Next Article in Journal
Melatonin Activates Anti-Inflammatory Features in Microglia in a Multicellular Context: Evidence from Organotypic Brain Slices and HMC3 Cells
Next Article in Special Issue
VEGFA Isoforms as Pro-Angiogenic Therapeutics for Cerebrovascular Diseases
Previous Article in Journal
Transcriptomic Effects on the Mouse Heart Following 30 Days on the International Space Station
Previous Article in Special Issue
Pathophysiology of Post-Traumatic Trigeminal Neuropathic Pain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Schizophrenia Animal Modeling with Epidermal Growth Factor and Its Homologs: Their Connections to the Inflammatory Pathway and the Dopamine System

1
Department of Molecular Neurobiology, Brain Research Institute, Niigata University, Niigata 951-8585, Japan
2
Department of Physiology, School of Medicine, Niigata University, Niigata 951-8122, Japan
3
Department of Physiological Sciences, School of Pharmaceutical Sciences, Wakayama Medical University, Wakayama 649-8156, Japan
*
Authors to whom correspondence should be addressed.
Biomolecules 2023, 13(2), 372; https://doi.org/10.3390/biom13020372
Submission received: 20 January 2023 / Revised: 10 February 2023 / Accepted: 12 February 2023 / Published: 15 February 2023

Abstract

:
Epidermal growth factor (EGF) and its homologs, such as neuregulins, bind to ErbB (Her) receptor kinases and regulate glial differentiation and dopaminergic/GABAergic maturation in the brain and are therefore implicated in schizophrenia neuropathology involving these cell abnormalities. In this review, we summarize the biological activities of the EGF family and its neuropathologic association with schizophrenia, mainly overviewing our previous model studies and the related articles. Transgenic mice as well as the rat/monkey models established by perinatal challenges of EGF or its homologs consistently exhibit various behavioral endophenotypes relevant to schizophrenia. In particular, post-pubertal elevation in baseline dopaminergic activity may illustrate the abnormal behaviors relevant to positive and negative symptoms as well as to the timing of this behavioral onset. With the given molecular interaction and transactivation of ErbB receptor kinases with Toll-like receptors (TLRs), EGF/ErbB signals are recruited by viral infection and inflammatory diseases such as COVID-19-mediated pneumonia and poxvirus-mediated fibroma and implicated in the immune–inflammatory hypothesis of schizophrenia. Finally, we also discuss the interaction of clozapine with ErbB receptor kinases as well as new antipsychotic development targeting these receptors.

1. Introduction

1.1. Biological Overviews of the Epidermal Growth Factor (EGF) Family

Prof. S. Cohen first purified EGF as an eyelid opening factor from the mouse submaximal gland with nerve growth factor (NGF) [1]. Subsequent studies verified that EGF is a potent cell growth factor for almost all epidermal cells and that hyper-signaling of EGF or EGF receptors is involved in cancer growth and metastasis [2,3]. EGF is abundant in the urine and amnion and is often referred to as urogastron [4]. NGF has become popular in developmental neuroscience with its marked trophic actions on the cell survival of neurons [5]. Although both EGF and NGF have the same historical and tissue origin, these factors have opened separate windows in new fields of science, leading to modern cancer biology and neurobiology, respectively.
Molecular cloning and protein sequencing allowed us to discover novel homologs or orthologs of EGF, which can evoke the same or similar activities [6,7,8] and include three subfamilies: the EGF group, neuregulin group, and virokine group (Figure 1). Despite the peripheral functions of the EGF family in previous studies, the high expressions of heparin-binding EGF-like growth factor (HB-EGF) and transforming growth factor alpha (TGFα) have been reported in the central nervous system (CNS) [9,10]. High levels of neuregulin-1 mRNA and protein are also detectable in brain neurons [11,12], suggesting the hidden functions of the EGF family in brain development and plasticity [13,14,15].
One of the peculiar characteristics of EGF signaling is its role in viral proliferation [17,18]. The poxvirus family encodes an EGF-like molecule in its genome that promotes host expansion after infection. The EGF-like factors of poxvirus are called virokines, which share common core cysteine repeats and their surrounding amino acid sequences with the EGF sequence, reacting with the same receptors for the EGF family [17,18]. Similarly, the papillomavirus induces host cell proliferation and produces poxes, modifying EGF receptor signaling in host cells [19].
Another peculiar feature of the EGF family has been elucidated in oncogenic biology as well as in immunology (Figure 2). It had been a mystery how tumorigenesis can be induced following chronic inflammation in the colon and lung [20,21]. EGF is one of the key molecules responsible for carcinogenesis resulting from chronic inflammation [20]. Toll-like receptor (TLR)-mediated or prostaglandin-triggered inflammation involves protein kinase C (PKC) signaling as well as ErbB kinase signaling [22,23,24]. The activation of PKC potentiates ADAM (a disintegrin and metalloprotease), eliciting ectodomain shedding of the EGF family’s precursors. The mature forms of active EGF homologs are released to stimulate proliferation of neighboring cells [25]. In turn, EGF/EGFR signaling in neighboring cells potently induces reactive oxygen species (ROS), activates STAT and NF-κB, and inactivates various phosphatases, leading to chronic inflammation [26] (Figure 2). Some reports propose direct molecular interactions between TLRs and ErbB receptor kinases as well [22,23,24]. These tight interactions and pathological connections between EGF signaling and the TLR system play a crucial role in current respiratory syncytial (RS)-virus-induced and COVID-19-induced cytokine storm and pneumonia [27,28]. Given the interactions between these systems, the EGF family and its signaling are augmented in the immune–inflammatory hypothesis of schizophrenia [29,30,31,32,33], potentially suggesting a substantial impact on the later risk of schizophrenia or other psychiatric diseases following the COVID-19 pandemic [34,35,36]. At present, there is no clear boundary between growth factors and inflammatory cytokines. Thus, members of the EGF family are currently classified into the category of cytokines as well.
Receptors of the EGF family are named ErbB1-4 or Her1-4 and belong to the receptor tyrosine kinase superfamily (Figure 1) [42,43]. Alternatively, EGF receptor is abbreviated to EGFR as well. A member of the EGF family binds to ErbB receptor kinases and evokes signal transduction via ErbB receptor homodimerization or heterodimerization [42,43]. In general, a ligand member in the EGF family binds to one of the ErbB1-4 receptors and recruits the second partner molecule from ErbB1-4 to induce their mutual trans-phosphorylation, evoking signal transduction from both ErbB molecules [43]. For example, EGF can evoke ErbB4 signaling by heterodimerization between the receptor of ErbB1 and the partner of ErbB4 [44].
As all neural stem cells are derived from the ectodermal layer, undifferentiated neural stem cells in the nervous system express high levels of ErbB1 or ErbB2 [45,46]. Following neural development and maturation, the level of ErbB1 is decreased, and ErbB4 expression is elevated, both of which are limited to midbrain dopaminergic neurons and GABAergic cells [46,47,48,49]. In the glial lineage, ErbB1 expression is replaced by ErbB3 during development. The interactions of neuregulins expressed on the neuronal surface with ErbB3 in oligodendrocytes promote their differentiation and myelination [50]. Accordingly, the EGF/neuregulin system contributes not only to the development of peripheral epidermal cells but also to the development and maturation of neurons and glial cells.
On the basis of these biological activities of the EGF family, we attempt to illustrate their neuropathologic contribution to schizophrenia as well as the phenotypic association of its model animals established by the members of the EGF family.

1.2. Phenotypic Association of The EGF Ligand Family with Schizophrenia

The association between this EGF family and schizophrenia was first reported in 2002 by two independent groups. We investigated neuropathological alterations in the EGF family and ErbB1-4 receptors in patients with schizophrenia [51]. We found that ErbB1 receptor levels are elevated specifically in the forebrain, whereas EGF content is reduced in the blood of patients with schizophrenia. The reduction in blood EGF concentrations is confirmed by later reports [52,53], although there is controversy still remaining [54]. In the same year of 2002, Decode Genetics Inc. found a genetic association between a neuregulin-1 haplotype and this illness by analyzing the national genome bank of Iceland [55]. The genetic association with neuregulin-1 is confirmed by later meta-analyses as well [56]. Groups in Finland reported positive interactions with the single-nucleotide polymorphism (SNP) of the EGF genome [57,58], but the other group failed to detect this genetic association in a Japanese population [59]. Groenestege et al. (2007) found that one family lineage of the EGFR-mutation-driven renal disease carries schizophrenia and other psychiatric diseases [60]. In addition to the above pioneering genetic study on neuregulin-1, the neuropathological association of neuregulin-1 with schizophrenia has also been investigated in postmortem brain and blood samples; Neuregulin-1 mRNA and its receptor protein ErbB4 are elevated in the brain [61,62,63]. Neuregulin-1 protein levels in the blood are increased in patients [64]. Those clinical studies on schizophrenia were followed by animal modeling and its translational studies [65].
In addition to the genetic hypothesis for schizophrenia, many epidemiological studies have raised various types of environmental hypotheses for schizophrenia, such as the immune–inflammatory hypothesis [31,66,67,68]. Among these, we tested the immune–inflammatory hypothesis to produce animal models for schizophrenia; prenatal or perinatal inflammatory cytokines perturb normal brain development and/or circuit organization, leading to abnormal neurocognition and emotions in schizophrenia [29,30,31,69,70]. Accordingly, we tested this hypothesis at the molecular level, challenging pregnant mice and neonatal rats with various inflammatory cytokines [71]. Other groups have tested the impact of fetal inflammation in pregnant mice, treating them with polyinosinic–polycytidylic acid (poly I:C) or lipopolysaccharide (LPS) [69,72,73], ligands of the Toll-like receptors. Of note, it is suggested that the maternal infection model established with poly I:C also involves abnormal ErbB signaling [33]. A pathological role of interleukin-6 (IL-6) is implicated in their behavioral deficits of the poly I:C-injection model, but we failed to induce behavioral impairments in their offspring with maternal and neonatal direct challenges of IL-6 [74,75]. In our cytokine challenge procedure, maternal administration of IL-2 produced the most prominent impact on offspring behaviors. In the neonatal treatment protocol, IL-1, neuregulin-1, and EGFR/ErbB1 ligands (EGF, TGFα, and epiregulin) exhibited the most remarkable effects on post-pubertal behaviors and cognitions [76,77,78,79].
Considering the strength of the behavioral impact and its reproducibility, we selected EGF and neuregulin-1 and characterized the animal models established by these cytokines. After publishing several results from the neuregulin-1 model, however, we ceased the direction of the neuregulin study and put our primary focus on the EGF animal model for the following reasons. EGF model rats or mice exhibit almost all the behavioral endophenotypes relevant to schizophrenia, including prepulse inhibition, social interaction, sensitivity to psychostimulants, and latent inhibition of fear learning, with high reproducibility [77,78,79,80,81,82,83]. In primate modeling, moreover, the neonatal EGF treatment of cynomolgus monkeys resulted in the post-pubertal emergence of behavioral abnormality [84]. A monkey challenged with EGF displayed visual hallucination-like behaviors and self-injury after five years of age. In contrast to the EGF model, one of the neuregulin-1 mouse models exhibited various face validities at the behavioral level but carried severe hearing deficits [79,80]. Thus, we expected future technical difficulty to distinguish direct and indirect effects of the hearing deficits on their cognitive functions.
In addition, our latest studies reveal that the EGF rat model exhibits most of the pathophysiological changes reported in patients with schizophrenia; electroencephalography abnormalities in duration mismatch negativity (MMN), frequency MMN, auditory steady-state response (ASSR), P300, auditory brain stem response (ABR), and functional inter-cortical connectivity [85,86,87]. In the following chapters, we summarize the dopaminergic role in the behavioral impairment of the EGF model and discuss the neuropathological implication of this model in the dopamine hypothesis for schizophrenia. As most of the above pathophysiological changes of the EGF models do not involve the dopaminergic system, we do not discuss their pathophysiology in this review.

2. Method for Literature Selection

We searched for previous publications using the keywords of our main topics {EGF, schizophrenia, rat/mice} and {EGF, schizophrenia, patients} in the NCBI PubMed database. The former search and the subsequent re-confirmation of their contents resulted in 38 papers, including 30 papers from our laboratory. Accordingly, the main logical flow of this review was constructed on the basis of our previous studies. In addition, clinical and genetic studies regarding EGF and schizophrenia total 14 papers, the contents of which are often confirmative, and only representative papers are discussed in this review. The other search was performed with the keywords {EGF/neuregulin, dopamine neuron}, resulting in 32 papers. Nineteen papers fitting the present discussion were adopted. To explain the backgrounds of technical terms and hypotheses, we additionally cited the literature relevant to the present discussions. Therefore, there is a limitation of our data explanation and discussions in this review. The present arguments are also hypothetical at this stage with the given uncertainty of the schizophrenia neuropathology. In this respect, this article is not a systemic review but a hypothesis-driven review.

3. Neurotrophic Functions of EGF and Neuregulins in Dopaminergic Neurons

3.1. Distributions of ErbB1 (EGFR) and ErbB4 in the Midbrain

The EGF receptor ErbB1 is distributed in dopaminergic neurons in the substantia nigra compacta, as well as in the ventral tegmental area [45,48]. Similarly, the neuregulin receptor ErbB4 is expressed in the same regions but not always in the same dopaminergic cell population [45,48]. Here, we discuss the neurotrophic actions of EGF and neuregulins on dopaminergic neurons located in these regions.
In culture, EGF enhances cell survival and neurite extension of dopaminergic neurons [48]. The neurotrophic activity of EGF is inhibited by EGF-neutralizing antibodies and by tyrosine kinase inhibitors for ErbB1 [48]. However, the neurotrophic activity of EGF was not blocked by glial cell line-derived neurotrophic factor (GDNF)-neutralizing antibody or an inhibitor of neurotrophin (TrkB1-Fc), suggesting a direct action of EGF on dopaminergic neurons. In vivo administration of ErbB1 kinase inhibitors attenuates postnatal elevation of tyrosine hydroxylase, verifying the neurotrophic role of endogenous EGF signaling in dopaminergic development [48].
GDNF is the most well-known factor for its neurotrophic action on dopaminergic neurons. This factor can prevent dopaminergic neurodegeneration in the animal models of Parkinson’s disease, and this GDNF action is mimicked by EGF [88,89,90,91,92,93]. Both factors are provided by striatal cells as retrograde factors that maintain dopaminergic neurons [94,95]. Neuregulin-1 also exerts similar neurotrophic activity in dopaminergic neurons in vivo [96,97]. However, knockout mice deficient of the ErbB4 gene exhibit normal neuroanatomy of dopamine neurons in the midbrain [98], although several controversies remain (see below).
Receptors for EGF and neuregulins, ErbB1 and ErbB4, are also expressed in several types of GABAergic populations, although this topic is not discussed in the present review paper. However, we need to point out the fact that the action of EGF on this cell population is opposite to that of dopaminergic neurons; it attenuates postnatal development or maturation of GABAergic neurons [99]. This contrasts with the positive neurotrophic actions of neuregulin-1 commonly found in the GABAergic cell populations [100,101,102].

3.2. Influences of EGF/Neuregulin Signaling on Dopamine Metabolism

In addition to the developmental action of EGF and neuregulins, these factors can exert functional effects on mature dopaminergic neurons in the adult stage. In vivo infusion of neuregulin-1 activates ErbB4 receptors in the prefrontal cortex, hippocampus, and dorsal striatum and elevates dopamine concentrations [103,104]. In addition, neuregulin/ErbB signaling plays regulatory roles in the expression of dopamine transporters as well as in glutamatergic synapse plasticity [103,105,106,107]. These phenomena are consistent with our previous observation of the neuregulin-1 model established by its neonatal challenges; neonatal treatment with neuregulin-1 elevates dopamine synthesis and release [79]. However, other studies on ErbB4 knockout mice have provided controversial results: the signal blockade of neuregulin-1/ErbB4 elevates striatal dopamine levels [108,109]. This appears to be in accordance with the report that the gene ablation of HB-EGF, a ligand for ErbB1/B4 receptors, resulted in the abnormal behaviors relevant to the dopaminergic system or schizophrenia pathology [110]. Therefore, the neuregulin/ErbB4 functions in dopaminergic regulation differ significantly among target cell populations, their developmental stages, or neuregulin-splicing variants [65,111].
In contrast to the physiological actions of neuregulins, the action of EGF on mature dopaminergic cells is relatively consistent across reports. EGF enhances ATP-triggered dopamine release from cultured PC12 cells [112]. In vivo administration of EGF, HB-EGF, or neuroegulin-1 elevates the concentrations of dopamine and its metabolite, 4-dihydroxyphenylacetic acid (DOPAC), in the striatum of Parkinson’s model animals [97,103,113] as well as in normal rats [114]. In a critical sense, however, it is uncertain whether the EGF-induced dopamine release is ascribed to enhanced dopamine synthesis or elevated vesicular release of this neurotransmitter.
We prepared rodent models of schizophrenia by subcutaneously injecting EGF or neuregulin-1 into neonatal rats and mice [77,78,79]. There was a marked increase in tyrosine hydroxylase in the whole forebrain regions during the postnatal administration of these factors [77,79]. Thus, the increased brain concentrations of dopamine and its metabolites presumably reflect the elevation of dopamine synthesis. At the adult stage following neonatal EGF treatment, however, we detected abnormal hyperdopaminergic innervation and higher dopamine release only in the globus pallidus. We speculate that the reason for the persistent hyperdopaminergic innervation to the globus pallidus is that the lateral area (tier) of the substantia nigra compacta, which mainly innervates the globus pallidus, is enriched with ErbB1/EGFR even at the adult stage [115].
In contrast to EGF model rats, schizophrenia model mice established by postnatal treatment with neuregulin-1 display a variety of behavioral abnormalities depending on the neuregulin-1 splicing variants administered [79,80]. In the model established by a full form of neuregulin-1β1, prefrontal hyperinnervation of A10 dopaminergic fibers is apparent, and dopamine levels are found elevated in the forebrain regions [79]. These neuregulin model mice show behavioral deficits relevant to schizophrenia. However, this mouse model exhibits a marked difference in acoustic hearing ability from the model established by another neuregulin-1 variant (a EGF core domain of neuregulin-1β) [80]. Behavioral responses to the NMDA receptor blocker MK801 also differed between the models established by two neuregulin-1 variants [65]. With the given difference among neuregulin-1 variants, the biological and behavioral effects of individual neuregulin-1 variants remain to be distinguished carefully.

3.3. Post-Pubertal Elevation of Dopaminergic Activity in the EGF/Neuregulin Models

One of the peculiar features of schizophrenia is that disease onset is limited to the pubertal and post-pubertal stages. Does this phenomenon correlate with dopaminergic abnormalities suggested in the EGF model? We examined the developmental alterations in the firing frequency and bursting ratio of midbrain dopaminergic neurons in EGF model rats and mice [116,117] (Figure 3). In general, midbrain dopaminergic activity and its synaptic innervation peak around puberty [118,119]. Before puberty, the firing frequency of the EGF model mice was indistinguishable or lower than that of control mice, whereas the frequency continued to increase and became higher after puberty. The burst ratio of dopaminergic firing subsided in control mice but not in EGF model mice after puberty. These phenomena agree with our pharmacological finding that SK (small conductance calcium-activated potassium) channel activity and sensitivity to apamin, implicated in the regulation of their burst firing, are reduced in EGF model mice only at the adult stage [116,117,118]. In addition, we speculate that the normal development of the local inhibitory system for dopamine neurons might also be impaired in this EGF model. In the neuregulin-1 mouse model, the attenuation factor has been identified as GABAergic inhibition from substantia nigra reticulata [120]. Disinhibition of dopaminergic neurons results in enhanced burst firing in the adult stage. Thus, almost all behavioral alterations in these models in mice, rats, and monkeys coincide with their post-pubertal abnormality of dopaminergic firing. In other animal models for schizophrenia, this coincidence is controversial among the reports. The neonatal ischemic model shows a similar increase in dopamine metabolism in the adult stage [121]. Similar abnormalities in the dopamine system or function are implicated in other animal models for schizophrenia, the prenatal methylazoxymethanol acetate (MAM) injection model and the prenatal poly I:C injection model [122,123]. More elaborate analyses regarding the temporal correlation between dopaminergic activity and behavioral alterations must be performed in these models to illustrate their association with the post-pubertal onset of schizophrenia.

4. Chronic Hyper-Dopaminergic Influences on Animal Models and Patients with Schizophrenia

As described above, EGF models exhibit hyperactivity of midbrain dopaminergic neurons in the substantia nigra compacta (the origin of the A9 pathway) and the ventral tegmental area (the origin of the A10 pathway) after puberty. In agreement with this, dopaminergic concentrations are elevated in the globus pallidus and medial prefrontal cortex—the target regions of these pathways. We investigated the behavioral impact of these hyperdopaminergic states in each pathway, distinguishing the resting and the stress-loaded conditions.

4.1. Interaction of A10 Dopaminergic Activity with Acoustic Prepulse Inhibition

The dopamine releasers of amphetamine and cocaine are known to decrease prepulse inhibition (PPI) of acoustic startle responses [124,125]. PPI reduction is induced by the non-selective dopamine receptor agonist apomorphine as well [126,127]. These reports indicated that hyperdopaminergic states are involved in the PPI reduction. Among the hyperdopaminergic regions in EGF model rats, our pharmacological interventions revealed that the globus pallidus plays a crucial role in PPI reduction [83] (Figure 4). The local infusion of a D2 receptor agonist, but not a D1 receptor agonist, into the globus pallidus of normal rats mimics the effects of EGF, that is, PPI reduction. Conversely, the local infusion of a D2 receptor antagonist, but not a D1 receptor antagonist, into the globus pallidus of EGF model rats ameliorated their PPI deficits [83]. The role of dopamine D2 receptor signaling in PPI regulation is also supported by previous pharmacological studies [128,129,130] and by gene ablation experiments of dopamine receptors [131]. However, in addition to our argument, the A10 dopaminergic pathway for the nucleus accumbens and the A9 pathway for the anteromedial striatum are also implicated in PPI regulation [132,133,134,135,136]. In these respects, controversies remain regarding the main dopamine pathway for PPI regulation and the mutual interactions of individual basal ganglia circuits.

4.2. Chronic Hyper-Dopaminergic States Impair Social Interaction

Antipsychotic treatment of EGF model rats with risperidone, but not with haloperidol, normalizes the dopaminergic hyperactivity of the A9 and A10 pathways [83] and their social interaction [137]. However, the pharmacological intervention of the A9 pathway does not affect the social interaction impairment of this model [83], suggesting that the basal hyperactivity of the A10 pathway stemming from the ventral tegmental is involved in this deficit.
We tested the role of the A10 pathway by chronically or acutely manipulating the firing activity of A10 dopaminergic neurons with the pharmacogenetics tool DREADD (Designer Receptors Exclusively Activated by Designer Drugs) [137,138]. DREADD-driven artificial normalization (i.e., a decrease) of A10 dopaminergic activity in EGF model rats ameliorates their social deficits [137]. Conversely, chronic upregulation of A10 dopaminergic activity in normal rats impairs social interaction [138]. Chronic dopamine elevation at the basal state negatively affects subsequent event-related dopamine responses and functions with two potential mechanisms (Figure 5). Our results suggest that the chronic and acute release of dopamine display opposite actions on the target functions. Chronic dopamine elevation at the basal state negatively affects subsequent dopamine responses and/functions. One of the molecular mechanisms is that with the higher baseline release of dopamine, the event-triggered fraction (i.e., the dynamic range) of dopamine release becomes lower. The other is that chronic dopamine release at the basal state induces the dopamine receptor internalization and the downregulation of receptor signaling, resulting in the attenuation of postsynaptic responses [139,140]. Therefore, acute stimulation of A10 dopaminergic neurons promotes social interactions [138,141]. Conversely, acute optogenetic suppression of A10 dopaminergic activity results in social interaction deficits [142]. In this respect, the event-triggered fraction (i.e., the dynamic range) of dopamine release determines the magnitude of event-related dopamine function and response. This argument suggests the possibility that dopaminergic hyperactivity at the resting state can be involved in parts of the negative symptoms of schizophrenia, such as stress vulnerability and anhedonia. A similar chronic abnormality in the A10 pathway would be implicated in social abnormalities in other animal models for schizophrenia, such as a chronic stress model with social defeats, a MAM-induced model, and a ventral hippocampus lesion model [122,143]. However, the antipsychotic responses of these models differ significantly [122,144].
In spite of the above argument, the medication effects of conventional antipsychotic drugs is limited on the negative symptoms of patients with schizophrenia [145,146,147,148]. Additional mechanisms would be required to fully illustrate the persistent nature of the depression-like negative symptoms in schizophrenia and in these animal models. Although the neuropathologic underpinnings of schizophrenia and depression are controversial, one of the potential explanations for the social deficits of the model animals would represent the neurodegenerative and cytotoxic actions of dopamine or inflammatory cytokines [149,150,151,152,153,154].
In general, the dopamine hypothesis for schizophrenia fails to illustrate the negative symptoms of this illness [145]. One is that typical antipsychotic drugs, which target dopamine D2 receptors, are ineffective against the negative symptoms of social withdrawal and anhedonia [145,146,147,148]. The chronic abuse of amphetamine or cocaine does not result in negative symptoms. However, the latest brain imaging studies have supported the dopamine hypothesis. Several PET studies have consistently reported a reduction in dopamine D1 receptor occupancy in the prefrontal cortex of patients with schizophrenia [155,156,157,158]. These studies also pointed out a negative correlation between occupancies and negative symptom scales across patients [155]. This reduction in dopamine receptor occupancy can be ascribed to either a decrease in receptor density or an increase in dopamine release. The latter explanation is supported by the following two findings. Incorporation of L-(beta 11C)-DOPA into the prefrontal cortex is elevated in patients with schizophrenia compared with that of control subjects [159]. A side product of dopamine synthesis, melanin, is highly deposited in the substantia nigra regions of patients with schizophrenia [160,161]. Although this explanation is not discrepant with our results from the EGF model, it would be better to verify whether the reduction in dopamine D1 receptor occupancy indeed stems from chronic upregulation of dopamine release.

5. Drug Development Targeting EGF/ErbB Signals

According to our initial postmortem findings in schizophrenia patients, ErbB1 receptors are upregulated, and EGF contents are downregulated in the forebrain regions of the patients [51]. This result suggests that abnormal EGF/ErbB1 signaling persists throughout life. However, it is unclear whether EGF/ErbB1 signals are enhanced or reduced in patients with schizophrenia. We attempted to answer this question using animal models. We chronically administered EGF into the ventricle of adult rat brains and monitored behavioral alterations in prepulse inhibition and social interaction [114]. Even in the adult stage, EGF infusion increases dopamine metabolism and PPI deficits [114]. This finding suggests the possibility that EGF model rats, established by perinatal EGF treatment, continue to harbor abnormal hyper-EGF/ErbB1 signaling in the brain.
To test this hypothesis, we administered specific inhibitors of EGF receptor kinase (i.e., gefitinib, erlotinib, and PD153035) to the lateral ventricle of EGF model rats [162]. These ErbB1 kinase inhibitors ameliorate the deficits in PPI and latent inhibition but not the social interactions of this model (Table 1). In parallel, the firing activity of midbrain dopaminergic neurons is normalized [162]. The antipsychotic action of these ErbB kinase inhibitors is not limited to the EGF model rats. These inhibitors also ameliorate several behavioral endophenotypes in rat models established by neonatal hippocampal lesioning [163]. Tadmor et al. (2018) tested another quinazoline ErbB inhibitor, JNJ28871063, and reported that this inhibitor ameliorates the phencyclidine (PCP)-induced decrease in social interaction, which is implicated in the negative symptoms of schizophrenia [164]. However, these ErbB kinase inhibitors have a limited capability to penetrate the blood–brain barrier. Alternatively, we discovered a compound that inhibits ErbB kinases and penetrates the blood–brain barrier: emodin [165]. Peripheral administration of emodin reduces PPI deficits and latent inhibition abnormalities in the EGF model.
Despite their antipsychotic potential, ErbB kinase inhibitors in the quinazoline family, which are prescribed to patients with cancer, are known to exert severe side effects such as interstitial pneumonia [167]. In this respect, it is challenging to develop new antipsychotic drugs targeting ErbB receptors (Table 1).
We also explored the possibility that the current antipsychotic drugs attenuate ErbB receptor kinases and contribute to their pharmacological profiles. Among many antipsychotic drugs, clozapine is known to be a unique atypical antipsychotic which exhibits antipsychotic efficacy on drug-tolerant patients with schizophrenia as well as on their negative symptoms and cognitive decline [168,169]. However, its prescription is strictly controlled, as clozapine often shows a severe side effect of agranulocytosis [168]. How does clozapine produce those favorable and unfavorable actions? We hypothesized that clozapine involves ErbB receptor signaling. To address this question, we exposed cultured brain tumor cells to low doses of clozapine and EGF [166]. Clozapine attenuated EGF-triggered growth and survival of tumor cells, similar to ErbB kinase inhibitors. To characterize the molecular interaction between clozapine and ErbB kinases, we tested the in vitro activity of clozapine using pure recombinant ErbB1-4 kinases. Tyrosine kinase activities of ErbB1-4 receptors were suppressed in vitro by sub-micromolar ranges of clozapine without any other components, indicating that clozapine directly acts on ErbB kinases and blocks their enzyme activity. These results indicate that the unique action of clozapine involves ErbB kinase inhibition, although several ambiguities and controversies still remain [170,171].

6. Provisional Conclusion

The members of the EGF family include a large variety of growth factors and cytokines expressed in the CNS, regulating dopaminergic development and functions. Their signal transduction is tightly connected to the immune inflammatory signaling stemming from TLRs. Accordingly, perinatal and prenatal infection and inflammation stimulate the TLR system to induce EGF/ErbB hyper-signaling, leading to abnormal dopamine firing and their phenotypic development/connections. Of note, the dopamine dysfunction of our EGF models reaches the peak at the post-pubertal stage. In particular, chronic hyperdopaminergic states at the adult resting state produce a negative impact on the event-triggered dopamine release and responses which are implicated in motivation and stress resilience.

7. Future Directions

Including the EGF family, various inflammatory cytokines and their signals are implicated in the molecular pathology of negative symptoms and/or depression [31,36,66,67,68]. A large variety of effective anti-inflammatory drugs have been developed, and some of those would be effective to ameliorate those psychiatric symptoms. These drugs mainly consist of small molecules and immunoglobulins targeting cytokine receptors and their signaling molecules [172]. In addition, the molecular structures of the ligand–ErbB receptor complex have been elucidated, which should hint at a new drug design targeting these receptors [173,174]. We hope that the present review will help future diversion of the pre-existing anti-inflammatory or anti-ErbB drugs as well as the development of novel antipsychotic drugs targeting these signal pathways.

Author Contributions

Original draft preparation and writing, H.S.; conceptualization, writing, review, and editing, H.N. (Hiroyuki Nawa); reference selection and proofreading, H.N. (Hisaaki Namba); reference selection and organization, M.T. and H.S. All authors have read and agreed to the published version of the manuscript.

Funding

The latest studies of our group were supported by Grant-in-Aid for Scientific Research on Innovative Area “Multiscale Brain” (HN; 18H05429, 18H05428), Grant-in-Aid for Challenging Exploratory Research (HN; 21K18242), and Grant-in-Aid for Scientific Research (B) (HN; 22H02728) provided by JSPS and/or MEXT in Japan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful for the contributions of our external collaborators (not shown) as well as for those of my lab. colleagues: Abe Y, Amada N, Araki K, Aoki H, Eda T, Furukawa K, Futamura T, Hanyu O, Hirano A, Horikawa HP, Ibaraki K, Imai C, Inaba H, Inamura N, Ishii K, Ishizuka Y, Iwakura Y, Jourdi H, Kai R, Kakiya N, Kanbe D, Kato T, Kawamura M, Kitayama E, Kobayashi Y, Komi E, Kumagi H, Mizuno M, Nagano T, Namba H, Narihara I, Narita E, Narisawa-Saito M, Nishikawa N, Okubo T, Otsu Y, Piao YS, Qi S, Sakai M, Sakai Y, Seki M, Shibuya M, Shishido Y, Sotoyama H, Sugai T, Takahashi M, Takasu Y, Takei N, Tanaka T, Tezuka T, Toyooka K, Tsuda N, Wang R, Watanabe Y, Xiong H, Yamaguchi T, Yanagi M, Yokomaku D, Yukawa T, Zheng Y, Zheng X, etc.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

EGF: epidermal growth factor; TLR: Toll-like receptor; NGF: nerve growth factor; HB-EGF: heparin-binding EGF-like growth factor; TGFα: transforming growth factor alpha; CNS: central nervous system; VGF: vaccinia virus growth factor; CGF: cowpox growth factor; SPGF: smallpox growth factor; SFGF: shope fibroma virus growth factor; MGF: myxoma virus growth factor; PKC: protein kinase C; ROS: reactive oxygen species; ADAM: a disintegrin and metalloprotease; LPA: lysophosphatidic acid; NOX: NADPH oxidase; GPCR: G-protein-coupled receptor; SNP: single-nucleotide polymorphism; poly I:C: polyinosinic–polycytidylic acid; LPS: lipopolysaccharide; IL: interleukin; ASSR: auditory steady-state response; ABR: auditory brain stem response; MMN: mismatch negativity; GDNF: glial cell line-derived neurotrophic factor; DOPAC: 4-dihydroxyphenylacetic acid; SK: small conductance calcium-activated potassium; MAM: methylazoxymethanol acetate; DA: dopamine; DOPA: L-3,4-dihydroxyphenylalanine; PPI: prepulse inhibition; STR; striatum; GP: globus pallidus; MSN: medium spiny neurons; SNc: substantia nigra pars compacta; SNr: substantia nigra pars reticulata; DREADD: designer receptors exclusively activated by designer drugs; CNO: clozapine-N-oxide; PCP: phencyclidine; LI: latent inhibition; VHL: ventral hippocampal lesion.

References

  1. Cohen, S. Origins of growth factors: NGF and EGF. J. Biol. Chem. 2008, 283, 33793–33797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Yarden, Y.; Pines, G. The ERBB network: At last, cancer therapy meets systems biology. Nat. Rev. Cancer 2012, 12, 553–563. [Google Scholar] [CrossRef] [PubMed]
  3. Kumar, A.; Petri, E.T.; Halmos, B.; Boggon, T.J. Structure and clinical relevance of the epidermal growth factor receptor in human cancer. J. Clin. Oncol. 2008, 26, 1742–1751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Gregory, H. In vivo aspects of urogastrone-epidermal growth factor. J. Cell Sci. Suppl. 1985, 3, 11–17. [Google Scholar] [CrossRef] [Green Version]
  5. Chao, M.V. Neurotrophins and their receptors: A convergence point for many signalling pathways. Nat. Rev. Neurosci. 2003, 4, 299–309. [Google Scholar] [CrossRef]
  6. Higashiyama, S.; Horikawa, M.; Yamada, K.; Ichino, N.; Nakano, N.; Nakagawa, T.; Miyagawa, J.; Matsushita, N.; Nagatsu, T.; Taniguchi, N.; et al. A novel brain-derived member of the epidermal growth factor family that interacts with ErbB3 and ErbB4. J. Biochem. 1997, 122, 675–680. [Google Scholar] [CrossRef]
  7. Harris, R.C.; Chung, E.; Coffey, J. The EGF Receptor Family: Biologic Mechanisms and Role in Cancer; Academic Press: London, UK, 2012; pp. 3–14. [Google Scholar]
  8. Vullhorst, D.; Ahmad, T.; Karavanova, I.; Keating, C.; Buonanno, A. Structural similarities between neuregulin 1-3 isoforms determine their subcellular distribution and signaling mode in central neurons. J. Neurosci. 2017, 37, 5232–5249. [Google Scholar] [CrossRef] [Green Version]
  9. Weickert, C.S.; Blum, M. Striatal TGF-alpha: Postnatal developmental expression and evidence for a role in the proliferation of subependymal cells. Brain Res. Dev. 1995, 86, 203–216. [Google Scholar] [CrossRef]
  10. Nakagawa, T.; Sasahara, M.; Hayase, Y.; Haneda, M.; Yasuda, H.; Kikkawa, R.; Higashiyama, S.; Hazama, F. Neuronal and glial expression of heparin-binding EGF-like growth factor in central nervous system of prenatal and early-postnatal rat. Brain Res. Dev. Brain Res. 1998, 108, 263–272. [Google Scholar] [CrossRef]
  11. Eilam, R.; Pinkas-Kramarski, R.; Ratzkin, B.J.; Segal, M.; Yarden, Y. Activity-dependent regulation of Neu differentiation factor/neuregulin expression in rat brain. Proc. Natl. Acad. Sci. USA 1998, 95, 1888–1893. [Google Scholar] [CrossRef]
  12. Kao, W.T.; Wang, Y.; Kleinman, J.E.; Lipska, B.K.; Hyde, T.M.; Weinberger, D.R.; Law, A.J. Common genetic variation in Neuregulin 3 (NRG3) influences risk for schizophrenia and impacts NRG3 expression in human brain. Proc. Natl. Acad. Sci. USA 2010, 107, 15619–15624. [Google Scholar] [CrossRef] [Green Version]
  13. Mei, L.; Xiong, W.-C. Neuregulin 1 in neural development, synaptic plasticity and schizophrenia. Nat. Rev. Neurosci. 2008, 9, 437–452. [Google Scholar] [CrossRef]
  14. Scalabrino, G. Epidermal growth factor in the CNS: A beguiling journey from integrated cell biology to multiple sclerosis. An extensive translationaloverview. Cell. Mol. Neurobiol. 2022, 42, 891–916. [Google Scholar] [CrossRef]
  15. Kwon, O.B.; Paredes, D.; Gonzalez, C.M.; Neddens, J.; Hernandez, L.; Vullhorst, D.; Buonanno, A. Neuregulin-1 regulates LTP at CA1 hippocampal synapses through activation of dopamine D4 receptors. Proc. Natl. Acad. Sci. USA 2008, 105, 15587–15592. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, R.; Iwakura, Y.; Araki, K.; Keino-Masu, K.; Masu, M.; Wang, X.Y.; Takei, N.; Higashiyama, S.; Nawa, H. ErbB2 dephosphorylation and anti-proliferative effects of neuregulin-1 in ErbB2-overexpressing cells; re-evaluation of their low-affinity interaction. Sci. Rep. 2013, 3, 1402. [Google Scholar] [CrossRef] [Green Version]
  17. Eppstein, D.A.; Marsh, Y.V.; Schreiber, A.B.; Newman, S.R.; Todaro, G.J.; Nestor, J.J. Epidermal growth factor receptor occupancy inhibits vaccinia virus infection. Nature 1985, 318, 663–665. [Google Scholar] [CrossRef]
  18. Kim, M.; Yang, H.; Kim, S.K.; Reche, P.A.; Tirabassi, R.S.; Hussey, R.E.; Chishti, Y.; Rheinwald, J.G.; Morehead, T.J.; Zech, T.; et al. Biochemical and functional analysis of smallpox growth factor (SPGF) and anti-SPGF monoclonal antibodies. J. Biol. Chem. 2004, 279, 25838–25848. [Google Scholar] [CrossRef]
  19. Conrad, M.; Goldstein, D.; Andresson, T.; Schlegel, R. The E5 protein of HPV-6, but not HPV-16, associates efficiently with cellular growth factor receptors. Virology 1994, 200, 796–800. [Google Scholar] [CrossRef]
  20. Greten, F.R.; Grivennikov, S.I. Inflammation and cancer: Triggers, mechanisms, and consequences. Immunity 2019, 51, 27–41. [Google Scholar] [CrossRef]
  21. Berasain, C.; Castillo, J.; Perugorria, M.J.; Latasa, M.U.; Prieto, J.; Avila, M.A. Inflammation and liver cancer: New molecular links. Ann. N. Y. Acad. Sci. 2009, 1155, 206–221. [Google Scholar] [CrossRef]
  22. Yamashita, M.; Chattopadhyay, S.; Fensterl, V.; Saikia, P.; Wetzel, J.L.; Sen, G.C. Epidermal growth factor receptor is essential for Toll-like receptor 3 signaling. Sci. Signal. 2012, 5, ra50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Fukata, M.; Chen, A.; Vamadevan, A.S.; Cohen, J.; Breglio, K.; Krishnareddy, S.; Hsu, D.; Xu, R.; Harpaz, N.; Dannenberg, A.J.; et al. Toll-like receptor-4 promotes the development of colitis-associated colorectal tumors. Gastroenterology 2007, 133, 1869–1881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Damiano, V.; Caputo, R.; Bianco, R.; D’Armiento, F.P.; Leonardi, A.; De Placido, S.; Bianco, A.R.; Agrawal, S.; Ciardiello, F.; Tortora, G. Novel toll-like receptor 9 agonist induces epidermal growth factor receptor (EGFR) inhibition and synergistic antitumor activity with EGFR inhibitors. Clin. Cancer Res. 2006, 12, 577–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Singh, A.B.; Harris, R.C. Autocrine, paracrine and juxtacrine signaling by EGFR ligands. Cell. Signal. 2005, 17, 1183–1193. [Google Scholar] [CrossRef] [PubMed]
  26. Shostak, K.; Chariot, A. EGFR and NF-κB: Partners in cancer. Trends Mol. Med. 2015, 21, 385–393. [Google Scholar] [CrossRef]
  27. Venkataraman, T.; Frieman, M.B. The role of epidermal growth factor receptor (EGFR) signaling in SARS coronavirus-induced pulmonary fibrosis. Antiviral Res. 2017, 143, 142–150. [Google Scholar] [CrossRef]
  28. Kalinowski, A.; Galen, B.T.; Ueki, I.F.; Sun, Y.; Mulenos, A.; Osafo-Addo, A.; Clark, B.; Joerns, J.; Liu, W.; Nadel, J.A.; et al. Respiratory syncytial virus activates epidermal growth factor receptor to suppress interferon regulatory factor 1-dependent interferon-lambda and antiviral defense in airway epithelium. Mucosal Immunol. 2018, 11, 958–967. [Google Scholar] [CrossRef] [Green Version]
  29. Watanabe, Y.; Someya, T.; Nawa, H. Cytokine hypothesis of schizophrenia pathogenesis: Evidence from human studies and animal models. Psychiatry Clin. Neurosci. 2010, 64, 217–230. [Google Scholar] [CrossRef]
  30. Khandaker, G.M.; Cousins, L.; Deakin, J.; Lennox, B.R.; Yolken, R.; Jones, P.B. Inflammation and immunity in schizophrenia: Implications for pathophysiology and treatment. Lancet Psychiatry 2015, 2, 258–270. [Google Scholar] [CrossRef] [Green Version]
  31. Ermakov, E.A.; Melamud, M.M.; Buneva, V.N.; Ivanova, S.A. Immune system abnormalities in schizophrenia: An integrative view and translational perspectives. Front. Psychiatry 2022, 13, 880568. [Google Scholar] [CrossRef]
  32. Nakamura, J.P.; Schroeder, A.; Gibbons, A.; Sundram, S.; Hill, R.A. Timing of maternal immune activation and sex influence schizophrenia-relevant cognitive constructs and neuregulin and GABAergic pathways. Brain. Behav. Immun. 2022, 100, 70–82. [Google Scholar] [CrossRef]
  33. Idrizi, R.; Malcolm, P.; Weickert, C.S.; Zavitsanou, K.; Sundram, S. Striatal but not frontal cortical up-regulation of the epidermal growth factor receptor in rats exposed to immune activation in utero and cannabinoid treatment in adolescence. Psychiatry Res. 2016, 240, 260–264. [Google Scholar] [CrossRef]
  34. Martins-Filho, P.R.; Tanajura, D.M.; Santos, H.P., Jr.; Santos, V.S. COVID-19 during pregnancy: Potential risk for neurodevelopmental disorders in neonates? Eur. J. Obstet. Gynecol. Reprod. Biol. 2020, 250, 255–256. [Google Scholar] [CrossRef]
  35. Figueiredo, C.P.; Fontes-Dantas, F.L.; da Poian, A.T.; Clarke, J.R. SARS-CoV-2-associated cytokine storm during pregnancy as a possible risk factor for neuropsychiatric disorder development in post-pandemic infants. Neuropharmacology 2021, 201, 108841. [Google Scholar] [CrossRef]
  36. Rhoades, R.; Solomon, S.; Johnson, C.; Teng, S. Impact of SARS-CoV-2 on host factors involved in mental disorders. Front. Microbiol. 2022, 13, 845559. [Google Scholar] [CrossRef]
  37. Yamaguchi, R.; Yamamoto, T.; Sakamoto, A.; Narahara, S.; Sugiuchi, H.; Yamaguchi, Y. Neutrophil elastase enhances IL-12p40 production by lipopolysaccharide-stimulated macrophages via transactivation of the PAR-2/EGFR/TLR4 signaling pathway. Blood Cells Mol. Dis. 2016, 59, 1–7. [Google Scholar] [CrossRef]
  38. Manea, A.; Tanase, L.I.; Raicu, M.; Simionescu, M. Jak/STAT signaling pathway regulates nox1 and nox4-based NADPH oxidase in human aortic smooth muscle cells. Arterioscler. Thromb. Vasc. Biol. 2010, 30, 105–112. [Google Scholar] [CrossRef] [Green Version]
  39. Ketharanathan, T.; Pereira, A.; Reets, U.; Walker, D.; Sundram, S. Brain Changes in NF-ΚB1 and epidermal growth factor system markers at peri-pubescence in the spiny mouse following maternal immune activation. Psychiatry Res. 2021, 295, 113564. [Google Scholar] [CrossRef]
  40. Zunke, F.; Rose-John, S. The shedding protease ADAM17: Physiology and pathophysiology. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 2059–2070. [Google Scholar] [CrossRef]
  41. Bernstein, H.-G.; Keilhoff, G.; Dobrowolny, H.; Lendeckel, U.; Steiner, J. From putative brain tumor marker to high cognitive abilities: Emerging roles of a disintegrin and metalloprotease (ADAM) 12 in the brain. J. Chem. Neuroanat. 2020, 109, 101846. [Google Scholar] [CrossRef]
  42. Roskoski, R. ErbB/HER protein-tyrosine kinases: Structures and small molecule inhibitors. Pharmacol. Res. 2014, 87, 42–59. [Google Scholar] [CrossRef] [PubMed]
  43. Iwakura, Y.; Nawa, H. ErbB1-4-dependent EGF/neuregulin signals and their cross talk in the central nervous system: Pathological implications in schizophrenia and Parkinson’s disease. Front. Cell. Neurosci. 2013, 7, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Liu, P.; Cleveland IV, T.E.; Bouyain, S.; Byrne, P.O.; Longo, P.A.; Leahy, D.J. A single ligand is sufficient to activate EGFR dimers. Proc. Natl. Acad. Sci. USA 2012, 109, 10861–10866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Abe, Y.; Namba, H.; Zheng, Y.; Nawa, H. In situ hybridization reveals developmental regulation of ErbB1-4 mRNA expression in mouse midbrain: Implication of ErbB receptors for dopaminergic neurons. Neuroscience 2009, 161, 95–110. [Google Scholar] [CrossRef] [PubMed]
  46. Seroogy, K.B.; Gall, C.M.; Lee, D.C.; Kornblum, H.I. Proliferative zones of postnatal rat brain express epidermal growth factor receptor mRNA. Brain Res. 1995, 670, 157–164. [Google Scholar] [CrossRef]
  47. Nagano, T.; Namba, H.; Abe, Y.; Aoki, H.; Takei, N.; Nawa, H. In vivo administration of epidermal growth factor and its homologue attenuates developmental maturation of functional excitatory synapses in cortical GABAergic neurons. Eur. J. Neurosci. 2007, 25, 380–390. [Google Scholar] [CrossRef]
  48. Iwakura, Y.; Zheng, Y.; Sibilia, M.; Abe, Y.; Piao, Y.S.; Yokomaku, D.; Wang, R.; Ishizuka, Y.; Takei, N.; Nawa, H. Qualitative and quantitative re-evaluation of epidermal growth factor-ErbB1 action on developing midbrain dopaminergic neurons in vivo and in vitro: Target-derived neurotrophic signaling (Part 1). J. Neurochem. 2011, 118, 45–56. [Google Scholar] [CrossRef]
  49. Vullhorst, D.; Neddens, J.; Karavanova, I.; Tricoire, L.; Petralia, R.S.; McBain, C.J.; Buonanno, A. Selective expression of ErbB4 in interneurons, but not pyramidal cells, of the rodent hippocampus. J. Neurosci. 2009, 29, 12255–12264. [Google Scholar] [CrossRef] [Green Version]
  50. Makinodan, M.; Rosen, K.M.; Ito, S.; Corfas, G. A critical period for social experience-dependent oligodendrocyte maturation and myelination. Science 2012, 337, 1357–1360. [Google Scholar] [CrossRef] [Green Version]
  51. Futamura, T.; Toyooka, K.; Iritani, S.; Niizato, K.; Nakamura, R.; Tsuchiya, K.; Someya, T.; Kakita, A.; Takahashi, H.; Nawa, H. Abnormal expression of epidermal growth factor and its receptor in the forebrain and serum of schizophrenic patients. Mol. Psychiatry 2002, 7, 673–682. [Google Scholar] [CrossRef]
  52. Zhang, X.; Xiao, W.; Chen, K.; Zhao, Y.; Ye, F.; Tang, X.; Du, X. Decreased serum EGF in first-episode and chronic schizophrenia patients: Negative correlation with psychopathology. Sci. Rep. 2020, 10, 6506. [Google Scholar] [CrossRef] [Green Version]
  53. Ikeda, Y.; Yahata, N.; Ito, I.; Nagano, M.; Toyota, T.; Yoshikawa, T.; Okubo, Y.; Suzuki, H. Low serum levels of brain-derived neurotrophic factor and epidermal growth factor in patients with chronic schizophrenia. Schizophr. Res. 2008, 101, 58–66. [Google Scholar] [CrossRef]
  54. Hashimoto, K.; Shimizu, E.; Komatsu, N.; Watanabe, H.; Shinoda, N.; Nakazato, M.; Kumakiri, C.; Okada, S.; Takei, N.; Iyo, M. No changes in serum epidermal growth factor levels in patients with schizophrenia. Psychiatry Res. 2005, 135, 257–260. [Google Scholar] [CrossRef]
  55. Stefansson, H.; Sigurdsson, E.; Steinthorsdottir, V.; Bjornsdottir, S.; Sigmundsson, T.; Ghosh, S.; Brynjolfsson, J.; Gunnarsdottir, S.; Ivarsson, O.; Chou, T.T.; et al. Neuregulin 1 and susceptibility to schizophrenia. Am. J. Hum. Genet. 2002, 71, 877–892. [Google Scholar] [CrossRef] [Green Version]
  56. Munafò, M.R.; Thiselton, D.L.; Clark, T.G.; Flint, J. Association of the NRG1 gene and schizophrenia: A meta-analysis. Mol. Psychiatry 2006, 11, 539–546. [Google Scholar] [CrossRef] [Green Version]
  57. Anttila, S.; Illi, A.; Kampman, O.; Mattila, K.M.; Lehtimäki, T.; Leinonen, E. Association of EGF polymorphism with schizophrenia in Finnish men. Neuroreport 2004, 15, 1215–1218. [Google Scholar] [CrossRef]
  58. Hänninen, K.; Katila, H.; Anttila, S.; Rontu, R.; Maaskola, J.; Hurme, M.; Lehtimäki, T. Epidermal growth factor a61g polymorphism is associated with the age of onset of schizophrenia in male patients. J. Psychiatr. Res. 2007, 41, 8–14. [Google Scholar] [CrossRef]
  59. Watanabe, Y.; Fukui, N.; Muratake, T.; Kaneko, N.; Someya, T. No association of EGF polymorphism with schizophrenia in a Japanese population. Neuroreport 2005, 16, 403–405. [Google Scholar] [CrossRef]
  60. Groenestege, W.M.T.; Thébault, S.; Van Der Wijst, J.; Van Den Berg, D.; Janssen, R.; Tejpar, S.; Van Den Heuvel, L.P.; Van Cutsem, E.; Hoenderop, J.G.; Knoers, N.V.; et al. Impaired basolateral sorting of pro-EGF causes isolated recessive renal hypomagnesemia. J. Clin. Investig. 2007, 117, 2260–2267. [Google Scholar] [CrossRef]
  61. Hahn, C.G.; Wang, H.Y.; Cho, D.S.; Talbot, K.; Gur, R.E.; Berrettini, W.H.; Bakshi, K.; Kamins, J.; Borgmann-Winter, K.E.; Siegel, S.J.; et al. Altered neuregulin 1-erbB4 signaling contributes to NMDA receptor hypofunction in schizophrenia. Nat. Med. 2006, 12, 824–828. [Google Scholar] [CrossRef]
  62. Law, A.J.; Lipska, B.K.; Weickert, C.S.; Hyde, T.M.; Straub, R.E.; Hashimoto, R.; Harrison, P.J.; Kleinman, J.E.; Weinberger, D.R. Neuregulin 1 transcripts are differentially expressed in schizophrenia and regulated by 5′ SNPs associated with the disease. Proc. Natl. Acad. Sci. USA 2006, 103, 6747–6752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Hashimoto, R.; Straub, R.E.; Weickert, C.S.; Hyde, T.M.; Kleinman, J.E.; Weinberger, D.R. Expression analysis of neuregulin-1 in the dorsolateral prefrontal cortex in schizophrenia. Mol. Psychiatry 2004, 9, 299–307. [Google Scholar] [CrossRef] [PubMed]
  64. Shibuya, M.; Komi, E.; Wang, R.; Kato, T.; Watanabe, Y.; Sakai, M.; Ozaki, M.; Someya, T.; Nawa, H. Measurement and comparison of serum neuregulin 1 immunoreactivity in control subjects and patients with schizophrenia: An influence of its genetic polymorphism. J. Neural Transm. 2010, 117, 887–895. [Google Scholar] [CrossRef] [PubMed]
  65. Nawa, H.; Sotoyama, H.; Iwakura, Y.; Takei, N.; Namba, H. Neuropathologic implication of peripheral neuregulin-1 and EGF signals in dopaminergic dysfunction and behavioral deficits relevant to schizophrenia: Their target cells and time window. Biomed. Res. Int. 2014, 2014, 697935. [Google Scholar] [CrossRef] [Green Version]
  66. Patlola, S.R.; Donohoe, G.; McKernan, D.P. The relationship between inflammatory biomarkers and cognitive dysfunction in patients with schizophrenia: A systematic review and meta-analysis. Prog. Neuropsychopharmacol. Biol. Psychiatry 2023, 121, 110668. [Google Scholar] [CrossRef]
  67. Goldsmith, D.R.; Bekhbat, M.; Mehta, N.D.; Felger, J.C. Inflammation-related functional and structural dysconnectivity as a pathway to psychopathology. Biol. Psychiatry 2023, 93, 405–418. [Google Scholar] [CrossRef]
  68. Ishikawa, Y.; Furuyashiki, T. The impact of stress on immune systems and its relevance to mental illness. Neurosci. Res. 2022, 175, 16–24. [Google Scholar] [CrossRef]
  69. Meyer, U. Prenatal poly(i:C) exposure and other developmental immune activation models in rodent systems. Biol. Psychiatry 2014, 75, 307–315. [Google Scholar] [CrossRef]
  70. Nawa, H.; Takei, N. Recent progress in animal modeling of immune inflammatory processes in schizophrenia: Implication of specific cytokines. Neurosci. Res. 2006, 56, 2–13. [Google Scholar] [CrossRef]
  71. Tohmi, M.; Tsuda, N.; Watanabe, Y.; Kakita, A.; Nawa, H. Perinatal inflammatory cytokine challenge results in distinct neurobehavioral alterations in rats: Implication in psychiatric of developmental origin. Neurosci. Res. 2004, 50, 67–75. [Google Scholar] [CrossRef]
  72. Shi, L.; Fatemi, S.H.; Sidwell, R.W.; Patterson, P.H. Maternal influenza infection causes marked behavioral and pharmacological changes in the offspring. J. Neurosci. 2003, 23, 297–302. [Google Scholar] [CrossRef]
  73. Fortier, M.È.; Joober, R.; Luheshi, G.N.; Boksa, P. Maternal exposure to bacterial endotoxin during pregnancy enhances amphetamine-induced locomotion and startle responses in adult rat offspring. J. Psychiatr. Res. 2004, 38, 335–345. [Google Scholar] [CrossRef]
  74. Aoki, H. Neurocognitive impairments of offspring induced by maternal cytokine challenges. Niigata Med. J. 2008, 122, 262–270. [Google Scholar]
  75. Hsiao, E.Y.; Patterson, P.H. Activation of the maternal immune system induces endocrine changes in the placenta via IL-6. Brain Behav. Immun. 2011, 25, 604–615. [Google Scholar] [CrossRef] [Green Version]
  76. Tsuda, N.; Mizuno, M.; Yamanaka, T.; Komurasaki, T.; Yoshimoto, M.; Nawa, H. Common behavioral influences of the ErbB1 ligands transforming growth factor alpha and epiregulin administered to mouse neonates. Brain Dev. 2008, 30, 533–543. [Google Scholar] [CrossRef]
  77. Futamura, T.; Kakita, A.; Tohmi, M.; Sotoyama, H.; Takahashi, H.; Nawa, H. Neonatal perturbation of neurotrophic signaling results in abnormal sensorimotor gating and social interaction in adults: Implication for epidermal growth factor in cognitive development. Mol. Psychiatry 2003, 8, 19–29. [Google Scholar] [CrossRef] [Green Version]
  78. Tohmi, M.; Tsuda, N.; Zheng, Y.; Mizuno, M.; Sotoyama, H.; Shibuya, M.; Kawamura, M.; Kakita, A.; Takahashi, H.; Nawa, H. The cellular and behavioral consequences of interleukin-1 alpha penetration through the blood-brain barrier of neonatal rats: A critical period for efficacy. Neuroscience 2007, 150, 234–250. [Google Scholar] [CrossRef]
  79. Kato, T.; Abe, Y.; Sotoyama, H.; Kakita, A.; Kominami, R.; Hirokawa, S.; Ozaki, M.; Takahashi, H.; Nawa, H. Transient exposure of neonatal mice to neuregulin-1 results in hyperdopaminergic states in adulthood: Implication in neurodevelopmental hypothesis for schizophrenia. Mol. Psychiatry 2011, 16, 307–320. [Google Scholar] [CrossRef] [Green Version]
  80. Kato, T.; Abe, Y.; Hirokawa, S.; Iwakura, Y.; Mizuno, M.; Namba, H.; Nawa, H. Neurobehavioral differences between mice receiving distinct neuregulin variants as neonates; impact on sensitivity to MK-801. Curr. Mol. Med. 2015, 15, 222–236. [Google Scholar] [CrossRef] [Green Version]
  81. Tohmi, M.; Tsuda, N.; Mizuno, M.; Takei, N.; Frankland, P.W.; Nawa, H. Distinct influences of neonatal epidermal growth factor challenge on adult neurobehavioral traits in four mouse strains. Behav. Genet. 2005, 35, 615–629. [Google Scholar] [CrossRef]
  82. Mizuno, M.; Malta, R.S.; Nagano, T.; Nawa, H. Conditioned place preference and locomotor sensitization after repeated administration of cocaine or methamphetamine in rats treated with epidermal growth factor during the neonatal period. Ann. N. Y. Acad. Sci. 2004, 1025, 612–618. [Google Scholar] [CrossRef] [PubMed]
  83. Sotoyama, H.; Zheng, Y.; Iwakura, Y.; Mizuno, M.; Aizawa, M.; Shcherbakova, K.; Wang, R.; Namba, H.; Nawa, H. Pallidal hyperdopaminergic innervation underlying D2 receptor-dependent behavioral deficits in the schizophrenia animal model established by EGF. PLoS ONE 2011, 6, e25831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Sakai, M.; Kashiwahara, M.; Kakita, A.; Nawa, H. An attempt of non-human primate modeling of schizophrenia with neonatal challenges of epidermal growth factor. J. Addict. Res. Ther. 2014, 5, 170. [Google Scholar]
  85. Jodo, E.; Inaba, H.; Narihara, I.; Sotoyama, H.; Kitayama, E.; Yabe, H.; Namba, H.; Eifuku, S.; Nawa, H. Neonatal exposure to an inflammatory cytokine, epidermal growth factor, results in the deficits of mismatch negativity in rats. Sci. Rep. 2019, 9, 7503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Inaba, H.; Kai, R.; Namba, H.; Sotoyama, H.; Jodo, E.; Nin, F.; Hibino, H.; Yabe, H.; Eifuku, S.; Horii, A.; et al. Perinatal epidermal growth factor signal perturbation results in the series of abnormal auditory oscillations and responses relevant to schizophrenia. Schizophr. Bull. Open 2021, 2, sgaa070. [Google Scholar] [CrossRef]
  87. Narihara, I.; Kitajo, K.; Namba, H.; Sotoyama, H.; Inaba, H.; Watanabe, D.; Nawa, H. Rat call-evoked electrocorticographic responses and intercortical phase synchrony impaired in a cytokine-induced animal model for schizophrenia. Neurosci. Res. 2022, 175, 62–72. [Google Scholar] [CrossRef]
  88. Iwakura, Y.; Piao, Y.S.; Mizuno, M.; Takei, N.; Kakita, A.; Takahashi, H.; Nawa, H. Influences of dopaminergic lesion on epidermal growth factor-ErbB signals in Parkinson’s disease and its model: Neurotrophic implication in nigrostriatal neurons. J. Neurochem. 2005, 93, 974–983. [Google Scholar] [CrossRef]
  89. Cohen, A.D.; Zigmond, M.J.; Smith, A.D. Effects of intrastriatal GDNF on the response of dopamine neurons to 6-hydroxydopamine: Time course of protection and neurorestoration. Brain Res. 2011, 1370, 80–88. [Google Scholar] [CrossRef] [Green Version]
  90. Hadjiconstantinou, M.; Fitkin, J.G.; Dalia, A.; Neff, N.H. Epidermal growth factor enhances striatal dopaminergic parameters in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated mouse. J. Neurochem. 1991, 57, 479–482. [Google Scholar] [CrossRef]
  91. Schneider, J.S.; DiStefano, L. Enhanced restoration of striatal dopamine concentrations by combined GM1 ganglioside and neurotrophic factor treatments. Brain Res. 1995, 674, 260–264. [Google Scholar] [CrossRef]
  92. Gerhardt, G.A.; Cass, W.A.; Huettl, P.; Brock, S.; Zhang, Z.; Gash, D.M. GDNF improves dopamine function in the substantia nigra but not the putamen of unilateral MPTP-lesioned rhesus monkeys. Brain Res. 1999, 817, 163–171. [Google Scholar] [CrossRef]
  93. Yu, Z.Q.; Zha, J.H.; Liu, H.M.; Ding, Y.X.; Wang, Y.Q.; Wang, H.J.; Gao, D.S. Effect of intranigral injection of GDNF and EGF on the survival and possible differentiation fate of progenitors and immature neurons in 6-OHDA-lesioned rats. Neurochem. Res. 2009, 34, 2089–2101. [Google Scholar] [CrossRef]
  94. Hidalgo-Figueroa, M.; Bonilla, S.; Gutierrez, F.; Pascual, A.; Lopez-Barneo, J. GDNF is predominantly expressed in the PV+ neostriatal interneuronal ensemble in normal mouse and after injury of the nigrostriatal pathway. J. Neurosci. 2012, 32, 864–872. [Google Scholar] [CrossRef] [Green Version]
  95. Kumar, A.; Kopra, J.; Varendi, K.; Porokuokka, L.L.; Panhelainen, A.; Kuure, S.; Marshall, P.; Karalija, N.; Härma, M.A.; Vilenius; et al. GDNF overexpression from the native locus reveals its role in the nigrostriatal dopaminergic system function. PLoS Genet. 2015, 11, e1005710. [Google Scholar] [CrossRef] [Green Version]
  96. Carlsson, T.; Schindler, F.R.; Höllerhage, M.; Depboylu, C.; Arias-Carrión, O.; Schnurrbusch, S.; Rösler, T.W.; Wozny, W.; Schwall, G.P.; Groebe, K.; et al. Systemic administration of neuregulin-1β1 protects dopaminergic neurons in a mouse model of Parkinson’s disease. J. Neurochem. 2011, 117, 1066–1074. [Google Scholar] [CrossRef]
  97. Depboylu, C.; Rösler, T.W.; de Andrade, A.; Oertel, W.H.; Höglinger, G.U. Systemically administered neuregulin-1β1 rescues nigral dopaminergic neurons via the ErbB4 receptor tyrosine kinase in MPTP mouse models of Parkinson’s disease. J. Neurochem. 2015, 133, 590–597. [Google Scholar] [CrossRef]
  98. Thuret, S.; Alavian, K.N.; Gassmann, M.; Lloyd, C.K.; Smits, S.M.; Smidt, M.P.; Klein, R.; Dyck, R.H.; Simon, H.H. The neuregulin receptor, ErbB4, is not required for normal development and adult maintenance of the substantia nigra pars compacta. J. Neurochem. 2004, 91, 1302–1311. [Google Scholar] [CrossRef]
  99. Namba, H.; Nagano, T.; Jodo, E.; Eifuku, S.; Horie, M.; Takebayashi, H.; Iwakura, Y.; Sotoyama, H.; Takei, N.; Nawa, H. Epidermal growth factor signals attenuate phenotypic and functional development of neocortical GABA neurons. J. Neurochem. 2017, 142, 886–900. [Google Scholar] [CrossRef] [Green Version]
  100. Abe, Y.; Namba, H.; Kato, T.; Iwakura, Y.; Nawa, H. Neuregulin-1 signals from the periphery regulate AMPA receptor sensitivity and expression in GABAergic interneurons in developing neocortex. J. Neurosci. 2011, 31, 5699–5709. [Google Scholar] [CrossRef] [Green Version]
  101. Ting, A.K.; Chen, Y.; Wen, L.; Yin, D.M.; Shen, C.; Tao, Y.; Liu, X.; Xiong, W.C.; Mei, L. Neuregulin 1 promotes excitatory synapse development and function in GABAergic interneurons. J. Neurosci. 2011, 31, 15–25. [Google Scholar] [CrossRef] [Green Version]
  102. Yang, J.M.; Zhang, J.; Chen, X.J.; Geng, H.Y.; Ye, M.; Spitzer, N.C.; Luo, J.H.; Duan, S.M.; Li, X.M. Development of GABA circuitry of fast-spiking basket interneurons in the medial prefrontal cortex of erbb4-mutant mice. J. Neurosci. 2013, 33, 19724–19733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Skirzewski, M.; Karavanova, I.; Shamir, A.; Erben, L.; Garcia-Olivares, J.; Shin, J.H.; Vullhorst, D.; Alvarez, V.A.; Amara, S.G.; Buonanno, A. ErbB4 signaling in dopaminergic axonal projections increases extracellular dopamine levels and regulates spatial/working memory behaviors. Mol. Psychiatry 2018, 23, 2227–2237. [Google Scholar] [CrossRef] [PubMed]
  104. Yurek, D.M.; Zhang, L.; Fletcher-Turner, A.; Seroogy, K.B. Supranigral injection of neuregulin1-beta induces striatal dopamine overflow. Brain Res. 2004, 1028, 116–119. [Google Scholar] [CrossRef] [PubMed]
  105. Roy, K.; Murtie, J.C.; El-Khodor, B.F.; Edgar, N.; Sardi, S.P.; Hooks, B.M.; Benoit-Marand, M.; Chen, C.; Moore, H.; O’Donnell, P.; et al. Loss of erbB signaling in oligodendrocytes alters myelin and dopaminergic function, a potential mechanism for neuropsychiatric disorders. Proc. Natl. Acad. Sci. USA 2007, 104, 8131–8136. [Google Scholar] [CrossRef] [Green Version]
  106. Ledonne, A.; Nobili, A.; Latagliata, E.C.; Cavallucci, V.; Guatteo, E.; Puglisi-Allegra, S.; D’Amelio, M.; Mercuri, N.B. Neuregulin 1 signalling modulates mGluR1 function inmesencephalic dopaminergic neurons. Mol. Psychiatry 2015, 20, 959–973. [Google Scholar] [CrossRef] [Green Version]
  107. Ledonne, A.; Mercuri, N.B. mGluR1-dependent long term depression in rodent midbrain dopamine neurons is regulated by neuregulin 1/ErbB signaling. Front. Mol. Neurosci. 2018, 11, 346. [Google Scholar] [CrossRef] [Green Version]
  108. Erben, L.; Welday, J.P.; Cronin, M.E.; Murphy, R.; Skirzewski, M.; Vullhorst, D.; Carroll, S.L.; Buonanno, A. Developmental, neurochemical, and behavioral analyses of ErbB4 Cyt-1 knockout mice. J. Neurochem. 2022, 161, 435–452. [Google Scholar] [CrossRef]
  109. Skirzewski, M.; Cronin, M.E.; Murphy, R.; Fobbs, W.; Kravitz, A.V.; Buonanno, A. ErbB4 null mice display altered mesocorticolimbic and nigrostriatal dopamine levels as well as deficits in cognitive and motivational behaviors. eNeuro 2020, 7. [Google Scholar] [CrossRef]
  110. Oyagi, A.; Oida, Y.; Kakefuda, K.; Shimazawa, M.; Shioda, N.; Moriguchi, S.; Kitaichi, K.; Nanba, D.; Yamaguchi, K.; Furuta, Y.; et al. Generation and characterization of conditional heparin-binding EGF-like growth factor knockout mice. PLoS ONE 2009, 4, e7461. [Google Scholar] [CrossRef] [Green Version]
  111. Golani, I.; Tadmor, H.; Buonanno, A.; Kremer, I.; Shamir, A. Disruption of the ErbB signaling in adolescence increases striatal dopamine levels and affects learning and hedonic-like behavior in the adult mouse. Eur. Neuropsychopharmacol. 2014, 24, 1808–1818. [Google Scholar] [CrossRef]
  112. Huang, C.M.; Kao, L.S. Nerve growth factor, epidermal growth factor, and insulin differentially potentiate ATP-induced [Ca2+]i rise and dopamine secretion in PC12 cells. J. Neurochem. 1996, 66, 124–130. [Google Scholar] [CrossRef]
  113. Farkas, L.M.; Krieglstein, K. Heparin-binding epidermal growth factor-like growth factor (HB-EGF) regulates survival of midbrain dopaminergic neurons. J. Neural Transm. 2002, 109, 267–277. [Google Scholar] [CrossRef]
  114. Mizuno, M.; Sotoyama, H.; Narita, E.; Kawamura, H.; Namba, H.; Zheng, Y.; Eda, T.; Nawa, H. A cyclooxygenase-2 inhibitor ameliorates behavioral impairments induced by striatal administration of epidermal growth factor. J. Neurosci. 2007, 27, 10116–10127. [Google Scholar] [CrossRef] [Green Version]
  115. Fallon, J.H.; Loughlin, S.E. Substantia nigra. In The Rat Nervous System; Pasxinos, G., Ed.; Academic Press: San Diego, CA, USA, 1995; p. 229. [Google Scholar]
  116. Namba, H.; Tomiyama, K.; Nawa, H. Abnormal development of nigral dopamine activities in a cytokine-induced schizophrenia model; implication for its postpubertal onset. Soc. Neurosci. Abstr. 2017, 47, 258.13. [Google Scholar]
  117. Namba, H.; Nawa, H. Post-pubertal difference in nigral dopaminergic cells firing in the schizophrenia model prepared by perinatal challenges of a cytokine, EGF. Neuroscience 2020, 441, 22–32. [Google Scholar] [CrossRef]
  118. McCutcheon, J.E.; Marinelli, M. Age matters. Eur. J. Neurosci. 2009, 29, 997–1014. [Google Scholar] [CrossRef]
  119. McCutcheon, J.E.; Conrad, K.L.; Carr, S.B.; Ford, K.A.; McGehee, D.S.; Marinelli, M. Dopamine neurons in the ventral tegmental area fire faster in adolescent rats than in adults. J. Neurophysiol. 2012, 108, 1620–1630. [Google Scholar] [CrossRef] [Green Version]
  120. Namba, H.; Okubo, T.; Nawa, H. Perinatal exposure to neuregulin-1 results in disinhibition of adult midbrain dopaminergic neurons: Implication in schizophrenia modeling. Sci. Rep. 2016, 6, 22606. [Google Scholar] [CrossRef] [Green Version]
  121. Giannopoulou, I.; Pagida, M.A.; Briana, D.D.; Panayotacopoulou, M.T. Perinatal hypoxia as a risk factor for psychopathology later in life: The role of dopamine and neurotrophins. Hormones 2018, 17, 25–32. [Google Scholar] [CrossRef] [Green Version]
  122. Lodge, D.J.; Grace, A.A. Divergent activation of ventromedial and ventrolateral dopamine systems in animal models of amphetamine sensitization and schizophrenia. Int. J. Neuropsychopharmacol. 2012, 15, 69–76. [Google Scholar] [CrossRef] [Green Version]
  123. Winter, C.; Djodari-Irani, A.; Sohr, R.; Morgenstern, R.; Feldon, J.; Juckel, G.; Meyer, U. Prenatal immune activation leads to multiple changes in basal neurotransmitter levels in the adult brain: Implications for brain disorders of neurodevelopmental origin such as schizophrenia. Int. J. Neuropsychopharmacol. 2009, 12, 513–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Swerdlow, N.R.; Mansbach, R.S.; Geyer, M.A.; Pulvirenti, L.; Koob, G.F.; Braff, D.L. Amphetamine disruption of prepulse inhibition of acoustic startle is reversed by depletion of mesolimbic dopamine. Psychopharmacology 1990, 100, 413–416. [Google Scholar] [CrossRef] [PubMed]
  125. Martinez, Z.A.; Ellison, G.D.; Geyer, M.A.; Swerdlow, N.R. Effects of sustained cocaine exposure on sensorimotor gating of startle in rats. Psychopharmacology 1999, 142, 253–260. [Google Scholar] [CrossRef] [PubMed]
  126. Swerdlow, N.R.; Keith, V.A.; Braff, D.L.; Geyer, M.A. Effects of spiperone, raclopride, SCH 23390 and clozapine on apomorphine inhibition of sensorimotor gating of the startle response in the rat. J. Pharmacol. Exp. Ther. 1991, 256, 530–536. [Google Scholar] [PubMed]
  127. Swerdlow, N.R.; Geyer, M.A. Clozapine and haloperidol in an animal model of sensorimotor gating deficits in schizophrenia. Pharmacol. Biochem. Behav. 1993, 44, 741–744. [Google Scholar] [CrossRef]
  128. Wan, F.J.; Taaid, N.; Swerdlow, N.R. Do D1/D2 interactions regulate prepulse inhibition in rats? Neuropsychopharmacology 1996, 14, 265–274. [Google Scholar]
  129. Geyer, M.A.; Krebs-Thomson, K.; Braff, D.L.; Swerdlow, N.R. Pharmacological studies of prepulse inhibition models of sensorimotor gating deficits in schizophrenia: A decade in review. Psychopharmacology 2001, 156, 117–154. [Google Scholar] [CrossRef]
  130. Ralph, R.J.; Paulus, M.P.; Fumagalli, F.; Caron, M.G.; Geyer, M.A. Prepulse inhibition deficits and perseverative motor patterns in dopamine transporter knock-out mice: Differential effects of D1 and D2 receptor antagonists. J. Neurosci. 2001, 21, 305–313. [Google Scholar] [CrossRef] [Green Version]
  131. Ralph, R.J.; Varty, G.B.; Kelly, M.A.; Wang, Y.M.; Caron, M.G.; Rubinstein, M.; Grandy, D.K.; Low, M.J.; Geyer, M.A. The dopamine D2, but not D3 or D4, receptor subtype is essential for the disruption of prepulse inhibition produced by amphetamine in mice. J. Neurosci. 1999, 19, 4627–4633. [Google Scholar] [CrossRef] [Green Version]
  132. Wan, F.J.; Geyer, M.A.; Swerdlow, N.R. Accumbens D2 modulation of sensorimotor gating in rats: Assessing anatomical localization. Pharmacol. Biochem. Behav. 1994, 49, 155–163. [Google Scholar] [CrossRef]
  133. Swerdlow, N.R.; Braff, D.L.; Masten, V.L.; Geyer, M.A. Schizophrenic-like sensorimotor gating abnormalities in rats following dopamine infusion into the nucleus accumbens. Psychopharmacology 1990, 101, 414–420. [Google Scholar] [CrossRef]
  134. Wan, F.J.; Swerdlow, N.R. Intra-accumbens infusion of quinpirole impairs sensorimotor gating of acoustic startle in rats. Psychopharmacology 1993, 113, 103–109. [Google Scholar] [CrossRef]
  135. Swerdlow, N.R.; Braff, D.L.; Geyer, M.A. GABAergic projection from nucleus accumbens to ventral pallidum mediates dopamine-induced sensorimotor gating deficits of acoustic startle in rats. Brain Res. 1990, 532, 146–150. [Google Scholar] [CrossRef]
  136. Swerdlow, N.R.; Caine, S.B.; Geyer, M.A. Regionally selective effects of intracerebral dopamine infusion on sensorimotor gating of the startle reflex in rats. Psychopharmacology 1992, 108, 189–195. [Google Scholar] [CrossRef]
  137. Sotoyama, H.; Namba, H.; Kobayashi, Y.; Hasegawa, T.; Watanabe, D.; Nakatsukasa, E.; Sakimura, K.; Furuyashiki, T.; Nawa, H. Resting-state dopaminergic cell firing in the ventral tegmental area negatively regulates affiliative social interactions in a developmental animal model of schizophrenia. Transl. Psychiatry 2021, 11, 236. [Google Scholar] [CrossRef]
  138. Sotoyama, H.; Inaba, H.; Iwakura, Y.; Namba, H.; Takei, N.; Sasaoka, T.; Nawa, H. The dual role of dopamine in the modulation of information processing in the prefrontal cortex underlying social behavior. FASEB J. 2022, 36, e22160. [Google Scholar] [CrossRef]
  139. Fauchey, V.; Jaber, M.; Caron, M.G.; Bloch, B.; Le Moine, C. Differential regulation of the dopamine D1, D2 and D3 receptor gene expression and changes in the phenotype of the striatal neurons in mice lacking the dopamine transporter. Eur. J. Neurosci. 2000, 12, 19–26. [Google Scholar] [CrossRef]
  140. Dumartin, B.; Jaber, M.; Gonon, F.; Caron, M.G.; Giros, B.; Bloch, B. Dopamine tone regulates D1 receptor trafficking and delivery in striatal neurons in dopamine transporter-deficient mice. Proc. Natl. Acad. Sci. USA 2000, 97, 1879–1884. [Google Scholar] [CrossRef] [Green Version]
  141. Gunaydin, L.A.; Grosenick, L.; Finkelstein, J.C.; Kauvar, I.V.; Fenno, L.E.; Adhikari, A.; Lammel, S.; Mirzabekov, J.J.; Airan, R.D.; Zalocusky, K.A. Natural neural projection dynamics underlying social behavior. Cell 2014, 157, 1535–1551. [Google Scholar] [CrossRef] [Green Version]
  142. Chaudhury, D.; Walsh, J.J.; Friedman, A.K.; Juarez, B.; Ku, S.M.; Koo, J.W.; Ferguson, D.; Tsai, H.C.; Pomeranz, L.; Christoffel, D.J.; et al. Rapid regulation of depression-related behaviours by control of midbrain dopamine neurons. Nature 2013, 493, 532–536. [Google Scholar] [CrossRef] [Green Version]
  143. Lipska, B.K.; Jaskiw, G.E.; Chrapusta, S.; Karoum, F.; Weinberger, D.R. Ibotenic acid lesion of the ventral hippocampus differentially affects dopamine and its metabolites in the nucleus accumbens and prefrontal cortex in the rat. Brain Res. 1992, 585, 1–6. [Google Scholar] [CrossRef] [PubMed]
  144. Rueter, L.E.; Ballard, M.E.; Gallagher, K.B.; Basso, A.M.; Curzon, P.; Kohlhaas, K.L. Chronic low dose risperidone and clozapine alleviate positive but not negative symptoms in the rat neonatal ventral hippocampal lesion model of schizophrenia. Psychopharmacology 2004, 176, 312–319. [Google Scholar] [CrossRef] [PubMed]
  145. Coyle, J.T. Glutamate and schizophrenia: Beyond the dopamine hypothesis. Cell Mol. Neurobiol. 2006, 26, 365–384. [Google Scholar] [CrossRef] [PubMed]
  146. Volavka, J.; Czobor, P.; Sheitman, B.; Lindenmayer, J.P.; Citrome, L.; McEvoy, J.P.; Cooper, T.B.; Chakos, M.; Lieberman, J.A. Clozapine, olanzapine, risperidone, and haloperidol in the treatment of patients with chronic schizophrenia and schizoaffective disorder. Am. J. Psychiatry 2002, 159, 255–262. [Google Scholar] [CrossRef] [PubMed]
  147. Tollefson, G.D.; Sanger, T.M. Negative symptoms: A path analytic approach to a double-blind, placebo- and haloperidol-controlled clinical trial with olanzapine. Am. J. Psychiatry 1997, 154, 466–474. [Google Scholar]
  148. Harrow, M.; Yonan, C.A.; Sands, J.R.; Marengo, J. Depression in schizophrenia: Are neuroleptics, akinesia, or anhedonia involved? Schizophr. Bull. 1994, 20, 327–338. [Google Scholar] [CrossRef] [Green Version]
  149. Segura-Aguilar, J.; Paris, I.; Muñoz, P.; Ferrari, E.; Zecca, L.; Zucca, F.A. Protective and toxic roles of dopamine in parkinson’s disease. J. Neurochem. 2014, 129, 898–915. [Google Scholar] [CrossRef]
  150. Ferrer, I.; Alcántara, S.; Ballabriga, J.; Olivé, M.; Blanco, R.; Rivera, R.; Carmona, M.; Berruezo, M.; Pitarch, S.; Planas, A.M. Transforming growth factor-alpha (TGF-Alpha) and epidermal growth factor-receptor (EGF-R) immunoreactivity in normal and pathologic brain. Prog. Neurobiol. 1996, 49, 99–123. [Google Scholar] [CrossRef]
  151. Qu, W.-S.; Liu, J.-L.; Li, C.-Y.; Li, X.; Xie, M.-J.; Wang, W.; Tian, D.-S. Rapidly activated epidermal growth factor receptor mediates lipopolysaccharide-triggered migration of microglia. Neurochem. Int. 2015, 90, 85–92. [Google Scholar] [CrossRef]
  152. Zhang, J.; Wu, J.; Fujita, Y.; Yao, W.; Ren, Q.; Yang, C.; Li, S.; Shirayama, Y.; Hashimoto, K. Antidepressant effects of TrkB ligands on depression-like behavior and dendritic changes in mice after inflammation. Int. J. Neuropsychopharmacol. 2014, 18, pyu077. [Google Scholar] [CrossRef]
  153. Furuyashiki, T.; Kitaoka, S. Neural mechanisms underlying adaptive and maladaptive consequences of stress: Roles of dopaminergic and inflammatory responses. Psychiatry Clin. Neurosci. 2019, 73, 669–675. [Google Scholar] [CrossRef] [Green Version]
  154. Wang, H.; He, Y.; Sun, Z.; Ren, S.; Liu, M.; Wang, G.; Yang, J. Microglia in depression: An overview of microglia in the pathogenesis and treatment of depression. J. Neuroinflamm. 2022, 19, 132. [Google Scholar] [CrossRef]
  155. Okubo, Y.; Suhara, T.; Suzuki, K.; Kobayashi, K.; Inoue, O.; Terasaki, O.; Someya, Y.; Sassa, T.; Sudo, Y.; Matsushima, E.; et al. Decreased prefrontal dopamine D1 receptors in schizophrenia revealed by PET. Nature 1997, 385, 634–636. [Google Scholar] [CrossRef]
  156. Stenkrona, P.; Matheson, G.J.; Halldin, C.; Cervenka, S.; Farde, L. D1-dopamine receptor availability in first-episode neuroleptic naive psychosis patients. Int. J. Neuropsychopharmacol. 2019, 22, 415–425. [Google Scholar] [CrossRef] [Green Version]
  157. Kosaka, J.; Takahashi, H.; Ito, H.; Takano, A.; Fujimura, Y.; Matsumoto, R.; Nozaki, S.; Yasuno, F.; Okubo, Y.; Kishimoto, T.; et al. Decreased binding of [11C]NNC112 and [11C]SCH23390 in patients with chronic schizophrenia. Life Sci. 2010, 86, 814–818. [Google Scholar] [CrossRef]
  158. Hirvonen, J.; van Erp, T.G.; Huttunen, J.; Aalto, S.; Någren, K.; Huttunen, M.; Lönnqvist, J.; Kaprio, J.; Cannon, T.D.; Hietala, J. Brain dopamine d1 receptors in twins discordant for schizophrenia. Am. J. Psychiatry 2006, 163, 1747–1753. [Google Scholar] [CrossRef]
  159. Lindström, L.H.; Gefvert, O.; Hagberg, G.; Lundberg, T.; Bergström, M.; Hartvig, P.; Långström, B. Increased dopamine synthesis rate in medial prefrontal cortex and striatum in schizophrenia indicated by L-(beta-11C) DOPA and PET. Biol. Psychiatry 1999, 46, 681–688. [Google Scholar] [CrossRef]
  160. Watanabe, Y.; Tanaka, H.; Tsukabe, A.; Kunitomi, Y.; Nishizawa, M.; Hashimoto, R.; Yamamori, H.; Fujimoto, M.; Fukunaga, M.; Tomiyama, N. Neuromelanin magnetic resonance imaging reveals increased dopaminergic neuron activity in the substantia nigra of patients with schizophrenia. PLoS ONE 2014, 9, e104619. [Google Scholar] [CrossRef] [Green Version]
  161. Shibata, E.; Sasaki, M.; Tohyama, K.; Otsuka, K.; Endoh, J.; Terayama, Y.; Sakai, A. Use of neuromelanin-sensitive MRI to distinguish schizophrenic and depressive patients and healthy individuals based on signal alterations in the substantia nigra and locus ceruleus. Biol. Psychiatry 2008, 64, 401–406. [Google Scholar] [CrossRef]
  162. Mizuno, M.; Sotoyama, H.; Namba, H.; Shibuya, M.; Eda, T.; Wang, R.; Okubo, T.; Nagata, K.; Iwakura, Y.; Nawa, H. ErbB inhibitors ameliorate behavioral impairments of an animal model for schizophrenia: Implication of their dopamine-modulatory actions. Transl. Psychiatry 2013, 3, e252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Mizuno, M.; Iwakura, Y.; Shibuya, M.; Zheng, Y.; Eda, T.; Kato, T.; Takasu, Y.; Nawa, H. Antipsychotic potential of quinazoline ErbB1 inhibitors in a schizophrenia model established with neonatal hippocampal lesioning. J. Pharmacol. Sci. 2010, 114, 320–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Tadmor, H.; Golani, I.; Doron, R.; Kremer, I.; Shamir, A. ErbB signaling antagonist ameliorates behavioral deficit induced by phencyclidine (PCP) in mice, without affecting metabolic syndrome markers. Prog. Neuropsychopharmacol. Biol. Psychiatry 2018, 82, 322–331. [Google Scholar] [CrossRef] [PubMed]
  165. Mizuno, M.; Kawamura, H.; Takei, N.; Nawa, H. The anthraquinone derivative Emodin ameliorates neurobehavioral deficits of a rodent model for schizophrenia. J. Neural Transm. 2008, 115, 521–530. [Google Scholar] [CrossRef]
  166. Kobayashi, Y.; Iwakura, Y.; Sotoyama, H.; Kitayama, E.; Takei, N.; Someya, T.; Nawa, H. Clozapine-dependent inhibition of EGF/neuregulin receptor (ErbB) kinases. Transl. Psychiatry 2019, 9, 181. [Google Scholar] [CrossRef] [Green Version]
  167. de Castro, G., Jr.; Awada, A. Side effects of anti-cancer molecular-targeted therapies (not monoclonal antibodies). Curr. Opin. Oncol. 2006, 18, 307–315. [Google Scholar] [CrossRef]
  168. Pandey, A.; Kalita, K.N. Treatment-resistant schizophrenia: How far have we traveled? Front. Psychiatry 2022, 13, 994425. [Google Scholar] [CrossRef]
  169. Leucht, S.; Corves, C.; Arbter, D.; Engel, R.R.; Li, C.; Davis, J.M. Second-generation versus first-generation antipsychotic drugs for schizophrenia: A meta-analysis. Lancet 2009, 373, 31–41. [Google Scholar] [CrossRef]
  170. Pereira, A.; Sugiharto-Winarno, A.; Zhang, B.; Malcolm, P.; Fink, G.; Sundram, S. Clozapine induction of ERK1/2 cell signalling via the EGF receptor in mouse prefrontal cortex and striatum is distinct from other antipsychotic drugs. Int. J. Neuropsychopharmacol. 2012, 15, 1149–1160. [Google Scholar] [CrossRef] [Green Version]
  171. Pereira, A.; Zhang, B.; Malcolm, P.; Sundram, S. Clozapine regulation of p90RSK and c-Fos signaling via the ErbB1-ERK pathway is distinct from olanzapine and haloperidol in mouse cortex and striatum. Prog. Neuropsychopharmacol. Biol. Psychiatry 2013, 40, 353–363. [Google Scholar] [CrossRef]
  172. Fitton, R.; Sweetman, J.; Heseltine-Carp, W.; van der Feltz-Cornelis, C. Anti-inflammatory medications for the treatment of mental disorders: A scoping review. Brain Behav. Immun. 2022, 26, 100518. [Google Scholar] [CrossRef]
  173. Diwanji, D.; Trenker, R.; Thaker, T.M.; Wang, F.; Agard, D.A.; Verba, K.A.; Jura, N. Structures of the HER2-HER3-NRG1β complex reveal a dynamic dimer interface. Nature 2021, 600, 339–343. [Google Scholar] [CrossRef] [PubMed]
  174. Haddad, Y.; Remes, M.; Adam, V.; Heger, Z. Toward structure-based drug design against the epidermal growth factor receptor (EGFR). Drug Discov. Today 2021, 26, 289–295. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Groups and members of the EGF family and their receptors. Many EGF-like peptides structurally form the EGF family. These peptides are released from membrane-anchored precursor proteins via ectodomain shedding. These peptides bind to their receptor ErbB1-4 (Her 1-4) and induce intracellular signaling, although their interaction with ErbB2 is modest [16]. EGF receptor is abbreviated to EGFR alternatively. The localization of these receptors is shown in this figure. Abbreviations: EGF (epidermal growth factor), HB-EGF (heparin-binding EGF-like factor), TGFα (transforming growth factor alpha), VGF (vaccinia virus growth factor), CGF (cowpox growth factor), SPGF (smallpox growth factor), SFGF (shope fibroma virus growth factor), MGF (myxoma virus growth factor).
Figure 1. Groups and members of the EGF family and their receptors. Many EGF-like peptides structurally form the EGF family. These peptides are released from membrane-anchored precursor proteins via ectodomain shedding. These peptides bind to their receptor ErbB1-4 (Her 1-4) and induce intracellular signaling, although their interaction with ErbB2 is modest [16]. EGF receptor is abbreviated to EGFR alternatively. The localization of these receptors is shown in this figure. Abbreviations: EGF (epidermal growth factor), HB-EGF (heparin-binding EGF-like factor), TGFα (transforming growth factor alpha), VGF (vaccinia virus growth factor), CGF (cowpox growth factor), SPGF (smallpox growth factor), SFGF (shope fibroma virus growth factor), MGF (myxoma virus growth factor).
Biomolecules 13 00372 g001
Figure 2. Interaction between Toll-like receptor (TLR) and EGF receptor tyrosine kinase (EGFR/ErbB1) in chronic inflammation and carcinogenesis. Many reports confirm the contribution of EGFR signaling to TLR-mediated chronic inflammation [37,38]. In addition, other studies indicate the requirement of EGFR in the TLR signaling itself [22,23,24]. EGFR signaling induces NADPH oxidase (NOX) activation resulting in the production of reactive oxygen species (ROS) and hydrogen peroxide, which inactivate protein phosphatases (SHP2, PTP1B, PTEN, etc.), elevating basal phosphorylation (P) or enhancing induced phosphorylation (P) of adaptor or signaling molecules: IRF, IkB, STAT, MAP, etc. Alternatively, ROS reacts with DNA, producing DNA mutation and leading to carcinogenesis. Elevation of their phosphorylation results in the DNA binding of NFkB, STAT, and AP-1 [39], leading to cell proliferation and gene expression of NOX, cyclooxygenase (COX), and cytokines and worsening inflammation. These promote the production of prostaglandins (PGE) and lysophosphatidic acid (LPA), which bind to their receptors of G-protein-coupled receptors (GPCRs). Activated protein kinase C (PKC) transactivates metalloproteases in a disintegrin and metalloprotease (ADAM) 9, 10, 12, or 17 [40,41]. ADAM cleaves the membrane-anchored EGF precursor or precursors of other members of the EGF family such as HB-EGF and TGFα and liberates mature EGF or its homolog, activating EGFR. These feed-forward and feed-back interactions between TLR and EGFR are implicated in the pathogenesis of chronic inflammation as well as in carcinogenesis.
Figure 2. Interaction between Toll-like receptor (TLR) and EGF receptor tyrosine kinase (EGFR/ErbB1) in chronic inflammation and carcinogenesis. Many reports confirm the contribution of EGFR signaling to TLR-mediated chronic inflammation [37,38]. In addition, other studies indicate the requirement of EGFR in the TLR signaling itself [22,23,24]. EGFR signaling induces NADPH oxidase (NOX) activation resulting in the production of reactive oxygen species (ROS) and hydrogen peroxide, which inactivate protein phosphatases (SHP2, PTP1B, PTEN, etc.), elevating basal phosphorylation (P) or enhancing induced phosphorylation (P) of adaptor or signaling molecules: IRF, IkB, STAT, MAP, etc. Alternatively, ROS reacts with DNA, producing DNA mutation and leading to carcinogenesis. Elevation of their phosphorylation results in the DNA binding of NFkB, STAT, and AP-1 [39], leading to cell proliferation and gene expression of NOX, cyclooxygenase (COX), and cytokines and worsening inflammation. These promote the production of prostaglandins (PGE) and lysophosphatidic acid (LPA), which bind to their receptors of G-protein-coupled receptors (GPCRs). Activated protein kinase C (PKC) transactivates metalloproteases in a disintegrin and metalloprotease (ADAM) 9, 10, 12, or 17 [40,41]. ADAM cleaves the membrane-anchored EGF precursor or precursors of other members of the EGF family such as HB-EGF and TGFα and liberates mature EGF or its homolog, activating EGFR. These feed-forward and feed-back interactions between TLR and EGFR are implicated in the pathogenesis of chronic inflammation as well as in carcinogenesis.
Biomolecules 13 00372 g002
Figure 3. Age-dependent alteration of firing activity of dopamine neurons. Baseline burst firing is reduced in the post-pubertal stage in parallel with the increase in SK channel expression in normal development. In contrast to the profile of normal animals, baseline burst firing is elevated and the expression of SK channels is reduced in the EGF model at their post-pubertal stage [117]. Accordingly, event-triggered and stress-evoked fractions of dopamine (DA) release are relatively suppressed with the given high baseline release and might be implicated in their stress vulnerability.
Figure 3. Age-dependent alteration of firing activity of dopamine neurons. Baseline burst firing is reduced in the post-pubertal stage in parallel with the increase in SK channel expression in normal development. In contrast to the profile of normal animals, baseline burst firing is elevated and the expression of SK channels is reduced in the EGF model at their post-pubertal stage [117]. Accordingly, event-triggered and stress-evoked fractions of dopamine (DA) release are relatively suppressed with the given high baseline release and might be implicated in their stress vulnerability.
Biomolecules 13 00372 g003
Figure 4. Abnormal basal ganglia circuits of the EGF model. In the EGF model, the globus pallidus receives hyper-dopamine innervation of the A9 pathway. Blockade of dopamine D2 receptors in the globus pallidus normalizes the local dopamine transmission and elevates GABA release from the terminals of medium spiny neurons (MSN). The following enhanced inhibition of GP neurons ameliorates the deficit in prepulse inhibition (PPI) [83]. STR: striatum; GP: globus pallidus; SNc: substantia nigra pars compacta; SNr: substantia nigra pars reticulata. Note; the two major circuits of direct and indirect pathways are present in the basal ganglia, counteractively regulating movement, attention, sensory filtering, etc. The direct and indirect pathways stem from the two types of MSN in STR which express dopamine D1 and D2 receptors, respectively.
Figure 4. Abnormal basal ganglia circuits of the EGF model. In the EGF model, the globus pallidus receives hyper-dopamine innervation of the A9 pathway. Blockade of dopamine D2 receptors in the globus pallidus normalizes the local dopamine transmission and elevates GABA release from the terminals of medium spiny neurons (MSN). The following enhanced inhibition of GP neurons ameliorates the deficit in prepulse inhibition (PPI) [83]. STR: striatum; GP: globus pallidus; SNc: substantia nigra pars compacta; SNr: substantia nigra pars reticulata. Note; the two major circuits of direct and indirect pathways are present in the basal ganglia, counteractively regulating movement, attention, sensory filtering, etc. The direct and indirect pathways stem from the two types of MSN in STR which express dopamine D1 and D2 receptors, respectively.
Biomolecules 13 00372 g004
Figure 5. The effect of basal dopamine release on behavioral regulation. Releases of dopamine (DA) are regulated by two independent modes: tonic and phasic activities of dopamine neurons. In the resting state (a baseline condition), tonic activity contributes to maintaining basal dopamine levels in the targets. Stressful events cause a phasic activity to trigger dopamine release and acutely increase dopamine levels. In contrast, a higher basal dopamine level at the resting state decreases the dynamic range of phasic dopamine efflux, attenuating the dopamine-dependent behavioral responses (Presynaptic Mechanism). Alternatively, the increased basal dopamine level causes the internalization of dopamine receptors in the targets and downregulates the receptor signaling. Even if the event-related phasic dopamine efflux occurs, the responses of the postsynaptic dopamine receptors are attenuated (Postsynaptic Mechanism).
Figure 5. The effect of basal dopamine release on behavioral regulation. Releases of dopamine (DA) are regulated by two independent modes: tonic and phasic activities of dopamine neurons. In the resting state (a baseline condition), tonic activity contributes to maintaining basal dopamine levels in the targets. Stressful events cause a phasic activity to trigger dopamine release and acutely increase dopamine levels. In contrast, a higher basal dopamine level at the resting state decreases the dynamic range of phasic dopamine efflux, attenuating the dopamine-dependent behavioral responses (Presynaptic Mechanism). Alternatively, the increased basal dopamine level causes the internalization of dopamine receptors in the targets and downregulates the receptor signaling. Even if the event-related phasic dopamine efflux occurs, the responses of the postsynaptic dopamine receptors are attenuated (Postsynaptic Mechanism).
Biomolecules 13 00372 g005
Table 1. Antipsychotic-like effects of drugs acting on ErbB receptor kinases.
Table 1. Antipsychotic-like effects of drugs acting on ErbB receptor kinases.
DrugTargetAssay ModelEffectsReferences
gefitinibErbB1EGF, VHLPPI, LI[162,163]
PD153035ErbB1/B2EGF, VHLPPI[162,163]
erlotinibErbB1VHLPPI[163]
JNJ28871063pan-ErbBPCPSocial Interaction[164]
emodintyrosine kinasesEGF, VHLPPI, LI[165]
clozapinepan-ErbBin vitro kinaseTyr phosphorylation[166]
Antipsychotic-like activities of the agents inhibiting ErbB receptor kinases are tested in various animal models for schizophrenia or the in vitro system. Abbreviations used: EGF: EGF model established by its perinatal challenges, VHL: a schizophrenia model established by ventral hippocampal lesioning, PCP: a model established by sub-chronic phencyclidine administration, PPI: prepulse inhibition, LI: latent inhibition of fear learning.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sotoyama, H.; Namba, H.; Tohmi, M.; Nawa, H. Schizophrenia Animal Modeling with Epidermal Growth Factor and Its Homologs: Their Connections to the Inflammatory Pathway and the Dopamine System. Biomolecules 2023, 13, 372. https://doi.org/10.3390/biom13020372

AMA Style

Sotoyama H, Namba H, Tohmi M, Nawa H. Schizophrenia Animal Modeling with Epidermal Growth Factor and Its Homologs: Their Connections to the Inflammatory Pathway and the Dopamine System. Biomolecules. 2023; 13(2):372. https://doi.org/10.3390/biom13020372

Chicago/Turabian Style

Sotoyama, Hidekazu, Hisaaki Namba, Manavu Tohmi, and Hiroyuki Nawa. 2023. "Schizophrenia Animal Modeling with Epidermal Growth Factor and Its Homologs: Their Connections to the Inflammatory Pathway and the Dopamine System" Biomolecules 13, no. 2: 372. https://doi.org/10.3390/biom13020372

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