Next Article in Journal
Machine-Learning Guided Discovery of Bioactive Inhibitors of PD1-PDL1 Interaction
Next Article in Special Issue
New Paradigms of Old Psychedelics in Schizophrenia
Previous Article in Journal
Regenerative Drug Discovery Using Ear Pinna Punch Wound Model in Mice
Previous Article in Special Issue
Current and Emerging Pharmacotherapeutic Interventions for the Treatment of Peripheral Nerve Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Brain Modulators in Neurodevelopment: Focus on Autism Spectrum Disorder and Associated Comorbidities

1
Department of Pharmacology & Therapeutics, College of Medicine and Health Sciences, United Arab Emirates University, Al Ain P.O. Box 17666, United Arab Emirates
2
Zayed Center for Health Sciences, United Arab Emirates University, Al Ain P.O. Box 17666, United Arab Emirates
3
Department of Biopharmaceutics and Clinical Pharmacy, School of Pharmacy, The University of Jordan, Amman P.O. Box 11942, Jordan
4
Department of Chemistry, College of Sciences, United Arab Emirates University, Al-Ain P.O. Box 15551, United Arab Emirates
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2022, 15(5), 612; https://0-doi-org.brum.beds.ac.uk/10.3390/ph15050612
Submission received: 20 April 2022 / Revised: 8 May 2022 / Accepted: 12 May 2022 / Published: 16 May 2022
(This article belongs to the Special Issue Therapeutic Agents for Neurological Disorders 2022)

Abstract

:
Autism spectrum disorder (ASD) and associated neurodevelopmental disorders share similar pathogenesis and clinical features. Pathophysiological changes in these diseases are rooted in early neuronal stem cells in the uterus. Several genetic and environmental factors potentially perturb neurogenesis and synaptogenesis processes causing incomplete or altered maturation of the brain that precedes the symptomology later in life. In this review, the impact of several endogenous neuromodulators and pharmacological agents on the foetus during pregnancy, manifested on numerous aspects of neurodevelopment is discussed. Within this context, some possible insults that may alter these modulators and therefore alter their role in neurodevelopment are high-lighted. Sometimes, a particular insult could influence several neuromodulator systems as is supported by recent research in the field of ASD and associated disorders. Dopaminergic hy-pothesis prevailed on the table for discussion of the pathogenesis of schizophrenia (SCH), atten-tion-deficit hyperactivity disorder (ADHD) and ASD for a long time. However, recent cumulative evidence suggests otherwise. Indeed, the neuromodulators that are dysregulated in ASD and comorbid disorders are as diverse as the causes and symptoms of this disease. Additionally, these neuromodulators have roles in brain development, further complicating their involvement in comorbidity. This review will survey the current understanding of the neuromodulating systems to serve the pharmacological field during pregnancy and to minimize drug-related insults in pa-tients with ASD and associated comorbidity disorders, e.g., SCH or ADHD.

1. Introduction

In the diagnostic and statistical manual of mental disorders-5 (DSM-5) and in 2013, autism with several related neuropsychiatric abnormalities has been classified into autism spectrum disorder (ASD) [1,2]. Genetic syndromes and idiopathic autism were also grouped together based on their shared diagnostic features of either impaired sociability, communication deficits, stereotyped behaviours, or in combination [3]. Moreover, ASD falls under a larger umbrella of neurodevelopmental diseases that share some similar symptoms, and more importantly, similar pathogenesis, evident by impaired brain development. Noteworthy, some of these disorders including schizophrenia (SCH), obsessive compulsive disorder, and epilepsy co-exist with ASD [4,5].
It has been demonstrated that pathophysiologic insults in neurodevelopmental disorders start early during prenatal neuron formation and maturation [6]. However, main symptoms will mostly appear later in life during childhood or adolescence. In parallel, underlying deficits in brain maturation and wiring are present from the first neuron formed to the complete adult brain. In fact, some autistic subjects can live seemingly normal lives despite the presence of the hidden impairment [6]. Moreover, and unlike males, autistic females usually have almost normal sociability that does not mirror the imbedded autistic brain [7]. Therefore, males appear to be more affected by ASD, and the disease strikes more severely in them [8]. While the underlying neurological deficits have been extensively studied since the recognition of ASD, much further research efforts are warranted to unveil the abnormalities in this complex disorder. Accordingly, approaches developed in studying genetic syndromes provided solid ground for understanding the underlying mechanisms for ASD stemming from genetic mutations [9]. Nevertheless, genetic factors only partially contribute to the etiology of ASD. In fact, around 90% of ASD cases are classified as idiopathic [9,10]. ASD literature can be divided in two approaches: (1) the study of the autistic adult brain with the aim to developing a “correcting” remedy for patients diagnosed with ASD, and (2) tracing the pathogenesis of the disease back to the womb to identify preventative solutions. Both of these approaches can be considered by analysing the contributions of central neuromodulator systems during development, and later when the disease symptoms are manifested. In this review, several selected studies describing the negative impact of several endogenous neuromodulators and pharmacological agents on the foetus during pregnancy, manifested in numerous aspects of neurodevelopment are discussed. In addition, this review will provide insights of the current understanding of the neuromodulating systems to serve the pharmacological field during pregnancy and to minimize drug-related insults in patients with ASD and associated comorbidity disorders, e.g., SCH or ADHD.

2. Neurodevelopment

2.1. Neurogenesis and Synaptogenesis

The journey of a neuronal tissue begins with the neuronal stem cells. These cells undergo proliferation and differentiation to generate neuronal and glial progenitor cells [11]. Further proliferation and differentiation of neuronal progenitor cells (NPCs) precede neuron cell formation and neurite outgrowth [11]. Some neurons are formed far from their ultimate destination, thus they need to undergo migration [12]. Neurites or projections outgrow (or sprout) to become dendrites or axons [13]. Through further maturation, axons elongate, and spines (early dendrites) protrude aided by several factors including excitatory input from neighbouring axons [14,15]. Communications, or synapses, are then formed between axons and dendrites. Chemical signalling through these synapses relies on neurotransmitters or neuromodulators stored in presynaptic compartments or vesicles until their release. Once released, they interact with postsynaptic dendritic receptors to induce excitation, inhibition, or modulation of the postsynaptic neurons [14,15]. Their receptors have different signalling pathways depending on subtypes of receptors as well as neuronal types. This process of neurogenesis and synaptogenesis starts from early gestation, with embryonic day 10 (E 10) in rodents and second trimester of pregnancy in humans (Figure 1). Further synaptogenesis and neurogenesis continue after birth and even into adulthood. For example, the hippocampus contains stem cells that upon activation can rejuvenate the neuronal formation [16].

2.2. Synaptic Plasticity

It has been reported that larger excitatory axonal input during development helps in the formation of the postsynaptic dendritic spines to build mature and functional synapses [29,30]. Later, these synapses are strengthened or weakened by the same principle in a process called synaptic plasticity. Indeed, the flow of ions through excitatory postsynaptic currents (EPSC) cause a build-up of excitatory postsynaptic potential (EPSP). Shortly following high frequency stimulation, EPSP amplitude at the activated synapses increases for a longer time compared to other types of stimulation, resulting in a long-term potentiation (LTP). On the other hand, inhibitory postsynaptic currents (IPSCs) were found to induce accumulating inhibitory postsynaptic potential, which upon a low frequency stimulation is capable of yielding long-term depression (LTD). At the functional and molecular levels, both LTP and LTD represent changes in synaptic strength and participate in neurodevelopment, synaptic plasticity, learning, and memory [29,30]. Additionally, LTP and LTD can occur at different synapses in the hippocampus, cortex, amygdala, and cerebellum [29,30]. Moreover, both LTP and LTD were well studied in glutamatergic synapses, however, the same principle can occur in any neurotransmitter system including gamma aminobutyric acid (GABA) neurons [29,30].

2.3. Neuroinflammation

It has been revealed that glial progenitor cells (GPCs) are formed in parallel with precursors of neurons, namely, NPCs [31]. Upon GPC differentiation, astrocytes, that are involved in immune activation and myelin sheet formation, are developed. Immune cells migrate to the brain during early pregnancy around the timing of early neurogenesis as depicted in Figure 1 [31]. Additionally, mast cells enter the CNS during development [32]. As immune cells, mast cells and microglia have an essential role of protecting the newly developing young neurons. This is done in association with astrocytes [32,33]. Nevertheless, the involvement of astrocytes in neurodevelopment is not only limited to mere guardiancy. These resident immune cells communicate with proliferating brain cells and directly participate in different aspects of neuro- and synaptogenesis [31,34,35]. Therefore, several cytokines and mediators, which are released jointly by those immune cells, influence cell proliferation and neurogenesis. Furthermore, microglia cells were found to secrete insulin growth factor 1 (IGF-1) to aid the growth of neuronal tissues [34]. Meanwhile, and in different brain areas, microglia cells activate the apoptosis and implement phagocytic engulfment of neuronal progenitor cells to mark the end of neurogenesis during the brain maturation process. This trimming and cropping activity of microglia cells has been found to also extend to synapses. Hence, microglia cells can sense the activity of synapses and in response remove inactive synapses for the perfection of neuronal circuits in a process referred to as synaptic pruning [31]. Therefore, active synapses will then be strengthened by early LTP activity as illustrated previously under synaptic plasticity. The function of microglia cells to modulate excitatory synaptic activity during maturation to prevent excitotoxicity is confirmed [34]. Despite their essential role in proper brain maturation, microglia cells, as described by Franzen et al, are a two-edged sword [36]. Severe infection, inflammation, or other environmental or genetic disturbances can trigger abnormal activation of immune activation that starts prenatally and progresses throughout life [32,37,38]. This pathogenic neuroinflammation—opposite to the above physiological neuroinflammation—secretes abnormal levels of several proinflammatory cytokines, which at these high levels are rather destructive on the process of neurogenesis [33] (Figure 2). Overall, pathogenic neuroinflammation can be the cause, i.e., inflammatory cytokines and effect, i.e., activated to respond to insults or damage, in such scenarios [38]. ASD in fact can be manifested as an imbalance between anti- and proinflammatory cytokines, favouring pathogenic neuroinflammation has been reported by several research groups [37,39]. Furthermore, children diagnosed with ASD have been found to show increased levels of neurotensin (NT), a neuromodulator known for its effects on mast cell activation, promoting the release of numerous proinflammatory mediators and eventually triggering neurotoxicity in ASD and epilepsy [32,37]. Another related anomality was previously described by monocyte chemoattractant protein-1 (MCP-1), which functions as mast cell chemoattractant, and was found to be elevated in brains of ASD patients [32]. On the other hand, large numbers of experimental studies have revealed that these cytokines play important roles in the development of ASD [40,41]. Furthermore, peripheral upregulated inflammation (e.g. IL-6 and IL-1β) has been found to worsen autistic symptoms in mice [42]. These proinflammatory cytokines cross the blood-brain barrier and activate microglia. Dysfunctional microglia have a negative effect on synaptic pruning, influencing the signal transmission of neurons, which contributes to ASD development—a brain disorder characterized by an imbalance between inhibitory and excitatory responses [42]. In addition, peripheral blood levels of macrophage migration inhibitory factor (MIF), eotaxin-1, MCP-1, and IL-8 were also markedly increased in ASD patients compared with controls [42]. In SCH, proinflammatory cytokines were found to reduce the level of the antiapoptotic and neurotrophic growth factor (brain-derived neurotrophic factor, BDNF) gene transcription [43]. In ASD, the neuroinflammatory reaction was described as being a long-lived process that persists after birth and specifically in the cerebellum [38]. Additionally, microglia cells were found to be additionally involved in neuroinflammatory responses to early changes in gut microbiota in several neurodevelopmental disorders [44]. On the other hand, the levels of IL-38 were found to be lower in the amygdala of children with ASD and seems to inhibit microglial inflammatory mediators [41].

2.4. Growth Factors, Synaptic Proteins and Intracellular Calcium

In addition to excitatory inputs, there are several proteins that were confirmed by their participation in growth of neurons and synapses, as well as in physical and morphological organization of synapses. The levels of these proteins was found to be dependent on several factors—primarily the development phase [45,46]. Examples of the proteins involved in growth of neurons and synapses include BDNF and members of the neurotrophin protein family. BDNF is produced from different cell types, including neurons, and interacts with receptors that are coupled to tyrosine kinase. The downstream signalling includes PI3K/Akt (phosphoinositide 3-kinases/Ak strain transforming) that in turn mediates cell survival and prevention of apoptosis [45]. Among other roles, BDNF was also reported to support neuron and microglia growth, neurite outgrowth, and LTP [45]. Moreover, BDNF gene expression was found to be controlled by neuron activity and consequently downstream signalling of neuromodulators [47,48].
Examples of the second type of structures, involved in the physical and morphological organization of synapses, are neurexin, neuroligin, and SHANK (SH3 and multiple ankyrin repeat domains proteins). They are primarily adhesion molecules and scaffolding proteins [49]. Accordingly, neuronal communication, including trans-synaptic signalling involving two families of cell-adhesion proteins, the presynaptic neurexins and the postsynaptic neuroligins, is one of the most recurrently affected pathways in neuropsychiatric disorders including ASD. Given the role of these proteins in determining synaptic function, abnormal synaptic plasticity, and failure to establish proper synaptic contacts might represent mechanisms underlying risk of ASD.
Intracellular calcium is known to be increased by several pathways, including ion channels associated with some excitatory neurotransmitters, especially glutamate through NMDA/AMPA (N-methyl-D-aspartate / α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid) receptors. Calcium was described to exhibit many roles inside the developing neurons including regulation of gene expression, neuron stem cell differentiation, apoptosis, neuron migration, neurite growth, spine formation, receptor maturation, and microglia function [50,51]. Expectedly, impaired calcium signalling was reported to be associated with neurodevelopmental disorders including ASD [50,52]. Interestingly, the release of BDNF was found to be dependent on intracellular calcium signalling [53]. Finally, it is common knowledge that intracellular calcium, through voltage gated calcium channels, is needed for any vesicular presynaptic release of any type of neuromodulators.

3. Endogenous Neuromodulators and Pharmacological Agents in Brain Development

Upon complete maturation of the brain in adulthood, neuronal systems and circuits are formed with complex synaptic interaction across brain regions. Neuromodulators at these synapses extensively participate in brain functions such as arousal, attention, cognition, emotions, and memory. Nevertheless, neuromodulators appear early in the prenatal brain before the outgrowth of any synapses [54]. Their receptors appear in the neuronal progenitor cells and many studies demonstrated their involvement in the neurodevelopmental process from neurogenesis to synaptic construction. Their receptors and functions, however, may differ between the developing and the mature brain [55]. By way of illustration, GABAA receptors are excitatory in the early brain development then shift to inhibitory later [55,56]. Furthermore, glutamatergic NMDA receptors action prevails in the foetal brain. Nevertheless, glutamate effect is mediated by AMPA and kinase receptor currents later in the developed brain [57]. Figure 1 illustrates the estimated ontogeny of several neuromodulators or their signalling systems in rodent brains that coincide with the early neurogenesis. Additionally, some neuromodulators were found to serve as communication pathways between immune cells and developing neurons as depicted by Figure 2.
ASD and related neurodevelopmental disorders have both genetic and environmental components in their pathogenesis. These components will ultimately converge into deficits in one or more neuromodulator signalling systems. Table 1 summarises examples of these environmental insults sorted by the affected neuromodulators. These genetic mutations afflict genes that encode for synaptic proteins or signalling pathways. Therefore, either the synaptic structure or function was found to be disrupted. The ultimate consequence of this disruption in synapses was reported to be altered or compromised neuromodulator systems that utilize these synapses to participate in overall brain function. Some of these mutations affect proteins that are not specific to certain synapses and therefore different neuromodulator signalling pathways are concerned. For example, the SHANK3 gene is mutated in some genetic and idiopathic forms or models of ASD which display deficits in signalling pathways of numerous brain neurotransmitters including dopamine, GABA, and glutamate [58,59], as SHANK3 belongs to a family of scaffolding proteins that are connected to the postsynaptic membrane to participate in synaptic signalling [60].

3.1. Neurotransmitters

3.1.1. Excitatory/Inhibitory Balance (Glutamate/GABA)

Brain homeostasis and function relies mainly on excitation/inhibition balance maintained by glutamatergic pyramidal neurons and GABAergic interneurons [91,92]. Furthermore, previous reports revealed that excitation/inhibition balance is maintained through compensatory mechanisms at the level of synaptic numbers and synaptic strength [93]. For example, inhibition of GABAergic synapses by bicuculline resulted in downregulation of excitatory currents to prevent an over-excitation status of cells [93]. Excitation/inhibition balance was also found to differ with age, brain maturation, and regions [93]. For example, mature cortex has a ratio of 80 to 20% excitation/inhibition [91,93]. Additionally, GABA is the major inhibitory neurotransmitter in the adult brain, however, it has an excitatory effect in the developing brain potentiating NMDA effect [21,55,56].
Interestingly, glutamate and GABA appear early in the embryonic life, indicating their role in cell proliferation, neurogenesis, and synaptogenesis [56,91,94,95]. Loss of functional mutation in a subunit of NMDA receptor, a receptor that was identified in severe ASD, alters neuronal differentiation [96]. Consequently, exposure to drugs that affect GABA/glutamate receptors during pregnancy can therefore interfere with brain maturation before affecting any excitatory/inhibitory balance later throughout development [56,91]. Several drugs, e.g., benzodiazepines or valproate, if administered during pregnancy, were found to predispose offspring to neurodevelopmental diseases, e.g., ASD [91,97]. Furthermore, alcohol consumption by pregnant women was found to depress GABAA receptor in an early stage of pregnancy, explaining the impaired growth of brains of offspring [55]. Likewise, in utero exposure to mGluR5 (metabotropic glutamate 5 receptor) or NMDA modulators (e.g., amantadine or ketamine) was reported to significantly increase the risk for ASD or other developmental disorders in neonates [98,99]. Noteworthy, the impaired excitatory/inhibitory balance was found to have its roots in every neurodevelopmental disorder including SCH, ASD, and epilepsy, to name a few [92,93]. Therefore, reduction in GABAergic neurotransmission or increment in glutamatergic circuits has been implicated in several neurodevelopmental disorders [93,100]. Nevertheless, reduction in glutamatergic function or hyperfunction of GABAergic neurotransmission can adversely affect neuron maturation and synaptogenesis. However, abnormal elevation in glutamatergic neurotransmission was found to induce neurotoxicity and consequently neurodevelopmental disorders including ASD and SCH [101,102]. On the other side, neuroinflammation could have a bidirectional role in this situation, namely, it may remove excess glutamate and prevent damage, or activate microglia cells and astrocytes that may release glutamate and proinflammatory factors with increased excitotoxicity, oxidative stress, and neuronal damage [32,101,103]. However, an appropriate amount of excitation was found to be necessary for brain maturation. Therefore, collective effects of NMDA glutamate receptors and early excitatory GABAA receptors were found in previous studies to mediate normal development of the brain [21,55,56]. Furthermore, hypofunction of glutamate receptors (AMPA, NMDA, and mGluR1-5) was found to be associated with impaired neuronal differentiation, maturation, synaptogenesis, and either synaptic plasticity, LTP, or in combination [56,104]. Such hypofunction was also reported to constitute a predisposition to neurodevelopmental disorders such as SCH and ASD [104].
Moreover, synaptogenesis and neurogenesis were found to continue after birth and in the adult brain in certain areas including the hippocampus [105]. Indeed, hippocampal LTP/LTD ratio was mostly found to depend particularly on glutamate signalling through NMDA receptors, signifying the effect of hypofunction of these glutamate receptors in impairing synaptic plasticity [106,107]. Therefore, several NMDA antagonists including phencyclidines, dizocilpine and ketamine were found to induce psychosis in numerous preclinical and clinical experiments [108,109]. Likewise, GABAergic hyperfunction was described to affect synaptogenesis and plasticity [110]. Accordingly, excitotoxicity was found to be injurious in the mature brain in several neurodevelopmental disorders such as epilepsy and ASD [106]. Remarkably, lower ability of NMDA receptors to activate GABAergic inhibitory interneurons was described to be adopted by NMDA hypofunction theory of SCH [111], demonstrating that there is an overall hyperexcitation between both neurotransmission systems [108]. In further support of the latter observations, reduced GABAergic signalling was well-documented in patients diagnosed with SCH [100]. Moreover, high chloride ion concentration inside immature neurons was found to drive GABAergic effect toward excitation before the shift in ion gradient later in the developed brain [112]. Interestingly, patients diagnosed with SCH were found to exhibit an altered expression of regulatory machinery that controls chloride gradient rendering GABA less inhibitory in the adult brain [112,113]. Further contraindications arise when both hyper- and hypoexcitability coincide. In maternal valproic acid-induced ASD in rats, elevated NMDA function was shown to reside in the young brain for some time before transforming into deficit in NMDA signalling and LTP in the adult brain [106]. Additionally, valproic acid was shown to induce ASD prenatally, yet it constitutes one of the treatment modalities of ASD and co-morbidities later in life [114,115]. Taken together, a well-balanced excitation is necessary for brain maturation and its proper functions.

3.1.2. Dopamine

The neurotransmitter dopamine and dopaminergic receptors appear early in the prenatal brain before the outgrowth of any synapses [116]. Moreover, once the dopaminergic neuronal system is formed, it undergoes changes until adulthood critical for development [116]. Interestingly, dopaminergic system maturation during adolescence coincides with onset of neuropsychiatric disease and substance abuse during this period [116]. Maturation of the excitatory/inhibitory balance in the frontal cortex was also found to be associated with dopaminergic neurotransmission, specifically dopamine D1 receptors [116]. The striatum and later the cortex are the earliest areas to receive dopaminergic innervation, and their development is significantly altered by factors influencing dopaminergic neurotransmission [116,117,118]. Dopamine signalling was reported to participate in several aspects of neurogenesis, neuron migration, and synaptogenesis [116,119], and its effect varies with receptor subtypes, brain regions, and age [116,119,120]. Furthermore, brain levels of dopamine have been found to influence neurogenesis, neuron migration, and synaptogenesis, since high or low dopaminergic signalling results in impaired plasticity of dopaminergic receptors [120]. This modulatory role of dopamine can be further highlighted by the opposite activities of neuroleptic and antipsychotic drugs in the developing versus developed brains. Consequently, exposure to different antipsychotics in the prenatal brain was reported to alter synaptogenesis while the same agents constitute the main treatment of ASD and SCH [55,119]. Overall, genetic, or environmental perturbation in the dopaminergic neurotransmission system has been associated with an array of neurodevelopmental disorders including ADHD, SCH, and ASD [120].

3.1.3. Serotonin

The neurotransmitter serotonin (5-HT) has an important role in neurodevelopment due to many reasons, including that (1) 5-HT is a trophic factor implicated in neural progenitor cell proliferation, synaptic plasticity, synapses, and circuit formation; (2) 5-HT was found to be supplied by the placenta in early pregnancy; and (3) that serotonergic neurons were found to be the first neuronal system to project to the forebrain in the second trimester in human pregnancy [121,122,123,124]. Furthermore, there is a period of 5-HT synthesis surge in the brain during childhood, and serotonin neuron projections and receptor expressions in the brain vary with age as a result. Denervation of serotonin neurons of an area in the mature brain was described to lead to damaged synapses and de-maturation of that area [121,122,123,124]. Noteworthy, SCH, Down’s syndrome, and ASD are neurodevelopmental diseases that were found to share the pathological basis of impaired serotoninergic neuron functions [121,122]. Interestingly, low level of lead exposure in neonates was reported to cause abnormality in the developed brain, due to damaging serotoninergic neurons [125]. Importantly, the well-known drug class of selective serotonin reuptake inhibitors (SSRIs) target serotonin transporters and thus, are capable of increasing the levels of 5-HT in the synaptic cleft [123]. Therefore, postnatal (through breastfeeding) exposure to SSRIs was found to adversely alter neurodevelopment and predispose neonates to ASD-like symptoms [123]. Bearing in mind that SSRIs are one of the most used treatment options for ASD and in similarity with valproic acid, which increases the risk of (or induce) ASD in the prenatal life but is a therapy in adulthood [126,127]. Furthermore, brain 5-HT levels were found to be disturbed during brain development in patients diagnosed with ASD compared to neurotypical individuals [123,124]. In this regard, a correlation was found between serotonin receptor subtype 7 (5-HT7R) and neurodevelopmental disorders including ASD, fragile X syndrome, and Rett syndrome. Therefore, modulators of 5-HT7R were found to improve altered behaviours in animal models and also to affect neuronal morphology via the 5-HT7Rs signalling pathway, signifying 5-HT7Rs to be a potential therapeutic target for the treatment of neurodevelopmental disorders [128].

3.1.4. Acetylcholine

The neurotransmitter acetylcholine, through muscarinic and nicotinic receptors, was also found to have a trophic role in many cell types including neurons [129]. In addition, acetylcholine was described to regulate neuronal circuits in the developing brain [130,131]. Accordingly, nicotinic receptors were found to mediate postnatal change in GABA receptors from excitatory to inhibitory maintaining balanced excitation [131]. Hence, deficits in cholinergic neurotransmission were observed in several neurodevelopmental disorders including ASD, ADHD, Tourette disorder, and SCH [132,133,134]. Furthermore, cholinergic deficits were linked to spontaneous arousal fluctuation, an early biomarker of ASD before the actual onset of symptoms [135]. On the other hand, elevated acetylcholine levels during neurodevelopment were reported to pose the same risk as deficiency. Notably, exposure to different organophosphates, synthetic compounds which are capable of blocking the acetylcholine metabolizing cholinesterase enzyme, was directly associated with several neurodevelopmental diseases [136]. Similarly, maternal inflammation was reported to predispose the offspring to ASD through incrementing the amount of cholinergic neurons in the foetal brain [137].

3.1.5. Norepinephrine

The locus coeruleus (LC) contains noradrenergic cell body clusters that send axonal extensions to a wide range of brain areas [19]. Pattern of LC neuron firing differs with age, as described above, partly because of postnatal impact of α2 receptors, which are auto-receptors. Both α- and β-adrenergic receptor subtypes were shown to be involved in growth, maturation, and formation of neurons and synapses through different signalling pathways [19]. Altered dendritic formation coupled to impairment in extracellular signal-regulated kinase (ERK) pathways, a downstream of α2 receptor activation, was previously observed in autistic Black and Tan Brachyury (BTBR) mice [19]. Abnormal α/β receptor signalling, polymorphism, or early exposure of their modulators, was linked to psychosis, cerebral palsy, ADHD, and ASD [19]. Postnatally, noradrenergic signalling was documented to moderate the modulation of attachment learning within the first weeks which influence behaviour later in adulthood [19]. The importance of this finding is that attachment deficit is an early sign of ASD [138]. Furthermore, disturbances in adrenergic system were found to perturb signalling of other neuronal systems such as dopamine and serotonin that are co-players in normal brain maturation [19]. Noteworthy, adrenergic neurons have a role in the modulation of microglia maturation and activity later in life, which may impact neuroinflammation and synaptic plasticity [19,139]. This is quite relevant as neuroinflammation was described to be one of the principal aetiologies in the pathogenesis of neurodevelopment disorders including ASD [140].

3.1.6. Histamine

For the brain neurotransmitter histamine, two sources of histamine in the brain, namely, neurons and mast cells, have been reported [141]. The histaminergic neurons were reported to originate early at E 13 in rodents, equivalent to second trimester in humans [141]. Early contribution of histamine to neurogenesis was described to be maintained by histamine supply from other neurons (i.e., noradrenergic) and mast cells [142]. In addition to neurogenesis, histamine was found to meditate synaptogenesis, synaptic maturation, synaptic plasticity, and neural circuit formation [142]. Interestingly, a previous preclinical research showed that induced histaminergic dysfunction was postulated as a mechanistic cause in valproic acid-induced ASD in a zebrafish model [143]. Histamine interacts with four different histamine receptor subtypes, namely, H1–H4 receptors which were reported to exhibit different roles in neurodevelopment. Interestingly, H3 heteroreceptors as mentioned above were described to control the release of histamine, as well as glutamate and several other neurotransmitters [142]. Moreover, histamine was found to interact with the polyamine binding site on NMDA receptors that are implicated in synaptic plasticity [141]. By interaction with H3 receptors, histamine shows many converging roles in neurodevelopment [142]. Notably, antagonists/inverse agonists targeting H3 receptors were suggested as a possible and promising pharmacological treatment that may tackle the underlying pathogenesis of several neurodevelopmental disorders, e.g., ASD in addition to symptomatic alleviation [126,127,144,145,146,147,148]. However, the success with these agents has been limited thus far to preclinical studies, as there are isoforms of H3Rs with different distributions that contribute to the varying or rather conflicting effects of H3R antagonists/inverse agonists [126,142,149]. Interestingly, mast cells being a non-neuronal source of histamine in the brain were reported to add to the multifaceted roles of histamine in neurodevelopment. Accordingly, crosstalk between neurons and immune cells was found to be essential for neurodevelopment despite the involvement of abnormal neuroinflammation in many neurodevelopmental diseases [150]. Indeed, histamine was reported to facilitate the release of neurotrophins (e.g., glial cell line-derived neurotrophic factor, GDNF) and several other anti-inflammatory factors (e.g., interleukin-10, IL-10) from astrocytes and microglia, preventing simultaneous proinflammatory factors such as transforming growth factor beta (TGF-β) [141,142,151]. However, histamine was also reported to be capable of facilitating proinflammatory responses of immune cells in the brain through H1 and H4 receptors [151]. Noteworthy, H1–H3 receptors are expressed in neurons as well as astrocytes, microglia, and mast cells [142]. However, the H4 receptor expression profile and possible role in neurodevelopment has been limited barely to only immune cells [142,152].

3.2. Neuropeptides in Brain Development

3.2.1. Oxytocin

The neuropeptide oxytocin and neurodevelopment are relevant topics in obstetrics since labour is facilitated by oxytocin administration. Additionally, concentration of oxytocin in the maternal blood was found to be increased naturally during the perinatal time for the same purpose. In this regard, four plausible questions arise: (1) can oxytocin be transferred from mother to foetal blood and more importantly to foetal brain? (2) is there any oxytocin receptor expression in the foetal brain? (3) is there any oxytocin production inside the immature brain? and finally, (4) how does oxytocin affect neurodevelopment? The following paragraphs will be devoted to answer these questions.
Oxytocin can be supplied to the foetus via the placenta to its blood [153]. Oxytocin’s entry into the brain exploits the state of immaturity of the blood brain barrier (BBB) in this period, as it was observed that immature BBB is permeable to oxytocin [153]. The significance of the latter finding is that it offers an explanation for the clinical observation of increased neurodevelopmental disorders in children delivered through oxytocin-facilitated labour [153]. Monks et al. reviewed several studies that concluded such observations. Therefore, natural oxytocin is needed in this period for proper brain development as well. In a previous preclinical study, null mutant male mice were found to differ from another null mutant mice, and this difference was dependent on the presence of oxytocin in their mothers. Null mutant mice delivered by null mutant dams were described to be more aggressive compared to their conspecifics delivered by heterozygous dams. Additionally, their behaviours were not affected if they were cross-fostered by wild-type dams [154]. Therefore, direct maternal transplacental transfer of oxytocin is important for adult behaviours. Moreover, perinatal oxytocin impact on neurodevelopment can be attributed to different factors. First, perinatal oxytocin contributes to the shift of GABAergic signalling to inhibitory before birth in rodents and thereby it modulates excitatory/inhibitory balance [155,156]. Second, oxytocin confers beneficial activities in maturation, neurogenesis, and synaptogenesis of hippocampus neurons in the newborn [156,157]. Finally, oxytocin prevents neuroinflammation during the perinatal period [158]. Taken together, perinatal oxytocin was found to be essential for shaping adult behaviours, but moderate levels are required. Noteworthy, oxytocin receptors can couple to Gi or Gq/11 and generate a plethora of downstream activities depending on the oxytocin levels that are present [159].
Oxytocin receptors were detected as early as E-12.5 in the mouse brain; however, oxytocin itself was not expressed until postnatal day 2 [25,154]. Excitingly, females displayed oxytocin mRNA at E 16.5 and as compared to males, where the mRNA appears much later [154]. Although it is not known when oxytocin will be expressed, it can be assumed that at least there is a gender specific development of oxytocinergic neurons in a reductionist’s viewpoint [154]. More broadly, it is also reasonable to assume a role of oxytocin in gender differences in neurodevelopment and consequently adult behaviour [154,160]. Additionally, the density of oxytocin receptors was found to be significantly higher in the female foetal brain [154]. Notably, oxytocin was found to regulate social behaviours and social recognition memory besides other roles [154,156,161,162,163]. Therefore, abnormalities in oxytocin levels would produce more sociability-impairing consequences in males than females. Interestingly, recent studies found a strong correlation between higher oxytocin receptor gene methylation (i.e., silencing) and either reduced IQ, social interaction, or in combination, scores in patients with neurodevelopment disorders [164,165].
Despite the fast move to clinical trials with oxytocin, little attention has been paid to the possibility that the oxytocin system in the brain is perturbed in these disorders and to what extent such perturbations may contribute to social behaviour deficits. However, large-scale whole-exome sequencing studies in subjects with ASD, along with biochemical and electrophysiological studies in animal models of the disorder, indicated several risk genes that have an essential role in brain synapses, suggesting that deficits in synaptic activity and plasticity underlie the pathophysiology in a considerable portion of these cases. Moreover, oxytocin has been repeatedly shown, both in vitro and in vivo, to alter synaptic properties and plasticity and to modulate neural activity in circuits that regulate social behaviours. Together, these observations concluded that failure of the oxytocin system during early development, as a direct or indirect consequence of genetic mutations, may influence social behaviours by altering synaptic activity and plasticity [111].

3.2.2. Opioids

Endogenous opioids consist of endorphins, enkephalins, and dynorphins [166]. To varying degrees, they all interact with three types of opioid receptors, namely, mu, delta, and kappa. All these receptors are associated with Gi proteins. Furthermore, opioid receptors were found to participate in pain perception between the brain and spinal cord [167]. Additionally, they were reported to mediate emotions and behaviours such as distress (emotional pain), anger, anxiety, sadness, motivation, and social affiliative actions [168]. Moreover and in rodents, endogenous opioids and their receptors were reported to be expressed at different time points during embryonic days, namely, between embryonic days 13 and 21, with delta receptors being the late receptor subtype to appear [169]. There is an interesting hypothesis reported early in the literature that assumed opioid excess in the brain is a cause or a factor of autistic features. A recent case report hypothesized a “high opioid tone” group of autistic patients [170,171,172]. However, the use of opioid receptor antagonist naltrexone to overcome such excess resulted in a significant reduction in hyperactivity without any improvement in the core symptoms of ASD [173]. Furthermore, details about the exact roles of endogenous opioids during development are scarce and mostly come from early literature. Yet, there are several reasons that support their involvement in neurodevelopment. First, preclinical and very recent clinical studies indicated that in utero exposure to exogenous opioids, such as morphine, significantly alter neurodevelopment of the neonates [174,175]. Mechanistically, prenatal exposure to an opioid agonist may decrease myelination and augment apoptosis of neurons and microglia [176,177]. In addition, opiates were found to decrease proliferation of cerebellar granular neurons in vitro and as expected could decrease cerebellar growth [178]. Secondly, a previous in vitro study in murine neuronal stem cells suggested that δ receptors have an important contribution to neurogenesis and neuroprotection [179]. Endogenous opioids were described as being negative growth factors during brain maturation, and opioid receptor antagonist naltrexone administration during pregnancy was found to increase proliferation of brain cells and overall brain weight [180]. Moreover, the abuse of some addictive agents or drugs (ethanol, and cocaine) that have neurodevelopmental consequences was clearly associated with elevated opioid signalling in different areas of the developing brain [79,167,181]. Therefore, neurodevelopmental changes produced by the addictive drugs during pregnancy can partially be explained by an opioid receptor upregulation mechanistically influencing neuron maturation[182]. Additional important consideration is that endogenous opioids were found to induce inhibitory sequela on maturation of prenatal neurons [180,182]. In harmony with the above-mentioned preclinical and clinical observations and following prenatal addictive drug exposure, dopaminergic neuronal function was found to be modified due to heightened opioid signalling at these neurons during development [79,181].

3.2.3. Cannabinoids

Two main endogenous or natural cannabinoids (eCB) were identified, namely, 2-arachidonoyl glycerol (2-AG) and anandamide (AEA) [183]. Both cannabinoids were reported not to be secreted in the conventional vascular pathway such as other classical neuromodulators [183]. They are released from postsynaptic neurons—and sometimes microglia—to control presynaptic activity and release of other neuromodulators such as glutamate and GABA among others as depicted in Figure 1 [183]. Consequently, 2-AG and AEA will then bind either cannabinoid receptors 1 (CBDRs1) or CBDRs2, which are coupled to Gi proteins. It was previously reported that neurons mostly express CBDRs1, and CBDRs2 are predominately found in immune cells including microglia and mast cells [183,184]. Importantly, in utero exposure to exogenous cannabinoids was found to produce deficits in maturation of the foetal brain which are strongly associated with SCH [185,186]. Do these deficits indicate a role of eCB signalling in neurodevelopment since exogenous cannabinoids are able to bind to some other “off-target” receptors? In humans, it has been reported that CBDRs1 are expressed as early as 14 weeks of gestation (equivalent to embryonic day 12 in rodents) and gradually grow in density to adulthood [187]. Similarly, mRNA of these receptors was detected around embryonic days 11–14 in rats. Many previous studies that utilized genetic or selective pharmacological manipulations of eCBD receptors concluded the involvement of this neuronal system in about each aspect of brain development. Hereafter, Alexandre et al. nicely reviewed roles of CBDRs1 and CBDRs2 in regulating proliferation of neuronal progenitor cells, neurogenesis, migration of neurons, and synaptogenesis [188]. In ASD children, significantly low levels of expression of N-acyl phosphatidylethanolamine-specific phospholipase D (NAPE-PLD), one of the synthesizing enzymes of AEA, were found compared to normal individuals [189]. Foetal exposure to cannabinoids impacts other neuronal systems. Therefore, cannabinoids were found to alter the expression of tyrosine hydroxylase and proenkephalin which are essential in the formation of dopamine and enkephalin, respectively [190]. Furthermore, CBDRs were found to be functional at an early stage of development during the time of ontogeny of other neuronal systems [191]. Additionally, CBDRs were found to be expressed in microglia and immune cells, especially CBDRs2, and several previous studies indicated the anti-inflammatory and neuroprotection roles of cannabinoids [184,189,192,193]. In agreement with other neuromodulator systems mentioned above, exogenous cannabinoid can be used to treat some neuropsychiatric disorders despite their damaging effects during prenatal stages. These agents could improve either behavioural anomalies, neuroinflammations, or both, in several animal models of ASD [194]. In clinical trials, amelioration of core symptoms of ASD by medicinal cannabinoids was found to be less pronounced than their seizure-inhibiting activity, since seizure is one significant co-morbidity in ASD [195]. Their activity on seizure control was found to stem from their modulatory roles on the release of GABA and glutamate and subsequent control of excitatory/inhibitory balance. Interestingly, cannabidiol has recently been approved by the Food and Drug Administration (FDA) and the European Medicine Agency (EMA) for the treatment of two ASD-associated epilepsy syndromes, namely, Lennox–Gastaut and Dravet syndromes [189].

3.2.4. Orexin

The two subtypes of orexin peptide, namely, orexin A and B, were found to have a neuronal system in the brain originating from lateral hypothalamus and spreading to most of the CNS [196]. Orexin peptides were detected at E 18 much later than their neurons that are expanding during E 12 [27,197]. Orexin’s late appearance and slow maturation during postnatal weeks was described to coincide with their main role in the sleep/wake cycle and feeding behaviour [27,197,198,199]. Moreover, orexins were reported to exhibit neurotrophic roles and to improve LTP in the developed adult hippocampus [196]. Furthermore, orexins were found to provide neuroprotective effects after hypoxic or ischemic attacks [200]. Furthermore, and through their receptors, orexins were shown to induce pro-survival changes in cultured cortical neurons and mediated mainly through orexin receptor 2 (OXR2). These cultured neurons are harvested from Wister rat embryonic brains on E 17 [201]. Thus, it is plausible to anticipate orexins’ involvement in the maturation of the embryonic brain. However, there is limited evidence about their participation in brain development. Hereafter, orexin A was found to be lower in serum levels in untreated ADHD children [202]. In patients with Prader–Willi neurodevelopmental syndrome, low level of orexin was also reported in their cerebrospinal fluid (CSF). Accordingly, the latter findings may suggest the involvement of orexins in the symptomatology of these diseases; however, their role in the pathogenesis cannot be clearly inferred [203]. Prenatal exposure to some agents capable of altering the orexin system, e.g., nicotine or ethanol, was shown to induce neurodevelopmental changes in adulthood. Consequently, nicotine intake during pregnancy was found to increase orexin neurons and their projections to reward-mediating dopaminergic cells at the ventral tegmental area (VTA) [77]. Furthermore, in utero exposure to ethanol was confirmed in its function to perturb the development of orexin neurons [89,90]. Therefore, the orexin system can significantly contribute to the behavioural changes in the offspring due to maternal nicotine or ethanol intake. In addition, orexin was found to noticeably modulate neurodevelopment through enhancing the release of GABA and glutamate in the prenatal brain of offspring of pre-treated rats [204].

4. Conclusions

Neurodevelopmental disorders including SCH, ADHD, and ASD have been found to incorporate impairments in several neuromodulator systems. Theses impairments interfere with proper maturation and vital brain functions, as well as postnatal development. The cause of such impairments can be attributed to either genetic mutations, environmental insults including pharmacological agents clinically used during pregnancy, or both. However, some of these pharmacological agents can be used later in life to correct the ailment in neuromodulator systems. Overall, understanding of these neuromodulator systems from prenatal to geriatric stages is essential in searching for the possible causes and potential pharmacological interventions for the management of ASD and associated comorbidity disorders, e.g., SCH or ADHD.

Author Contributions

A.K.S. and B.S. were responsible for the study concept, design, and acquisition and analysis of data. A.K.S. and B.S. drafted the manuscript. A.A., A.M. and S.A. critically revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Office of Graduate Studies and Research of UAE University (31R233) as well as Zayed-Center for Health Sciences (12M099).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The Office of Graduate Studies and Research of UAE University as well as Zayed-Center for Health Sciences are thanked for the support provided to BS with funds (31R233 and 12M099).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

2-AG2-Arachidonoylglycerol
5-HTSerotonin
ADHDAttention deficit hyperactivity disorder
AEAAnandamide
AktAk strain transforming also known as protein kinase B
AMPAα-Amino-3-hydroxy-5-methyl-4-isoxazole propionic acid
ASD
BDNF
Autism spectrum disorder
Brain-derived neurotrophic factor
BTBRBlack and Tan Brachyury
CBRCannabinoid receptor
CSFCerebrospinal fluid
DSM
DAG
Diagnostic and statistical manual of mental disorders
Diaceylglycerol
eCBEndogenous cannabinoids
E or EDEmbryonic day
EMAEuropean Medicine Agency
EPSPExcitatory postsynaptic potential
ERKExtracellular signal-regulated kinase
FDAFood and Drug Administration in USA
GABAGamma aminobutyric acid
GDNFGlial cell line-derived neurotrophic factor
GiInhibitory G-protein
GluGlutamate
GPCGlial progenitor cell
Gs
IGF-1
Stimulatory G protein
Insulin growth factor 1
IL-6
IL-8
IL-β
IL-10
IPSCs
LC
Interleukin-6
Interleukin-8
Interleukin-beta
Interleukin-10
Inhibitory postsynaptic currents
Locus coeruleus
LTDLong-term depression
LTPLong-term potentiation
mGluR
MCP-1
Metabotropic glutamate receptor
Monocyte chemoattractant protein 1,
MIFMacrophage migration inhibitory factor
NAPE-PLDN-acyl phosphatidylethanolamine-specific phospholipase D
NMDAN-methyl-D-aspartate
NPC
NT
OXR2
Neuronal progenitor cells
Neurotensin
Orexin receptor 2
PI3KsPhosphoinositide 3-kinases
SCHSchizophrenia
SHANKSH3 and multiple ankyrin repeat domains protein
SSRIsSelective serotonin reuptake inhibitors
TGF-βTransforming growth factor beta
VTAVentral tegmental area

References

  1. ASA. Neurodevelopmental Disorders: DSM-5® Selections; American Psychiatric Publishing: Washington, DC, USA, 2015. [Google Scholar]
  2. Posar, A.; Visconti, P. Neurodevelopmental disorders between past and future. J. Pediatr. Neurosci. 2017, 12, 301–302. [Google Scholar] [CrossRef] [PubMed]
  3. Taheri, A.; Perry, A.; Factor, D.C. Brief report: A further examination of the DSM. 5 autism spectrum disorder criteria in practice. J. Dev. Disabil. 2014, 20, 116. [Google Scholar]
  4. Hwang, Y.I.; Srasuebkul, P.; Foley, K.R.; Arnold, S.; Trollor, J.N. Mortality and cause of death of Australians on the autism spectrum. Autism Res. 2019, 12, 806–815. [Google Scholar] [CrossRef] [PubMed]
  5. Ivanović, I. Psychiatric comorbidities in children with ASD: Autism Centre experience. Front. Psychiatry 2021, 12, 673169. [Google Scholar] [CrossRef] [PubMed]
  6. Ingersoll, B.; Wainer, A. The Broader Autism Phenotype. In Handbook of Autism and Pervasive Developmental Disorders, 4th ed.; Volume Diagnosis, Development, and Brain Mechanisms; Volkmar, F.R., Paul, R., Rogers, S.J., Pelphrey, K.A., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014. [Google Scholar]
  7. Hull, L.; Petrides, K.; Mandy, W. The female autism phenotype and camouflaging: A narrative review. Rev. J. Autism Dev. Disord. 2020, 7, 306–317. [Google Scholar] [CrossRef] [Green Version]
  8. McLennan, J.D.; Lord, C.; Schopler, E. Sex differences in higher functioning people with autism. J. Autism Dev. Disord. 1993, 23, 217–227. [Google Scholar] [CrossRef]
  9. Zafeiriou, D.I.; Ververi, A.; Dafoulis, V.; Kalyva, E.; Vargiami, E. Autism spectrum disorders: The quest for genetic syndromes. Am. J. Med. Genet. Part B Neuropsychiatr. Genet. 2013, 162, 327–366. [Google Scholar] [CrossRef]
  10. Grabrucker, A.M. Environmental factors in autism. Front. Psychiatry 2013, 3, 118. [Google Scholar] [CrossRef] [Green Version]
  11. Urbán, N.; Guillemot, F. Neurogenesis in the embryonic and adult brain: Same regulators, different roles. Front. Cell. Neurosci. 2014, 8, 396. [Google Scholar] [CrossRef] [Green Version]
  12. Kriegstein, A.R.; Noctor, S.C. Patterns of neuronal migration in the embryonic cortex. Trends Neurosci. 2004, 27, 392–399. [Google Scholar] [CrossRef]
  13. Miller, K.E.; Suter, D.M. An integrated cytoskeletal model of neurite outgrowth. Front. Cell. Neurosci. 2018, 12, 447. [Google Scholar] [CrossRef] [PubMed]
  14. McAllister, A.K. Dynamic Aspects of CNS Synapse Formation. Annu. Rev. Neurosci. 2007, 30, 425–450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Yoshihara, Y.; De Roo, M.; Muller, D. Dendritic spine formation and stabilization. Curr. Opin. Neurobiol. 2009, 19, 146–153. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, X.; van Praag, H. Steps towards standardized quantification of adult neurogenesis. Nat. Commun. 2020, 11, 4275. [Google Scholar] [CrossRef]
  17. Atluri, P.; Fleck, M.; Shen, Q.; Mah, S.; Stadfelt, D.; Barnes, W.; Goderie, S.; Temple, S.; Schneider, A. Functional nicotinic acetylcholine receptor expression in stem and progenitor cells of the early embryonic mouse cerebral cortex. Dev. Biol. 2001, 240, 143–156. [Google Scholar] [CrossRef]
  18. Höhmann, C.F.; Pert, C.C.; Ebner, F.F. Development of cholinergic markers in mouse forebrain. II. Muscarinic receptor binding in cortex. Dev. Brain Res. 1985, 23, 243–253. [Google Scholar] [CrossRef]
  19. Saboory, E.; Ghasemi, M.; Mehranfard, N. Norepinephrine, neurodevelopment and behavior. Neurochem. Int. 2020, 135, 104706. [Google Scholar] [CrossRef]
  20. Popolo, M.; McCarthy, D.M.; Bhide, P.G. Influence of dopamine on precursor cell proliferation and differentiation in the embryonic mouse telencephalon. Dev. Neurosci. 2004, 26, 229–244. [Google Scholar] [CrossRef] [Green Version]
  21. Herlenius, E.; Lagercrantz, H. Development of neurotransmitter systems during critical periods. Exp. Neurol. 2004, 190, 8–21. [Google Scholar] [CrossRef]
  22. Ritter, L.M.; Unis, A.S.; Meador-Woodruff, J.H. Ontogeny of ionotropic glutamate receptor expression in human fetal brain. Dev. Brain Res. 2001, 127, 123–133. [Google Scholar] [CrossRef]
  23. Hebebrand, J.; Hofmann, D.; Reichelt, R.; Schnarr, S.; Knapp, M.; Propping, P.; Födisch, H. Early ontogeny of the central benzodiazepine receptor in human embryos and fetuses. Life Sci. 1988, 43, 2127–2136. [Google Scholar] [CrossRef]
  24. Kinnunen, A.; Lintunen, M.; Karlstedt, K.; Fukui, H.; Panula, P. In situ detection of H1-receptor mRNA and absence of apoptosis in the transient histamine system of the embryonic rat brain. J. Comp. Neurol. 1998, 394, 127–137. [Google Scholar] [CrossRef]
  25. Tribollet, E.; Charpak, S.; Schmidt, A.; Dubois-Dauphin, M.; Dreifuss, J. Appearance and transient expression of oxytocin receptors in fetal, infant, and peripubertal rat brain studied by autoradiography and electrophysiology. J. Neurosci. 1989, 9, 1764–1773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Rius, R.A.; Barg, J.; Bem, W.T.; Coscia, C.J.; Loh, Y.P. The prenatal developmental profile of expression of opioid peptides and receptors in the mouse brain. Dev. Brain Res. 1991, 58, 237–241. [Google Scholar] [CrossRef] [Green Version]
  27. Steininger, T.L.; Kilduff, T.S.; Behan, M.; Benca, R.M.; Landry, C.F. Comparison of hypocretin/orexin and melanin-concentrating hormone neurons and axonal projections in the embryonic and postnatal rat brain. J. Chem. Neuroanat. 2004, 27, 165–181. [Google Scholar] [CrossRef]
  28. Ogawa, S.; Iriguchi, M.; Lee, Y.-A.; Yoshikawa, S.; Goto, Y. Atypical social rank recognition in autism spectrum disorder. Sci. Rep. 2019, 9, 15657. [Google Scholar] [CrossRef] [PubMed]
  29. Maffei, A. Long-term potentiation and long-term depression. In Oxford Research Encyclopedia of Neuroscience; Oxford University Press: Oxford, UK, 2018. [Google Scholar]
  30. Luscher, C.; Malenka, R.C. NMDA Receptor-Dependent Long-Term Potentiation and Long-Term Depression (LTP/LTD). Cold Spring Harb. Perspect. Biol. 2012, 4, a005710. [Google Scholar] [CrossRef] [Green Version]
  31. Ransohoff, R.M.; Schafer, D.; Vincent, A.; Blachère, N.E.; Bar-Or, A. Neuroinflammation: Ways in which the immune system affects the brain. Neurotherapeutics 2015, 12, 896–909. [Google Scholar] [CrossRef] [Green Version]
  32. Hendriksen, E.; van Bergeijk, D.; Oosting, R.S.; Redegeld, F.A. Mast cells in neuroinflammation and brain disorders. Neurosci. Biobehav. Rev. 2017, 79, 119–133. [Google Scholar] [CrossRef]
  33. DiSabato, D.J.; Quan, N.; Godbout, J.P. Neuroinflammation: The devil is in the details. J. Neurochem. 2016, 139, 136–153. [Google Scholar] [CrossRef] [Green Version]
  34. Meyer, U. Developmental neuroinflammation and schizophrenia. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2013, 42, 20–34. [Google Scholar] [CrossRef] [PubMed]
  35. Slezak, M.; Pfrieger, F.W. New roles for astrocytes: Regulation of CNS synaptogenesis. Trends Neurosci. 2003, 26, 531–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Franzen, R.; Bouhy, D.; Schoenen, J. Nervous system injury: Focus on the inflammatory cytokine ‘granulocyte-macrophage colony stimulating factor’. Neurosci. Lett. 2004, 361, 76–78. [Google Scholar] [CrossRef] [PubMed]
  37. Theoharides, T.C.; Zhang, B. Neuro-inflammation, blood-brain barrier, seizures and autism. J. Neuroinflamm. 2011, 8, 168. [Google Scholar] [CrossRef] [Green Version]
  38. Vargas, D.L.; Nascimbene, C.; Krishnan, C.; Zimmerman, A.W.; Pardo, C.A. Neuroglial activation and neuroinflammation in the brain of patients with autism. Ann. Neurol. Off. J. Am. Neurol. Assoc. Child Neurol. Soc. 2005, 57, 67–81. [Google Scholar] [CrossRef]
  39. Nakagawa, Y.; Chiba, K.J. Involvement of neuroinflammation during brain development in social cognitive deficits in autism spectrum disorder and schizophrenia. J. Pharmacol. Exp. Ther. 2016, 358, 504–515. [Google Scholar] [CrossRef] [Green Version]
  40. Nadeem, A.; Ahmad, S.F.; Al-Harbi, N.O.; Al-Ayadhi, L.Y.; Sarawi, W.; Attia, S.M.; Bakheet, S.A.; Alqarni, S.A.; Ali, N.; AsSobeai, H.M. Imbalance in pro-inflammatory and anti-inflammatory cytokines milieu in B cells of children with autism. Mol. Immunol. 2022, 141, 297–304. [Google Scholar] [CrossRef]
  41. Tsilioni, I.; Pantazopoulos, H.; Conti, P.; Leeman, S.E.; Theoharides, T.C. IL-38 inhibits microglial inflammatory mediators and is decreased in amygdala of children with autism spectrum disorder. Proc. Natl. Acad. Sci. USA 2020, 117, 16475–16480. [Google Scholar] [CrossRef]
  42. Zhao, H.; Zhang, H.; Liu, S.; Luo, W.; Jiang, Y.; Gao, J. Association of peripheral blood levels of cytokines with autism spectrum disorder: A meta-analysis. Front. Psychiatry 2021, 12, 1006. [Google Scholar] [CrossRef]
  43. Giacobbo, B.L.; Doorduin, J.; Klein, H.C.; Dierckx, R.A.; Bromberg, E.; de Vries, E.F. Brain-derived neurotrophic factor in brain disorders: Focus on neuroinflammation. Mol. Neurobiol. 2019, 56, 3295–3312. [Google Scholar] [CrossRef] [Green Version]
  44. Rea, K.; Dinan, T.G.; Cryan, J.F. The microbiome: A key regulator of stress and neuroinflammation. Neurobiol. Stress 2016, 4, 23–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kowiański, P.; Lietzau, G.; Czuba, E.; Waśkow, M.; Steliga, A.; Moryś, J. BDNF: A key factor with multipotent impact on brain signaling and synaptic plasticity. Cell. Mol. Neurobiol. 2018, 38, 579–593. [Google Scholar] [CrossRef] [PubMed]
  46. Monteiro, P.; Feng, G. SHANK proteins: Roles at the synapse and in autism spectrum disorder. Nat. Rev. Neurosci. 2017, 18, 147–157. [Google Scholar] [CrossRef] [PubMed]
  47. Suzuki, S.; Koshimizu, H.; Adachi, N.; Matsuoka, H.; Fushimi, S.; Ono, J.; Ohta, K.-I.; Miki, T. Functional interaction between BDNF and mGluR II in vitro: BDNF down-regulated mGluR II gene expression and an mGluR II agonist enhanced BDNF-induced BDNF gene expression in rat cerebral cortical neurons. Peptides 2017, 89, 42–49. [Google Scholar] [CrossRef]
  48. Jaggar, M.; Vaidya, V.A. 5-HT 2A receptors and BDNF regulation: Implications for psychopathology. In 5-HT2A Receptors in the Central Nervous System; Springer: Berlin/Heidelberg, Germany, 2018; pp. 395–438. [Google Scholar]
  49. Trobiani, L.; Meringolo, M.; Diamanti, T.; Bourne, Y.; Marchot, P.; Martella, G.; Dini, L.; Pisani, A.; De Jaco, A.; Bonsi, P. The neuroligins and the synaptic pathway in Autism Spectrum Disorder. Neurosci. Biobehav. Rev. 2020, 119, 37–51. [Google Scholar] [CrossRef]
  50. Glaser, T.; Sampaio, V.F.A.; Lameu, C.; Ulrich, H. Calcium signalling: A common target in neurological disorders and neurogenesis. Semin. Cell Dev. Biol. 2019, 95, 25–33. [Google Scholar] [CrossRef]
  51. Konur, S.; Ghosh, A. Calcium signaling and the control of dendritic development. Neuron 2005, 46, 401–405. [Google Scholar] [CrossRef] [Green Version]
  52. Lohmann, C. Calcium signaling and the development of specific neuronal connections. Prog. Brain Res. 2009, 175, 443–452. [Google Scholar] [CrossRef]
  53. Brigadski, T.; Leßmann, V. The physiology of regulated BDNF release. Cell Tissue Res. 2020, 382, 15–45. [Google Scholar] [CrossRef]
  54. Nguyen, L.; Rigo, J.-M.; Rocher, V.; Belachew, S.; Malgrange, B.; Rogister, B.; Leprince, P.; Moonen, G. Neurotransmitters as early signals for central nervous system development. Cell Tissue Res. 2001, 305, 187–202. [Google Scholar] [CrossRef]
  55. Herlenius, E.; Lagercrantz, H. Neurotransmitters and neuromodulators during early human development. Early Hum. Dev. 2001, 65, 21–37. [Google Scholar] [CrossRef]
  56. Medvedeva, V.P.; Pierani, A. How do electric fields coordinate neuronal migration and maturation in the developing cortex? Front. Cell Dev. Biol. 2020, 8, 580657. [Google Scholar] [CrossRef] [PubMed]
  57. Choudhury, P.R.; Lahiri, S.; Rajamma, U. Glutamate mediated signaling in the pathophysiology of autism spectrum disorders. Pharmacol. Biochem. Behav. 2012, 100, 841–849. [Google Scholar] [CrossRef] [PubMed]
  58. Peça, J.; Feliciano, C.; Ting, J.T.; Wang, W.; Wells, M.F.; Venkatraman, T.N.; Lascola, C.D.; Fu, Z.; Feng, G. Shank3 mutant mice display autistic-like behaviours and striatal dysfunction. Nature 2011, 472, 437–442. [Google Scholar] [CrossRef] [Green Version]
  59. Bariselli, S.; Tzanoulinou, S.; Glangetas, C.; Prévost-Solié, C.; Pucci, L.; Viguié, J.; Bezzi, P.; O’Connor, E.C.; Georges, F.; Lüscher, C. SHANK3 controls maturation of social reward circuits in the VTA. Nat. Neurosci. 2016, 19, 926–934. [Google Scholar] [CrossRef] [Green Version]
  60. Herbert, M.R. SHANK3, the synapse, and autism. N. Engl. J. Med. 2011, 365, 173–175. [Google Scholar] [CrossRef] [Green Version]
  61. Zimmerman, R.; Smith, A.; Fech, T.; Mansour, Y.; Kulesza, R.J. In utero exposure to valproic acid disrupts ascending projections to the central nucleus of the inferior colliculus from the auditory brainstem. Exp. Brain Res. 2020, 238, 551–563. [Google Scholar] [CrossRef]
  62. Lenart, J.; Augustyniak, J.; Lazarewicz, J.W.; Zieminska, E. Altered expression of glutamatergic and GABAergic genes in the valproic acid-induced rat model of autism: A screening test. Toxicology 2020, 440, 152500. [Google Scholar] [CrossRef]
  63. Silvestrin, R.B.; Bambini-Junior, V.; Galland, F.; Bobermim, L.D.; Quincozes-Santos, A.; Abib, R.T.; Zanotto, C.; Batassini, C.; Brolese, G.; Gonçalves, C.-A. Animal model of autism induced by prenatal exposure to valproate: Altered glutamate metabolism in the hippocampus. Brain Res. 2013, 1495, 52–60. [Google Scholar] [CrossRef] [Green Version]
  64. Miyazaki, K.; Narita, N.; Narita, M. Maternal administration of thalidomide or valproic acid causes abnormal serotonergic neurons in the offspring: Implication for pathogenesis of autism. Int. J. Dev. Neurosci. 2005, 23, 287–297. [Google Scholar] [CrossRef]
  65. Dufour-Rainfray, D.; Vourc’h, P.; Le Guisquet, A.-M.; Garreau, L.; Ternant, D.; Bodard, S.; Jaumain, E.; Gulhan, Z.; Belzung, C.; Andres, C.R. Behavior and serotonergic disorders in rats exposed prenatally to valproate: A model for autism. Neurosci. Lett. 2010, 470, 55–59. [Google Scholar] [CrossRef] [PubMed]
  66. Dai, Y.-C.; Zhang, H.-F.; Schön, M.; Böckers, T.M.; Han, S.-P.; Han, J.-S.; Zhang, R. Neonatal oxytocin treatment ameliorates autistic-like behaviors and oxytocin deficiency in valproic acid-induced rat model of autism. Front. Cell. Neurosci. 2018, 12, 355. [Google Scholar] [CrossRef] [PubMed]
  67. Kellogg, C.K.; Primus, R.J.; Bitran, D.J. Sexually dimorphic influence of prenatal exposure to diazepam on behavioral responses to environmental challenge and on gamma-aminobutyric acid (GABA)-stimulated chloride uptake in the brain. J. Pharmacol. Exp. Ther. 1991, 256, 259–265. [Google Scholar] [PubMed]
  68. Abekawa, T.; Ito, K.; Nakagawa, S.; Koyama, T. Prenatal exposure to an NMDA receptor antagonist, MK-801 reduces density of parvalbumin-immunoreactive GABAergic neurons in the medial prefrontal cortex and enhances phencyclidine-induced hyperlocomotion but not behavioral sensitization to methamphetamine in postpubertal rats. Psychopharmacology 2007, 192, 303–316. [Google Scholar] [CrossRef]
  69. Alonazi, M.A.; El Gezeery, A.; El-Ansary, A.; Bhat, R.S. Neurochemical changes in the brain after fetal exposure to fluoxetine, a serotonin reuptake inhibitor (SRI) in rodents. Curr. Proteom. 2021, 18, 499–504. [Google Scholar] [CrossRef]
  70. Lawrence, R.H.; Palumbo, M.C.; Freeman, S.M.; Guoynes, C.D.; Bales, K.L. Developmental fluoxetine exposure alters behavior and neuropeptide receptors in the prairie vole. Front. Behav. Neurosci. 2020, 14, 215. [Google Scholar] [CrossRef]
  71. Rosengarten, H.; Quartermain, D. Effect of prenatal administration of haloperidol, risperidone, quetiapine and olanzapine on spatial learning and retention in adult rats. Pharmacol. Biochem. Behav. 2002, 72, 575–579. [Google Scholar] [CrossRef]
  72. Previc, F.H. Prenatal influences on brain dopamine and their relevance to the rising incidence of autism. Med. Hypotheses 2007, 68, 46–60. [Google Scholar] [CrossRef]
  73. Zhao, T.; Chen, Y.; Sun, Z.; Shi, Z.; Qin, J.; Lu, J.; Li, C.; Ma, D.; Zhou, L.; Song, X. Prenatal sevoflurane exposure causes neuronal excitatory/inhibitory imbalance in the prefrontal cortex and neurofunctional abnormality in rats. Neurobiol. Dis. 2020, 146, 105121. [Google Scholar] [CrossRef]
  74. Borges, S.; Coimbra, B.; Soares-Cunha, C.; Pêgo, J.M.; Sousa, N.; Rodrigues, A.J. Dopaminergic modulation of affective and social deficits induced by prenatal glucocorticoid exposure. Neuropsychopharmacology 2013, 38, 2068–2079. [Google Scholar] [CrossRef] [Green Version]
  75. Alkam, T.; Mamiya, T.; Kimura, N.; Yoshida, A.; Kihara, D.; Tsunoda, Y.; Aoyama, Y.; Hiramatsu, M.; Kim, H.-C.; Nabeshima, T.J.P. Prenatal nicotine exposure decreases the release of dopamine in the medial frontal cortex and induces atomoxetine-responsive neurobehavioral deficits in mice. Psychopharmacology 2017, 234, 1853–1869. [Google Scholar] [CrossRef] [PubMed]
  76. Dwyer, J.B.; Cardenas, A.; Franke, R.M.; Chen, Y.; Bai, Y.; Belluzzi, J.D.; Lotfipour, S.; Leslie, F.M. Prenatal nicotine sex-dependently alters adolescent dopamine system development. Transl. Psychiatry 2019, 9, 304. [Google Scholar] [CrossRef] [PubMed]
  77. Morgan, A.J.; Harrod, S.B.; Lacy, R.T.; Stanley, E.M.; Fadel, J.R. Intravenous prenatal nicotine exposure increases orexin expression in the lateral hypothalamus and orexin innervation of the ventral tegmental area in adult male rats. Drug Alcohol Depend. 2013, 132, 562–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Stanwood, G.; Washington, R.; Shumsky, J.; Levitt, P. Prenatal cocaine exposure produces consistent developmental alterations in dopamine-rich regions of the cerebral cortex. Neuroscience 2001, 106, 5–14. [Google Scholar] [CrossRef]
  79. Chai, L.; Choi, W.S.; Rönnekleiv, O.K. Maternal cocaine treatment alters dynorphin and enkephalin mRNA expression in brains of fetal rhesus macaques. J. Neurosci. Res. 1997, 17, 1112. [Google Scholar] [CrossRef]
  80. Chai, L.; Bosch, M.A.; Moore, J.M.; Rønnekleiv, O.K. Chronic prenatal cocaine treatment down-regulates μ-opioid receptor mRNA expression in the brain of fetal Rhesus Macaque. Neurosci. Lett. 1999, 261, 45–48. [Google Scholar] [CrossRef]
  81. Kim, P.; Park, J.H.; Choi, C.S.; Choi, I.; Joo, S.H.; Kim, M.K.; Kim, S.Y.; Kim, K.C.; Park, S.H.; Kwon, K.J. Effects of ethanol exposure during early pregnancy in hyperactive, inattentive and impulsive behaviors and MeCP2 expression in rodent offspring. Neurochem. Res. 2013, 38, 620–631. [Google Scholar] [CrossRef]
  82. Carneiro, L.M.; Diógenes, J.P.L.; Vasconcelos, S.M.; Aragão, G.F.; Noronha, E.C.; Gomes, P.B.; Viana, G.S.J. Behavioral and neurochemical effects on rat offspring after prenatal exposure to ethanol. Neurotoxicology Teratol. 2005, 27, 585–592. [Google Scholar] [CrossRef]
  83. Facciol, A.; Bailleul, C.; Nguyen, S.; Chatterjee, D.; Gerlai, R. Developmental stage-dependent deficits induced by embryonic ethanol exposure in zebrafish: A neurochemical analysis. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2020, 99, 109859. [Google Scholar] [CrossRef]
  84. Waddell, J.; Mooney, S.M. Choline and working memory training improve cognitive deficits caused by prenatal exposure to ethanol. Nutrients 2017, 9, 1080. [Google Scholar] [CrossRef]
  85. Smiley, J.F.; Bleiwas, C.; Canals-Baker, S.; Williams, S.Z.; Sears, R.; Teixeira, C.M.; Wilson, D.A.; Saito, M. Neonatal ethanol causes profound reduction of cholinergic cell number in the basal forebrain of adult animals. Alcohol 2021, 97, 1–11. [Google Scholar] [CrossRef] [PubMed]
  86. Holman, P.J.; Ellis, L.; Morgan, E.; Weinberg, J. Prenatal alcohol exposure disrupts male adolescent social behavior and oxytocin receptor binding in rodents. Horm. Behav. 2018, 105, 115–127. [Google Scholar] [CrossRef] [PubMed]
  87. He, F.-Q.; Zhang, J.; Guo, X. Prenatal ethanol exposure increases depressive-like behavior and central estrogen receptor α and oxytocin expressions in adult female Mandarin Voles. Zool. Stud. 2012, 51, 1–11. [Google Scholar]
  88. Arias, C.; Chotro, M.G. Increased palatability of ethanol after prenatal ethanol exposure is mediated by the opioid system. Pharmacol. Biochem. Behav. 2005, 82, 434–442. [Google Scholar] [CrossRef] [PubMed]
  89. Collier, A.D.; Halkina, V.; Min, S.S.; Roberts, M.Y.; Campbell, S.D.; Camidge, K.; Leibowitz, S.F. Embryonic ethanol exposure affects the early development, migration, and location of hypocretin/orexin neurons in zebrafish. Alcohol. Clin. Exp. Res. 2019, 43, 1702–1713. [Google Scholar] [CrossRef] [PubMed]
  90. Collier, A.D.; Yasmin, N.; Khalizova, N.; Campbell, S.; Onoichenco, A.; Fam, M.; Albeg, A.S.; Leibowitz, S.F. Sexually dimorphic and asymmetric effects of embryonic ethanol exposure on hypocretin/orexin neurons as related to behavioral changes in zebrafish. Sci. Rep. 2021, 11, 16078. [Google Scholar] [CrossRef]
  91. Di Cristo, G. Development of cortical GABAergic circuits and its implications for neurodevelopmental disorders. Clin. Genet. 2007, 72, 1–8. [Google Scholar] [CrossRef]
  92. De Caro, C.; Iannone, L.F.; Citraro, R.; Striano, P.; De Sarro, G.; Constanti, A.; Cryan, J.F.; Russo, E. Can we ‘seize’ the gut microbiota to treat epilepsy? Neurosci. Biobehav. Rev. 2019, 107, 750–764. [Google Scholar] [CrossRef]
  93. Gatto, C.L.; Broadie, K. Genetic controls balancing excitatory and inhibitory synaptogenesis in neurodevelopmental disorder models. Front. Synaptic Neurosci. 2010, 2, 4. [Google Scholar] [CrossRef] [Green Version]
  94. Miranda-Contrerasa, P.B.-D.L.; Mendoza-Briceñoa, R.V.; Palacios-Prüa, Z.P.-C.E. Prenatal and postnatal contents of amino acid neurotransmitters in mouse parietal cortex. Dev. Neurosci. 2003, 25, 366–374. [Google Scholar]
  95. Luján, R.; Shigemoto, R.; López-Bendito, G. Glutamate and GABA receptor signalling in the developing brain. Neuroscience 2005, 130, 567–580. [Google Scholar] [CrossRef] [PubMed]
  96. Bell, S.; Maussion, G.; Jefri, M.; Peng, H.; Theroux, J.-F.; Silveira, H.; Soubannier, V.; Wu, H.; Hu, P.; Galat, E. Disruption of GRIN2B impairs differentiation in human neurons. Stem Cell Rep. 2018, 11, 183–196. [Google Scholar] [CrossRef] [PubMed]
  97. Roullet, F.I.; Lai, J.K.; Foster, J.A. In utero exposure to valproic acid and autism—a current review of clinical and animal studies. Neurotoxicol. Teratol. 2013, 36, 47–56. [Google Scholar] [CrossRef] [PubMed]
  98. van Gelder, M.M.; Bos, J.H.; Roeleveld, N.; de Jong-van den Berg, L.T.W. Drugs associated with teratogenic mechanisms. Part I: Dispensing rates among pregnant women in the Netherlands, 1998–2009. Hum. Reprod. 2014, 29, 161–167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Gandhi, R.; Luk, K.C.; Rymar, V.V.; Sadikot, A.F. Group I mGluR5 metabotropic glutamate receptors regulate proliferation of neuronal progenitors in specific forebrain developmental domains. J. Neurochem. 2008, 104, 155–172. [Google Scholar] [CrossRef]
  100. Gogolla, N.; LeBlanc, J.J.; Quast, K.B.; Südhof, T.C.; Fagiolini, M.; Hensch, T.K. Common circuit defect of excitatory-inhibitory balance in mouse models of autism. J. Neurodev. Disord. 2009, 1, 172–181. [Google Scholar] [CrossRef] [Green Version]
  101. Olloquequi, J.; Cornejo-Córdova, E.; Verdaguer, E.; Soriano, F.X.; Binvignat, O.; Auladell, C.; Camins, A. Excitotoxicity in the pathogenesis of neurological and psychiatric disorders: Therapeutic implications. J. Psychopharmacol. 2018, 32, 265–275. [Google Scholar] [CrossRef] [PubMed]
  102. Essa, M.; Braidy, N.; Vijayan, K.; Subash, S.; Guillemin, G.J. Excitotoxicity in the pathogenesis of autism. Neurotox. Res. 2013, 23, 393–400. [Google Scholar] [CrossRef]
  103. Viviani, B.; Boraso, M.; Marchetti, N.; Marinovich, M. Perspectives on neuroinflammation and excitotoxicity: A neurotoxic conspiracy? Neurotoxicology 2014, 43, 10–20. [Google Scholar] [CrossRef]
  104. Nakamoto, M.; Nalavadi, V.; Epstein, M.P.; Narayanan, U.; Bassell, G.J.; Warren, S.T. Fragile X mental retardation protein deficiency leads to excessive mGluR5-dependent internalization of AMPA receptors. Proc. Natl. Acad. Sci. USA 2007, 104, 15537–15542. [Google Scholar] [CrossRef] [Green Version]
  105. Bovetti, S.; Gribaudo, S.; Puche, A.C.; De Marchis, S.; Fasolo, A. From progenitors to integrated neurons: Role of neurotransmitters in adult olfactory neurogenesis. J. Chem. Neuroanat. 2011, 42, 304–316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Martin, H.G.S.; Manzoni, O. Late onset deficits in synaptic plasticity in the valproic acid rat model of autism. Front. Cell. Neurosci. 2014, 8, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Barnes, J.R.; Mukherjee, B.; Rogers, B.C.; Nafar, F.; Gosse, M.; Parsons, M.P. The relationship between glutamate dynamics and activity-dependent synaptic plasticity. J. Neurosci. 2020, 40, 2793–2807. [Google Scholar] [CrossRef]
  108. Kehrer, C.; Maziashvili, N.; Dugladze, T.; Gloveli, T. Altered excitatory-inhibitory balance in the NMDA-hypofunction model of schizophrenia. Front. Mol. Neurosci. 2008, 1, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Lahti, A.C.; Weiler, M.A.; Michaelidis, B.T.; Parwani, A.; Tamminga, C.A. Effects of ketamine in normal and schizophrenic volunteers. Neuropsychopharmacology 2001, 25, 455–467. [Google Scholar] [CrossRef] [Green Version]
  110. Jacob, T.C.; Wan, Q.; Vithlani, M.; Saliba, R.S.; Succol, F.; Pangalos, M.N.; Moss, S.J. GABAA receptor membrane trafficking regulates spine maturity. Proc. Natl. Acad. Sci. USA 2009, 106, 12500–12505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Rajamani, K.T.; Wagner, S.; Grinevich, V.; Harony-Nicolas, H. Oxytocin as a Modulator of Synaptic Plasticity: Implications for Neurodevelopmental Disorders. Front. Synaptic Neurosci. 2018, 10, 17. [Google Scholar] [CrossRef] [Green Version]
  112. Arion, D.; Lewis, D.A. Altered expression of regulators of the cortical chloride transporters NKCC1 and KCC2 in Schizophrenia. Arch. Gen. Psychiatry 2011, 68, 21. [Google Scholar] [CrossRef]
  113. Hyde, T.M.; Lipska, B.K.; Ali, T.; Mathew, S.V.; Law, A.J.; Metitiri, O.E.; Straub, R.E.; Ye, T.; Colantuoni, C.; Herman, M.M.; et al. Expression of GABA signaling molecules KCC2, NKCC1, and GAD1 in cortical development and schizophrenia. J. Neurosci. 2011, 31, 11088–11095. [Google Scholar] [CrossRef] [Green Version]
  114. Mabunga, D.F.N.; Gonzales, E.L.T.; Kim, J.-W.; Kim, K.C.; Shin, C.Y. Exploring the validity of valproic acid animal model of autism. Exp. Neurobiol. 2015, 24, 285. [Google Scholar] [CrossRef] [Green Version]
  115. Plioplys, A.V. Autism: Electroencephalogram abnormalities and clinical improvement with valproic acid. Arch. Pediatrics Adolesc. Med. 1994, 148, 220–222. [Google Scholar] [CrossRef] [PubMed]
  116. Money, K.; Stanwood, G. Developmental origins of brain disorders: Roles for dopamine. Front. Cell. Neurosci. 2013, 7, 260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Ozawa, K.; Hashimoto, K.; Kishimoto, T.; Shimizu, E.; Ishikura, H.; Iyo, M. Immune activation during pregnancy in mice leads to dopaminergic hyperfunction and cognitive impairment in the offspring: A neurodevelopmental animal model of schizophrenia. Biol. Psychiatry 2006, 59, 546–554. [Google Scholar] [CrossRef]
  118. Vuillermot, S.; Weber, L.; Feldon, J.; Meyer, U. A longitudinal examination of the neurodevelopmental impact of prenatal immune activation in mice reveals primary defects in dopaminergic development relevant to schizophrenia. J. Neurosci. Methods 2010, 30, 1270–1287. [Google Scholar] [CrossRef] [PubMed]
  119. Spencer, G.; Klumperman, J.; Syed, N. Neurotransmitters and neurodevelopment. Role of dopamine in neurite outgrowth, target selection and specific synapse formation. Perspect. Dev. Neurobiol. 1998, 5, 451–467. [Google Scholar]
  120. Areal, L.B.; Blakely, R.D. Neurobehavioral changes arising from early life dopamine signaling perturbations. Neurochem. Int. 2020, 137, 104747. [Google Scholar] [CrossRef] [PubMed]
  121. Azmitia, E. Serotonin neurons, neuroplasticity, and homeostasis of neural tissue. Neuropsychopharmacology 1999, 21, S33–S45. [Google Scholar] [CrossRef]
  122. Lesch, K.-P.; Waider, J. Serotonin in the modulation of neural plasticity and networks: Implications for neurodevelopmental disorders. Neuron 2012, 76, 175–191. [Google Scholar] [CrossRef] [Green Version]
  123. Kepser, L.-J.; Homberg, J.R. The neurodevelopmental effects of serotonin: A behavioural perspective. Behav. Brain Res. 2015, 277, 3–13. [Google Scholar] [CrossRef]
  124. Chugani, D.C. Role of altered brain serotonin mechanisms in autism. Mol. Psychiatry 2002, 7, S16–S17. [Google Scholar] [CrossRef] [Green Version]
  125. Tang, H.W.; Huel, G.; Campagna, D.; Hellier, G.; Boissinot, C.; Blot, P. Neurodevelopmental evaluation of 9-month-old infants exposed to low levels of lead in utero: Involvement of monoamine neurotransmitters. J. Appl. Toxicol. 1999, 19, 167–172. [Google Scholar] [CrossRef]
  126. Eissa, N.; Al-Houqani, M.; Sadeq, A.; Ojha, S.K.; Sasse, A.; Sadek, B. Current enlightenment about etiology and pharmacological treatment of autism spectrum disorder. Front. Neurosci. 2018, 12, 304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Eissa, N.; Jayaprakash, P.; Azimullah, S.; Ojha, S.K.; Al-Houqani, M.; Jalal, F.Y.; Łażewska, D.; Kieć-Kononowicz, K.; Sadek, B. The histamine H3R antagonist DL77 attenuates autistic behaviors in a prenatal valproic acid-induced mouse model of autism. Sci. Rep. 2018, 8, 13077. [Google Scholar] [CrossRef] [PubMed]
  128. Lee, J.; Avramets, D.; Jeon, B.; Choo, H. Modulation of serotonin receptors in neurodevelopmental disorders: Focus on 5-HT7 receptor. Molecules 2021, 26, 3348. [Google Scholar] [CrossRef]
  129. Court, J.; Perry, E.; Spurden, D.; Griffiths, M.; Kerwin, J.; Morris, C.; Johnson, M.; Oakley, A.; Birdsall, N.; Clementi, F. The role of the cholinergic system in the development of the human cerebellum. Dev. Brain Res. 1995, 90, 159–167. [Google Scholar] [CrossRef]
  130. Abreu-Villaça, Y.; Filgueiras, C.C.; Manhães, A.C. Developmental aspects of the cholinergic system. Behav. Brain Res. 2011, 221, 367–378. [Google Scholar] [CrossRef]
  131. Bruel-Jungerman, E.; Lucassen, P.J.; Francis, F. Cholinergic influences on cortical development and adult neurogenesis. Behav. Brain Res. 2011, 221, 379–388. [Google Scholar] [CrossRef]
  132. Deutsch, S.I.; Burket, J.A. An evolving therapeutic rationale for targeting the α 7 nicotinic acetylcholine receptor in autism spectrum disorder. In Behavioral Pharmacology of the Cholinergic System; Springer: Berlin/Heidelberg, Germany, 2020; pp. 167–208. [Google Scholar]
  133. Poppi, L.A.; Ho-Nguyen, K.T.; Shi, A.; Daut, C.T.; Tischfield, M.A. Recurrent implication of striatal cholinergic interneurons in a range of neurodevelopmental, neurodegenerative, and neuropsychiatric disorders. Cells 2021, 10, 907. [Google Scholar] [CrossRef]
  134. Adams, C.E.; Stevens, K.E. Evidence for a role of nicotinic acetylcholine receptors in schizophrenia. Front. Biosci. 2007, 12, 55–4772. [Google Scholar] [CrossRef] [Green Version]
  135. Artoni, P.; Piffer, A.; Vinci, V.; Leblanc, J.; Nelson, C.A.; Hensch, T.K.; Fagiolini, M. Deep learning of spontaneous arousal fluctuations detects early cholinergic defects across neurodevelopmental mouse models and patients. Proc. Natl. Acad. Sci. USA 2020, 117, 23298–23303. [Google Scholar] [CrossRef] [Green Version]
  136. Mostafalou, S.; Abdollahi, M. The link of organophosphorus pesticides with neurodegenerative and neurodevelopmental diseases based on evidence and mechanisms. Toxicology 2018, 409, 44–52. [Google Scholar] [CrossRef] [PubMed]
  137. Ni, L.; Acevedo, G.; Muralidharan, B.; Padala, N.; To, J.; Jonakait, G.M. Toll-like receptor ligands and CD154 stimulate microglia to produce a factor(s) that promotes excess cholinergic differentiation in the developing rat basal forebrain: Implications for neurodevelopmental disorders. Pediatric Res. 2007, 61, 15–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Vivanti, G.; Nuske, H.J. Autism, attachment, and social learning: Three challenges and a way forward. Behav. Brain Res. 2017, 325, 251–259. [Google Scholar] [CrossRef] [PubMed]
  139. Stowell, R.D.; Sipe, G.O.; Dawes, R.P.; Batchelor, H.N.; Lordy, K.A.; Whitelaw, B.S.; Stoessel, M.B.; Bidlack, J.M.; Brown, E.; Sur, M. Noradrenergic signaling in the wakeful state inhibits microglial surveillance and synaptic plasticity in the mouse visual cortex. Nat. Neurosci. 2019, 22, 1782–1792. [Google Scholar] [CrossRef] [PubMed]
  140. Eissa, N.; Sadeq, A.; Sasse, A.; Sadek, B. Role of Neuroinflammation in Autism Spectrum Disorder and the Emergence of Brain Histaminergic System. Lessons Also for BPSD? Front. Pharmacol. 2020, 11, 886. [Google Scholar] [CrossRef]
  141. Haas, H.L.; Sergeeva, O.A.; Selbach, O. Histamine in the nervous system. Physiol. Rev. 2008, 88, 1183–1241. [Google Scholar] [CrossRef]
  142. Carthy, E.; Ellender, T. Histamine, neuroinflammation and neurodevelopment: A review. Front. Neurosci. 2021, 15, 680214. [Google Scholar] [CrossRef]
  143. Baronio, D.; Puttonen, H.A.; Sundvik, M.; Semenova, S.; Lehtonen, E.; Panula, P. Embryonic exposure to valproic acid affects the histaminergic system and the social behaviour of adult zebrafish (Danio rerio). Br. J. Pharmacol. 2018, 175, 797–809. [Google Scholar] [CrossRef] [Green Version]
  144. Mahmood, D. Histamine H3 receptors and its antagonism as a novel mechanism for antipsychotic effect: A current preclinical & clinical perspective. Int. J. Health Sci. 2016, 10, 564. [Google Scholar]
  145. Eissa, N.; Venkatachalam, K.; Jayaprakash, P.; Falkenstein, M.; Dubiel, M.; Frank, A.; Reiner-Link, D.; Stark, H.; Sadek, B. The multi-targeting ligand ST-2223 with histamine H3 receptor and dopamine D2/D3 receptor antagonist properties mitigates autism-like repetitive behaviors and brain oxidative stress in mice. Int. J. Mol. Sci. 2021, 22, 1947. [Google Scholar] [CrossRef]
  146. Sadek, B.; Saad, A.; Subramanian, D.; Shafiullah, M.; Lazewska, D.; Kiec-Kononowiczc, K. Anticonvulsant and procognitive properties of the non-imidazole histamine H3 receptor antagonist DL77 in male adult rats. Neuropharmacology 2016, 106, 46–55. [Google Scholar] [CrossRef] [PubMed]
  147. Brown, J.W.; Whitehead, C.A.; Basso, A.M.; Rueter, L.E.; Zhang, M. Preclinical evaluation of non-imidazole histamine H3 receptor antagonists in comparison to atypical antipsychotics for the treatment of cognitive deficits associated with schizophrenia. Int. J. Neuropsychopharmacol. 2013, 16, 889–904. [Google Scholar] [CrossRef] [PubMed]
  148. Ellenbroek, B.A.; Ghiabi, B. The other side of the histamine H3 receptor. Trends Neurosci. 2014, 37, 191–199. [Google Scholar] [CrossRef] [PubMed]
  149. Bhowmik, M.; Khanam, R.; Vohora, D. Histamine H3receptor antagonists in relation to epilepsy and neurodegeneration: A systemic consideration of recent progress and perspectives. Br. J. Pharmacol. 2012, 167, 1398–1414. [Google Scholar] [CrossRef] [Green Version]
  150. Cowan, M.; Petri, W.A., Jr. Microglia: Immune regulators of neurodevelopment. Front. Immunol. 2018, 9, 2576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Zhang, W.; Zhang, X.; Zhang, Y.; Qu, C.; Zhou, X.; Zhang, S. Histamine induces microglia activation and the release of proinflammatory mediators in rat brain via H1R or H4R. J. Neuroimmune Pharmacol. 2020, 15, 280–291. [Google Scholar] [CrossRef]
  152. Ahmad, S.F.; Nadeem, A.; Ansari, M.A.; Bakheet, S.A.; Al-Mazroua, H.A.; Khan, M.R.; Alasmari, A.F.; Alanazi, W.A.; Sobeai, H.M.A.; Attia, S.M. The histamine-4 receptor antagonist JNJ7777120 prevents immune abnormalities by inhibiting RORγt/T-bet transcription factor signaling pathways in BTBR T+ Itpr3tf/J mice exposed to gamma rays. Mol. Immunol. 2019, 114, 561–570. [Google Scholar] [CrossRef]
  153. Monks, D.; Palanisamy, A. Oxytocin: At birth and beyond. A systematic review of the long-term effects of peripartum oxytocin. Anaesthesia 2021, 76, 1526–1537. [Google Scholar] [CrossRef]
  154. Tamborski, S.; Mintz, E.; Caldwell, H. Sex differences in the embryonic development of the central oxytocin system in mice. J. Neuroendocrinol. 2016, 28, 1–7. [Google Scholar] [CrossRef]
  155. Wu, Z.; Xie, C.; Kuang, H.; Wu, J.; Chen, X.; Liu, H.; Liu, T. Oxytocin mediates neuroprotection against hypoxic-ischemic injury in hippocampal CA1 neuron of neonatal rats. Neuropharmacology 2021, 187, 108488. [Google Scholar] [CrossRef]
  156. El Falougy, H.; Filova, B.; Ostatnikova, D.; Bacova, Z.; Bakos, J. Neuronal morphology alterations in autism and possible role of oxytocin. Endocr. Regul. 2019, 53, 46–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Filova, B.; Reichova, A.; Zatkova, M.; Srancikova, A.; Bukatova, S.; Bacova, Z.; Bakos, J. Expression of synaptic proteins in the hippocampus is modulated by neonatal oxytocin treatment. Neurosci. Lett. 2020, 725, 134912. [Google Scholar] [CrossRef] [PubMed]
  158. Mairesse, J.; Zinni, M.; Pansiot, J.; Hassan-Abdi, R.; Demene, C.; Colella, M.; Charriaut-Marlangue, C.; Novais, A.R.B.; Tanter, M.; Maccari, S. Oxytocin receptor agonist reduces perinatal brain damage by targeting microglia. Glia 2019, 67, 345–359. [Google Scholar] [CrossRef] [PubMed]
  159. Pekarek, B.T.; Hunt, P.J.; Arenkiel, B.R. Oxytocin and sensory network plasticity. Front. Neurosci. 2020, 14, 30. [Google Scholar] [CrossRef] [PubMed]
  160. Aulino, E.A.; Caldwell, H.K. Pharmacological manipulation of oxytocin receptor signaling during mouse embryonic development results in sex-specific behavioral effects in adulthood. Horm. Behav. 2021, 135, 105026. [Google Scholar] [CrossRef]
  161. Nunes, A.R.; Gliksberg, M.; Varela, S.A.; Teles, M.; Wircer, E.; Blechman, J.; Petri, G.; Levkowitz, G.; Oliveira, R.F. Developmental effects of oxytocin neurons on social affiliation and processing of social information. J. Neurosci. 2021, 41, 8742–8760. [Google Scholar] [CrossRef]
  162. Williams, A.V.; Duque-Wilckens, N.; Ramos-Maciel, S.; Campi, K.L.; Bhela, S.K.; Xu, C.K.; Jackson, K.; Chini, B.; Pesavento, P.A.; Trainor, B.C. Social approach and social vigilance are differentially regulated by oxytocin receptors in the nucleus accumbens. Neuropsychopharmacology 2020, 45, 1423–1430. [Google Scholar] [CrossRef]
  163. Fineberg, S.K.; Ross, D.A. Oxytocin and the social brain. Biol. Psychiatry 2017, 81, e19. [Google Scholar] [CrossRef] [Green Version]
  164. Siu, M.T.; Goodman, S.J.; Yellan, I.; Butcher, D.T.; Jangjoo, M.; Grafodatskaya, D.; Rajendram, R.; Lou, Y.; Zhang, R.; Zhao, C. DNA methylation of the oxytocin receptor across neurodevelopmental disorders. J. Autism Dev. Disord. 2021, 51, 3610–3623. [Google Scholar] [CrossRef]
  165. Moerkerke, M.; Bonte, M.-L.; Daniels, N.; Chubar, V.; Alaerts, K.; Steyaert, J.; Boets, B. Oxytocin receptor gene (OXTR) DNA methylation is associated with autism and related social traits–A systematic review. Res. Autism Spectr. Disord. 2021, 85, 101785. [Google Scholar] [CrossRef]
  166. Boggess, T.; Risher, W.C. Clinical and basic research investigations into the long-term effects of prenatal opioid exposure on brain development. J. Neurosci. Res. 2020, 100, 396–409. [Google Scholar] [CrossRef] [PubMed]
  167. Méndez, M.; Hernández-Fonseca, K.; Abate, P. Prenatal ethanol exposure and enkephalinergic neurotransmission. In Opioid Hormones; Litwack, G., Ed.; Elsevier Science: Amsterdam, The Netherlands, 2019; Volume 111, pp. 313–337. [Google Scholar]
  168. Nummenmaa, L.; Tuominen, L. Opioid system and human emotions. Br. J. Pharmacol. 2018, 175, 2737–2749. [Google Scholar] [CrossRef] [PubMed]
  169. Georges, F.; Normand, E.; Bloch, B.; Le Moine, C. Opioid receptor gene expression in the rat brain during ontogeny, with special reference to the mesostriatal system: An in situ hybridization study. Dev. Brain Res. 1998, 109, 187–199. [Google Scholar] [CrossRef]
  170. Pellissier, L.P.; Gandía, J.; Laboute, T.; Becker, J.A.; Le Merrer, J. μ opioid receptor, social behaviour and autism spectrum disorder: Reward matters. Br. J. Pharmacol. 2018, 175, 2750–2769. [Google Scholar] [CrossRef]
  171. Tarnowska, K.; Gruczyńska–Sękowska, E.; Kowalska, D.; Majewska, E.; Kozłowska, M.; Winkler, R. The opioid excess theory in autism spectrum disorders-is it worth investigating further? Crit. Rev. Food Sci. Nutr. 2021, 1–14. [Google Scholar] [CrossRef]
  172. Anugu, V.; Ringhisen, J.; Johnson, B. Case Report: Cause and treatment of “high opioid tone” autism. Front. Psychol. 2021, 12, 657952. [Google Scholar] [CrossRef]
  173. Roy, A.; Roy, M.; Deb, S.; Unwin, G.; Roy, A. Are opioid antagonists effective in attenuating the core symptoms of autism spectrum conditions in children: A systematic review. J. Intellect. Disabil. Res. 2015, 59, 293–306. [Google Scholar] [CrossRef]
  174. Lutz, P.-E.; Kieffer, B.L. Opioid receptors: Distinct roles in mood disorders. Trends Neurosci. 2013, 36, 195–206. [Google Scholar] [CrossRef] [Green Version]
  175. Monnelly, V.J.; Hamilton, R.; Chappell, F.M.; Mactier, H.; Boardman, J.P. Childhood neurodevelopment after prescription of maintenance methadone for opioid dependency in pregnancy: A systematic review and meta-analysis. Dev. Med. Child Neurol. 2019, 61, 750–760. [Google Scholar] [CrossRef] [Green Version]
  176. Conradt, E.; Crowell, S.E.; Lester, B.M. Early life stress and environmental influences on the neurodevelopment of children with prenatal opioid exposure. Neurobiol. Stress 2018, 9, 48–54. [Google Scholar] [CrossRef]
  177. Eschenroeder, A.C.; Vestal-Laborde, A.A.; Sanchez, E.S.; Robinson, S.E.; Sato-Bigbee, C. Oligodendrocyte responses to buprenorphine uncover novel and opposing roles of μ-opioid-and nociceptin/orphanin FQ receptors in cell development: Implications for drug addiction treatment during pregnancy. Glia 2012, 60, 125–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Hauser, K.F.; Houdi, A.A.; Turbek, C.S.; Elde, R.P.; Iii, W.M. Opioids intrinsically inhibit the genesis of mouse cerebellar granule neuron precursors in vitro: Differential impact of μ and δ receptor activation on proliferation and neurite elongation. Eur. J. Neurosci. 2000, 12, 1281–1293. [Google Scholar] [CrossRef] [PubMed]
  179. Narita, M.; Kuzumaki, N.; Miyatake, M.; Sato, F.; Wachi, H.; Seyama, Y.; Suzuki, T. Role of δ-opioid receptor function in neurogenesis and neuroprotection. J. Neurochem. 2006, 97, 1494–1505. [Google Scholar] [CrossRef] [PubMed]
  180. McLaughlin, P.J.; Tobias, S.W.; Lang, C.M.; Zagon, I.S. Chronic exposure to the opioid antagonist naltrexone during pregnancy: Maternal and offspring effects. Physiol. Behav. 1997, 62, 501–508. [Google Scholar] [CrossRef]
  181. Wille-Bille, A.; Bellia, F.; García, A.M.J.; Miranda-Morales, R.S.; D’Addario, C.; Pautassi, R.M. Early exposure to environmental enrichment modulates the effects of prenatal ethanol exposure upon opioid gene expression and adolescent ethanol intake. Neuropharmacology 2020, 165, 107917. [Google Scholar] [CrossRef] [PubMed]
  182. Zagon, I.S.; Tobias, S.W.; McLaughlin, P.J. Endogenous opioids and prenatal determinants of neuroplasticity. In Brain Plasticity; Filogamo, G., Vernadakis, A., Gremo, F., Privat, A.M., Timiras, P.S., Eds.; Springer: Boston, MA, USA, 1997; Volume 429, pp. 289–303. [Google Scholar]
  183. Gomes, T.M.; da Silva, D.D.; Carmo, H.; Carvalho, F.; Silva, J.P. Epigenetics and the endocannabinoid system signaling: An intricate interplay modulating neurodevelopment. Pharmacol. Res. 2020, 162, 105237. [Google Scholar] [CrossRef] [PubMed]
  184. Walter, L.; Stella, N. Cannabinoids and neuroinflammation. Br. J. Pharmacol. 2004, 141, 775–785. [Google Scholar] [CrossRef] [Green Version]
  185. Zamberletti, E.; Rubino, T.; Parolaro, D. Therapeutic potential of cannabidivarin for epilepsy and autism spectrum disorder. Pharmacol. Ther. 2021, 226, 107878. [Google Scholar] [CrossRef]
  186. Nashed, M.G.; Hardy, D.B.; Laviolette, S.R. Prenatal Cannabinoid Exposure: Emerging Evidence of Physiological and Neuropsychiatric Abnormalities. Front. Psychiatry 2021, 11, 1577. [Google Scholar] [CrossRef]
  187. Mato, S.; Del Olmo, E.; Pazos, A. Ontogenetic development of cannabinoid receptor expression and signal transduction functionality in the human brain. Eur. J. Neurosci. 2003, 17, 1747–1754. [Google Scholar] [CrossRef]
  188. Alexandre, J.; Carmo, H.; Carvalho, F.; Silva, J.P. Synthetic cannabinoids and their impact on neurodevelopmental processes. Addict. Biol. 2020, 25, e12824. [Google Scholar] [CrossRef] [PubMed]
  189. Carbone, E.; Manduca, A.; Cacchione, C.; Vicari, S.; Trezza, V. Healing autism spectrum disorder with cannabinoids: A neuroinflammatory story. Neurosci. Biobehav. Rev. 2020, 121, 128–143. [Google Scholar] [CrossRef] [PubMed]
  190. Fernández-Ruiz, J.; De Fonseca, F.R.; Navarro, M.; Ramos, J. Maternal cannabinoid exposure and brain development: Changes in the ontogeny of dopaminergic neurons. In Marijuana/Cannabinoids: Neurobiology and Neurophysiology, 1st ed.; Murphy, L., Bartke, A., Eds.; CRC Press: Boca Raton, FL, USA, 2019; pp. 119–164. [Google Scholar]
  191. Fernández-Ruiz, J.; Berrendero, F.; Hernández, M.L.; Ramos, J.A. The endogenous cannabinoid system and brain development. Trends Neurosci. 2000, 23, 14–20. [Google Scholar] [CrossRef]
  192. Lisboa, S.F.; Niraula, A.; Resstel, L.B.; Guimaraes, F.S.; Godbout, J.P.; Sheridan, J.F. Repeated social defeat-induced neuroinflammation, anxiety-like behavior and resistance to fear extinction were attenuated by the cannabinoid receptor agonist WIN55, 212–2. Neuropsychopharmacology 2018, 43, 1924–1933. [Google Scholar] [CrossRef] [PubMed]
  193. García-Baos, A.; Alegre-Zurano, L.; Cantacorps, L.; Martín-Sánchez, A.; Valverde, O. Role of cannabinoids in alcohol-induced neuroinflammation. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 104, 110054. [Google Scholar] [CrossRef]
  194. Nezgovorova, V.; Ferretti, C.; Taylor, B.; Shanahan, E.; Uzunova, G.; Hong, K.; Devinsky, O.; Hollander, E. Potential of cannabinoids as treatments for autism spectrum disorders. J. Psychiatr. Res. 2021, 137, 194–201. [Google Scholar] [CrossRef]
  195. Fusar-Poli, L.; Cavone, V.; Tinacci, S.; Concas, I.; Petralia, A.; Signorelli, M.S.; Díaz-Caneja, C.M.; Aguglia, E. Cannabinoids for people with ASD: A systematic review of published and ongoing studies. Brain Sci. 2020, 10, 572. [Google Scholar] [CrossRef]
  196. Chieffi, S.; Carotenuto, M.; Monda, V.; Valenzano, A.; Villano, I.; Precenzano, F.; Tafuri, D.; Salerno, M.; Filippi, N.; Nuccio, F.; et al. Orexin system: The key for a healthy life. Front. Physiol. 2017, 8, 357. [Google Scholar] [CrossRef] [Green Version]
  197. Ogawa, Y.; Kanda, T.; Vogt, K.; Yanagisawa, M. Anatomical and electrophysiological development of the hypothalamic orexin neurons from embryos to neonates. J. Comp. Neurol. 2017, 525, 3809–3820. [Google Scholar] [CrossRef]
  198. Aran, A.; Shors, I.; Lin, L.; Mignot, E.; Schimmel, M.S.J.S. CSF levels of hypocretin-1 (orexin-A) peak during early infancy in humans. Sleep 2012, 35, 187–191. [Google Scholar] [CrossRef] [Green Version]
  199. Dickinson, H.; Walker, D.W.; Castillo-Melendez, M. Onset of feeding at birth—Perinatal development of the hypothalamic mechanisms that induce appetite and feeding in the newborn. Neurosci. Lett. 2008, 436, 1–6. [Google Scholar] [CrossRef] [PubMed]
  200. Sokołowska, P.; Urbańska, A.; Biegańska, K.; Wagner, W.; Ciszewski, W.; Namiecińska, M.; Zawilska, J.B. Orexins protect neuronal cell cultures against hypoxic stress: An involvement of Akt signaling. J. Mol. Neurosci. 2014, 52, 48–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Sokołowska, P.; Urbańska, A.; Namiecińska, M.; Biegańska, K.; Zawilska, J.B. Orexins promote survival of rat cortical neurons. Neurosci. Lett. 2012, 506, 303–306. [Google Scholar] [CrossRef] [PubMed]
  202. Baykal, S.; Albayrak, Y.; Durankuş, F.; Güzel, S.; Abbak, Ö.; Potas, N.; Beyazyüz, M.; Karabekiroğlu, K.; Donma, M.M. Decreased serum orexin A levels in drug-naive children with attention deficit and hyperactivity disorder. Neurol. Sci. 2019, 40, 593–602. [Google Scholar] [CrossRef] [PubMed]
  203. Omokawa, M.; Ayabe, T.; Nagai, T.; Imanishi, A.; Omokawa, A.; Nishino, S.; Sagawa, Y.; Shimizu, T.; Kanbayashi, T. Decline of CSF orexin (hypocretin) levels in Prader–Willi syndrome. Am. J. Med. Genet. Part A 2016, 170, 1181–1186. [Google Scholar] [CrossRef] [PubMed]
  204. Van Den Pol, A.N.; Patrylo, P.R.; Ghosh, P.K.; Gao, X.B. Lateral hypothalamus: Early developmental expression and response to hypocretin (orexin). J. Comp. Neurol. 2001, 433, 349–363. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Timeline of ontogeny of neurodevelopment processes and neuromodulator systems in rodents and human in days. Top is the expected timeline in human in weeks. E = embryonic day in rodents. GD: gestation day in human. Each neuromodulator ontogeny E number is based on first expression of receptors and if possible, receptor binding ability. If receptor information is not available, neuron ontogeny is noted. Ontogeny related to detection of synthesizing enzymes, neuromodulator or mRNA levels is not considered in this graph. The use of receptor ontogeny is to predict if external pharmacological agents could have any activity. Ontogeny of the neuromodulator itself could be before or after receptor ontogeny. References for each neuromodulator are as follows: acetylcholine nicotinic [17], acetylcholine muscarinic [18], norepinephrine [19], dopamine [20], serotonin [21], glutamate ionotropic [22], GABA [23], histamine [24], oxytocin [25], opioid μ and κ receptors [26], orexin [27,28]. The figure was created in Biorender.com and was licensed for publication in this journal.
Figure 1. Timeline of ontogeny of neurodevelopment processes and neuromodulator systems in rodents and human in days. Top is the expected timeline in human in weeks. E = embryonic day in rodents. GD: gestation day in human. Each neuromodulator ontogeny E number is based on first expression of receptors and if possible, receptor binding ability. If receptor information is not available, neuron ontogeny is noted. Ontogeny related to detection of synthesizing enzymes, neuromodulator or mRNA levels is not considered in this graph. The use of receptor ontogeny is to predict if external pharmacological agents could have any activity. Ontogeny of the neuromodulator itself could be before or after receptor ontogeny. References for each neuromodulator are as follows: acetylcholine nicotinic [17], acetylcholine muscarinic [18], norepinephrine [19], dopamine [20], serotonin [21], glutamate ionotropic [22], GABA [23], histamine [24], oxytocin [25], opioid μ and κ receptors [26], orexin [27,28]. The figure was created in Biorender.com and was licensed for publication in this journal.
Pharmaceuticals 15 00612 g001
Figure 2. Graphic representation of synapses and microglia interaction during neurodevelopment and possible roles of neuromodulators. R—neuromodulator receptors, 2-AG—2-arachidonoylglycerol, AEA—anandamide, DAGL—diacylglycerol lipase, NAPE-PLD—N-acyl phosphatidylethanolamine phospholipase D, DAG-diaceylglycerol, AA—arachidonic acid, GLU—glutamate, DA—dopamine, NE—norepinephrine, HA—histamine, ACh—acetylcholine, 5-HT—serotonin, OXT—oxytocin. The figure was created in Biorender.com and was licensed for publication in this journal.
Figure 2. Graphic representation of synapses and microglia interaction during neurodevelopment and possible roles of neuromodulators. R—neuromodulator receptors, 2-AG—2-arachidonoylglycerol, AEA—anandamide, DAGL—diacylglycerol lipase, NAPE-PLD—N-acyl phosphatidylethanolamine phospholipase D, DAG-diaceylglycerol, AA—arachidonic acid, GLU—glutamate, DA—dopamine, NE—norepinephrine, HA—histamine, ACh—acetylcholine, 5-HT—serotonin, OXT—oxytocin. The figure was created in Biorender.com and was licensed for publication in this journal.
Pharmaceuticals 15 00612 g002
Table 1. The neuromodulatory effects of prenatal exposure to pharmacological agents and addictive substances in animal models.
Table 1. The neuromodulatory effects of prenatal exposure to pharmacological agents and addictive substances in animal models.
DrugAffected System Preclinical Outcomes References
ValproateGABA
Glutamate Serotonin
Oxytocin
Auditory dysfunction relevant to ASD
Autism-like behaviour
Autism-like behaviour
Autism-like behaviour
Autism-like behaviour
[61]
[62,63]
[64,65]
[66]
Diazepam GABAAnxiety
Sex-specific changes in social interaction
[67]
Ketamine GABA Model of treatment-resistant schizophrenia[68]
Selective serotonin reuptake inhibitors (SSRIs)Serotonin
Dopamine
Noradrenaline
Oxytocin
Neurodevelopmental changes, warning against intake during pregnancy
elevated anxiety + low sociability
[69]
[70]
Antipsychotics Dopamine Reduced spatial learning and reduced postnatal activity of dopaminergic neurons[71,72]
Sevoflurane Glutamate/GABA balanceImpaired learning
Depression-like behaviour
[73]
GlucocorticoidsDopamine Altered sociability
Anhedonia
Depression-like behaviour
[74]
Nicotine Dopamine
Orexin
Disturbed dopaminergic activity
ADHD-like behaviour
Increased drug abuse
[75,76]
[77]
Amphetamine Dopamine Altered activity of dopaminergic neurons[72]
Cocaine Dopamine
Opioids
Disrupted development and activity of dopaminergic neurons
Disrupted opioid expression in dopaminergic neurons
[72,78]
[79,80]
Ethanol Dopamine, Serotonin
Acetylcholine
Oxytocin
Opioids
Orexin
ADHD-like behaviour
Depressive behaviour
Cognitive deficit
Impaired sociability
Impairment in working memory and reversal learning
Impaired attentional set shifting
Deficit in social interaction and depression-like behaviour
Increased likability of ethanol taste
Ethanol abuse
[81,82,83]
[84,85]
[86,87]
[88]
[89,90]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saad, A.K.; Akour, A.; Mahboob, A.; AbuRuz, S.; Sadek, B. Role of Brain Modulators in Neurodevelopment: Focus on Autism Spectrum Disorder and Associated Comorbidities. Pharmaceuticals 2022, 15, 612. https://0-doi-org.brum.beds.ac.uk/10.3390/ph15050612

AMA Style

Saad AK, Akour A, Mahboob A, AbuRuz S, Sadek B. Role of Brain Modulators in Neurodevelopment: Focus on Autism Spectrum Disorder and Associated Comorbidities. Pharmaceuticals. 2022; 15(5):612. https://0-doi-org.brum.beds.ac.uk/10.3390/ph15050612

Chicago/Turabian Style

Saad, Ali K., Amal Akour, Abdulla Mahboob, Salahdein AbuRuz, and Bassem Sadek. 2022. "Role of Brain Modulators in Neurodevelopment: Focus on Autism Spectrum Disorder and Associated Comorbidities" Pharmaceuticals 15, no. 5: 612. https://0-doi-org.brum.beds.ac.uk/10.3390/ph15050612

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