neurons and allow efficient data processing in the whole central nervous system (CNS). In this framework, a balanced modulation of synaptic plasticity processes, through direct regulation of neuron–neuron interaction, is mandatory to orchestrate the correct topography of functional brain networks [18]. Specifically, neurotransmitter receptors, scaffold proteins, intracellular signaling molecules, and other elements comprised in the PSD, are essential to ensure proper communication between neurons [31]. Otherwise, altered synaptic plasticity may disrupt individual interactions among neurons, turning in functional dysconnectivity among entire brain regions [18].
Efforts to understand how synaptic changes can affect the overall brain activity as well as the connectivity of specific brain regions have pointed to the role of structural and functional rearrangement at the dendritic spine, where qualitative and quantitative alteration of PSD proteins (for instance, the NMDAR and the partner scaffold PSD-95) may modulate the synaptic strength, influence the reciprocal interaction between cluster of neurons (or of glia cells), and in turn impact the connectivity of functionally related brain regions [32].
Alterations in processes that regulate synaptic plasticity and PSD architecture may affect brain connectivity, disrupting the synchronized firing of those neurons organized in a high-frequency oscillation network (30–90 Hz/s), known as gamma waves, involved in cognition and memory functions [33,34,35]. Studies have associated altered gamma oscillations with neurological and psychiatric disorders, including Alzheimer’s disease, Parkinson’s disease, and schizophrenia [36,37]. Moreover, gamma oscillations were demonstrated to be modulated by antipsychotic treatment in preclinical studies [38,39,40]. In particular, risperidone and clozapine but not haloperidol were able to decrease phencyclidine (PCP)-induced prefrontal and cortical-hippocampal hypersynchrony without restoring hippocampal and circuit desynchronization probably responsible for their antipsychotic effect [40].
With regard to the pathophysiology of psychotic disorders, postmortem studies have unveiled significant differences in PSD proteins and presynaptic markers in schizophrenia patients compared to normal controls [41,42,43,44]. A computational model network was adopted to simulate how variability in excitatory synaptic strength between parvalbumin interneurons could regulate gamma power [44]. The variability in excitatory synaptic strength was computed by quantifying the vesicular glutamate transporter 1 (VGlut1) and PSD-95 protein levels in the postmortem brain of 20 matched pairs of controls and schizophrenia subjects [44].
Martin et al. reported the effects of olanzapine treatment on dysregulated gene expression in a non-human primate model of schizophrenia [45]. PCP-exposed primates showed abnormalities in 146 transcripts compared to controls, and olanzapine administration was able to reverse about the 34% of PCP-induced changes in gene expression [45]. In particular, several genes coding for synaptic and PSD proteins were found down-regulated in the pharmacological model of psychosis and enriched following antipsychotic treatment [45].
In this framework, schizophrenia has been conceptualized both as a synaptopathy and a dysconnectivity syndrome, in which improper processing of synaptic inputs and loss of signal integration between brain areas may account for psychopathological manifestations [17]. Moreover, a stable and robust re-organization of neuronal and regional interconnections in a pathological manner may support the persistence of psychotic symptoms [18]. The study of the effect of antipsychotics on functional brain connectivity and synaptic plasticity is, therefore, necessary to understand their real impact on the neurobiological core of clinical manifestations. By directly controlling confounding factors, preclinical models may provide further information on the molecular modifications induced by antipsychotics and their modulation of functional connectivity. Lastly, animal studies can help in developing translational models of prediction based on molecular parameters of postsynaptic density and interregional connections to foresee the efficacy and side effects of novel compounds.

4. The Postsynaptic Density: A Molecular Hub Involved in Neuronal Communication and IEGs Expression

The PSD is an electron-dense thickening underneath the glutamatergic synapse [15] consisting of packed proteins forming a disc-shaped molecular platform whose width is a few hundred nm and whose thickness is ∼30–50 nm [46,47]. It consists of approximately 1500 molecules including neurotransmitter receptors (e.g., α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor—AMPAR-, N-methyl-D-aspartate receptor—NMDAR-, metabotropic glutamate receptors—mGluR-), adaptors and adhesion molecules (N-cadherin, neuroligins), and scaffolding proteins (e.g., Shank, Homer, PSD-95, GKAP) that have been proven to form condensates at physiological concentrations [16,48].
The PSD has been conceived as a structural and functional hub that is responsible for modeling the architecture of the dendritic spine and regulating the propagation of intracellular signals in response to various neuronal or environmental stimuli, resulting in modifications of synaptic plasticity and metaplasticity [49], and therefore relevant to the pathophysiology of several neuropsychiatric disorders [50]. Specifically, the role of PSD is to cluster postsynaptic receptors in proximity to released neurotransmitters to ensure a proper translation of the information flow through the synaptic cleft [51]. At the postsynaptic site, the activation of the glutamatergic receptors enhances multiple downstream signaling cascades, such as cAMP-dependent protein kinase (PKA), mitogen-activated protein kinase (MAPK), phospholipase C (PLC), calmodulin-dependent protein kinase II and IV (CaMKII/IV), phosphoinositide 3-kinase (PI3K), and Protein kinase B (PKB/Akt), that converge on cAMP-response element binding protein (CREB) phosphorylation [52,53]. The phosphorylation of CREB at Ser133 is mandatory to induce IEGs transcription, upon the binding to the cAMP response element (CRE) sequence of the promoters of its target genes and the recruitment of CREB-binding protein (CBP) [54].
IEGs comprise a large category of genes, whose expression is induced rapidly and transiently after postsynaptic activation and include, among others, c-fos, zif268, brain-derived neurotrophic factor (BDNF), Narp, Homer1a, and Arc [19]. Beyond their capability to fine-tune plastic and metaplastic processes involved in the regulation of long-term potentiation (LTP) and depression (LTD), IEGs represent a noteworthy molecular marker of synaptic integrity and postsynaptic activation due to their fleeting pattern of expression [19]. For the purposes of this paper, we dwelled on IEGs regional expression, with special regard to topographic changes induced by typical and atypical antipsychotics.
Over the well-known effect of glutamatergic neurotransmission on synaptic plasticity, other pathways have been variously related to PSD (Figure 3). In particular, the dopamine D1 receptor (D1R) was seen to interact directly with the NMDA receptor subunits 1 (NR1) and indirectly with the NMDAR by the interposition of PSD-95 (Figure 3) [55]. Moreover, the serotonergic 2A receptor (5-HT2AR) was found to physically engage with multiple PDZ protein-1 (MUPP1) and PSD-95 at the dendritic spines of cortical pyramidal neurons (Figure 3) [56]. Therefore, not only glutamatergic but also serotonergic and dopaminergic pathways, as well as adrenergic and muscarinic receptors, can interact and modulate the composition of the PSD, with consequent changes in processes that regulate synaptic plasticity and interneural communicability [57,58]. Thus, typical and atypical antipsychotics through individual pharmacodynamic profiles could diversely impact PSD changes and exhibit specific spatial effects on brain regions. This issue appears to be particularly relevant for defining antipsychotics-based differences in the modulation of interregional functional connectivity.
Aberrant processes of assembling and disassembling PSD structures have been conceptualized as the possible neurobiological basis for psychosis as well as for the delivery of antipsychotic effects. In fact, multiple lines of evidence have suggested that psychiatric disorders, including schizophrenia and affective disorders, may be related to the disruption of synaptic structure, function, and capability to make connections between different brain regions [59,60,61]. In this regard, a recent meta-analysis examined postsynaptic elements in the postmortem brain tissue of patients affected by schizophrenia compared to healthy controls and found an overall decrease in dendritic spine density (DSD) in cortical but not subcortical regions [13]. A significant reduction in the PSD-95 and mGluR1protein levels as well as a decrease in the ratio between Homer1b/c and Homer1a isoforms were found in the hippocampus of schizophrenia patients compared to controls [62]. Other studies proved an alteration of PSD-95 protein expression also in the dorsolateral prefrontal cortex (DLPFC) of patients with schizophrenia [63,64]. Through the application of a machine learning-based algorithm, a recent postmortem study identified multiple clusters of molecular markers of schizophrenia, including several proteins involved in the regulation of synaptic plasticity processes [41]. The above-mentioned defects detected in the PSD of psychotic patients and even others, such as the loss of neuropil, reduced cell size and synaptic counts, and a decrease in the protein expression of synaptophysin, strongly support the neurodevelopmental hypothesis for schizophrenia [14]. Overall, alterations in molecular components of the PSD in schizophrenia patients could result in an improper clustering of postsynaptic receptors, with a following impairment in the neuron–neuron interaction and whole-brain functional connectivity.
The extent to which antipsychotics modulate receptor activity is essential to ensure proper processes of neurite outgrowth, synaptogenesis, and spine organization [65]. In this framework, antipsychotics have shown a different capability to regulate PSD composition and spine architecture by exhibiting different pharmacodynamic profiles and acting through epigenetic mechanisms [65,66,67,68,69]. Specifically, aripiprazole and clozapine were found to increase PSD-95 protein levels, the number of dendritic spines, and the number of PSD puncta in cortical neuron cultures [70]. Otherwise, haloperidol had opposite effects on synaptic architecture by reducing PSD-95 levels, the number of spines, and the number of PSD puncta [70]. In another study, the administration of haloperidol, but not clozapine, was associated with an increase in the number of perforated synapses, which are characterized by a discontinuous density when observed in transversal sections, within the caudate nucleus [71]. On the other hand, haloperidol showed to significantly reduce the average volume of dendritic spines in layer VI neurons of the PFC, whereas clozapine increased the number of symmetric axospinous synapses [72]. Furthermore, olanzapine and aripiprazole, but not haloperidol, increased PSD-95 protein levels in rat hippocampal neuronal cultures under both normal and toxic conditions [73]. In prefrontal cortical neurons, olanzapine was shown to prevent PCP-induced reduction in neurite length and branches as well as the decrease in both synaptophysin and PSD-95 mRNA and protein levels through a neuregulin-mediated mechanism [74]. The application of the D2R-specific agonist quinpirole impaired the synaptic architecture of cortical neuron cultures by reducing spine branches and length, along with a decrease in PSD-95 and synaptophysin levels, especially in animals showing a dysfunctional copy of the disrupted schizophrenia 1 (DISC1) gene [75]. In the same study, the pre-treatment with aripiprazole and haloperidol significantly prevented the quinpirole-induced changes in spine morphology, with a more effective impact exerted by aripiprazole [75]. Interestingly, the two antipsychotics could act through different intracellular signaling, with aripiprazole, but not haloperidol, able to increase pAkt-Thr308 levels in cortical neurons [75]. In monocyte-derived-neuronal-like cells (MDNCs) isolated from schizophrenia patients and healthy controls, incubation with haloperidol showed no capability to modulate the structure and the pruning of primary and secondary neurites [76]. Therefore, the effects of antipsychotics on synaptic architecture could depend on the specific brain region examined, with relative differences between cortical and subcortical areas [71,72,73]. In addition, antipsychotics could restore, at least in part, a working pattern of brain functional connectivity through the regulation of the synaptic architecture and spine morphology as well as PSD composition and IEGs expression, with differences between typical and atypical compounds [65,70,71,72,73,75].
Given their role as markers of postsynaptic activity, IEGs may represent a remarkable candidate to investigate how acute or chronic antipsychotics administration could be responsible for setting the molecular framework for long-term changes at the PSD level, resulting in clinical phenotypes, brain morphological variations, and treatment responses [77,78,79]. In addition, we represented, in Figure 4, the chemical structure of all antipsychotics and drugs with potential antipsychotic effects discussed in the present review by consulting the PubChem database [80].

5. Antipsychotics-Mediated IEGs Regulation: Toward a Topographical Analysis of Genes Expression Patterns

Several studies attempted to explore the effect of individual antipsychotics on the IEGs spatial pattern of expression in a preclinical setting. Behind their molecular tasks and their capability to influence synaptic plasticity and metaplasticity, IEGs expression was used specifically to obtain topographical information on regional brain activation. We gained evidence on antipsychotics-induced increase or decrease in IEGs expression, with regard to time of exposure, region of interest, and type of IEG involved, as shown in Table 1. Antipsychotic action on PSD proteins’ assembly and interaction may trigger changes in dendritic spines and therefore influence trans-synaptic communication between neurons [70,81].
Multiple differences were found in the capability of individual molecules to modulate neuronal activity in the animal brain, especially between typical and atypical antipsychotics.
Of interest, haloperidol, a first-generation antipsychotic with a strong affinity for the D2R [82], was found to regulate striatal activity by increasing the expression of IEGs, such as Homer1a, Arc, c-fos, zif-268, Ania-3, and Nurr77 in both an acute and chronic paradigm [79,83,84,85,86,87,88,89,90,91,92,93,94,95]. Inconsistent evidence linked haloperidol activity with the modulation of IEGs in cortical regions [96], showing a decrease in Homer1a, Arc, and Nurr77 expression in the somatosensory, cingulate, insular, motor, and medial agranular cortices [79,92,93,97] and an increase of Nurr77 and Krox-20 in the PFC [83,88] and of zif-268 and Homer1a in the somatosensory, cingulate, insular, motor, and medial agranular cortices [85,86,94].
Amisulpride, an “atypical” second-generation antipsychotic, that is almost a pure D2/D3R antagonist with relevant presynaptic action, was found to effectively treat both positive and, even if with less degree, negative symptoms in psychosis, inducing only a few extrapyramidal side effects through the capability to enhance dopaminergic neurotransmission [98]. In preclinical studies, the acute administration of amisulpride was shown to increase IEGs expression in both cortical and subcortical regions [85,94].
Three atypical antipsychotics, risperidone, lurasidone, and ziprasidone, were found to increase the expression of arc, zif-268, Homer1a, and c-fos in the striatum, hippocampus, PFC, cingulate cortex, and other cortical regions, with specific patterns of expression depending on the delivered dose [86,99,100,101,102].
Repetitive administrations of quetiapine, olanzapine, and clozapine were associated with an overall increase in BDNF expression, especially in the hippocampus and PFC [103,104,105,106,107,108]. Otherwise, acute administrations of clozapine and olanzapine demonstrated the capability to modulate both cortical and striatal regions and induced a heterogenous pattern of IEGs modifications, based on the specific gene regulated, the dose, and the brain area involved [79,83,86,87,88,89,90,91,93,97,99,109].
In addition, asenapine also was found to differently modulate IEGs expression according to the time of exposure. Specifically, an elevation in c-fos, zif-268, and Homer1a was detected in the striatum and motor, insular, somatosensory, medial agranular, and cingulate cortices after acute injection, whereas chronic administration was associated with a decrease in Homer1a and arc expression in cortical regions and an increase in the striatum [79,93].
The chronic administration of aripiprazole, a third-generation antipsychotic with a D2R partial agonism, showed to reduce BDNF expression in the hippocampus and, simultaneously, raise arc mRNA levels in the striatum and PFC [110].
A new compound with antipsychotic effects, SEP-363856 (SEP-856), was tested in a preclinical setting to define the spatial profile of IEGs induction [111]. SEP-856 was found to increase the expression of Arc, c-Fos, Zif-268, and Npas4 in the hippocampus and PFC through a possible agonism at 5-HT1AR and trace amine-associated (TAAR1) receptors [111].
A comprehensive summary of the specific IEGs expression pattern induced by single antipsychotics is provided in Table 1, with consideration of the brain region affected, the administration paradigm, and the type of IEG included.
Table 1. The topographical pattern of IEGs expression induced by antipsychotics.
Table 1. The topographical pattern of IEGs expression induced by antipsychotics.
Antipsychotic DrugAdministrationAuthorsBrain RegionIEGs Expression
AmisulprideAcutede Bartolomeis et al., 2013 [85]Caudate-Putamenc-fos,Arc,Homer1a,Zif-268
Cingulate CortexZif-268
MCZif-268,Ania-3
de Bartolomeis et al., 2016 [94]SS,
Cingulate Cortex,
M2,
MC,
NAc,
Caudate-Putamen
Homer1a
AripiprazoleChronicLuoni et al., 2014 [110]HippocampusBDNF
PFC,
Striatum
Arc
AsenapineAcutede Bartolomeis et al., 2015 [79]Caudate-Putamen,
NAc Shell and Core
Arc,c-fos,zif-268,Homer1a
SS,
IC
Homer1a
M2,
MC,
Cingulate cortex
Homer1a,Arc
ChronicBuonaguro et al., 2017 [93]StriatumHomer1a,Arc
IC,
M2,
MC,
SS
Homer1a,Arc
ClozapineAcuteBeaudry et al., 2000 [83]PFC,
Cingulate Cortex,
NAc Shell,
Dorsolateral Striatum
Nurr77
Nguyen et al., 1992 [90]StriatumZif-268
MacGibbon et al., 1994 [89]Striatum,
NAc,
Ventral Septum
c-fos,Jun B,Krox24,FRA
Langlois et al., 2001 [97]SSNurr77
Cochran et al., 2002 [84]PFCc-fos,fos B,fra2
Mediodorsal thalamusc-fos
Polese et al., 2002 [87]NAcc-fos,Homer1a
Caudate-PutamenHomer1a
Werme et al., 2000 [91]NAc Shellc-fos,Nor1,Nurr77
Robbins et al., 2008 [88]NAcc-fos
Hippocampusc-jun
Striatumc-fos,sgk-1,Narp
Thalamusc-fos,arc
Pinna et al., 2019 [99]PL,
IL,
PFC
c-fos
Chronic Kontkanen et al., 2002 [105]PFC,
FC,
Cingulate Cortex
c-fos,c-jun,junB
Kim et al., 2012 [108]Whole brainBDNF
Verma et al., 2007 [106]PFCc-fos
Rizig et al., 2012 [103]Whole brainBDNF
Bai et al., 2003 [104]HippocampusBDNF
HaloperidolAcuteBeaudry et al., 2000 [83]PFC,
NAc Shell,
Dorsolateral Striatum
Nurr77
Cochran et al., 2002 [84]Mediodorsal thalamusc-fos
NAc Shellfra2
de Bartolomeis et al., 2015 [79]NAc Shell and Core,
Caudate-Putamen,
Lateral septum
Homer1a,Arc,c-fos,zif-268
SS,
M2,
MC,
Cingulate cortex,
IC
Arc
Robbins et al., 2008 [88]Thalamus↑ fra-1, ↓ BDNF
Striatumc-fos, ↑ Krox-20,arc
NAc Shell and Corec-fos, ↑ Krox-20,arc,c-jun
PFCKrox-20
De Bartolomeis et al., 2013 [85]Cingulate CortexHomer1a
Caudate-Putamen,
NAc Shell and Core
Ania-3,c-fos,Homer1a,Zif-268,Arc,Norbin
IC,
MC,
SS
Zif-268
Iasevoli et al., 2010 [95]Caudate-Putamen,
NAc Shell and Core
Ania-3,arc,Homer1a
Cingulate Cortex,
M2
Ania-3,Homer1a
MC,
SS,
IC
Homer1a
De Bartolomeis et al., 2016 [94]Caudate-Putamen,
MC
Homer1a
MacGibbon et al., 1994 [89]Striatum,
NAc,
Ventral Septum
c-fos,Jun B,Krox24,FRA
Nguyen et al., 1992 [90]StriatumZif-268,c-fos
Polese et al., 2002 [87]NAc,
Caudate-Putamen
c-fos,Homer1a
Langlois et al., 2001 [97]SSNurr77
Werme et al., 2000 [91]NAc Shell and Core,
Caudate-Putamen
c-fos,Nor1,Nurr77
Dell’Aversano et al., 2009 [92]Caudate-Putamen,
NAc Core
Ania-3,Homer1a
Parietal CortexHomer1a
Iasevoli et al., 2010 [86]StriatumHomer1a
Iasevoli et al., 2010 [86]StriatumHomer1a
Transient treatmentSamaha et al., 2008 [112]Striatumc-fos
ChronicBuonaguro et al., 2017 [93] Caudate-Putamen,
NAc
Homer1a
IC,
MC,
M2,
SS
Homer1a,Arc
Fumagalli et al., 2009 [109]StriatumArc
Verma et al., 2007 [106]PFCc-fos,Zif-268
Iasevoli et al., 2010 [86]Striatum,
Cingulate Cortex,
MC,
M2,
SS,
IC
Homer1a
Iasevoli et al., 2010 [86]Caudate-Putamen,
NAc
Homer1a,Ania-3
SS,
MC,
M2,
Cingulate Cortex
Homer1a
Parikh et al., 2004 [113]HippocampusBDNF
LurasidoneAcuteLuoni et al., 2014 [101]StriatumArc,zif-268
HippocampusArc,Npas4
ChronicLuoni et al., 2014 [101]Hippocampus,
PFC,
Striatum
Arc,zif-268
Fumagalli et al., 2012 [102]PFC,
Hippocampus
BDNF
OlanzapineAcuteDe Bartolomeis et al., 2015 [79]M2,
MC,
Cingulate cortex,
IC
Homer1a
Caudate-Putamen,
NAc Shell
c-fos,zif-268,Homer1a
Iasevoli et al., 2010 [86]StriatumHomer1a
Pinna et al., 2019 [99]BSTL,
IPAC,
SLEA,
CeA
c-fos
Buonauguro et al., 2017 [93]Striatum,
NAc Shell
Homer1a,Arc
SS,
IC,
MC,
M2,
Cingulate Cortex
Arc
Fumagalli et al., 2009 [109]Striatum,
FC
Arc
ChronicBai et al., 2003 [104]HippocampusBDNF
QuetiapineChronicPark et al., 2006 [107]HippocampusBDNF
RisperidoneAcuteRobinet et al., 2001 [100]Striatumc-fos
Iasevoli et al., 2010 [86]StriatumHomer1a
Pinna et al., 2019 [99]PL,
IL,
PFC
c-fos
ChronicRobinet et al., 2001 [100]Striatumc-fos
SertindoleAcuteIasevoli et al., 2010 [95]SS,
IC
Homer1a
ZiprasidoneAcuteIasevoli et al., 2010 [86]Striatum,
SS,
MC,
Cingulate cortex,
IC
Homer1a
ChronicIasevoli et al., 2010 [86]Striatum,
MC,
IC
Homer1a
SEP-856AcuteBegni et al., 2021 [111]PFCArc, ↑ c-Fos, ↑ Egr1, ↑ Npas4
HippocampusArc, ↑ c-Fos
↑ = Gene expression is up-regulated; ↓ = gene expression is down-regulated; BSTL = bed nucleus of stria terminalis lateral division; CeA = central amygdaloid; FC = frontal cortex; IC = insular cortex; IL = infralimbic; IPAC = interstitial nucleus of the posterior limb of the anterior commissure; M2= medial agranular cortex; MC = motor cortex; NAc = nucleus accumbens; PFC = prefrontal cortex; PL= prelimbic; SLEA = sublenticular extended amygdale; nucleus; SS = somatosensory cortex.

6. Methods to Study Brain Functional Connectivity in Translational Models: A Brief Overview of Pros and Cons

Even though there are several methods to explore the impact of antipsychotic drugs on brain functional connectivity in preclinical models, two are the most widely applied: functional magnetic resonance imaging (fMRI) and, on the other hand, electrophysiology/electroencephalogram (EEG), which assess interregional patterns of co-activation and neuronal synchrony by measuring brain blood flow and electrical activity, respectively [114,115,116].
However, these methods show several limitations. First, fMRI measures regional differences in cerebral blood flow, indirectly reflecting presynaptic activity and not postsynaptic modifications, which are more representative of those neuroplastic processes altered in schizophrenia [117]. In addition, fMRI acquisitions can be affected by small head movements and often require sedation [118]. On the other hand, sedative drugs can significantly affect functional connectivity and provide different patterns of co-activation depending on the specific compounds adopted, thus compromising the consistency and reproducibility of individual studies [118]. Although feasible in multiple settings and characterized by elevated temporal resolution, EEG applications may suffer from reduced topographical resolution with difficulty in detecting the exact location where the activity is generated and are characterized by a poor signal-to-noise ratio, which severely impairs the sensitivity and specificity of the data [119].
Moreover, human studies have difficulties in exploring the effect of different antipsychotics on network connectivity, and antipsychotic treatment is rather considered a confounding variable. Except for research conducted on drug-naive participants, it may be challenging to distinguish between the structural and functional changes induced by antipsychotic agents and those attributable to the disease itself. In this context, preclinical research may enable us to address these shortcomings, by analyzing changes in functional connectivity after pharmacological challenges in animals.
To overcome these limitations, two studies explored changes in functional connectivity after antipsychotic administration by directly assessing IEGs expression through an in situ hybridization (ISH) protocol and generating group-specific correlation matrices to obtain details on differential patterns of co-activation between brain areas [8,9]. The simultaneous detection of molecular markers of postsynaptic activity and spatial information offers a unique opportunity to relate alterations in neuronal plasticity to connectivity dysregulation and represents a novel method for testing the effectiveness of antipsychotics in preclinical models.
Furthermore, novel molecular techniques, such as ISH chain reaction (HCR), will probably enable a step forward in the research of the neurobiology underlying psychiatric disorders and help shed light on antipsychotics’ mechanism of action due to the possibility of matching topographical information with transcriptomic or proteomic data [120].

7. Antipsychotics-Mediated Modulation of Functional Connectivity: Preliminary Evidence from Preclinical Studies

Several antipsychotics and even new molecules in experimentation were tested in preclinical models to assess their putative impact on functional connectivity, especially but not exclusively in brain networks disrupted by the injection of drugs that induce a psychotic-like phenotype, as outlined in Table 2.
Of interest, the administration of N-methyl-D-aspartate receptor (NMDAR) antagonists such as PCP, MK-801, and ketamine [121,122] to properly model manifestations resembling schizophrenia in animals determined a disconnection between and within the neuronal activity of the hippocampus and prefrontal cortex (PFC). PCP was found to induce different patterns of brain activity depending on its levels, with high doses globally impairing functional connectivity and low doses affecting mainly the hippocampal and frontal regions [123]. The alteration in the hippocampal-PFC pathway is particularly relevant to the physiopathology of schizophrenia, as this pathway is involved in the regulation of learning, behavioral flexibility, novelty processing, and memory and has been found to be impaired in patients with schizophrenia [124,125,126,127].
Two electrophysiological studies found an overall dysregulation in the activity of pyramidal cells of the PFC in rats and mice treated with PCP, including alterations in the neuronal firing rate, reduced cortical synchrony, and a decrease in slow cortical oscillations [128,129]. In both studies, the co-administration of antipsychotics was effective in reversing PCP-induced modifications. Precisely, haloperidol and clozapine were both able to counteract the alterations in pyramidal cell firing and the loss in cortical synchrony [128]. By the complementary application of histochemistry techniques, clozapine was also seen to prevent the PCP-mediated induction of c-fos levels in pyramidal neurons of the PFC, showing the capability to modulate the glutamatergic output and IEG expression [128]. Clozapine ability to regulate the neuronal activity in the PFC has been ascribed to its partial agonism at the serotonin 1A receptor (5-HT1AR), as highlighted by the use of mice knockout for the serotonin receptor [129]. Similarly, chlorpromazine, a first-generation antipsychotic, was effective in preventing the excitatory action of ketamine on the EEG bands of the hippocampus, with an even more pronounced effect after the addition of alpha-lipoic acid [130]. These findings were supported by a recent study, in which haloperidol, clozapine, and risperidone effectively prevented the PCP-induced desynchronization of the PFC in rats [127]. In the same study, the effect of a novel compound EGIS-11150, an inhibitor of adrenergic alpha1 and alpha2c, and serotonergic 5-HT2AR and 5-HT7R receptors, were evaluated after both PCP and ketamine administrations [127]. EGIS-11150 not only showed to fully tackle the EEG modifications induced by ketamine and PCP but also to increase the hippocampus-PFC coherence and reverse the inhibition of long-term potentiation (LTP) induced by stressors more than clozapine and haloperidol [127].
Moreover, the effects on brain network organization may be quite different between typical and atypical antipsychotics, with selective neural mechanisms underlying drugs’ therapeutic action [40,131]. In a murine model of psychosis, the PCP injection was responsible for an increase and decrease in the neuronal synchronization of the PFC and hippocampus, respectively, with an impairment in the overall circuit connectivity [40]. These alterations along with pathological behavioral phenotypes were partially reversed by atypical but not typical antipsychotics through a presumptive major activity on cortical serotonergic receptors [40,131,132,133]. These results were replicated by the application of an in vivo amperometry technique in a ketamine-induced psychotic-like model, showing the ability of clozapine but not haloperidol to restore a physiological pattern of the O2 signal in the PFC and medial ventral striatum [134]. In another study, the mid-dose of atypical antipsychotics with stronger affinity to 5-HT2A, 5-HT2C, and 5-HT1A receptors showed a positive association with increased striatal–PFC functional connectivity, whereas the affinity for the D2R was not related to any connectivity measure, as detected by fMRI application [135]. However, haloperidol, a strong D2R blocker, was effective in modulating functional connectivity between the substantia nigra and several brain regions, including the cingulate and prefrontal cortices, hippocampus, ventral pallidum, and motor cortex, as detected by fMRI [136]. These results were partially confirmed by an ISH study assessing Homer1a gene expression in 33 regions of interest (ROIs) and performing a network analysis, which showed reduced intercorrelations between brain areas, such as cingulate cortex, anterior insula, caudate-putamen, and nucleus accumbens, as well as enhanced interactivity between cortical and striatal regions, and within caudate putamen subdivisions [8]. A previous study assessed the expression of zif268, an IEG, in 83 ROIs after haloperidol administration via an ISH protocol and found increased intercorrelations between the dorsolateral striatum and thalamus, and between different subdivisions of the thalamus [9]. Taken together, these results highlight the ability of compounds with different pharmacodynamic profiles to induce specific patterns of brain activation, with a stronger effect of atypical antipsychotics compared to typical ones on PFC connectivity via the modulation of serotonergic receptors in disrupted networks.
Beyond the serotonergic modulation of hippocampal and PFC activity, further evidence pointed to the glutamatergic pathway as a key regulator of brain functional connectivity. Specifically, the increase in cortical gamma band oscillations observed after the administration of MK-801, an NMDAR antagonist, was reversed not only by haloperidol and clozapine but also by riluzole, a glutamate release inhibitor, and LY379268, a mGlu2/3R agonist [137]. In another study, the detrimental effects on brain networks due to 5-methoxy-N,N-dimethyltryptamine (5-MeO-DMT), a natural hallucinogen contained in Ayahuasca, were evaluated through a combined electrophysiology and fMRI protocol and included dysregulation of pyramidal neuron activity and reduced power of low-frequency cortical oscillations in rodent PFC, which were partially reversed by haloperidol, clozapine, risperidone, and even LY379268, a mGlu2/3R agonist [138]. Additional evidence pointed to mGlu2/3R agonists as putative regulators of brain functional connectivity. In this regard, other studies highlighted the ability of mGlu2/3R agonists, such as TASP0443294, LY379268, and LY404039, to partially increase EEG coherence and counteract the basal gamma hyperactivity detected in the frontal regions of rats treated with NMDAR antagonists [139,140,141,142,143]. To differentiate the impact of mGlu2Rs and mGlu3Rs on brain functional networks, Wood and colleagues employed Han Wistar rats, which lack mGlu2R but not mGlu3R [144]. The mGlu2/3R agonist, LY354740, showed to reverse psychotic-like behaviors induced by amphetamine and PCP administration in control Wistar but not in Han Wistar rats. In the same experimental setting, the mGlu2/3R agonist LY379268 differentially modified the network oscillatory activity, reducing the power of high-frequency gamma bands in control Wistar but not in Han Wistar rats [144]. In addition, regulators of the NMDAR have been experimented in clinical trials and may account for beneficial antipsychotic effects also by modulating brain connectivity [6]. Specifically, SSR504734, an inhibitor of the glycine transporter-1 (GlyT1) could enhance the NMDAR-mediated glutamatergic neurotransmission by increasing extracellular glycine levels and was shown to significantly reduce the ketamine-induced increase of brain metabolic activity in the PFC, cingulate cortex, hippocampus, nucleus accumbens, and anteroventral thalamic nucleus [145]. Moreover, SSR504734 reversed the increase of the absolute power of the alpha1 band in rats treated with MK-801, together with a reduction in psychotic-like behaviors [145]. Overall, these findings suggest a strong impact of mGlu2R and NMDAR modulators on brain functional connectivity with a presumptive antipsychotic effect in a translational and clinical context.
Further, antipsychotic effects were reported after the administration of VU0152100, a positive allosteric modulator of the M4 muscarinic receptor, which reverted amphetamine-mediated psychotic-like behaviors and, simultaneously, tuned regional patterns of functional connectivity in the nucleus accumbens, retrosplenial cortex, motor cortex, hippocampus, and medial thalamus [146].
For a comprehensive and systematized insight into the overall topographical modifications of brain activity induced by single antipsychotics, please see Table 2.
Table 2. Antipsychotics modulation of brain functional connectivity in a preclinical setting.
Table 2. Antipsychotics modulation of brain functional connectivity in a preclinical setting.
AuthorMethodsAntipsychoticsResults
Kargieman et al.,
2007 [128]
Electrophysiology, ISH (c-fos)Haloperidol, ClozapineHaloperidol and clozapine reversed PCP-induced alterations in pyramidal cell firing and the loss in cortical synchrony of limbic PFC. Clozapine prevented the PCP-mediated induction of c-fos levels in pyramidal neurons of the PFC.
Kargieman et al.,
2012 [129]
ElectrophysiologyClozapineClozapine counteracted the PCP-mediated reduction in slow cortical oscillations in the PFC of both wild-type and 5-HT2AR knockout mice but not of 5HT1AR mice.
Sampaio et al.,
2017 [130]
EEGChlorpromazine (+/− alpha-lipoic acid)Chlorpromazine prevented the ketamine-induced increase in the average spectral power of gamma low-bands. The combination of chlorpromazine with alpha-lipoic acid potentiated the inhibitory effects on gamma low-band oscillations.
M. Spedding et al.,
2022 [127]
EEGClozapine, Risperidone, Haloperidol, EGIS 11150EGIS 11,150 increased Hippocampal-PFC coherence and the 8–9 Hz theta band of the EEG power spectrum. Clozapine, risperidone, haloperidol, and EGIS 11,150 prevented the desynchronization effect of PCP or ketamine on the power spectra of the PFC.
C. Delgado-Sallent et al.,
2021 [40]
EEGRisperidone, Clozapine, Haloperidol, M100907, 8-OH-DPATPCP increased PFC synchronization, impaired hippocampal synchronization, and disrupted circuit connectivity. Risperidone and clozapine, but not haloperidol, reduced PCP-induced prefrontal hypersynchrony whereas none of the substances restored hippocampal and circuit desynchronization. Also, 5-HT2AR antagonism by M100907 and 5-HT1AR agonism by 8-OH-DPAT were effective in reducing prefrontal hypersynchrony.
Gener et al.,
2019 [131]
EEGRisperidone, 8-OH-DPAT, M100907, HaloperidolRisperidone, 8-OH-DPAT, M100907, and haloperidol reduced locomotor activity of mice along with suppression of neural spiking, power of theta and gamma oscillations in PFC and hippocampus, and reduction of PFC-Hippocampal theta phase synchronization.
Sebban et al.,
2002 [132]
EEGClozapine, M100907Clozapine and M100907 antagonized the decrease in power between 5 and 30 Hz but not the increase in power at 1–3 Hz in PFC caused by PCP.
Feinberg et al.,
1998 [133]
EEGHaloperidolHaloperidol potentiates the EEG slowing of MK-801 but it blocks its motor effects and its neurotoxic vacuolization.
Jennifer Li et al.,
2014 [134]
ElectrophysiologyClozapine, HaloperidolClozapine but not haloperidol attenuated the ketamine-induced increase in O2 signal and PFC-mVS coherence.
F. Tollens et al.,
2018 [135]
fMRIAmisulpride, Risperidone, OlanzapineThe mid-dose of antipsychotic compounds with stronger affinity to serotonin 5-HT2AR, 5-HT2CR, and 5-HT1AR was associated with increased PFC-striatal functional connectivity, whereas no correlation was observed for D2R.
N. Gass et al.,
2013 [136]
fMRIHaloperidolHaloperidol reduced the functional connectivity between the substantia nigra and several brain regions, including the cingulate cortex, PFC, posterodorsal hippocampus, ventral pallidum, and motor cortex.
A. Barone et al.,
2021 [8]
ISH (Homer1a)HaloperidolHaloperidol reduced intercorrelations between brain areas, such as the cingulate cortex, anterior insula, caudate-putamen, and nucleus accumbens, as well as enhanced interactivity between cortical and striatal regions, and within caudate putamen subdivisions.
A. L. Wheeler,
2014 [9]
ISH (zif268)HaloperidolHaloperidol increased intercorrelations between the dorsolateral striatum and thalamus, and between different subdivisions of the thalamus.
T. Hiyoshi et al.,
2014 [137]
EEGHaloperidol, Clozapine, Riluzole, LY379268Haloperidol, Clozapine, Riluzole, and LY379268 reversed the increase in cortical gamma band oscillations observed after the administration of MK-801.
S. Riga et al.,
2014 [138]
Electrophysiology/fMRIHaloperidol, Clozapine, Risperidone, LY379268, M100907, WAY100635The natural hallucinogen 5-MeO-DMT increased the firing rate of 51% of neurons and reduced that of 35% of pyramidal cells of the PFC as well as decreased the BOLD signal in the PFC and V1. Clozapine and haloperidol reversed the effects on the pyramidal discharge of 5-MeO-DMT. Haloperidol, Clozapine, Risperidone, LY379268, M100907, and WAY100635 antagonized the reduction in the power of low-frequency cortical oscillations induced by 5-MeO-DMT.
H. Hikichi et al.,
2015 [139]
ElectrophysiologyTASP0443294TASP0443294 inhibited methamphetamine-induced hyperlocomotion in rats and reduced the basal gamma hyperactivity in the PFC exerted by ketamine.
T. Hiyoshi et al.,
2014 [140]
Electrophysiology LY379268, JNJ16259685LY379268 and JNJ16259685 decreased the ketamine-induced basal gamma hyperactivity in the rat frontal cortex.
A. Ahnaou et al.,
2016 [141]
Electrophysiology Risperidone, Olanzapine, LY404039
Risperidone, Olanzapine, and LY404039 attenuated alterations in auditory-evoked potentials and gamma/alpha oscillations induced by amphetamine or PCP.
M. Fujáková et al.,
2014 [142]
EEGLY379268LY379268 completely prevented the ketamine-induced power increase in high-frequency bands and partially improved the decrease in EEG coherence exerted by ketamine. LY379268 also reversed the hyperlocomotion due to ketamine administration.
N. C. Jones et al.,
2012 [143]
EEGClozapine, Haloperidol, LY379268 Clozapine, Haloperidol, and LY379268 inhibited ketamine-induced hyperlocomotion, although only LY379268 prevented the ketamine-induced elevation in gamma power.
Wood et al.,
2018 [144]
EEGLY354740, LY379268LY354740 reversed psychotic-like behaviors induced by amphetamine and PCP in control Wistar but not in Han Wistar rats. LY379268 modified the network oscillatory activity, reducing the power of high-frequency gamma bands in control Wistar but not in Han Wistar rats.
Depoortère et al.,
2005 [145]
ElectrophysiologySSR504734 SSR504734 reduced the ketamine-induced increase of brain metabolic activity in the PFC, cingulate cortex, hippocampus, nucleus accumbens, and anteroventral thalamic nucleus and reversed the increase of the absolute power of the alpha1 band in rats treated with MK-801, along with a decrease in psychotic-like behaviors.
Byun et al.,
2014 [146]
phMRIVU0152100VU0152100 antagonized amphetamine-mediated psychotic-like behaviors and modulated regional patterns of functional connectivity in the nucleus accumbens, retrosplenial cortex, motor cortex, hippocampus, and medial thalamus, together with a reversion of the amphetamine-induced elevation of extracellular dopamine levels in the nucleus accumbens and caudate-putamen.
5-HTR = serotonin receptor; BOLD = blood oxygenation level dependent; D2R = dopamine D2 receptor; EEG = electroencephalogram; fMRI = functional magnetic resonance; mVS = medial ventral striatum; PCP = phencyclidine; PFC = prefrontal cortex; phMRI = pharmacological magnetic resonance; V1 = visual cortex.

8. Discussion

In the present systematic review, we attempted to dissect the neurobiological basis of antipsychotics’ synaptic effects, by highlighting the close relationship between synaptic plasticity and functional connectivity and provide a comprehensive topographical analysis of antipsychotic-induced modifications in brain activity.
Alterations at the PSD level may impair the capability of neurons to communicate with each other and may account for an overall dysconnectivity between brain regions [18]. On the other hand, modifications in brain networks involved in the regulation of human behavior and cognition may directly result in psychotic symptoms [17]. Thus, the biological chain linking disruptions in synaptic plasticity with psychopathological manifestations through an impairment in functional connectivity may be explicative of the altered processes underlying schizophrenia. In this regard, functional but not structural imaging could mirror, on the macroscale, PSD alterations localized at the molecular level and macroscopic impairments in anatomical connections should be considered mainly for their consequences on brain activity [18,28].
Moreover, a global disruption in brain networks, including both increases and decreases in functional connectivity, as detected within the DMN and salience network, may underlie psychopathological manifestations in schizophrenia patients [10,11,147,148,149]. Of interest, preclinical studies showed that NMDAR antagonists, used to reproduce psychotic-like behaviors in animals, could induce both hypo- and hypersynchrony between the same cortical regions, translationally mimicking the same features that can be tracked down in the dysregulated networks of schizophrenia neurobiology [40,129,133,137]. In fact, a stable and robust re-organization of brain networks in a pathological conformation may sustain the persistency of psychotic disorders and account for resistance to canonical antipsychotics [18]. Thus, early identification and treatment of psychotic disorders could improve clinical outcomes and responsiveness to antipsychotics probably also by preventing the consolidation of altered brain connections [150,151].
A replicated feature found in the included preclinical studies was the detrimental consequence of NMDAR antagonists, such as PCP, MK-801, and ketamine, on prefrontal and hippocampal activity [40,127,145]. Given the key role in regulating high-level cognitive functions, including learning, behavioral flexibility, novelty processing, and memory, alterations in the hippocampus–PFC pathway have a remarkable translational value [127]. In this framework, dysconnectivity within networks comprising both hippocampus and PFC was detected in patients affected by schizophrenia and proposed as a possible driver of impairments in the working memory domain [152,153]. During working memory tasks, hippocampal deactivation and coordinated dorsolateral PFC enhancement are necessary to ensure the proper abolition of interfering cognitive processes [154]. In addition, ketamine has been demonstrated to disrupt striatal–PFC connectivity [134,145], which was also unveiled in patients affected by schizophrenia and psychotic bipolar disorders [155]. Taken together, these findings highlight the feasibility of NMDAR antagonism to model some aspects of psychosis’ pathophysiology, which represents the first stage in testing the efficacy of canonical and uncanonical antipsychotic agents.
Overall, antipsychotics were found to effectively antagonize the psychotic-like phenotype induced in a preclinical setting by the administration of NMDAR antagonists, along with modifications in the functional connectivity and IEGs expression profiles [128]. In an attempt to shed light on the specific pattern of molecular changes induced by individual antipsychotics, several differences between first- and second-generation compounds were outlined, as detailed in Table 1 and Table 2. Typical antipsychotics, such as haloperidol, showed to modulate especially but not exclusively striatal activity, and this modulation, at least in part, was associated with an elevation in the IEGs expression after both acute and chronic challenges [86,93,95]. Otherwise, haloperidol was demonstrated to inconsistently modulate cortical activity by both increasing and decreasing IEGs levels [79,85]. On the other hand, atypical antipsychotics, including quetiapine, olanzapine, and clozapine, were effective in regulating the cortical expression of IEGs, specifically BDNF in the hippocampus and PFC [103,104,105,106,107,108], as well as affecting subcortical stations, including dorsal and ventral striatum [87,88]. Clozapine, the only drug approved for treatment-resistant schizophrenia [156], was found to fine-tune cortical functional connectivity, neuronal synchrony in high-gamma bands, and PFC-medial ventral striatum coherence by reverting the alterations induced by PCP and ketamine in preclinical models via modulation of serotonin receptors [40,127,129,134]. Therefore, the beneficial action of first-generation compounds on psychotic symptoms may rely on their capability to modulate subcortical-cortical networks by tackling the dopamine receptors sited at the striatal pathways, whereas clozapine and other atypical antipsychotics may mainly target functional circuits linking the hippocampus and ventral striatum with the PFC [8,9,128,129,134]. Even novel molecules with presumptive antipsychotic efficacy, such as modulators of the NMDAR, mGluR, adrenergic, serotonergic, and muscarinic receptors, were tested in a preclinical setting and showed to be able to counteract the effects of psychomimetic compounds by affecting the structure of functional networks and cortical synchrony [127,140,145,146].
Clinical studies conducted in patients affected by schizophrenia have assessed the capability of antipsychotics to modulate differently brain activity and functional networks [157,158,159,160,161]. Specifically, monotherapy with atypical antipsychotics, such as risperidone, in drug-naive first-episode psychosis patients, showed to significantly increase functional connectivity, measured by resting-state fMRI, between the cingulate cortex and the PFC, along with improvement in positive symptoms [157]. In addition, clozapine efficacy was proven to be correlated to modifications in the functional connectivity between caudate and frontal and parietal regions, without any significant changes in striatal dopamine functioning [158]. Clozapine was also found to enhance power in the alpha, delta, and theta bands in the EEG of patients diagnosed with schizophrenia, with a more pronounced effect in frontal areas [159]. In schizophrenia patients, positron emission tomography (PET) with the application of [15O]-water highlighted differences between clozapine and haloperidol in the modulation of regional cerebral blood flow [160]. Clozapine was more effective in increasing cingulate cortex and PFC blood flow, whereas haloperidol had a more significant effect on striatal regions [160]. Overall, these findings confirmed the expectations inferred from preclinical studies. These results suggest tracking down functional connectivity for putative early detection in patients with psychosis to potentially help in selecting individuals who may benefit from specific drugs that can precisely target those regions involved. Future prospective clinical trials will help define predictive models of response/resistance to antipsychotics based on brain connectivity information and combined with genetic and clinical factors [162,163,164,165,166].
For the purpose of translating the present results into clinical practice terms, some limitations must be kept in mind, including the presence of a small number of papers on the topic, the high heterogeneity among studies, and the ones intrinsic to animal models. In fact, animal models cannot fully mirror the complexity of human behaviors and only indirectly reflect psychotic phenotypes, especially for cognitive dysfunction and negative symptoms of schizophrenia [167]. The pharmacological treatments modeling psychotic-like behaviors could affect different neurotransmitter pathways and represent a further source for heterogeneity in terms of antipsychotics-induced IEGs expression, brain connectivity, and effectiveness in counteracting behavioral abnormalities [168]. In this regard, most studies included in the present review evaluated changes in fMRI and electrophysiological parameters exerted by antipsychotics in murine models of psychosis, especially but not exclusively set by the administration of NMDAR antagonists. However, antipsychotics’ effects on functional connectivity have been assessed also after amphetamine administrations as well as in animals that did not model psychotic disorders, as presented in Table 2. Although the specific drug used to mirror psychosis in the preclinical setting may affect the variability of the results, it should be remembered that patients with schizophrenia also showed alterations in different brain pathways, as highlighted by PET studies [169,170,171,172,173,174,175,176,177,178]. Thus, the application of drugs capable to reproduce changes in different neurotransmitter systems, including the dopaminergic and glutamatergic ones, is relevant in terms of translatability and clinical implication, through the opportunity to explore the effects of antipsychotics on different backgrounds and by providing the basis for tailoring the therapeutic choice according to the patient-specific characteristics. For what concerns IEGs expression, most studies explored the consequences of antipsychotic administrations in animals without psychotic-like phenotypes, as shown in Table 1. In this case, the high variability detected between studies could be mostly explained by the combination of multiple factors. Among others, different drug doses, acute or chronic administrations, and genes’ specificity were the main variables involved in determining heterogeneity [78,79,85,87,179].
Since animal models never fully reflect the disease state observed in patients with psychosis, neuroimaging application in humans remains essential to study in vivo the modifications specific to the pathology and those exerted by antipsychotics. Thus, preclinical and clinical models may provide complementary information, which is indispensable to explore psychotic disorders from different perspectives.
Further studies will be needed to precisely define the different effects of each antipsychotic to modulate interregional activity in the animal and human brain, grouping results according to different drug doses, genes considered, and administration paradigms. The application of novel techniques, such as HCR RNA-ISH, will be a further step toward the understanding of the molecular changes induced by antipsychotics and their potential impact on psychopathological manifestations, due to the unique opportunity to associate topographic and transcriptomic information [120].
Despite the limitations, the preclinical setting may offer a fertile background for testing the possible effectiveness of novel compounds with putative antipsychotic activity by directly targeting the neurobiological core of psychosis, such as alterations in synaptic plasticity and brain functional connectivity. The preclinical investigation could be instrumental to overcome some limitations of human studies, including confounding variables, the issue of undirect measures of brain activity, and the low-spatial resolution resulting from fMRI and EEG applications. The opportunity to explore the specific pattern of molecular changes induced by different compounds characterized by different pharmacodynamic profiles represents a prerequisite for creating predictive models of drug response or resistance. Lastly, preclinical studies could provide translational findings for the development of more focused therapeutic strategies, for instance, by matching information on the topography of brain activity induced by different antipsychotics with patient-specific alterations in functional connectivity profiles.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/biomedicines10123183/s1, the search strings are provided in the Supplementary Materials.

Author Contributions

Conceptualization: A.d.B.; methodology: A.d.B. and G.D.S.; writing—original draft preparation: A.d.B., A.B., A.C., M.C., G.D.S., B.M. and L.V.; writing—review and editing: A.d.B. and G.D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

A.d.B. has received unrestricted research support from Janssen (Beerse, Belgium), Lundbeck (Copenhagen, Denmark), and Otsuka (Chiyoda-ku, Tokyo, Japan) and lecture honoraria for educational meeting from Chiesi (Parma, Italy), Lundbeck (Copenhagen, Denmark), Roche (Basel, Switzerland), Sunovion (Marlborough, Massachusetts, U.S.A.), Vitria (Menlo Park, California, U.S.A.), Recordati (Milan, Italy), Angelini (Rome, Italy), and Takeda (Chuo-ku, Osaka, Japan); he has served on advisory boards for Eli Lilly (Indianapolis, Indiana, U.S.A.), Janssen (Beerse, Belgium) Lundbeck (Copenhagen, Denmark), Otsuka (Chiyoda-ku, Tokyo, Japan), Roche (Basel, Switzerland), Vitria (Menlo Park, California, USA), Chiesi (Parma, Italy), Recordati (Milan, Italy), Angelini (Rome, Italy), Takeda (Chuo-ku, Osaka, Japan). No activity is related directly or indirectly to the present manuscript content. All the other authors declare no conflict of interest.

References

  1. Bowden, C.L. Atypical antipsychotic augmentation of mood stabilizer therapy in bipolar disorder. J. Clin. Psychiatry 2005, 66, 12–19. [Google Scholar] [PubMed]
  2. Haddad, P.M.; Correll, C.U. The acute efficacy of antipsychotics in schizophrenia: A review of recent meta-analyses. Ther. Adv. Psychopharmacol. 2018, 8, 303–318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. de Greef, R.; Maloney, A. Dopamine D2 occupancy as a biomarker for antipsychotics: Quantifying the relationship with efficacy and extrapyramidal symptoms. Aaps J. 2011, 13, 121–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. de Bartolomeis, A.; Barone, A. Present and future antipsychotic drugs: A systematic review of the putative mechanisms of action for efficacy and a critical appraisal under a translational perspective. Pharmacol. Res. 2022, 176, 106078. [Google Scholar] [CrossRef] [PubMed]
  5. de Bartolomeis, A.; Vellucci, L. Clozapine’s multiple cellular mechanisms: What do we know after more than fifty years? A systematic review and critical assessment of translational mechanisms relevant for innovative strategies in treatment-resistant schizophrenia. Pharmacol. Ther. 2022, 236, 108236. [Google Scholar] [PubMed]
  6. de Bartolomeis, A.; Vellucci, L. Rational and Translational Implications of D-Amino Acids for Treatment-Resistant Schizophrenia: From Neurobiology to the Clinics. Biomolecules 2022, 12, 909. [Google Scholar] [CrossRef] [PubMed]
  7. Lieberman, J.A.; Tollefson, G.D.; Charles, C.; Zipursky, R.; Sharma, T.; Kahn, R.S.; Keefe, R.S.E.; Green, A.I.; Gur, R.E.; McEvoy, J.; et al. Antipsychotic Drug Effects on Brain Morphology in First-Episode Psychosis. Arch. Gen. Psychiatry 2005, 62, 361–370. [Google Scholar]
  8. Barone, A.; Signoriello, S.; Latte, G.; Vellucci, L.; Giordano, G.; Avagliano, C.; Buonaguro, E.F.; Marmo, F.; Tomasetti, C.; Iasevoli, F.; et al. Modulation of glutamatergic functional connectivity by a prototypical antipsychotic: Translational inference from a postsynaptic density immediate-early gene-based network analysis. Behav. Brain Res. 2021, 404, 113160. [Google Scholar] [CrossRef]
  9. Wheeler, A.L.; Creed, M.C.; Voineskos, A.N.; Nobrega, J.N. Changes in brain functional connectivity after chronic haloperidol in rats: A network analysis. Int. J. Neuropsychopharmacol. 2014, 17, 1129–1138. [Google Scholar] [CrossRef] [Green Version]
  10. Huang, H.; Chen, C.; Rong, B.; Wan, Q.; Chen, J.; Liu, Z.; Zhou, Y.; Wang, G.; Wang, H. Resting-state functional connectivity of salience network in schizophrenia and depression. Sci. Rep. 2022, 12, 11204. [Google Scholar]
  11. Zhou, Y.; Liang, M.; Tian, L.; Wang, K.; Hao, Y.; Liu, H.; Liu, Z.; Jiang, T. Functional disintegration in paranoid schizophrenia using resting-state fMRI. Schizophr. Res. 2007, 97, 194–205. [Google Scholar] [PubMed]
  12. Zhang, D.; Guo, L.; Hu, X.; Li, K.; Zhao, Q.; Liu, T. Increased cortico-subcortical functional connectivity in schizophrenia. Brain Imaging Behav. 2011, 6, 27–35. [Google Scholar] [CrossRef] [PubMed]
  13. Berdenis van Berlekom, A.; Muflihah, C.H. Synapse Pathology in Schizophrenia: A Meta-analysis of Postsynaptic Elements in Postmortem Brain Studies. Schizophr. Bull. 2020, 46, 374–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Marenco, S.; Weinberger, D.R. The neurodevelopmental hypothesis of schizophrenia: Following a trail of evidence from cradle to grave. Dev. Psychopathol. 2000, 12, 501–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Palay, S.L. Synapses in the central nervous system. J. Cell Biol. 1956, 2, 193–202. [Google Scholar] [CrossRef]
  16. Sheng, M.; Hoogenraad, C.C. The Postsynaptic Architecture of Excitatory Synapses: A More Quantitative View. Annu. Rev. Biochem. 2007, 76, 823–847. [Google Scholar] [CrossRef] [Green Version]
  17. Friston, K.J. The disconnection hypothesis. Schizophr. Res. 1998, 30, 115–125. [Google Scholar] [CrossRef]
  18. Obi-Nagata, K.; Temma, Y.; Hayashi-Takagi, A. Synaptic functions and their disruption in schizophrenia: From clinical evidence to synaptic optogenetics in an animal model. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2019, 95, 179–197. [Google Scholar]
  19. Minatohara, K.; Akiyoshi, M.; Okuno, H. Role of Immediate-Early Genes in Synaptic Plasticity and Neuronal Ensembles Underlying the Memory Trace. Front. Mol. Neurosci. 2016, 8, 78. [Google Scholar] [CrossRef] [Green Version]
  20. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. Syst. Rev. 2021, 88, 105906. [Google Scholar]
  21. Sporns, O. Network attributes for segregation and integration in the human brain. Curr. Opin. Neurobiol. 2013, 23, 162–171. [Google Scholar] [CrossRef] [PubMed]
  22. Sporns, O. Structure and function of complex brain networks. Dialogues Clin. Neurosci. 2013, 15, 247–262. [Google Scholar] [CrossRef] [PubMed]
  23. Stanley, M.L.; Moussa, M.N.; Paolini, B.M.; Lyday, R.G.; Burdette, J.H.; Laurienti, P.J. Defining nodes in complex brain networks. Front. Comput. Neurosci. 2013, 7, 169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Tsai, M.-L.; Wang, C.-C.; Lee, F.-C.; Peng, S.-J.; Chang, H.; Tseng, S.-H. Resting-State EEG Functional Connectivity in Children with Rolandic Spikes with or without Clinical Seizures. Biomedicines 2022, 10, 1553. [Google Scholar] [CrossRef] [PubMed]
  25. Skudlarski, P.; Jagannathan, K.; Anderson, K.; Stevens, M.C.; Calhoun, V.D.; Skudlarska, B.A.; Pearlson, G. Brain Connectivity Is Not Only Lower but Different in Schizophrenia: A Combined Anatomical and Functional Approach. Biol. Psychiatry 2010, 68, 61–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. White, T.P.; Wigton, R.; Joyce, D.W.; Collier, T.; Fornito, A.; Shergill, S.S. Dysfunctional Striatal Systems in Treatment-Resistant Schizophrenia. Neuropsychopharmacology 2016, 41, 1274–1285. [Google Scholar] [CrossRef]
  27. Komatsu, H.; Watanabe, E.; Fukuchi, M. Psychiatric Neural Networks and Precision Therapeutics by Machine Learning. Biomedicines 2021, 9, 403. [Google Scholar] [CrossRef]
  28. Seguin, C.; Mansour, S.L.; Sporns, O.; Zalesky, A.; Calamante, F. Network communication models narrow the gap between the modular organization of structural and functional brain networks. NeuroImage 2022, 257, 119323. [Google Scholar] [CrossRef]
  29. Bassett, D.S.; Sporns, O. Network neuroscience. Nat. Neurosci. 2017, 20, 353–364. [Google Scholar]
  30. Mochida, S. Mechanisms of Synaptic Vesicle Exo- and Endocytosis. Biomedicines 2022, 10, 1593. [Google Scholar] [CrossRef]
  31. Meldolesi, J. Post-Synapses in the Brain: Role of Dendritic and Spine Structures. Biomedicines 2022, 10, 1859. [Google Scholar] [CrossRef]
  32. Bassi, M.S.; Iezzi, E.; Gilio, L.; Centonze, D.; Buttari, F. Synaptic Plasticity Shapes Brain Connectivity: Implications for Network Topology. Int. J. Mol. Sci. 2019, 20, 6193. [Google Scholar] [CrossRef] [Green Version]
  33. Warm, D.; Schroer, J.; Sinning, A. Gabaergic Interneurons in Early Brain Development: Conducting and Orchestrated by Cortical Network Activity. Front. Mol. Neurosci. 2022, 14, 807969. [Google Scholar] [PubMed]
  34. Salkoff, D.B.; Zagha, E.; Yüzgeç, O.; McCormick, D.A. Synaptic Mechanisms of Tight Spike Synchrony at Gamma Frequency in Cerebral Cortex. J. Neurosci. 2015, 35, 10236–10251. [Google Scholar] [CrossRef] [Green Version]
  35. Bitzenhofer, S.H.; Sieben, K.; Siebert, K.D.; Spehr, M.; Hanganu-Opatz, I.L. Oscillatory Activity in Developing Prefrontal Networks Results from Theta-Gamma-Modulated Synaptic Inputs. Cell Rep. 2015, 11, 486–497. [Google Scholar] [CrossRef]
  36. Guan, A.; Wang, S.; Huang, A.; Qiu, C.; Li, Y.; Li, X.; Wang, J.; Wang, Q.; Deng, B. The role of gamma oscillations in central nervous system diseases: Mechanism and treatment. Front. Cell. Neurosci. 2022, 16, 962957. [Google Scholar] [CrossRef] [PubMed]
  37. Lu, Z.; Wang, H.; Gu, J.; Gao, F. Association between abnormal brain oscillations and cognitive performance in patients with bipolar disorder: Molecular mechanisms and clinical evidence. Synapse 2022, 76, e22247. [Google Scholar] [CrossRef]
  38. Ahnaou, A.; Huysmans, H.; Van De Casteele, T.; Drinkenburg, W.H.I.M. Cortical high gamma network oscillations and connectivity: A translational index for antipsychotics to normalize aberrant neurophysiological activity. Transl. Psychiatry 2017, 7, 1285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Hudson, M.; Rind, G.; O’Brien, T.; Jones, N.C. Reversal of evoked gamma oscillation deficits is predictive of antipsychotic activity with a unique profile for clozapine. Transl. Psychiatry 2016, 6, e784. [Google Scholar] [CrossRef]
  40. Delgado-Sallent, C.; Nebot, P.; Gener, T.; Fath, A.B.; Timplalexi, M.; Puig, M.V. Atypical, but not typical, antipsychotic drugs reduce hypersynchronized prefrontal-hippocampal circuits during psychosis-like states in mice: Contribution of 5-HT2A and 5-HT1A receptors. Cereb. Cortex 2021, 32, 3472–3487. [Google Scholar] [CrossRef]
  41. De Rosa, A.; Fontana, A.; Nuzzo, T.; Garofalo, M.; Di Maio, A.; Punzo, D.; Copetti, M.; Bertolino, A.; Errico, F.; Rampino, A.; et al. Machine Learning algorithm unveils glutamatergic alterations in the post-mortem schizophrenia brain. Schizophrenia 2022, 8, 8. [Google Scholar] [CrossRef]
  42. Barksdale, K.A.; Lahti, A.C.; Roberts, R.C. Synaptic Proteins in the Postmortem Anterior Cingulate Cortex in Schizophrenia: Relationship to Treatment and Treatment Response. Neuropsychopharmacology 2014, 39, 2095–2103. [Google Scholar] [CrossRef]
  43. Schoonover, K.E.; McCollum, L.A.; Roberts, R.C. Protein Markers of Neurotransmitter Synthesis and Release in Postmortem Schizophrenia Substantia Nigra. Neuropsychopharmacology 2017, 42, 540–550. [Google Scholar] [CrossRef] [Green Version]
  44. Chung, D.W.; Geramita, M.A.; Lewis, D.A. Synaptic Variability and Cortical Gamma Oscillation Power in Schizophrenia. Am. J. Psychiatry 2022, 179, 277–287. [Google Scholar] [CrossRef]
  45. Martin, M.V.; Mirnics, K.; Nisenbaum, L.K.; Vawter, M.P. Olanzapine Reversed Brain Gene Expression Changes Induced by Phencyclidine Treatment in Non-Human Primates. Complex Psychiatry 2015, 1, 82–93. [Google Scholar] [CrossRef]
  46. Cohen, R.S.; Blomberg, F.; Berzins, K.; Siekevitz, P. The structure of postsynaptic densities isolated from dog cerebral cortex: I. overall morphology and protein composition. J. Cell Biol. 1977, 74, 181–203. [Google Scholar] [CrossRef] [Green Version]
  47. Harris, K.M.; Weinberg, R.J. Ultrastructure of Synapses in the Mammalian Brain. Cold Spring Harb. Perspect. Biol. 2012, 4, a005587. [Google Scholar] [CrossRef] [Green Version]
  48. Zeng, M.; Chen, X.; Guan, D.; Xu, J.; Wu, H.; Tong, P.; Zhang, M. Reconstituted Postsynaptic Density as a Molecular Platform for Understanding Synapse Formation and Plasticity. Cell 2018, 174, 1172–1187.e16. [Google Scholar] [PubMed] [Green Version]
  49. Kaizuka, T.; Takumi, T. Postsynaptic density proteins and their involvement in neurodevelopmental disorders. J. Biochem. 2018, 163, 447–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. de Bartolomeis, A.; Latte, G. Glutamatergic postsynaptic density protein dysfunctions in synaptic plasticity and dendritic spines morphology: Relevance to schizophrenia and other behavioral disorders pathophysiology, and implications for novel therapeutic approaches. Mol. Neurobiol. 2014, 49, 484–511. [Google Scholar] [CrossRef] [PubMed]
  51. Keith, D.; El-Husseini, A. Excitation control: Balancing PSD-95 function at the synapse. Front. Mol. Neurosci. 2008, 1, 4. [Google Scholar] [CrossRef] [Green Version]
  52. Cammarota, M.; Bevilaqua, L.R.; Ardenghi, P.; Paratcha, G.; de Stein, M.L.; Izquierdo, I.; Medina, J.H. Learning-associated activation of nuclear MAPK, CREB and Elk-1, along with Fos production, in the rat hippocampus after a one-trial avoidance learning: Abolition by NMDA receptor blockade. Mol. Brain Res. 2000, 76, 36–46. [Google Scholar] [CrossRef]
  53. Zheng, F.; Wang, H. NMDA-mediated and self-induced bdnf exon IV transcriptions are differentially regulated in cultured cortical neurons. Neurochem. Int. 2009, 54, 385–392. [Google Scholar] [CrossRef] [Green Version]
  54. Wang, H.; Xu, J.; Lazarovici, P.; Quirion, R.; Zheng, W. cAMP Response Element-Binding Protein (CREB): A Possible Signaling Molecule Link in the Pathophysiology of Schizophrenia. Front. Mol. Neurosci. 2018, 11, 255. [Google Scholar] [CrossRef] [Green Version]
  55. Zhang, J.; Xu, T.X. PSD-95 uncouples dopamine-glutamate interaction in the D1/PSD-95/NMDA receptor complex. J. Neurosci. 2009, 29, 2948–2960. [Google Scholar] [CrossRef] [Green Version]
  56. Zhang, G.; Stackman, R.W., Jr. The role of serotonin 5-HT2A receptors in memory and cognition. Front. Pharmacol. 2015, 6, 225. [Google Scholar] [CrossRef] [Green Version]
  57. Hu, L.A.; Tang, Y. Beta 1-adrenergic receptor association with PSD-Inhibition of receptor internalization and facilitation of beta 1-adrenergic receptor interaction with N-methyl-D-aspartate receptors. J. Biol. Chem. 2000, 275, 38659–38666. [Google Scholar] [CrossRef] [Green Version]
  58. Giessel, A.J.; Sabatini, B.L. M1 Muscarinic Receptors Boost Synaptic Potentials and Calcium Influx in Dendritic Spines by Inhibiting Postsynaptic SK Channels. Neuron 2010, 68, 936–947. [Google Scholar] [CrossRef] [Green Version]
  59. Bristot, G.; De Bastiani, M.A.; Pfaffenseller, B.; Kapczinski, F.; Kauer-Sant’anna, M. Gene Regulatory Network of Dorsolateral Prefrontal Cortex: A Master Regulator Analysis of Major Psychiatric Disorders. Mol. Neurobiol. 2020, 57, 1305–1316. [Google Scholar] [CrossRef]
  60. Guan, F.; Ni, T.; Zhu, W.; Williams, L.K.; Cui, L.-B.; Li, M.; Tubbs, J.; Sham, P.-C.; Gui, H. Integrative omics of schizophrenia: From genetic determinants to clinical classification and risk prediction. Mol. Psychiatry 2021, 27, 113–126. [Google Scholar] [CrossRef]
  61. Heavner, W.E.; Lautz, J.D.; Speed, H.E.; Gniffke, E.P.; Immendorf, K.B.; Welsh, J.P.; Baertsch, N.A.; Smith, S.E.P. Remodeling of the Homer-Shank interactome mediates homeostatic plasticity. Sci. Signal. 2021, 14, eabd7325. [Google Scholar] [CrossRef]
  62. Matosin, N.; Fernandez-Enright, F. Molecular evidence of synaptic pathology in the CA1 region in schizophrenia. NPJ Schizophr. 2016, 2, 16022. [Google Scholar] [CrossRef] [Green Version]
  63. Funk, A.J.; Mielnik, C.A.; Koene, R.; Newburn, E.; Ramsey, A.J.; Lipska, B.K.; McCullumsmith, R.E. Postsynaptic Density-95 Isoform Abnormalities in Schizophrenia. Schizophr. Bull. 2017, 43, 891–899. [Google Scholar] [CrossRef]
  64. Chung, D.W.; Fish, K.N.; Lewis, D.A. Pathological Basis for Deficient Excitatory Drive to Cortical Parvalbumin Interneurons in Schizophrenia. Am. J. Psychiatry 2016, 173, 1131–1139. [Google Scholar] [CrossRef] [Green Version]
  65. Huang, X.-F.; Song, X. Effects of antipsychotic drugs on neurites relevant to schizophrenia treatment. Med. Res. Rev. 2019, 39, 386–403. [Google Scholar] [CrossRef]
  66. de Bartolomeis, A.; Barone, A. The Homer1 family of proteins at the crossroad of dopamine-glutamate signaling: An emerging molecular “Lego” in the pathophysiology of psychiatric disorders. A systematic review and translational insight. Neurosci. Biobehav. Rev. 2022, 136, 104596. [Google Scholar] [CrossRef]
  67. de Bartolomeis, A.; Barone, A. Linking Inflammation, Aberrant Glutamate-Dopamine Interaction, and Post-synaptic Changes: Translational Relevance for Schizophrenia and Antipsychotic Treatment: A Systematic Review. Mol. Neurobiol. 2022, 59, 6460–6501. [Google Scholar] [CrossRef]
  68. Swathy, B.; Banerjee, M. Understanding epigenetics of schizophrenia in the backdrop of its antipsychotic drug therapy. Epigenomics 2017, 9, 721–736. [Google Scholar] [CrossRef] [Green Version]
  69. Ovenden, E.S.; McGregor, N.W.; Emsley, R.A.; Warnich, L. DNA methylation and antipsychotic treatment mechanisms in schizophrenia: Progress and future directions. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2018, 81, 38–49. [Google Scholar] [CrossRef]
  70. Takaki, M.; Kodama, M.; Mizuki, Y.; Kawai, H.; Yoshimura, B.; Kishimoto, M.; Sakamoto, S.; Okahisa, Y.; Yamada, N. Effects of the antipsychotics haloperidol, clozapine, and aripiprazole on the dendritic spine. Eur. Neuropsychopharmacol. 2018, 28, 610–619. [Google Scholar] [CrossRef]
  71. Meshul, C.K.; Janowsky, A.; Casey, D.E.; Stallbaumer, R.K.; Taylor, B. Coadministration of haloperidol and SCH-23390 prevents the increase in “perforated”synapses due to either drug alone. Neuropsychopharmacology 1992, 7, 285–293. [Google Scholar]
  72. Vincent, S.L.; McSparren, J.; Wang, R.Y.; Benes, F.M. Evidence for ultrastructural changes in cortical axodendritic synapses following long-term treatment with haloperidol or clozapine. Neuropsychopharmacology 1991, 5, 147–155. [Google Scholar]
  73. Park, S.W.; Lee, C.H.; Cho, H.Y.; Seo, M.K.; Lee, J.G.; Lee, B.J.; Seol, W.; Kee, B.S.; Kim, Y.H. Effects of antipsychotic drugs on the expression of synaptic proteins and dendritic outgrowth in hippocampal neuronal cultures. Synapse 2013, 67, 224–234. [Google Scholar] [CrossRef]
  74. Zhang, Q.; Yu, Y.; Huang, X.-F. Olanzapine Prevents the PCP-induced Reduction in the Neurite Outgrowth of Prefrontal Cortical Neurons via NRG1. Sci. Rep. 2016, 6, 19581. [Google Scholar] [CrossRef] [Green Version]
  75. Zheng, P.; Hu, M.; Xie, Y.; Yu, Y.; Jaaro-Peled, H.; Huang, X.-F. Aripiprazole and haloperidol protect neurite lesions via reducing excessive D2R-DISC1 complex formation. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2019, 92, 59–69. [Google Scholar] [CrossRef]
  76. Bellon, A.; Feuillet, V.; Cortez-Resendiz, A.; Mouaffak, F.; Kong, L.; Hong, L.E.; De Godoy, L.; Jay, T.M.; Hosmalin, A.; Krebs, M.-O. Dopamine-induced pruning in monocyte-derived-neuronal-like cells (MDNCs) from patients with schizophrenia. Mol. Psychiatry 2022, 27, 2787–2802. [Google Scholar] [CrossRef]
  77. Engmann, O.; Giralt, A.; Girault, J.-A. Acute drug-induced spine changes in the nucleus accumbens are dependent on β-adducin. Neuropharmacology 2016, 110, 333–342. [Google Scholar] [CrossRef]
  78. Buonaguro, E.F.; Tomasetti, C.; Chiodini, P.; Marmo, F.; Latte, G.; Rossi, R.; Avvisati, L.; Iasevoli, F.; de Bartolomeis, A. Postsynaptic density protein transcripts are differentially modulated by minocycline alone or in add-on to haloperidol: Implications for treatment resistant schizophrenia. J. Psychopharmacol. 2017, 31, 406–417. [Google Scholar] [CrossRef]
  79. de Bartolomeis, A.; Iasevoli, F. Progressive recruitment of cortical and striatal regions by inducible postsynaptic density transcripts after increasing doses of antipsychotics with different receptor profiles: Insights for psychosis treatment. Eur. Neuropsychopharmacol. 2015, 25, 566–582. [Google Scholar] [CrossRef]
  80. Kim, S.; Chen, J.; Cheng, T.; Gindulyte, A.; He, J.; He, S.; Li, Q.; Shoemaker, B.A.; Thiessen, P.A.; Yu, B.; et al. PubChem in 2021: New data content and improved web interfaces. Nucleic Acids Res. 2021, 49, D1388–D1395. [Google Scholar] [CrossRef]
  81. Tomasetti, C.; Dell’Aversano, C.; Iasevoli, F.; Marmo, F.; de Bartolomeis, A. The acute and chronic effects of combined antipsychotic–mood stabilizing treatment on the expression of cortical and striatal postsynaptic density genes. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2011, 35, 184–197. [Google Scholar] [CrossRef] [PubMed]
  82. Solmi, M.; Murru, A.; Pacchiarotti, I.; Undurraga, J.; Veronese, N.; Fornaro, M.; Stubbs, B.; Monaco, F.; Vieta, E.; Seeman, M.V.; et al. Safety, tolerability, and risks associated with first- and second-generation antipsychotics: A state-of-the-art clinical review. Ther. Clin. Risk Manag. 2017, 13, 757–777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Beaudry, G.; Langlois, M.-C.; Weppe, I.; Rouillard, C.; Lévesque, D. Contrasting Patterns and Cellular Specificity of Transcriptional Regulation of the Nuclear Receptor Nerve Growth Factor-Inducible B by Haloperidol and Clozapine in the Rat Forebrain. J. Neurochem. 2002, 75, 1694–1702. [Google Scholar] [CrossRef] [Green Version]
  84. Cochran, S.; McKerchar, C.; Morris, B.; Pratt, J. Induction of differential patterns of local cerebral glucose metabolism and immediate-early genes by acute clozapine and haloperidol. Neuropharmacology 2002, 43, 394–407. [Google Scholar] [CrossRef]
  85. de Bartolomeis, A.; Marmo, F. Imaging brain gene expression profiles by antipsychotics: Region-specific action of amisulpride on postsynaptic density transcripts compared to haloperidol. Eur. Neuropsychopharmacol. 2013, 23, 1516–1529. [Google Scholar] [CrossRef]
  86. Iasevoli, F.; Ambesi-Impiombato, A.; Fiore, G.; Panariello, F.; Muscettola, G.; de Bartolomeis, A. Pattern of acute induction of Homer1a gene is preserved after chronic treatment with first- and second-generation antipsychotics: Effect of short-term drug discontinuation and comparison with Homer1a-interacting genes. J. Psychopharmacol. 2010, 25, 875–887. [Google Scholar] [CrossRef]
  87. Polese, D.; de Serpis, A.A. Homer 1a Gene Expression Modulation by Antipsychotic Drugs: Involvement of the Glutamate Metabotropic System and Effects of D-Cycloserine. Neuropsychopharmacology 2002, 27, 906–913. [Google Scholar] [CrossRef] [Green Version]
  88. Robbins, M.; Critchlow, H.; Lloyd, A.; Cilia, J.; Clarke, J.; Bond, B.; Jones, D.; Maycox, P. Differential expression of IEG mRNA in rat brain following acute treatment with clozapine or haloperidol: A semi-quantitative RT—PCR study. J. Psychopharmacol. 2008, 22, 536–542. [Google Scholar] [CrossRef]
  89. MacGibbon, G.; Lawlor, P.; Bravo, R.; Dragunow, M. Clozapine and haloperidol produce a differential pattern of immediate early gene expression in rat caudate-putamen, nucleus accumbens, lateral septum and islands of Calleja. Mol. Brain Res. 1994, 23, 21–32. [Google Scholar] [CrossRef] [PubMed]
  90. Nguyen, T.V.; Kosofsky, B.E.; Birnbaum, R.; Cohen, B.M.; Hyman, S.E. Differential expression of c-fos and zif268 in rat striatum after haloperidol, clozapine, and amphetamine. Proc. Natl. Acad. Sci. USA 1992, 89, 4270–4274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Werme, M.; Ringholm, A.; Olson, L.; Brené, S. Differential patterns of induction of NGFI-B, Nor1 and c-fos mRNAs in striatal subregions by haloperidol and clozapine. Brain Res. 2000, 863, 112–119. [Google Scholar] [CrossRef] [PubMed]
  92. Dell’Aversano, C.; Tomasetti, C.; Iasevoli, F.; de Bartolomeis, A. Antipsychotic and antidepressant co-treatment: Effects on transcripts of inducible postsynaptic density genes possibly implicated in behavioural disorders. Brain Res. Bull. 2009, 79, 123–129. [Google Scholar] [CrossRef] [PubMed]
  93. Buonaguro, E.F.; Iasevoli, F.; Marmo, F.; Eramo, A.; Latte, G.; Avagliano, C.; Tomasetti, C.; de Bartolomeis, A. Re-arrangements of gene transcripts at glutamatergic synapses after prolonged treatments with antipsychotics: A putative link with synaptic remodeling. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2017, 76, 29–41. [Google Scholar] [CrossRef] [PubMed]
  94. de Bartolomeis, A.; Marmo, F. Switching antipsychotics: Imaging the differential effect on the topography of postsynaptic density transcripts in antipsychotic-naïve vs. antipsychotic-exposed rats. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2016, 70, 24–38. [Google Scholar] [CrossRef] [PubMed]
  95. Iasevoli, F.; Fiore, G.; Cicale, M.; Muscettola, G.; de Bartolomeis, A. Haloperidol induces higher Homer1a expression than risperidone, olanzapine and sulpiride in striatal sub-regions. Psychiatry Res. 2010, 177, 255–260. [Google Scholar] [CrossRef]
  96. Seo, M.K.; Lee, C.H.; Cho, H.Y.; You, Y.S.; Lee, B.J.; Lee, J.G.; Park, S.W.; Kim, Y.H. Effects of antipsychotic drugs on the expression of synapse-associated proteins in the frontal cortex of rats subjected to immobilization stress. Psychiatry Res. 2015, 229, 968–974. [Google Scholar] [CrossRef]
  97. Langlois, M.-C.; Beaudry, G.; Zekki, H.; Rouillard, C.; Lévesque, D. Impact of antipsychotic drug administration on the expression of nuclear receptors in the neocortex and striatum of the rat brain. Neuroscience 2001, 106, 117–128. [Google Scholar] [CrossRef]
  98. Lecrubier, Y. Is amisulpride an ‘atypical’ atypical antipsychotic agent? Int. Clin. Psychopharmacol. 2000, 15, S21–S26. [Google Scholar]
  99. Pinna, A.; Costa, G.; Contu, L.; Morelli, M. Fos expression induced by olanzapine and risperidone in the central extended amygdala. Eur. J. Pharmacol. 2019, 865, 172764. [Google Scholar] [CrossRef]
  100. Robinet, E.A.; Geurts, M.; Maloteaux, J.-M.; Pauwels, P.J. Chronic treatment with certain antipsychotic drugs preserves upregulation of regulator of G-protein signalling 2 mRNA in rat striatum as opposed to c-fos mRNA. Neurosci. Lett. 2001, 307, 45–48. [Google Scholar] [CrossRef]
  101. Luoni, A.; Rocha, F.F.; Riva, M.A. Anatomical specificity in the modulation of activity-regulated genes after acute or chronic lurasidone treatment. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2014, 50, 94–101. [Google Scholar] [CrossRef] [PubMed]
  102. Fumagalli, F.; Calabrese, F.; Luoni, A.; Bolis, F.; Racagni, G.; Riva, M.A. Modulation of BDNF expression by repeated treatment with the novel antipsychotic lurasidone under basal condition and in response to acute stress. Int. J. Neuropsychopharmacol. 2011, 15, 235–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Rizig, M.; McQuillin, A.; Ng, A.; Robinson, M.; Harrison, A.; Zvelebil, M.; Hunt, S.P.; Gurling, H.M. A gene expression and systems pathway analysis of the effects of clozapine compared to haloperidol in the mouse brain implicates susceptibility genes for schizophrenia. J. Psychopharmacol. 2012, 26, 1218–1230. [Google Scholar] [CrossRef] [PubMed]
  104. Bai, O.; Chlan-Fourney, J.; Bowen, R.; Keegan, D.; Li, X.-M. Expression of brain-derived neurotrophic factor mRNA in rat hippocampus after treatment with antipsychotic drugs. J. Neurosci. Res. 2002, 71, 127–131. [Google Scholar] [CrossRef] [PubMed]
  105. Kontkanen, O.; Lakso, M.; Wong, G.; Castr, E. Chronic Antipsychotic Drug Treatment Induces Long-lasting Expression of fos and jun Family Genes and Activator Protein 1 Complex in the Rat Prefrontal Cortex. Neuropsychopharmacology 2002, 27, 152–162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Verma, V.; Lim, E.P.; Han, S.P.; Nagarajah, R.; Dawe, G.S. Chronic high-dose haloperidol has qualitatively similar effects to risperidone and clozapine on immediate-early gene and tyrosine hydroxylase expression in the rat locus coeruleus but not medial prefrontal cortex. Neurosci. Res. 2007, 57, 17–28. [Google Scholar] [CrossRef]
  107. Park, S.-W.; Lee, S.-K.; Kim, J.-M.; Yoon, J.-S.; Kim, Y.-H. Effects of quetiapine on the brain-derived neurotrophic factor expression in the hippocampus and neocortex of rats. Neurosci. Lett. 2006, 402, 25–29. [Google Scholar] [CrossRef] [PubMed]
  108. Kim, H.-W.; Cheon, Y.; Modi, H.R.; Rapoport, S.I.; Rao, J.S. Effects of chronic clozapine administration on markers of arachidonic acid cascade and synaptic integrity in rat brain. Psychopharmacology 2012, 222, 663–674. [Google Scholar] [CrossRef]
  109. Fumagalli, F.; Frasca, A.; Racagni, G.; Riva, M.A. Antipsychotic drugs modulate Arc expression in the rat brain. Eur. Neuropsychopharmacol. 2009, 19, 109–115. [Google Scholar] [CrossRef]
  110. Luoni, A.; Fumagalli, F.; Racagni, G.; Riva, M.A. Repeated aripiprazole treatment regulates Bdnf, Arc and Npas4 expression under basal condition as well as after an acute swim stress in the rat brain. Pharmacol. Res. 2014, 80, 1–8. [Google Scholar] [CrossRef] [PubMed]
  111. Begni, V.; Sanson, A.; Luoni, A.; Sensini, F.; Grayson, B.; Munni, S.; Neill, J.; Riva, M. Towards Novel Treatments for Schizophrenia: Molecular and Behavioural Signatures of the Psychotropic Agent SEP-363856. Int. J. Mol. Sci. 2021, 22, 4119. [Google Scholar] [CrossRef] [PubMed]
  112. Samaha, A.-N.; Reckless, G.E.; Seeman, P.; Diwan, M.; Nobrega, J.N.; Kapur, S. Less Is More: Antipsychotic Drug Effects Are Greater with Transient Rather Than Continuous Delivery. Biol. Psychiatry 2008, 64, 145–152. [Google Scholar] [CrossRef]
  113. Parikh, V.; Khan, M.M.; Mahadik, S.P. Olanzapine counteracts reduction of brain-derived neurotrophic factor and TrkB receptors in rat hippocampus produced by haloperidol. Neurosci. Lett. 2004, 356, 135–139. [Google Scholar] [CrossRef] [PubMed]
  114. Gianlorenco, A.C.L.; de Melo, P.S. Electroencephalographic Patterns in taVNS: A Systematic Review. Biomedicines 2022, 10, 2208. [Google Scholar] [CrossRef]
  115. Bakhshayesh, H.; Fitzgibbon, S.; Janani, A.S.; Grummett, T.S.; Pope, K. Detecting synchrony in EEG: A comparative study of functional connectivity measures. Comput. Biol. Med. 2018, 105, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Kim, J.; Kim, Y.-K. Crosstalk between Depression and Dementia with Resting-State fMRI Studies and Its Relationship with Cognitive Functioning. Biomedicines 2021, 9, 82. [Google Scholar] [CrossRef]
  117. Jueptner, M.; Weiller, C. Review: Does Measurement of Regional Cerebral Blood Flow Reflect Synaptic Activity?—Implications for PET and fMRI. NeuroImage 1995, 2, 148–156. [Google Scholar] [CrossRef] [PubMed]
  118. Mandino, F.; Cerri, D.H.; Garin, C.M.; Straathof, M.; Van Tilborg, G.A.F.; Chakravarty, M.M.; Dhenain, M.; Dijkhuizen, R.M.; Gozzi, A.; Hess, A.; et al. Animal Functional Magnetic Resonance Imaging: Trends and Path Toward Standardization. Front. Neuroinform. 2020, 13, 78. [Google Scholar] [CrossRef]
  119. Repovš, G. Dealing with Noise in EEG Recording and Data Analysis. Inform. Med. Slov. 2010, 15, 18–25. [Google Scholar]
  120. Moses, L.; Pachter, L. Museum of spatial transcriptomics. Nat. Methods 2022, 19, 534–546. [Google Scholar] [CrossRef]
  121. Adell, A. Brain NMDA Receptors in Schizophrenia and Depression. Biomolecules 2020, 10, 947. [Google Scholar] [CrossRef] [PubMed]
  122. Littlejohn, G.; Guymer, E. Modulation of NMDA Receptor Activity in Fibromyalgia. Biomedicines 2017, 5, 15. [Google Scholar] [CrossRef] [Green Version]
  123. Paasonen, J.; Salo, R.A.; Ihalainen, J.; Leikas, J.V.; Savolainen, K.; Lehtonen, M.; Forsberg, M.M.; Gröhn, O. Dose-response effect of acute phencyclidine on functional connectivity and dopamine levels, and their association with schizophrenia-like symptom classes in rat. Neuropharmacology 2017, 119, 15–25. [Google Scholar] [CrossRef] [PubMed]
  124. Park, A.J.; Harris, A.Z.; Martyniuk, K.M.; Chang, C.-Y.; Abbas, A.I.; Lowes, D.C.; Kellendonk, C.; Gogos, J.A.; Gordon, J.A. Reset of hippocampal–prefrontal circuitry facilitates learning. Nature 2021, 591, 615–619. [Google Scholar] [CrossRef]
  125. Bähner, F.; Meyer-Lindenberg, A. Hippocampal–prefrontal connectivity as a translational phenotype for schizophrenia. Eur. Neuropsychopharmacol. 2017, 27, 93–106. [Google Scholar] [CrossRef] [PubMed]
  126. Tamminga, C.A. The neurobiology of cognition in schizophrenia. J. Clin. Psychiatry 2006, 67, e11. [Google Scholar] [CrossRef]
  127. Spedding, M.; Sebban, C.; Jay, T.M.; Rocher, C.; Tesolin-Decros, B.; Chazot, P.; Schenker, E.; Szénási, G.; Lévay, G.I.; Megyeri, K.; et al. Phenotypical Screening on Neuronal Plasticity in Hippocampal-Prefrontal Cortex Connectivity Reveals an Antipsychotic with a Novel Profile. Cells 2022, 11, 1181. [Google Scholar] [CrossRef] [PubMed]
  128. Kargieman, L.; Santana, N.; Mengod, G.; Celada, P.; Artigas, F. Antipsychotic drugs reverse the disruption in prefrontal cortex function produced by NMDA receptor blockade with phencyclidine. Proc. Natl. Acad. Sci. USA 2007, 104, 14843–14848. [Google Scholar] [CrossRef] [Green Version]
  129. Kargieman, L.; Riga, M.; Artigas, F.; Celada, P. Clozapine Reverses Phencyclidine-Induced Desynchronization of Prefrontal Cortex through a 5-HT1A Receptor-Dependent Mechanism. Neuropsychopharmacology 2011, 37, 723–733. [Google Scholar] [CrossRef]
  130. Sampaio, L.R.L.; Borges, L.T.N.; Barbosa, T.M.; Matos, N.C.B.; Lima, R.D.F.; de Oliveira, M.N.; Gularte, V.N.; Patrocínio, M.C.A.; Macêdo, D.; Vale, O.C.D.; et al. Electroencephalographic study of chlorpromazine alone or combined with alpha-lipoic acid in a model of schizophrenia induced by ketamine in rats. J. Psychiatr. Res. 2017, 86, 73–82. [Google Scholar] [CrossRef]
  131. Gener, T.; Campo, A.T.; Alemany-González, M.; Nebot, P.; Delgado-Sallent, C.; Chanovas, J.; Puig, M.V. Serotonin 5-HT1A, 5-HT2A and dopamine D2 receptors strongly influence prefronto-hippocampal neural networks in alert mice: Contribution to the actions of risperidone. Neuropharmacology 2019, 158, 107743. [Google Scholar] [CrossRef] [PubMed]
  132. Sebban, C.; Tesolin-Decros, B.; Ciprian-Ollivier, J.; Perret, L.; Spedding, M. Effects of phencyclidine (PCP) and MK 801 on the EEGq in the prefrontal cortex of conscious rats; antagonism by clozapine, and antagonists of AMPA-, α1- and 5-HT2A -receptors. J. Cereb. Blood Flow Metab. 2002, 135, 65–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Feinberg, I.; Campbell, I.G. Haloperidol potentiates the EEG slowing of MK-801 despite blocking its motor effects: Implications for the PCP model of schizophrenia. Neuroreport 1998, 9, 2189–2193. [Google Scholar] [CrossRef] [PubMed]
  134. Li, J.; Ishiwari, K.; Conway, M.W.; Francois, J.; Huxter, J.; Lowry, J.P.; Schwarz, A.J.; Tricklebank, M.; Gilmour, G. Dissociable Effects of Antipsychotics on Ketamine-Induced Changes in Regional Oxygenation and Inter-Regional Coherence of Low Frequency Oxygen Fluctuations in the Rat. Neuropsychopharmacology 2014, 39, 1635–1644. [Google Scholar] [CrossRef] [Green Version]
  135. Tollens, F.; Gass, N.; Becker, R.; Schwarz, A.; Risterucci, C.; Künnecke, B.; Lebhardt, P.; Reinwald, J.; Sack, M.; Weber-Fahr, W.; et al. The affinity of antipsychotic drugs to dopamine and serotonin 5-HT2 receptors determines their effects on prefrontal-striatal functional connectivity. Eur. Neuropsychopharmacol. 2018, 28, 1035–1046. [Google Scholar] [CrossRef]
  136. Gass, N.; Schwarz, A.J.; Sartorius, A.; Cleppien, D.; Zheng, L.; Schenker, E.; Risterucci, C.; Meyer-Lindenberg, A.; Weber-Fahr, W. Haloperidol modulates midbrain-prefrontal functional connectivity in the rat brain. Eur. Neuropsychopharmacol. 2013, 23, 1310–1319. [Google Scholar] [CrossRef]
  137. Hiyoshi, T.; Kambe, D.; Karasawa, J.; Chaki, S. Involvement of glutamatergic and GABAergic transmission in MK-801-increased gamma band oscillation power in rat cortical electroencephalograms. Neuroscience 2014, 280, 262–274. [Google Scholar] [CrossRef]
  138. Riga, M.S.; Soria, G.; Tudela, R.; Artigas, F.; Celada, P. The natural hallucinogen 5-MeO-DMT, component of Ayahuasca, disrupts cortical function in rats: Reversal by antipsychotic drugs. Int. J. Neuropsychopharmacol. 2014, 17, 1269–1282. [Google Scholar] [CrossRef] [Green Version]
  139. Hikichi, H.; Hiyoshi, T.; Marumo, T.; Tomishima, Y.; Kaku, A.; Iida, I.; Urabe, H.; Tamita, T.; Yasuhara, A.; Karasawa, J.-I.; et al. Antipsychotic profiles of TASP0443294, a novel and orally active positive allosteric modulator of metabotropic glutamate 2 receptor. J. Pharmacol. Sci. 2015, 127, 352–361. [Google Scholar] [CrossRef]
  140. Hiyoshi, T.; Hikichi, H.; Karasawa, J.-I.; Chaki, S. Metabotropic glutamate receptors regulate cortical gamma hyperactivities elicited by ketamine in rats. Neurosci. Lett. 2014, 567, 30–34. [Google Scholar] [CrossRef] [PubMed]
  141. Ahnaou, A.; Biermans, R.; Drinkenburg, W.H. Modulation of mGlu2 Receptors, but Not PDE10A Inhibition Normalizes Pharmacologically-Induced Deviance in Auditory Evoked Potentials and Oscillations in Conscious Rats. PLoS ONE 2016, 11, e0147365. [Google Scholar] [CrossRef]
  142. Fujáková, M.; Páleníček, T.; Brunovský, M.; Gorman, I.; Tylš, F.; Kubešová, A.; Řípová, D.; Krajča, V.; Horáček, J. The effect of ((−)-2-oxa-4-aminobicyclo[3.1.0]hexane-2,6-dicarboxylic acid (LY379268), an mGlu2/3 receptor agonist, on EEG power spectra and coherence in ketamine model of psychosis. Pharmacol. Biochem. Behav. 2014, 122, 212–221. [Google Scholar] [CrossRef] [PubMed]
  143. Jones, N.C.; Reddy, M. Acute administration of typical and atypical antipsychotics reduces EEG γ power, but only the preclinical compound LY379268 reduces the ketamine-induced rise in γ power. Int. J. Neuropsychopharmacol. 2012, 15, 657–668. [Google Scholar] [CrossRef] [Green Version]
  144. Wood, C.M.; Wafford, K.A.; McCarthy, A.P.; Hewes, N.; Shanks, E.; Lodge, D.; Robinson, E. Investigating the role of mGluR2 versus mGluR3 in antipsychotic-like effects, sleep-wake architecture and network oscillatory activity using novel Han Wistar rats lacking mGluR2 expression. Neuropharmacology 2018, 140, 246–259. [Google Scholar] [CrossRef] [PubMed]
  145. Depoortère, R.; Dargazanli, G.; Estenne-Bouhtou, G.; Coste, A.; Lanneau, C.; Desvignes, C.; Poncelet, M.; Héaulme, M.; Santucci, V.; Decobert, M.; et al. Neurochemical, Electrophysiological and Pharmacological Profiles of the Selective Inhibitor of the Glycine Transporter-1 SSR504734, a Potential New Type of Antipsychotic. Neuropsychopharmacology 2005, 30, 1963–1985. [Google Scholar] [CrossRef] [PubMed]
  146. Byun, N.E.; Grannan, M.; Bubser, M.; Barry, R.L.; Thompson, A.; Rosanelli, J.; Gowrishankar, R.; Kelm, N.D.; Damon, S.; Bridges, T.M.; et al. Antipsychotic Drug-Like Effects of the Selective M4 Muscarinic Acetylcholine Receptor Positive Allosteric Modulator VU0152100. Neuropsychopharmacology 2014, 39, 1578–1593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Wang, H.; Guo, W.; Liu, F.; Wang, G.; Lyu, H.; Wu, R.; Chen, J.; Wang, S.; Li, L.; Zhao, J. Patients with first-episode, drug-naive schizophrenia and subjects at ultra-high risk of psychosis shared increased cerebellar-default mode network connectivity at rest. Sci. Rep. 2016, 6, 26124. [Google Scholar] [CrossRef] [Green Version]
  148. Jang, J.H.; Jung, W.H.; Choi, J.-S.; Choi, C.-H.; Kang, D.-H.; Shin, N.Y.; Hong, K.S.; Kwon, J.S. Reduced prefrontal functional connectivity in the default mode network is related to greater psychopathology in subjects with high genetic loading for schizophrenia. Schizophr. Res. 2011, 127, 58–65. [Google Scholar] [CrossRef]
  149. Bluhm, R.L.; Miller, J.; Lanius, R.A.; Osuch, E.A.; Boksman, K.; Neufeld, R.W.J.; Theberge, J.; Schaefer, B.; Williamson, P. Spontaneous Low-Frequency Fluctuations in the BOLD Signal in Schizophrenic Patients: Anomalies in the Default Network. Schizophr. Bull. 2007, 33, 1004–1012. [Google Scholar] [CrossRef]
  150. Yada, Y.; Yoshimura, B.; Kishi, Y. Correlation between delay in initiating clozapine and symptomatic improvement. Schizophr. Res. 2015, 168, 585–586. [Google Scholar] [CrossRef]
  151. Carbon, M.; Correll, C.U. Clinical predictors of therapeutic response to antipsychotics in schizophrenia. Dialogues Clin. Neurosci. 2014, 16, 505–524. [Google Scholar] [CrossRef] [PubMed]
  152. Meyer-Lindenberg, A.; Poline, J.-B.; Kohn, P.D.; Holt, J.L.; Egan, M.F.; Weinberger, D.R.; Berman, K.F. Evidence for Abnormal Cortical Functional Connectivity During Working Memory in Schizophrenia. Am. J. Psychiatry 2001, 158, 1809–1817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Benetti, S.; Mechelli, A.; Picchioni, M.; Broome, M.; Williams, S.; McGuire, P. Functional integration between the posterior hippocampus and prefrontal cortex is impaired in both first episode schizophrenia and the at risk mental state. Brain 2009, 132, 2426–2436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Meyer-Lindenberg, A.S.; Olsen, R.K.; Kohn, P.D.; Brown, T.; Egan, M.F.; Weinberger, D.R.; Berman, K.F. Regionally Specific Disturbance of Dorsolateral Prefrontal–Hippocampal Functional Connectivity in Schizophrenia. Arch. Gen. Psychiatry 2005, 62, 379–386. [Google Scholar] [CrossRef] [Green Version]
  155. Karcher, N.R.; Rogers, B.P.; Woodward, N.D. Functional Connectivity of the Striatum in Schizophrenia and Psychotic Bipolar Disorder. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2019, 4, 956–965. [Google Scholar] [CrossRef]
  156. Correll, C.U.; Agid, O.; Crespo-Facorro, B.; de Bartolomeis, A.; Fagiolini, A.; Seppälä, N.; Howes, O.D. A Guideline and Checklist for Initiating and Managing Clozapine Treatment in Patients with Treatment-Resistant Schizophrenia. CNS Drugs 2022, 36, 659–679. [Google Scholar] [CrossRef]
  157. Zong, X.; Hu, M.; Pantazatos, S.P.; Mann, J.J.; Wang, G.; Liao, Y.; Liu, Z.-C.; Liao, W.; Yao, T.; Li, Z.; et al. A Dissociation in Effects of Risperidone Monotherapy on Functional and Anatomical Connectivity within the Default Mode Network. Schizophr. Bull. 2018, 45, 1309–1318. [Google Scholar] [CrossRef]
  158. Blazer, A.; Chengappa, K.N.R.; Foran, W.; Parr, A.C.; Kahn, C.E.; Luna, B.; Sarpal, D.K. Changes in corticostriatal connectivity and striatal tissue iron associated with efficacy of clozapine for treatment-resistant schizophrenia. Psychopharmacology 2022, 239, 2503–2514. [Google Scholar] [CrossRef]
  159. MacCrimmon, D.; Brunet, D.; Criollo, M.; Galin, H.; Lawson, J.S. Clozapine Augments Delta, Theta, and Right Frontal EEG Alpha Power in Schizophrenic Patients. ISRN Psychiatry 2012, 2012, 596486. [Google Scholar] [CrossRef]
  160. Lahti, A.C.; Holcomb, H.H.; Weiler, M.A.; Medoff, D.R.; Tamminga, C.A. Functional effects of antipsychotic drugs: Comparing clozapine with haloperidol. Biol. Psychiatry 2003, 53, 601–608. [Google Scholar] [CrossRef]
  161. Iasevoli, F.; D’Ambrosio, L.; Ciccarelli, M.; Barone, A.; Gaudieri, V.; Cocozza, S.; Pontillo, G.; Brunetti, A.; Cuocolo, A.; de Bartolomeis, A.; et al. Altered Patterns of Brain Glucose Metabolism Involve More Extensive and Discrete Cortical Areas in Treatment-resistant Schizophrenia Patients Compared to Responder Patients and Controls: Results from a Head-to-Head 2-[18F]-FDG-PET Study. Schizophr. Bull. 2022, in press. [Google Scholar] [CrossRef] [PubMed]
  162. Arranz, M.J.; Salazar, J.; Hernández, M.H. Pharmacogenetics of antipsychotics: Clinical utility and implementation. Behav. Brain Res. 2020, 401, 113058. [Google Scholar] [CrossRef] [PubMed]
  163. Barone, A.; De Prisco, M.; Altavilla, B.; Avagliano, C.; Balletta, R.; Buonaguro, E.F.; Ciccarelli, M.; D’Ambrosio, L.; Giordano, S.; Latte, G.; et al. Disorganization domain as a putative predictor of Treatment Resistant Schizophrenia (TRS) diagnosis: A machine learning approach. J. Psychiatr. Res. 2022, 155, 572–578. [Google Scholar] [CrossRef] [PubMed]
  164. Iasevoli, F.; Avagliano, C.; Altavilla, B.; Barone, A.; Ciccarelli, M.; D’Ambrosio, L.; Francesco, D.N.; Razzino, E.; Fornaro, M.; de Bartolomeis, A. Evaluation of a few discrete clinical markers may predict categorization of actively symptomatic non-acute schizophrenia patients as treatment resistant or responders: A study by ROC curve analysis and multivariate analyses. Psychiatry Res. 2018, 269, 481–493. [Google Scholar] [CrossRef]
  165. Iasevoli, F.; Avagliano, C.; Altavilla, B.; Barone, A.; D’Ambrosio, L.; Matrone, M.; Francesco, D.N.; Razzino, E.; De Bartolomeis, A. Disease Severity in Treatment Resistant Schizophrenia Patients Is Mainly Affected by Negative Symptoms, Which Mediate the Effects of Cognitive Dysfunctions and Neurological Soft Signs. Front. Psychiatry 2018, 9, 553. [Google Scholar] [CrossRef]
  166. Iasevoli, F.; Razzino, E.; Altavilla, B.; Avagliano, C.; Barone, A.; Ciccarelli, M.; D’Ambrosio, L.; Matrone, M.; Milandri, F.; Francesco, D.N.; et al. Relationships between early age at onset of psychotic symptoms and treatment resistant schizophrenia. Early Interv. Psychiatry 2021, 16, 352–362. [Google Scholar] [CrossRef]
  167. Neill, J.C.; Barnes, S.; Cook, S.; Grayson, B.; Idris, N.F.; McLean, S.L.; Snigdha, S.; Rajagopal, L.; Harte, M.K. Animal models of cognitive dysfunction and negative symptoms of schizophrenia: Focus on NMDA receptor antagonism. Pharmacol. Ther. 2010, 128, 419–432. [Google Scholar] [CrossRef] [Green Version]
  168. Białoń, M.; Wąsik, A. Advantages and Limitations of Animal Schizophrenia Models. Int. J. Mol. Sci. 2022, 23, 5968. [Google Scholar] [CrossRef]
  169. Caravaggio, F.; Porco, N.; Kim, J.; Carmona, E.T.; Brown, E.; Iwata, Y.; Nakajima, S.; Gerretsen, P.; Remington, G.; Graff-Guerrero, A. Measuring Amphetamine-Induced Dopamine Release in Humans: A Comparative Meta-Analysis of [11C]-Raclopride and [11C]-(+)-PHNO Studies. Biol. Psychiatry 2021, 89, S94–S95. [Google Scholar] [CrossRef]
  170. Howes, O.D.; Bose, S.K.; Turkheimer, F.; Valli, I.; Egerton, A.; Valmaggia, L.R.; Murray, R.M.; McGuire, P. Dopamine Synthesis Capacity Before Onset of Psychosis: A Prospective [18F]-DOPA PET Imaging Study. Am. J. Psychiatry 2011, 168, 1311–1317. [Google Scholar] [CrossRef] [Green Version]
  171. Slifstein, M.; Kegeles, L.S.; Xu, X.; Thompson, J.L.; Urban, N.; Castrillon, J.; Hackett, E.; Bae, S.-A.; Laruelle, M.; Abi-Dargham, A. Striatal and extrastriatal dopamine release measured with PET and [18F] fallypride. Synapse 2009, 64, 350–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Breier, A.; Su, T.-P.; Saunders, R.; Carson, R.E.; Kolachana, B.S.; de Bartolomeis, A.; Weinberger, D.R.; Weisenfeld, N.; Malhotra, A.K.; Eckelman, W.C.; et al. Schizophrenia is associated with elevated amphetamine-induced synaptic dopamine concentrations: Evidence from a novel positron emission tomography method. Proc. Natl. Acad. Sci. USA 1997, 94, 2569–2574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Demjaha, A.; Murray, R.; McGuire, P.; Kapur, S.; Howes, O. Dopamine Synthesis Capacity in Patients with Treatment-Resistant Schizophrenia. Am. J. Psychiatry 2012, 169, 1203–1210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kim, E.; Howes, O.D.; Veronese, M.; Beck, K.; Seo, S.; Park, J.W.; Lee, J.S.; Lee, Y.-S.; Kwon, J.S. Presynaptic Dopamine Capacity in Patients with Treatment-Resistant Schizophrenia Taking Clozapine: An [18F]DOPA PET Study. Neuropsychopharmacology 2016, 42, 941–950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Egerton, A.; Murphy, A.; Donocik, J.; Anton, A.; Barker, G.J.; Collier, T.; Deakin, B.; Drake, R.; Eliasson, E.; Emsley, R.; et al. Dopamine and Glutamate in Antipsychotic-Responsive Compared with Antipsychotic-Nonresponsive Psychosis: A Multicenter Positron Emission Tomography and Magnetic Resonance Spectroscopy Study (STRATA). Schizophr. Bull. 2020, 47, 505–516. [Google Scholar] [CrossRef] [PubMed]
  176. Mouchlianitis, E.; Bloomfield, M.; Law, V.; Beck, K.; Selvaraj, S.; Rasquinha, N.; Waldman, A.D.; Turkheimer, F.; Egerton, A.; Stone, J.; et al. Treatment-Resistant Schizophrenia Patients Show Elevated Anterior Cingulate Cortex Glutamate Compared to Treatment-Responsive. Schizophr. Bull. 2016, 42, 744–752. [Google Scholar] [CrossRef]
  177. Huang, L.-C.; Lin, S.-H.; Tseng, H.-H.; Chen, K.C.; Abdullah, M.; Yang, Y.K. Altered glutamate level and its association with working memory among patients with treatment-resistant schizophrenia (TRS): A proton magnetic resonance spectroscopy study. Psychol. Med. 2022; in press. [Google Scholar] [CrossRef]
  178. Matrone, M.; Kotzalidis, G.D.; Romano, A.; Bozzao, A.; Cuomo, I.; Valente, F.; Gabaglio, C.; Lombardozzi, G.; Trovini, G.; Amici, E.; et al. Treatment-resistant schizophrenia: Addressing white matter integrity, intracortical glutamate levels, clinical and cognitive profiles between early- and adult-onset patients. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 114, 110493. [Google Scholar] [CrossRef]
  179. Iasevoli, F.; Buonaguro, E.F. The Effects of Antipsychotics on the Synaptic Plasticity Gene Homer1a Depend on a Combination of Their Receptor Profile, Dose, Duration of Treatment, and Brain Regions Targeted. Int. J. Mol. Sci. 2020, 21, 5555. [Google Scholar] [CrossRef]
Figure 1. Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) flow chart. The diagram outlines the databases queried, the number of titles/abstracts screened, the full-text documents retrieved, and the number of included and excluded studies. Ti & ab, title and abstract [20].
Figure 1. Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) flow chart. The diagram outlines the databases queried, the number of titles/abstracts screened, the full-text documents retrieved, and the number of included and excluded studies. Ti & ab, title and abstract [20].
Biomedicines 10 03183 g001
Figure 2. Graphical representation of network attributes. The circles indicate nodes; the nodes of the same color are grouped into modules; the dotted lines represent the hedges. Please, see the corresponding text for a full acknowledgment of network properties. Created with BioRender.com, accessed on 24 November 2022.
Figure 2. Graphical representation of network attributes. The circles indicate nodes; the nodes of the same color are grouped into modules; the dotted lines represent the hedges. Please, see the corresponding text for a full acknowledgment of network properties. Created with BioRender.com, accessed on 24 November 2022.
Biomedicines 10 03183 g002
Figure 3. Interaction of presynaptic serotonergic, dopaminergic, and glutamatergic terminals with postsynaptic density proteins. 5-HTP, 5-Hydroxytryptophan; 5-HT, 5-hydroxytryptamine; VMAT2, Vesicular monoamine transporter 2; MAO, Monoamine oxidase; SERT, Serotonin transporter; L-DOPA, Levodopa; COMT, Catechol-O-methyltransferase; MAO-B, Monoamine oxidase B; DAT, Dopamine transporter; D2R, Dopamine receptor D2; TCA cycle, Tricarboxylic acid cycle; SNAT1, Sodium-coupled neutral amino acid transporters; EAAT-2, Excitatory amino acid transporter 2; 5-HT2AR, 5-hydroxytryptamine 2 A receptor; MUPP1, multiple PDZ protein 1; PSD-95, Postsynaptic density protein 95; D1R, Dopamine receptor D1; NMDAR, N-methyl-D-aspartate receptor; AMPAR, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; CAMKII, Ca2+/calmodulin-dependent protein kinase II; GKAP, Guanylate kinase-associated protein; TANK, TRAF Family Member Associated NFKB Activator; SHANK, SH3 and multiple ankyrin repeat domains protein; F-ACTIN, filamentous-actin; mGluR5, Metabotropic glutamate receptor 5; TRPC, Transient receptor potential channels; IP3R, Inositol trisphosphate receptor. Created with BioRender.com, accessed on 26 October 2022.
Figure 3. Interaction of presynaptic serotonergic, dopaminergic, and glutamatergic terminals with postsynaptic density proteins. 5-HTP, 5-Hydroxytryptophan; 5-HT, 5-hydroxytryptamine; VMAT2, Vesicular monoamine transporter 2; MAO, Monoamine oxidase; SERT, Serotonin transporter; L-DOPA, Levodopa; COMT, Catechol-O-methyltransferase; MAO-B, Monoamine oxidase B; DAT, Dopamine transporter; D2R, Dopamine receptor D2; TCA cycle, Tricarboxylic acid cycle; SNAT1, Sodium-coupled neutral amino acid transporters; EAAT-2, Excitatory amino acid transporter 2; 5-HT2AR, 5-hydroxytryptamine 2 A receptor; MUPP1, multiple PDZ protein 1; PSD-95, Postsynaptic density protein 95; D1R, Dopamine receptor D1; NMDAR, N-methyl-D-aspartate receptor; AMPAR, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; CAMKII, Ca2+/calmodulin-dependent protein kinase II; GKAP, Guanylate kinase-associated protein; TANK, TRAF Family Member Associated NFKB Activator; SHANK, SH3 and multiple ankyrin repeat domains protein; F-ACTIN, filamentous-actin; mGluR5, Metabotropic glutamate receptor 5; TRPC, Transient receptor potential channels; IP3R, Inositol trisphosphate receptor. Created with BioRender.com, accessed on 26 October 2022.
Biomedicines 10 03183 g003
Figure 4. Graphical representation of the chemical structure of all antipsychotics and drugs with potential antipsychotic effects discussed in the present review.
Figure 4. Graphical representation of the chemical structure of all antipsychotics and drugs with potential antipsychotic effects discussed in the present review.
Biomedicines 10 03183 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

de Bartolomeis, A.; De Simone, G.; Ciccarelli, M.; Castiello, A.; Mazza, B.; Vellucci, L.; Barone, A. Antipsychotics-Induced Changes in Synaptic Architecture and Functional Connectivity: Translational Implications for Treatment Response and Resistance. Biomedicines 2022, 10, 3183. https://doi.org/10.3390/biomedicines10123183

AMA Style

de Bartolomeis A, De Simone G, Ciccarelli M, Castiello A, Mazza B, Vellucci L, Barone A. Antipsychotics-Induced Changes in Synaptic Architecture and Functional Connectivity: Translational Implications for Treatment Response and Resistance. Biomedicines. 2022; 10(12):3183. https://doi.org/10.3390/biomedicines10123183

Chicago/Turabian Style

de Bartolomeis, Andrea, Giuseppe De Simone, Mariateresa Ciccarelli, Alessia Castiello, Benedetta Mazza, Licia Vellucci, and Annarita Barone. 2022. "Antipsychotics-Induced Changes in Synaptic Architecture and Functional Connectivity: Translational Implications for Treatment Response and Resistance" Biomedicines 10, no. 12: 3183. https://doi.org/10.3390/biomedicines10123183

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop