Next Article in Journal
The Utilization of the SaLux19-Based Loop-Mediated Isothermal Amplification (LAMP) Assay for the Rapid and Sensitive Identification of Minute Amounts of a Biological Specimen
Previous Article in Journal
Predictors of Mortality in Acute Myocardial Infarction Complicated by Cardiogenic Shock despite Intra-Aortic Balloon Pump: Opportunities for Advanced Mechanical Circulatory Support in Asia
Previous Article in Special Issue
Cardiovascular Disease from Pathophysiology to Risk Estimation: Is Inflammation Estimated through Perivascular Attenuation on Computed Tomography the Key?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Pathophysiological Underpinnings of Gamma-Band Alterations in Psychiatric Disorders

by
Annalisa Palmisano
1,2,3,*,
Siddhartha Pandit
2,
Carmelo L. Smeralda
2,4,
Ilya Demchenko
5,6,
Simone Rossi
4,
Lorella Battelli
7,8,
Davide Rivolta
3,
Venkat Bhat
5,6,9 and
Emiliano Santarnecchi
2,10
1
Chair of Lifespan Developmental Neuroscience, Faculty of Psychology, TUD Dresden University of Technology, 01069 Dresden, Germany
2
Precision Neuroscience and Neuromodulation Program, Gordon Center for Medical Imaging, Massachusetts General Hospital, Harvard Medical School, Boston, MA 02115, USA
3
Department of Education, Psychology, and Communication, University of Bari Aldo Moro, 70121 Bari, Italy
4
Siena Brain Investigation and Neuromodulation (SI-BIN) Laboratory, Department of Medicine, Surgery and Neuroscience, Neurology and Clinical Neurophysiology Section, University of Siena, 53100 Siena, Italy
5
Interventional Psychiatry Program, St. Michael’s Hospital—Unity Health Toronto, Toronto, ON M5B 1W8, Canada
6
Institute of Medical Science, Temerty Faculty of Medicine, University of Toronto, Toronto, ON M5S 1A1, Canada
7
Berenson-Allen Center for Noninvasive Brain Stimulation, Department of Neurology, Beth Israel Deaconess Medical Center, Harvard Medical School, Boston, MA 02215, USA
8
Center for Neuroscience and Cognitive Systems@UniTn, Istituto Italiano di Tecnologia, 38068 Rovereto, Italy
9
Department of Psychiatry, Temerty Faculty of Medicine, University of Toronto, Toronto, ON M5S 1A1, Canada
10
Department of Neurology and Radiology, Massachusetts General Hospital, Boston, MA 02114, USA
*
Author to whom correspondence should be addressed.
Life 2024, 14(5), 578; https://doi.org/10.3390/life14050578
Submission received: 5 February 2024 / Revised: 4 April 2024 / Accepted: 6 April 2024 / Published: 30 April 2024
(This article belongs to the Special Issue Physiology and Pathology: Feature Review Papers)

Abstract

:
Investigating the biophysiological substrates of psychiatric illnesses is of great interest to our understanding of disorders’ etiology, the identification of reliable biomarkers, and potential new therapeutic avenues. Schizophrenia represents a consolidated model of γ alterations arising from the aberrant activity of parvalbumin-positive GABAergic interneurons, whose dysfunction is associated with perineuronal net impairment and neuroinflammation. This model of pathogenesis is supported by molecular, cellular, and functional evidence. Proof for alterations of γ oscillations and their underlying mechanisms has also been reported in bipolar disorder and represents an emerging topic for major depressive disorder. Although evidence from animal models needs to be further elucidated in humans, the pathophysiology of γ-band alteration represents a common denominator for different neuropsychiatric disorders. The purpose of this narrative review is to outline a framework of converging results in psychiatric conditions characterized by γ abnormality, from neurochemical dysfunction to alterations in brain rhythms.

1. Introduction

Electric current generation within the brain results from multiple sources (i.e., synaptic activity, intra- and extracellular ionic fluxes through voltage-gated and ligand-gated channels, and intrinsic membrane potential resonance) [1,2]. Extracellular electric fields’ spatiotemporal summation gives rise to neural oscillations, which are considered to reflect neuronal populations’ activity [3]. It is still debated whether neural oscillations are merely an epiphenomenon of the coordinated neuronal activity or actively contribute to state-dependent informational inflow, outflow, and integration across different brain regions [4]. Nevertheless, brain rhythms hold functional significance as they arise from brain dynamics underlying cognitive processes [5]. Consistent with this, neural oscillations’ distinct frequency components have been associated with unique brain functions and states [6], ranging from basic sensory processes to higher-order cognitive operations [7,8,9,10].
Among the spectrum of the characteristic frequency bands, gamma (γ) oscillations (30–80 Hz) have been related to higher brain functions (e.g., the attentive processing of information and active maintenance of memory contents) [11,12,13]. Disturbances in γ rhythms have been reported in multiple neuropsychiatric conditions [2,14,15], even prior to symptom onset [16]. Schizophrenia (ScZ) represents a prime example of brain rhythm disruption, with γ oscillations recognized as a core substrate of clinical symptoms and cognitive deficits [17]. Similar evidence has been reported for bipolar disorder (BD) and, less robustly, major depressive disorder (MDD).
Gamma-band alterations are hypothesized to result from altered cortical excitatory/inhibitory (E/I) balance [18,19,20,21,22]. The subsequent shift towards epileptic hyperexcitability [22,23] has been found to characterize multiple neurodevelopmental [24], neurodegenerative [25], and psychiatric diseases [26]. The proper functioning and stability of brain circuits are based on the temporal balance between excitatory and inhibitory inputs from glutamatergic pyramidal cells and inhibitory GABAergic interneurons, respectively [27], with intracortical networks of specialized GABAergic interneurons (i.e., with fast, strong, and shunting inhibitory synapses) orchestrating the generation of fast oscillations [28]. The functional assessment of the E/I imbalance (e.g., via the aperiodic 1/f-like component of the neural power spectrum) shows an association between a shift towards excitation and a decrease in high-frequency activity oscillations [29,30,31]. E/I imbalance caused by a predominant GABAegic dysfunction has been proposed as a unifying hallmark of different neuropsychiatric diseases [32,33,34,35,36,37].
At the cellular level, γ oscillations are thought to arise from the activity of parvalbumin-positive (PV+) cells, a subpopulation of GABAergic interneurons exhibiting fast-spiking (FS) properties [38,39,40], with in vitro and in vivo studies showing the firing of PV+ cells to be robustly coupled to γ oscillations [40,41,42,43,44]. Critical for this purpose are α1-subunit-containing GABAA receptors, which display fast kinetics and are abundant in PV+ interneuron synapses [45,46]. PV+ interneuron maturation is crucial for the onset of critical periods (i.e., time windows of enhanced plasticity) [47]. Indeed, a delayed or extended period of circuit instability in the developing brain—that is, the E/I imbalance—is consistent with a loss of PV+ function in psychiatric disorders [48].
Interneuronal dysfunction has been associated with the impaired or delayed development of the perineuronal nets (PNNs), the predominant extracellular matrix (ECM) aggregates responsible for neuronal structural support, synaptic functions, and plasticity [49,50,51,52,53]. Immunolabeling evidence with markers for PNNs’ chondroitin sulfate proteoglycans (CSPGs) shows that PNNs mainly encompass GABAergic PV+ cells [54,55,56,57] and regulate FS firing [58,59]. The behavioral or in vitro manipulation of neuronal activity during critical periods has been shown to reduce the PNNs, which suggests that their expression is activity-dependent and tied to both the development of PV+ neurons and the establishment of the E/I balance in neural circuits [60,61,62] (Figure 1).
Oxidative stress and neuroinflammation are thought to affect the structural and functional integrity of PNNs by mediating excessive microglia activation and the release of ECM-cleaving enzymes, such as matrix metalloproteinases (MMPs) [63,64]. Research on PNN modulators in psychiatric conditions showed that abnormal rates of PNN-degrading proteases, such as metalloproteinase-9 (MMP-9) [65], or a diminished expression of reelin (i.e., a key protein of the PNNs expressed in GABAergic neurons and secreted into PNNs [66,67]) negatively impact PV+ cell-mediated activity [68,69,70,71,72]. Excessive PNN cleavage, in turn, leads to the formation of reactive oxygen species (ROS), exacerbating microglia activation and neuroinflammation [73,74]. Evidence from animal models points to PNN abnormalities as the main source of antioxidant defence failure [75], with subsequent cortical alterations attributable to the loss of PV+ interneurons [58,76,77].
The outlined evidence delineates a pathophysiological cascade of abnormal γ rhythms in various neuropsychiatric conditions resulting from abnormalities in PV+ interneurons, the disruption of PNNs, and neuroinflammation. Despite not constituting a complete etiological explanation, this model of pathogenesis is supported by converging molecular, cellular, and functional evidence.
Worth noting, ScZ, BD, and MDD are quite complex conditions that result from the interplay of various etiological factors such as genetic predisposition [78,79], environmental influences [80,81], neurotransmitter dysregulation [82,83,84], and structural brain abnormalities [85,86,87]. The purpose of the current narrative review is to provide a pathophysiological framework of γ-band disturbances and the underlying mechanisms in ScZ, BD, and MDD, with evidence of alterations from the micro- (i.e., cellular, molecular) to the macro-scale (i.e., brain oscillations).

2. Schizophrenia-Related Pathophysiological Cascade

ScZ represents the most common and complex form of functional psychosis and the seventh-most-costly medical illness in our society, affecting about 0.3 percent of the population worldwide [88,89]. It is characterized by a combination of symptoms classified into ‘positive’ (e.g., delusions, hallucinations) and ‘negative’ (e.g., social withdrawal, apathy, impaired judgment) [90]. The etiology and pathogenesis remain unclear due, in part, to the heterogeneity of the ScZ phenotype and the lack of clear pathological markers such as those that provide reference points in the study of other neuropsychiatric disorders [91,92].

2.1. Neuroinflammation and Redox Dysregulation in Schizophrenia

One hypothesis for ScZ etiology is that perinatal infections (e.g., due to spirochetes such as Borrelia burgdorferi, and viruses such as Epstein–Barr virus) might trigger a neuroinflammatory response that persists during adulthood [93,94,95,96,97]. Indeed, a growing body of evidence has revealed signs of systemic and central nervous system inflammation in ScZ patients [98,99,100,101], such as an increase in inflammatory markers (e.g., neutrophil-lymphocyte ratio, proinflammatory cytokines) [96,102] and evidence for microglial (i.e., the brain macrophage) overactivation [103,104,105]. The inflammatory overproduction of ROS increases cellular oxidative stress, leading to neurotoxic and detrimental effects on brain development and the maturation of PV+ neurons [106,107].
Excessive oxidative stress directly affects PNNs by overstepping their protective capacity, which in turn impacts PV+ cells [108,109]. Given their high energy demand [110], the maturation of PV+ interneurons and their proper functioning require a well-regulated antioxidant system and seem to be particularly vulnerable to ROS overproduction [111,112]. Indeed, early-life inflammatory conditions cause a persistent decrease in hippocampal and prefrontal cortical PV+ cells in animal models [113,114] and determine susceptibility to ScZ [115,116,117]. Consistently, post-mortem studies have reported increased oxidative stress in the prefrontal cortex, anterior cingulate cortex, caudate region, and hippocampus in subtypes of ScZ [118,119,120,121,122].

2.2. Perineuronal Net Abnormalities in Schizophrenia

PNN alterations have been solidly implicated in the pathophysiology of ScZ [57]. Post-mortem studies reveal alterations in PNN structure and biological/biochemical composition in the deep amygdala nuclei and entorhinal cortex [123,124,125]. The evidence for massive CSPG abnormalities (i.e., the main components of mature PNNs) in glial cells further supports the pivotal role of ECM–glial interactions in the pathogenesis of this disease [126]. Consistently, extracellular analyses of the cortex of ScZ patients show significant alterations, such as water diffusion anisotropy in tissue, together with markers for neuroinflammation and microglia activation [127]. Noteworthy, the PTPRZ1 gene encoding for the CSPG phosphacan has been implicated in susceptibility for ScZ [128], as well as growth factors affecting astrocytic CSPG expression [129].
ECM alterations in the brain tissue of ScZ patients have also been detected in the neocortex, with evidence for reduced PNN density in layers three and five of the PFC [130]. Despite Enwright and colleagues [131] reporting normal PNN density in the dorsolateral prefrontal cortex (DLPFC), the fluorescence intensity of Wisteria floribunda-Agglutinin (WFA) labeling and immunohistochemical staining for aggrecan (i.e., the primary CSPG of PNNs) was lower around PV+ interneurons in ScZ brains. The role of PNNs in prefrontal regions comprises the regulation of the synaptic refinement that occurs through the peripubertal period until late adolescence and early adulthood, with its dysfunction compromising the experience-dependent consolidation of synaptic connectivities [132,133,134]. This cascade of effects temporally matches with the usual age of onset for ScZ symptomatology, thus providing further support for the hypothesis that PNN abnormalities primarily affecting PV+ cells exert a core role in the impairment of cortical integrity and stability in ScZ [135,136].

2.3. PV+ Interneuron Alterations in Schizophrenia

Alterations in PV+ interneurons have been consistently implicated in the pathophysiology of ScZ, with evidence pointing towards altered GABA synthesis, abnormal transmission, and structural abnormalities [137,138,139]. Proof for GABAergic inhibitory circuitry alterations in ScZ mainly derives from molecular evidence on gene expression in the PFC—for instance, investigating the mRNA expression of GAD67, a GABA-synthesizing enzyme [140,141]. Indeed, Hashimoto and colleagues [140] found reduced PV expression in the brain of ScZ patients, which was correlated with a decrease in the density of neurons containing GAD67 mRNA. The co-localization of the reduction in PV- and GAD67-mRNA expression supports the idea that GABA neurotransmission mediated by PV+ interneurons is altered in ScZ patients [142].
Multiple studies also revealed abnormalities involving GABAA postsynaptic receptors containing the α1 subunit, which is the class of GABA receptors exhibiting the fastest kinetics [45,143]. Lower expression of the mRNA for the α1 subunit has been detected in the neocortex, particularly in layers three and four of the PFC of ScZ patients [144,145]. This specific molecular deficit results in the reduced density and strength of PV+ inhibitory inputs to the pyramidal cells [146]. An increase in the mRNA for the GABAA receptor α2 subunits has also been detected in the PFC of patients with ScZ and interpreted as a compensatory response to the deficient GABA activity and release due to lower levels of GAD67 [144,147]. The lower density of the GABA membrane transporter GAT-1 has also been found in patients’ PFC [148], suggesting a reduced reuptake of this neurotransmitter. This might be linked to the altered inputs from the medial dorsal thalamus, which are critical for initiating and maintaining γ activity [149,150].
Reduced density of hippocampal PV+ interneurons has been reported in mice lacking the tyrosine kinase ErbB4 receptor for Neuregulin-1 (NRG1), whose expression is highly enriched in both the chandelier and basket subclasses of PV+ cells in the PFC, hippocampus, and amygdala [151]. Both ErbB4 and NRG1 are susceptibility genes for ScZ [152]. Interestingly, PV+ cell alteration in this model is associated with impaired network inhibition and reduced γ oscillatory power [153]. In vivo, the loss of NRG1/ErbB4 signaling in the developing brain causes a reduction in the number of GABAergic interneurons in the postnatal cortex [154], and inhibition of NRG1/ErbB4 signaling in PV+ interneurons of the PV-ErbB4−/− mice is associated with reduced GABAergic activity and the E/I imbalance [155]. Zhang and Reynold found the relative densities of hippocampal PV-immunoreactive cells in post-mortem brains to be deficient in ScZ [156]. This evidence can be linked to the role of PV+ interneurons in the generation of hippocampal oscillations [157,158]. Moreover, this reduction was unrelated to antipsychotic drug treatment, age, or duration of illness.
Mice expressing a truncated version of the ScZ risk gene DISC1 have reduced PV immunoreactivity in the PFC and hippocampus [159,160]. Interestingly, the DISC1 pathway is convergent with those of other disease-related genes, including NRG1 and β-amyloid precursor protein (APP) during cortical development [161]. Even though beyond the scope of this review, it is worth noting that APP gene mutation is associated with an early onset familiar form of Alzheimer’s disease [162], a neurodegenerative disease in which PV+ dysfunction and brain rhythms disturbances have been widely described [37,163].
Despite results on the relative density of PV+ cells in ScZ not always being consistent (e.g., Glausier and colleagues reporting lower PV protein levels and unchanged density of cell inputs [141]), the extensive evidence from molecular studies corroborates the core role of GABAergic PV+ interneuron dysfunction in the pathophysiology of ScZ.

2.4. Gamma-Band Oscillatory Alterations in Schizophrenia

The disturbance of neural oscillations in the γ-band is considered a core pathophysiological feature of ScZ, particularly in the generation of cognitive dysfunction [164,165]. Multiple animal models of ScZ support disturbances in γ oscillations—for instance, in terms of abnormalities in the rhythmic network activity, as well as proneness to hippocampal epileptiform activity [153,166]. Lodge and colleagues [167] showed that a decreased density of PV+ interneurons throughout the medial prefrontal cortex (mPFC) and ventral subiculum correlates with a reduced γ-band response to a conditioned tone during a latent inhibition paradigm. Similarly, Nguyen and colleagues [168] found that the inhibition of PV+ or GAD65 neurons alters network oscillatory activity in the ventral hippocampus (vHPC) of rats and generates sensorimotor deficits [169]. The transgenic mouse in the study from Cho and colleagues [170] (i.e., impaired Dlx5 and Dlx6 transcription factors involved in the development of PV+ interneurons) exhibited a reduction in evoked γ-band oscillations, as well as cognitive inflexibility. Notably, optogenetic stimulation of cortical PV+ neurons in the γ frequency (40Hz) was capable of restoring cognitive flexibility. In an animal model of first-episode psychosis, spatial memory deficits were accompanied by disturbances in hippocampal θ-γ oscillations (i.e., reduction, uncoupling) and long-term potentiation mechanisms [171].
Alterations of γ-band activity in ScZ patients have been demonstrated with different experimental paradigms (i.e., spontaneous and evoked activity/responses). For instance, Grent-’t-Jong and colleagues [172] found individuals meeting clinical high-risk criteria for ScZ to exhibit increased γ power at rest, while both first-episode and chronic patients showed aberrant spectral power at both low and high γ-band frequencies. Moreover, the shift in E/I-balance toward augmented excitation found in high-risk subjects correlated with increased occipital γ power. The differential oscillatory patterns across illness stages are of great interest in the field of biomarkers research. For instance, increased excitation and resting-state γ disturbances might constitute spectral markers to predict the onset of psychosis [173]. Consistent with this, both patients and their unaffected siblings were found to exhibit reduced γ activity in the parietal cortex relative to non-affected individuals [174]. Also, impaired perceptual integration and disorganized symptoms in ScZ are accompanied by a pronounced reduction in spectral power in the γ-band [175]. The sub-anesthetic administration of ketamine (i.e., N-methyl-D-aspartate receptor (NMDA-R) antagonist leading to psychopathology as observed in ScZ) in healthy volunteers was found to induce dysregulated γ activity in thalamo-cortical networks [172,176].
In addition to abnormalities in spontaneous oscillations, altered stimuli-evoked γ responses have been reported with both visual [177] and auditory paradigms [178]. An EEG-fMRI study demonstrated a γ-band-specific network (involving bilateral auditory cortices, thalamus, and frontal brain regions) to be less active in subjects at high risk for ScZ, together with worse task performance [179]. This has been linked to a specific dysfunction in the frontotemporal network responsible for γ-phase synchronization and indexed by the ‘gamma-band auditory steady-state response’ (ASSR), which represents a robust biomarker that is increasingly studied in neuropsychiatric disorders as a reflection of deficient γ-band oscillations [19,180]. Abnormal ASSRs have been reported in first-episode and chronic ScZ patients, with specific frequency patterns associated with the severity of hallucinatory experiences [181,182]. Also, visually induced γ-band activity has been shown to correlate with genetic risk for ScZ [183]. This further strengthens the idea of neuronal firing-pattern dysfunction as a pathophysiological factor, rather than a consequence, in the development of this disorder.
The literature investigating oscillatory brain networks in ScZ points towards alterations in both local and long-range oscillatory network dynamics [175,176,184]. As such, patients with ScZ exhibit spatial functional hyper- or disconnectivity [174], as well as disturbances in the temporal structure of γ-band oscillations [185,186] (e.g., abnormalities in the Default Mode Network [DMN] have been reported and associated with decreased coherence in the γ-band at rest [187,188]). Transcranial magnetic stimulation (TMS), which activates cortical neurons via time-varying magnetic fields eliciting action potentials, can be coupled with electroencephalography (EEG) to investigate cortical γ modulation in ScZ [189]. Ferrarelli and colleagues [190] found that ScZ patients exhibit decreased EEG-evoked responses in the γ-band when TMS is applied over the frontal cortex relative to healthy controls. Also, repetitive TMS (rTMS) has been applied over the DLPFC to normalize γ oscillations and cognitive performance in patients [32]. Indeed, an extensive body of research has adopted neuroimaging approaches such as multimodal non-invasive brain stimulation techniques (NIBS) to investigate the neurophysiological correlates of abnormal brain activity, as well as for interventional purposes. Promising results emerge from the recent use of transcranial Alternating Current Stimulation (tACS) [191,192,193]. This technique allows for frequency-specific entrainment of endogenous neural oscillations and, potentially, induction of long-term synaptic plasticity [194]. When applied at the γ frequency over prefrontal areas, it has been shown to improve positive and negative symptoms, cognitive status, and well-being in ScZ patients [195].
A complex clinical phenotype arises from the described pathophysiological scenario. We believe that the study of γ brain rhythms and the underlying molecular and cellular mechanisms might provide a bridge between the neural and psychopathological levels. See Figure 2 for an overview of the outlined framework in ScZ.

3. Bipolar Disorder-Related Pathophysiological Cascade

BD is a recurrent chronic disorder affecting 1% of the world’s population [201]. Patients experience episodes of depression (i.e., low mood and related symptoms) and episodes of either mania (i.e., elated or irritable mood, or both) or hypomania (i.e., less severe or less-protracted symptoms of mania) [202]. Based on the extent and severity of mood elevations, BD is classified along a spectrum from unipolar to bipolar II or bipolar I [203]. Given the high rate of clinical comorbidities and the complexity of diagnosis, a further understanding of the underlying pathogenesis would support the development of reliable biomarkers for diagnosis and treatment [204,205].

3.1. Neuroinflammation and Redox Dysregulation in Bipolar Disorder

In the last decade, several studies have evaluated whether inflammatory and redox dysregulation processes might be involved in the pathophysiology of BD [206]. As reported for ScZ, prenatal exposure to infectious or inflammatory insults might contribute to the etiology of BD. As such, studies on animal models of prenatal maternal immune activation (MIA) showed that immune activation leads to the transgenerational transmission of pathological traits and behavioral alterations [207,208]. Mice with abnormal prefrontal levels of glutathione—the major cellular redox regulator and antioxidant—exhibit behavioral, morphological, electrophysiological, and neurochemical alterations analogous to those observed in patients. At the neuronal level, redox imbalance causes the impairment of PV+ interneurons and abnormal neuronal synchronization [209]. Increased concentrations of neuroinflammatory markers (e.g., Interleukin (IL)-1β, IL-6, IL-10) have been reported in the frontal cortex and hippocampus of a mouse model of mania [210].
Mitochondrial dysfunction has been extensively detected in BD patients [121,211,212] and linked to high levels of oxygen consumption and ROS overproduction [213]. Oxidative stress resulting from dysfunctional mitochondria leads to metabolic disturbances in neurons and glia, cell damage, and further oxidative stress, stimulating the generation of proinflammatory cytokines [206,214]. Proof for peripheral proinflammatory status (e.g., increased Tumor Necrosis Factor-α, TNF-α) in patients during mania [215] and its link with other markers of BD (e.g., abnormalities in circadian rhythmicity, metabolism) suggest that inflammatory mechanisms play a crucial role in the pathophysiology of this disease.

3.2. Perineuronal Net Abnormalities in Bipolar Disorder

Various lines of evidence support ECM abnormalities in BD. A post-mortem study by Knable and colleagues reported decreased reelin together with PV+ cell reduction in the brains of BD patients [216]. Further evidence revealed reduced glial cell immunoreactivity to the PNN’s CSPG aggrecan in the amygdala of BD patients compared to healthy subjects, together with lower PNN clusters [124]. Indeed, structural changes in BD brains have been detected in multiple amygdala nuclei in terms of volume, neuronal density, and size [217]. Steullet and colleagues [218] found a decrease in PNNs in the thalamic reticular nucleus of BD brains, with no effects on the duration of illness or the age of onset, together with a decrease in the number of PV+ cells and altered neuronal activity. Similar results emerged in transgenic mice with redox dysregulation compared with the wild-type mice [218]. A lower density of PNNs was also detected in the DLPFC of BD patients [219]. Despite the outlined converging evidence on PNN contribution to disease onset, the literature also shows inconsistencies [130]. Arguably, the differences could be quantitative rather than qualitative, and deficits could be region-specific or based on disease-specific timelines.
A specific mechanism of ECM molecular alterations in BD involves PNN regulation by homeoprotein Otx2, which controls the onset and closure of critical periods by triggering PNN expression and maturation [220]. Otx2 accumulation in PV+ cells via PNN interaction is required to maintain the non-plastic state of the adult cerebral cortex [221]. Notably, polymorphisms in Otx2 have been associated with BD and might be implicated in alterations of PNN-PV+ assembly dynamics in this disease [222]. The impact of Otx2 dysregulation on PNN formation might be amplified by redox dysregulation: decreased PNN formation due to redox imbalance prevents the internalization of Otx2 by PV+ cells, which in turn leads to a reduction of the PNN staining [223,224].
Genomic evidence supports the association of the ScZ risk gene Neurocan (NCAN) and specific mania symptoms in BD [225,226], supporting a genotype overlap between BD and ScZ [227]. Importantly, Neurocan is one of the main ECM proteoglycans mediating both the proper formation of PNNs and its proteins during brain development and the amount of GAD65/67 in GABAergic synapses [228].

3.3. PV+ Interneuron Alterations in Bipolar Disorder

As for the involvement of PV+ interneurons in BD, animal models of ankyrin-G isoform imbalance affecting PV+ cells exhibit a lack of axonal voltage-gated sodium channels, altered PV+ firing thresholds, and a diminished action-potential dynamic range [229]. Notably, there were behavioral changes in a mice model of BD patients. Evidence from post-mortem studies demonstrates the GAD65/67 enzymes for GABA synthesis are downregulated in the cerebellum and hippocampus of BD subjects compared to controls [230,231]. Guidotti and colleagues [66] also reported a reduced PFC and cerebellar expression of both GAD67 protein and GAD mRNA. Remarkably, the authors found a decreased expression of reelin, supporting the relevance of both GABAergic and ECM abnormalities in BD.
Fatemi and colleagues [232] also reported protein reduction for both the GABBR1 and GABBR2 subunits of the GABAB receptor in the lateral cerebella of BD brains (as in ScZ and MDD). As such, GABAB receptors are responsible for presynaptic and postsynaptic inhibition in PV+ interneurons and contribute to the dynamic modulation of their activity during oscillations [233]. Another study demonstrated the GABABR alterations in the PFC of BD subjects to be similar to those in ScZ [234]. Further neurochemical evidence consists of a reduced cortical concentration of GABA in the occipital cortex of recovered BD patients [235]. Moreover, low GABA and GAD concentrations have been reported in patients’ cerebrospinal fluid and plasma, respectively [236,237]. Disconfirmations in the detection of low GABA levels appear to be marginal and related to the limitations in techniques (e.g., regional analysis through magnetic resonance spectroscopy [MRS] resulting in high measurement variance and low specificity) [238].
Despite fewer immunochemical and cytoarchitectural studies being available for PV+ interneuron alterations in BD compared to ScZ, abundant evidence reveals the reduced density of cortical GABAergic interneurons in the anterior cingulate cortex, PFC, and parahippocampal region of BD brains [239,240,241]. In addition to density, the spatial distribution of GABAergic interneurons was found to be altered [242]. Steullet and colleagues [218] tested the disruption of cortico-thalamo-cortical circuits in BD with a combination of human post-mortem and rodent studies on PV+ neurons: both transgenic mice and patients’ brains showed a decrease in the density of PV+ cells. Indeed, BD subjects exhibit significantly reduced PV mRNA in different cortical layers of the PFC [243]. Some studies revealed GABAergic and PV+ abnormalities in BD that are not detected in ScZ (e.g., the density of PV+ interneurons in the entorhinal cortex) [244]. Despite the fact that GABAergic deficits in BD have been found to generally match those in psychosis, they might differ in cortical localization (e.g., deficits in hippocampal PV+ are more profound in ScZ patients) [245] and severity [246].

3.4. Gamma-Band Oscillatory Alterations in Bipolar Disorder

Abnormalities in γ-frequency band activity are hypothesized to be integral to the pathophysiology of BD. Results from studies on mice models of maternal immune activation (MIA) (i.e., a psychiatric disease risk factor) provided evidence for a selectively reduced transmission in PV+ interneurons as well as for the disruption of γ activity in the mPFC, with concurrent behavioral alterations (i.e., anxiety) [207]. Prospective birth cohort studies highlight the associations between MIA and offspring risk of BD [247], as well as ScZ [248]. This also highlights the above-mentioned relevance of immune-related inflammatory mechanisms in the etiology of psychiatric disorders. Further evidence for γ abnormalities derives from the Clock-Δ19 mice, a model for alteration of the CLOCK gene conveying risk for BD. Specifically, the phasic entrainment of the nucleus accumbens (NAC) low γ (30–55 Hz) to delta (δ, 1–4 Hz) oscillations was found to be impaired in the Clock-Δ19 mice compared to wild-type mice [249]. Interestingly, the transgenic mice mania-like behavior ameliorated after the administration of lithium, a medication commonly used to manage similar behavioral manifestations in human patients suffering from BD [250].
Human studies show disturbances of synchronization and oscillations within brain networks measured with EEG at rest [251]. Specifically, changes in network characteristics for the γ-band were associated with the severity of patient symptoms. Özerdem and colleagues [252,253] investigated long-distance event-related γ coherence during a visual oddball paradigm with light stimulation in patients with BD. The authors reported disturbances in functional long-range connectivity as compared to healthy controls (e.g., patients’ lower coherence values in response to non-target stimuli). Moreover, γ oscillations were also investigated through auditory oddball and paired-click paradigms administered to BD patients, monozygotic BD twins, unaffected relatives, and healthy controls. Results showed reduced γ power in patients as compared to unaffected relatives [254]. Moreover, low and high γ ASSRs (that have been suggested to reflect the efficiency of inhibitory interneuronal activity) and ASSR γ synchronization were found to be reduced in BD patients [255,256,257]. Interestingly, induced and evoked γ responses to visual grating stimuli (i.e., stimuli known to induce primary visual γ oscillations) were investigated in subjects suffering from schizoaffective BD, a clinical syndrome with features of both BD and ScZ. Results showed that subjects with schizoaffective BD have increased visual cortical γ power as compared to controls [258]. Similarly to ScZ, the literature on BD investigating γ disturbances supports its role as a valuable candidate biomarker for this disorder [259].
The majority of studies adopting brain stimulation in BD investigate cortical inhibitory deficits (i.e., short-interval cortical inhibition [SICI], cortical silent period [SP], and interhemispheric inhibition [IHI]) [260,261], while the evidence for NIBS targeting γ oscillations is scarce. In a TMS-EEG study assessing event-related spectral perturbations (ERSPs), Canali and colleagues showed that TMS stimulation results in patients’ lower activation of γ activity in the frontal cortico-thalamo-cortical circuits as compared to healthy controls [262]. Interestingly, the same finding emerged for ScZ and MDD. TMS-EEG has also been adopted to assess brain oscillatory activity before and after antidepressant treatments. Canali et al. assessed ERSPs in TMS/EEG sessions before and after sleep deprivation and light therapy exposure, showing that patients exhibit lower activation of the frontal γ-band response compared to healthy participants. Furthermore, TMS stimulation did not result in changes in the γ-band response in patients, irrespectively of remission occurrence [263]. This could suggest that a reduced γ response to external perturbation in the PFC might represent an endophenotype, not modifiable with pharmacological therapy, of psychotic dysfunction [264].
The outlined literature suggests that the pathophysiology of γ activity alterations might support the identification of novel markers of BD and the development of valuable therapeutic trajectories. See Figure 3 for an overview of the outlined framework in BD.

4. Major Depressive Disorder-Related Pathophysiological Cascade

Depression is an extremely common mood disorder accounting for the highest proportion of the global burden attributable to mental illness [265]. MDD, the most common type of depression, consists of a distinct collection of symptoms, including lowered mood, loss of interest or pleasure (anhedonia), and persistent feelings of guilt, as well as somatic disturbances such as impaired sleep, appetite, energy levels, and sexual functioning [266]. Even though several biological mechanisms with a possible role in MDD etiology and progression have been identified [267], a unified pathophysiology of the disease is still lacking.

4.1. Neuroinflammation and Redox Dysregulation in Major Depressive Disorder

Compared to ScZ and BD, no infectious insults seem to be involved in the onset of MDD [268]. However, neuroinflammation seems to participate in the genesis and development of this disorder (e.g., via ROS contribution to excessive MMPs activation perpetuating inflammation, and aberrant microglial activity [100,269,270,271]). Excessive inflammatory cytokines have been detected and linked to synaptic plasticity disruption and damage in MDD [272,273]. A recent etiological study demonstrated that a feed-forward mechanism occurs between chronic inflammation (e.g., elevated TNF-α) and elevated activity of protein kinase glycogen synthase kinase-3 (GSK-3) in MDD. This protein’s functions have extensive implications in neurodegenerative diseases such as Alzheimer’s disease [274,275], in which neuroinflammation, together with PNN, PV+ cells, and γ activity alterations, represents core pathophysiological players [276]. Thus, it has been hypothesized that a stress-meditated cascade involving GSK-3 activation and inflammatory cytokines dysregulation promotes the development of MDD [277]. Importantly, animal model studies demonstrated that the inhibition of GSK-3 with lithium treatment modulates the γ-band activity [278], and GSK-3 inhibition ameliorates deficits in spatial WM and increases evoked γ power in the primary auditory cortex [279]. Depression-like behavior and decline in PV+ cells induced by chronic unpredicted mild stress (CUMS) in mice have been linked to neuroinflammatory dysregulation [280]. Interestingly, inducible nitric oxide synthase (iNOS) (i.e., a major downstream mediator of inflammation) appears to be upregulated in CUMS models [281,282].

4.2. Perineuronal Net Abnormalities in Major Depressive Disorder

Evidence for GABAergic pathology in MDD has been linked to PNN alterations and abnormal deposition, with subsequent impairment of PV+ cell activity and pyramidal cell hyperexcitability [283,284]. Interestingly, it has been proposed that the mechanisms underlying the effect of antidepressant medications may involve metalloproteinases (MMPs), which impact PNN structures around PV+ interneurons when their levels are elevated, with subsequent disruption of the E/I balance [57,285]. This appears to be the case with the serotonin and norepinephrine reuptake inhibitor venlafaxine, which increases hippocampal levels of MMP-9 protein and subsequently attenuates PNNs [286]. This effect can be explained by the ability of monoamine reuptake inhibitors to increase the expression of PNN-degrading proteases such as MMP-9 [287], as well as to impact γ activity [283].
Yu and colleagues [284] investigated PNN density in the prelimbic cortex in animal models of CUMS. Depressive-like behaviors emerged after CUMS, accompanied by a decrease in PNN+ cell density and aggrecan expression. Moreover, in the same study, rats were divided into two subpopulations based on the exhibited behaviors: vulnerable versus resilient. The density of PNNs and the expression level of neurocan in the vulnerable group were decreased compared to the control and resilient groups. Behavioral examination of locomotion showed that PNN density and neurocan levels were lower in rats, defined as “low responders” in the new context, compared to “high responders”. An increase in the number of PNNs surrounding PV+ interneurons has also been detected [288]. Thus, animal models of depression suggest stress-induced changes in ECM/PNNs and PV+ interneurons [289], with the direction of these alterations largely depending on stress recency and timing [290]. Specifically, a biphasic temporal regulation (i.e., both increase and decrease) of ECM/PNNs is observed after the stress [291]. In light of the above-mentioned research, there is emerging evidence for the role of PNNs in depression despite inconsistencies attributable to the specificity of different stress models.
The literature on PNN disturbances in MDD patients consists of studies on molecular alterations. Some evidence has been provided for decreased reelin expression in the PFC, hippocampus, cerebellum, and blood of MDD patients, as in BD and ScZ [69,292,293]. Furthermore, similar to ScZ, increased levels of MMP-9 have been reported in patient blood samples during depressed states [71,294]. Further studies are needed to understand the molecular underpinnings of PNN alternations in MDD. Despite the evidence from human studies being scarce, the intriguing results on the impact of treatments for mood disorders on ECM molecules and PNN composition from animal models imply a need for further investigation.

4.3. PV+ Interneuron Alterations in Major Depressive Disorder

Accumulating evidence suggests that the GABAergic system is altered in the brain of patients with MDD. The available molecular knowledge mainly derives from stress models [295]. For instance, CUMS leads to a reduction of GAD67 levels in both rat PFC and hippocampus, as well as in vitro cortical neurons [296]. PV+ cells also decreased in the dentate gyrus, hippocampus, and mPFC after exposure to long-term psychosocial stress [297,298]. Interestingly, electrophysiological analyses revealed reduced frequencies of spontaneous IPSCs, decreased release probability of GABAergic synapses, and alterations in GABAB receptor-mediated signaling. In turn, pyramidal neurons showed a higher excitability [299]. Kigawa and colleagues [300] also found that combined exposure to prenatal and adolescent stress reduces PV+ interneurons. Interestingly, epigenetic changes in the GABAergic system seem to be involved in stress-related models exhibiting depression-like behaviors [301]. The outlined findings from chronic stress provide strong support for a fronto-limbic PV+ involvement in the emotional and cognitive symptoms of MDD.
Clinical evidence suggests that MDD is associated with reduced levels of GABA in the plasma, cerebrospinal fluid, and neocortex [302,303,304]. Moreover, effective medications have been shown to impact the GABAergic system by reversing low levels of GABA in the occipital cortex [305,306]. The literature from human studies on PV+ alterations is, however, inconclusive. A post-mortem study investigating PV+ and CB+ cells in the DLPFC and orbitofrontal cortex (OFC) in MDD brains reported a reduction of CB+ cell density and size compared to the control group, especially in the DLPFC region, while a trend toward reduction emerged for PV+ interneurons [307]. Reduced density of CB+ cells has also been reported in patients’ somatosensory cortex, while PV+ interneurons showed a smaller reduction that did not reach statistical significance [308]. Sibille and colleagues [243] found that MDD patients have reduced SST+ mRNA expression in the DLPFC, and no alterations in other transcripts. Similarly, Zhang & Reynolds [156] reported no significant reduction in PV+ neurons compared to BD and ScZ patients.
The outlined findings from human studies provide some evidence for abnormalities in distinct subpopulations of GABAergic interneurons in MDD. However, the framework is not definitive, possibly due to methodological limitations and study population specifics (e.g., death by suicide, antidepressant medication use at the time of death). It might be argued that GABAergic abnormalities qualitatively differ to some extent among neuropsychiatric disorders, potentially in correlation with macro-level manifestations [246,309].

4.4. Gamma-Band Oscillatory Alterations in Major Depressive Disorder

Brain oscillatory alterations in the γ-band represent an emerging topic in MDD research [310], with evidence deriving both from animal models and patients. Deep brain stimulation (DBS) of the ventral tegmental area (VTA) in a mouse model of depression was shown to induce an increase in the high γ-band phase coherence, which was correlated with an improvement in depressive-like behavior [311]. Also, Khalid and colleagues [312] demonstrated that remission from depressive behavior induced by chronic restraint stress in mice is associated with the restoration of γ activity. Riga and colleagues [313] showed that the administration of Vortioxetine, a multimodal antidepressant drug with pro-cognitive actions on GABAergic signaling, modulates γ oscillations in the paraventricular thalamic nucleus and prevents recognition memory deficits. Moreover, Vortioxetine appears to prevent the γ coherence deficits in the prelimbic nucleus–dorsal hippocampus pathway.
Electrophysiological evidence in humans showed that rsEEG metrics for γ rhythms discriminate healthy controls from MDD patients [314] and predict behavioral responses (i.e., abnormal responses to task errors, potentially due to lower γ activity within regions subserving affective and/or motivational responses to salient cues) [315]. Also, it was found that during the maintenance phase of WM tasks, MDD patients exhibit less occipital γ and that the severity of depression correlates with this alteration [316]. Increased γ-band EEG in a sustained semantic information processing task was detected in MDD patients, while a decrease throughout the task emerged in patients with ScZ, suggesting qualitative differences in γ-band disturbances between these disorders for elaborative emotional processing (i.e., higher physiological reactivity to negative information in depressed patients) [317]. Distinct patterns and locations of oscillatory alterations have been found to distinguish MDD from other disorders during emotional tasks (e.g., γ activity in the parietal and left posterior temporal cortex as an index to differentiate BD and MDD patients during implicit emotional tasks) [318,319].
Repetitive TMS paradigms over the DLPFC in MDD have been adopted as a treatment for pharmacoresistant patients to ameliorate symptoms [320] and to restore brain network functional connectivity among nodes including the left insula, right superior and inferior frontal gyrus, right inferior parietal lobule, and amygdala [321]. Paire-pulse protocols revealed increased cortico-cortical inhibition [322], supporting the hypothesis of GABAergic tone reduction in this disorder [323]. However, evidence from brain stimulation studies on γ oscillations remains scarce. Canali and colleagues [262] reported reduced TMS-evoked frontal γ-band frequencies after TMS over the premotor area in MDD patients (as well as BD and ScZ) compared to healthy subjects. Another study showed that two weeks of high-frequency rTMS over the left DLPFC of MDD patients determines an increase in frontal resting γ power and central-temporal θ-γ coupling at electrode sites. Furthermore, increased frontal γ power correlated with improved scores on the Hamilton Rating Scale for Depression (HAM-D17) and Beck Depression Inventory (BDI), while the increased central coupling correlated with a reduced number of errors on the Wisconsin Card Sorting Test (WCST) [324]. These findings are not in line with a recent rTMS study in which increased anterior-central γ power correlated with worsened depressive symptoms. These inconsistencies might be attributed to the complexity of γ oscillation dynamics over the course of TMS, as well as to individual differences in brain responses [325]. As such, MDD might be characterized by either elevated or decreased γ oscillation power, with varying values of “optimal control” among patients [310].
The outlined literature provides some evidence for γ-band abnormalities as a relevant feature, as well as a potential biomarker and therapeutic target for DBS and NIBS treatments in MDD [326]. Further research is needed to clarify the dynamics and patterns of γ alterations in relation to clinical manifestations. See Figure 4 for an overview of the outlined framework in MDD.

5. Discussion

5.1. Gamma-Band-Related Pathophysiology

Alterations in brain oscillations represent a core aspect of different psychiatric disorders [328]. Rhythmic disturbances in ScZ have been extensively reported and reviewed, with γ oscillations recognized as a substrate of both clinical symptoms and cognitive impairments [329]. Furthermore, the processes underlying this physiological dysfunction could be incorporated into a pathophysiological framework summarized as follows. Gamma-band alterations result from an imbalance between excitation and inhibition in the brain [330], particularly from weakened PV+ cells’ inhibitory inputs onto the pyramidal neurons [142]. Cortical hyperexcitability and interneuron dysfunction have also been associated with PNN impairment [57], which disrupts interneuronal plasticity and creates the environment for chronic neuroinflammation, with detrimental effects on PV+ cells [65,331]. Similar proofs for γ alterations and the underlying pathophysiology have been reported in BD, and growing evidence corroborates this cascade in MDD. Both rodent and human studies support the delineated multi-dimensional explanation of brain dysfunction from PNN environment disruption to PV+ alterations and oscillatory disturbances.
The evidence for the involvement of GABAergic interneurons mainly involves PV+ cells, which represent the main source of γ oscillation [38], both in microcircuits [43] and at the network level [41]. However, evidence for CB+ and SST+ interneurons dysfunction and/or reduction is also available from studies on ScZ, BD, and particularly MDD [242,245]. Noteworthy, some studies have shown that the activity of non-PV+ GABAergic interneurons may also be involved in fast oscillations, particularly when considering the network activity [332]. Despite the evidence for a prominent role of PV+ interneurons in γ activity, an unequivocal link between specific cell types and the production of rhythms at a certain frequency would be simplistic. For instance, FS bistratified cells exhibit a strong coupling with the hippocampal γ oscillations cycle [333]. SST+ interneurons seem to be involved in γ rhythms, as demonstrated by the decrease in the power of slow γ oscillations following SST+ cell disruption, as for PV+ cells. Moreover, photoexcitation on SST+ interneurons seems to generate γ oscillations. Indeed, complementary roles for SST+ and PV+ interneurons have been hypothesized in the generation of γ rhythms [157]. At the neurochemical level, SST+ and CB+ cells exhibit features predominantly attributed to PV+ neurons (i.e., expression of the voltage-gated K+ channel subunits Kv3, characterized by short and rapid currents and immunoreactivity to PNN markers) [334,335,336]. This suggests that some non-PV neurons may exhibit fast-firing properties [68].
The second aspect worthy of discussion is represented by the association of GABAergic with glutamatergic abnormalities in determining the E/I imbalance [20]. A strong correlation emerges between levels of glutamate+glutamine (Glx) and GABA from magnetic resonance spectroscopy (MRS) databases [337]. Glutamatergic medications (i.e., Ketamine) potentiate GABAergic synapses by blocking NMDA-R located on GABA neurons [271,338]. As introduced earlier, the E/I imbalance causing impaired γ rhythm stems from glutamate and GABA coexistent dysfunction and has been robustly reported in psychiatric and neurological disorders [32,34,339]. Altered levels of Glx, reduced pyramidal cell density, and attenuated NMDAR neurotransmission characterize ScZ, with differences depending on disease stage and treatment history [340,341]. Importantly, a reduction and/or hypo-activity of NMDAR has been documented and linked to GABAergic circuits dysfunction [342]. Indeed, NMDAR impairments generate abnormalities in levels of GAD67 in PV+ interneurons, as well as in PV expression [140,196], supporting the role of NMDAR-mediated glutamatergic signaling in GABAergic interneuronal development and behavior [343]. Aberrant glutamatergic modulation of GABA interneurons is also evident in BD, in which abnormal levels of Glx [344] and increased cortical glutamate levels [235] have been detected. Consistent with the GABAergic-glutamatergic modulation of GABA neurotransmission, Woo and colleagues [345] showed a decreased density of GAD67 mRNA–containing neurons in BD patients and an even higher reduction of cells co-expressing the NMDA NR(2A) subunit [346]. Also, MDD patients exhibit increased levels of glutamate and altered NMDARs function in the PFC, which has been specifically linked to altered DMN connectivity [347,348]. Crucially, it has been hypothesized that GABAergic deficit–induced changes in neural excitability and reduced conversion of glutamate to GABA may be responsible for the increased glutamate concentrations found in the brain of psychiatric patients [349]. However, given the evidence for reduced rather than increased brain content of glutamate in psychiatric patients [350], the relationship between GABAergic and glutamatergic transmission changes needs to be further elucidated.

5.2. Evidence of Gamma-Band-Related Pathophysiology in Other Mental Disorders

The literature supporting the pathophysiological framework outlined in this narrative review suggests that γ abnormality and its underlying cellular and molecular alterations are abundant in psychosis and mood disorders. Some pieces of evidence on psychiatric conditions other than ScZ, BD, and MDD can be mentioned to corroborate the relevance of the delineated framework in psychiatry. For instance, recent work from Briones and colleagues [351] revealed PV+ cells and their surrounding PNNs to be altered in the dorsal striatum of mice models of excessive repetitive behavior (i.e., relevant for Obsessive-compulsive Disorder and other neuropsychiatric disorders). Interestingly, this was linked to abnormalities in inhibitory postsynaptic potentials, reduced γ power, and altered dendritic spine density, with normalization of repetitive behavior following the restoration of PNNs and PV+ cells. Given the heterogeneity of the presented literature, particularly in terms of methodological variations across studies (e.g., immunolabeling techniques, mice model types, medications in patients’ samples), caution should be applied when interpreting and linking results. This also points to the need for further investigation of the outlined pathophysiological framework.
A relevant point that should be discussed is the MDD evidence, abundantly relying on mice models of stress (e.g., CUMS) [284]. Although the outlined literature refers to depression-like behaviors [301] and their link with γ oscillations [313], results from these models might be relevant for a wide span of psychiatric conditions, including generalized anxiety, panic, social and separation anxiety, agoraphobia, post-traumatic stress, and OCD [352]. Moreover, early life adversity events, that is, the case of environmental stress models, have been linked to the development of both anxiety disorders and clinical depression [353]. Interestingly, evidence for the involvement of PV+ interneurons in anxiety emerged from a mouse model of mutant Erbin (i.e., a protein expressed in cortical and hippocampal PV+ interneurons regulating fast excitatory synaptic transmission) [354], with inhibition of PV+ cells resulting in anxiogenic effects and reduced excitatory postsynaptic responses in the amygdala [355]). Further studies are needed to elucidate the pathophysiology of PV-related γ dysfunction in different stress-related disorders.
Deficient neural oscillatory patterns in the γ frequency band have also been reported in Autism Spectrum Disorder, together with evidence for a reduction in the PV+ expression [356,357] as well as PNN alterations in mice models [358,359]. Neuroinflammation has also been detected and associated with reduced GABA levels [360,361]. Given the recent evidence for the outlined pathophysiology involving PNNs and PV+ interneurons, neural oscillations in the γ band are receiving growing attention as a potential target in Autism Spectrum Disorder [34,362]. The outlined evidence corroborates the relevance of the γ-related pathophysiology in psychiatry as a valuable topic for further investigation of disease biomarkers and the potential development of new therapeutics.
The pathophysiological cascade outlined in this review has recently gained attention also in the field of dementia, particularly Alzheimer’s Disease [363,364]. Indeed, degradation of PNN and neuroinflammation have been reported in multiple neurodegenerative diseases, as well as E/I imbalance and hyperexcitability subserving γ-band oscillatory alterations [25,365,366]. These have been attributed to PV+ interneurons’ dysfunction [367,368], which seems to also increase brain vulnerability to β-amyloid (Aβ) deposition [369]. Accordingly, a growing body of research is currently investigating the therapeutic effects of γ-induction via NIBS or multisensory stimulation on multiple pathological facets of neurodegeneration. Promising results emerge from both animal models [370,371,372] and studies on patients (i.e., behavioural improvements, restoration of intracortical connectivity, increased brain perfusion, and reduced tau burden) [373,374,375,376,377].

5.3. Conclusions: Brain Oscillatory Biomarkers and Non-Invasive Treatments

In his pioneering work, Rhythms of the Brain, Buzsáky remarked that
‘[…] gamma oscillations, and neuronal oscillations in general, are not “independent” events that impose timing on neuronal spikes but rather are a reflection of self-organized interactions of those neurons’.
[378]
Investigating the molecular underpinnings of altered brain oscillations offers great insights into disease mechanisms involving these disturbances. Given the heterogeneity and complexity of psychiatric disorders, this might support the search for valuable biomarkers, which represents one of the major challenges in the field of precision medicine. Indeed, oscillatory alterations, as in the γ rhythms, offer a dynamic picture of brain disturbances, even before the full development of the disorder is evident [328]. In this regard, MEG and EEG-based research has attracted increased interest in the identification of markers for neuropsychiatric disorders [379]. We believe two reasons delineated this scenario. First, abnormal brain oscillations have the potential to be adopted as biomarkers [380]. Secondly, the recent combination of EEG with NIBS allows oscillatory activity and other cortical properties underlying brain rhythms (e.g., excitability, inhibition, connectivity, and plasticity) to be directly modulated and causally investigated [12]. As already mentioned, the TMS-EEG methodology has shown its potential to uncover biological targets in psychiatric illnesses and, not least, to advance our understanding of biological, physiological, and clinical features integrated into a comprehensive pathological framework [381].
As also shown in the field of dementia [382], abundant evidence from the last two decades points to the use of NIBS as a promising avenue for developing effective interventions in psychiatry [383]. As such, rTMS is already an established interventional technique for MDD, and further evidence is available in BD and ScZ, especially for the treatment of hallucinations, which have also been treated with tDCS (See [384] for an overview of NIBS treatments in psychiatric disorders) [385]. Promising results have recently emerged from clinical trials adopting tACS for the restoration of disturbed brain oscillations in ScZ [386] and MDD [387]. In addition to its potential therapeutic role, brain responses to tACS could serve as a possible tool for determining the diagnosis, classification, and prognosis of psychiatric disorders [388,389]. Further methodological development of neuromodulation targeting brain oscillations benefits from the elucidation of the pathophysiology of illness, from physiological alteration to its neurobiological underpinnings.
Future studies would further elucidate the presented framework and provide evidence for causal relationships between common pathophysiological factors across psychiatric disorders. Indeed, potential confounding factors (e.g., medication effects, substance use, comorbidities) on the presented neuropathological framework should be further addressed together with the impact of genetic predisposition and environmental influences.

Author Contributions

Conceptualization, A.P. and E.S.; writing—original draft preparation, A.P.; writing—review and editing, S.P., C.L.S., I.D., S.R., L.B., D.R. and V.B.; supervision, E.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Buzsáki, G.; Anastassiou, C.A.; Koch, C. The Origin of Extracellular Fields and Currents—EEG, ECoG, LFP and Spikes. Nat. Rev. Neurosci. 2012, 13, 407–420. [Google Scholar] [CrossRef] [PubMed]
  2. Mathalon, D.H.; Sohal, V.S. Neural Oscillations and Synchrony in Brain Dysfunction and Neuropsychiatric Disorders: It’s About Time. JAMA Psychiatry 2015, 72, 840–844. [Google Scholar] [CrossRef]
  3. Jia, X.; Kohn, A. Gamma Rhythms in the Brain. PLOS Biol. 2011, 9, e1001045. [Google Scholar] [CrossRef] [PubMed]
  4. Jones, S.R. When Brain Rhythms Aren’t ‘Rhythmic’: Implication for Their Mechanisms and Meaning. Curr. Opin. Neurobiol. 2016, 40, 72–80. [Google Scholar] [CrossRef] [PubMed]
  5. Ward, L.M. Synchronous Neural Oscillations and Cognitive Processes. Trends Cogn. Sci. 2003, 7, 553–559. [Google Scholar] [CrossRef] [PubMed]
  6. Başar, E.; Başar-Eroglu, C.; Karakaş, S.; Schürmann, M. Gamma, Alpha, Delta, and Theta Oscillations Govern Cognitive Processes. Int. J. Psychophysiol. 2001, 39, 241–248. [Google Scholar] [CrossRef] [PubMed]
  7. Buzsáki, G.; Draguhn, A. Neuronal Oscillations in Cortical Networks. Science 2004, 304, 1926. [Google Scholar] [CrossRef] [PubMed]
  8. Fries, P.; Nikolić, D.; Singer, W. The Gamma Cycle. Trends Neurosci. 2007, 30, 309–316. [Google Scholar] [CrossRef]
  9. Uhlhaas, P.; Pipa, G.; Lima, B.; Melloni, L.; Neuenschwander, S.; Nikolić, D.; Singer, W. Neural Synchrony in Cortical Networks: History, Concept and Current Status. Front. Integr. Neurosci. 2009, 3, 17. [Google Scholar] [CrossRef]
  10. Uhlhaas, P.J.; Roux, F.; Rodriguez, E.; Rotarska-Jagiela, A.; Singer, W. Neural Synchrony and the Development of Cortical Networks. Trends Cogn. Sci. 2010, 14, 72–80. [Google Scholar] [CrossRef] [PubMed]
  11. Gray, C.M.; Singer, W. Stimulus-Specific Neuronal Oscillations in Orientation Columns of Cat Visual Cortex. Proc. Natl. Acad. Sci. USA 1989, 86, 1698–1702. [Google Scholar] [CrossRef] [PubMed]
  12. Herrmann, C.S.; Strüber, D.; Helfrich, R.F.; Engel, A.K. EEG Oscillations: From Correlation to Causality. Int. J. Psychophysiol. 2016, 103, 12–21. [Google Scholar] [CrossRef] [PubMed]
  13. Ribary, U.; Ioannides, A.A.; Singh, K.D.; Hasson, R.; Bolton, J.P.; Lado, F.; Mogilner, A.; Llinás, R. Magnetic Field Tomography of Coherent Thalamocortical 40-Hz Oscillations in Humans. Proc. Natl. Acad. Sci. USA 1991, 88, 11037–11041. [Google Scholar] [CrossRef] [PubMed]
  14. Kitchigina, V.F. Alterations of Coherent Theta and Gamma Network Oscillations as an Early Biomarker of Temporal Lobe Epilepsy and Alzheimer’s Disease. Front. Integr. Neurosci. 2018, 12, 36. [Google Scholar] [CrossRef] [PubMed]
  15. Uhlhaas, P.J.; Singer, W. High-Frequency Oscillations and the Neurobiology of Schizophrenia. Dialogues Clin. Neurosci. 2013, 15, 301–313. [Google Scholar] [CrossRef] [PubMed]
  16. Tada, M.; Nagai, T.; Kirihara, K.; Koike, S.; Suga, M.; Araki, T.; Kobayashi, T.; Kasai, K. Differential Alterations of Auditory Gamma Oscillatory Responses between Pre-Onset High-Risk Individuals and First-Episode Schizophrenia. Cereb. Cortex 2016, 26, 1027–1035. [Google Scholar] [CrossRef]
  17. Sun, Y.; Farzan, F.; Barr, M.S.; Kirihara, K.; Fitzgerald, P.B.; Light, G.A.; Daskalakis, Z.J. Gamma Oscillations in Schizophrenia: Mechanisms and Clinical Significance. Brain Res. 2011, 1413, 98–114. [Google Scholar] [CrossRef] [PubMed]
  18. Deleuze, C.; Bhumbra, G.S.; Pazienti, A.; Lourenço, J.; Mailhes, C.; Aguirre, A.; Beato, M.; Bacci, A. Strong Preference for Autaptic Self-Connectivity of Neocortical PV Interneurons Facilitates Their Tuning to γ-Oscillations. PLoS Biol. 2019, 17, 1–26. [Google Scholar] [CrossRef] [PubMed]
  19. Koshiyama, D.; Kirihara, K.; Tada, M.; Nagai, T.; Fujioka, M.; Ichikawa, E.; Ohta, K.; Tani, M.; Tsuchiya, M.; Kanehara, A.; et al. Electrophysiological Evidence for Abnormal Glutamate-GABA Association Following Psychosis Onset. Transl. Psychiatry 2018, 8, 211. [Google Scholar] [CrossRef]
  20. Rujescu, D.; Bender, A.; Keck, M.; Hartmann, A.M.; Ohl, F.; Raeder, H.; Giegling, I.; Genius, J.; McCarley, R.W.; Möller, H.-J.; et al. A Pharmacological Model for Psychosis Based on N-Methyl-D-Aspartate Receptor Hypofunction: Molecular, Cellular, Functional and Behavioral Abnormalities. Biol. Psychiatry 2006, 59, 721–729. [Google Scholar] [CrossRef] [PubMed]
  21. Scotti-Muzzi, E.; Chile, T.; Moreno, R.; Pastorello, B.F.; da Costa Leite, C.; Henning, A.; Otaduy, M.C.G.; Vallada, H.; Soeiro-de-Souza, M.G. ACC Glu/GABA Ratio Is Decreased in Euthymic Bipolar Disorder I Patients: Possible in Vivo Neurometabolite Explanation for Mood Stabilization. Eur. Arch. Psychiatry Clin. Neurosci. 2021, 271, 537–547. [Google Scholar] [CrossRef]
  22. Tatti, R.; Haley, M.S.; Swanson, O.K.; Tselha, T.; Maffei, A. Neurophysiology and Regulation of the Balance Between Excitation and Inhibition in Neocortical Circuits. Biol. Psychiatry 2017, 81, 821–831. [Google Scholar] [CrossRef] [PubMed]
  23. Kirischuk, S. Keeping Excitation–Inhibition Ratio in Balance. Int. J. Mol. Sci. 2022, 23, 5746. [Google Scholar] [CrossRef]
  24. Foss-Feig, J.H.; Adkinson, B.D.; Ji, J.L.; Yang, G.; Srihari, V.H.; McPartland, J.C.; Krystal, J.H.; Murray, J.D.; Anticevic, A. Searching for Cross-Diagnostic Convergence: Neural Mechanisms Governing Excitation and Inhibition Balance in Schizophrenia and Autism Spectrum Disorders. Biol. Psychiatry 2017, 81, 848–861. [Google Scholar] [CrossRef] [PubMed]
  25. Maestú, F.; de Haan, W.; Busche, M.A.; DeFelipe, J. Neuronal Excitation/Inhibition Imbalance: Core Element of a Translational Perspective on Alzheimer Pathophysiology. Ageing Res. Rev. 2021, 69, 101372. [Google Scholar] [CrossRef] [PubMed]
  26. O’Donnell, P. (Ed.) Increased Cortical Excitability as a Critical Element in Schizophrenia Pathophysiology. In Cortical Deficits in Schizophrenia: From Genes to Function; Springer: Boston, MA, USA, 2008; pp. 219–236. ISBN 978-0-387-74351-6. [Google Scholar]
  27. Marín, O. Interneuron Dysfunction in Psychiatric Disorders. Nat. Rev. Neurosci. 2012, 13, 107–120. [Google Scholar] [CrossRef] [PubMed]
  28. Bartos, M.; Vida, I.; Jonas, P. Synaptic Mechanisms of Synchronized Gamma Oscillations in Inhibitory Interneuron Networks. Nat. Rev. Neurosci. 2007, 8, 45–56. [Google Scholar] [CrossRef] [PubMed]
  29. Wilkinson, C.L.; Nelson, C.A. Increased Aperiodic Gamma Power in Young Boys with Fragile X Syndrome Is Associated with Better Language Ability. Mol. Autism 2021, 12, 17. [Google Scholar] [CrossRef] [PubMed]
  30. Ahmad, J.; Ellis, C.; Leech, R.; Voytek, B.; Garces, P.; Jones, E.; Buitelaar, J.; Loth, E.; dos Santos, F.P.; Amil, A.F.; et al. From Mechanisms to Markers: Novel Noninvasive EEG Proxy Markers of the Neural Excitation and Inhibition System in Humans. Transl. Psychiatry 2022, 12, 467. [Google Scholar] [CrossRef] [PubMed]
  31. Molina, J.L.; Voytek, B.; Thomas, M.L.; Joshi, Y.B.; Bhakta, S.G.; Talledo, J.A.; Swerdlow, N.R.; Light, G.A. Memantine Effects on Electroencephalographic Measures of Putative Excitatory/Inhibitory Balance in Schizophrenia. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2020, 5, 562–568. [Google Scholar] [CrossRef] [PubMed]
  32. Farzan, F.; Barr, M.S.; Levinson, A.J.; Chen, R.; Wong, W.; Fitzgerald, P.B.; Daskalakis, Z.J. Evidence for Gamma Inhibition Deficits in the Dorsolateral Prefrontal Cortex of Patients with Schizophrenia. Brain 2010, 133, 1505–1514. [Google Scholar] [CrossRef] [PubMed]
  33. Fritschy, J.-M. Epilepsy, E/I Balance and GABAA Receptor Plasticity. Front. Mol. Neurosci. 2008, 1, 5. [Google Scholar] [CrossRef] [PubMed]
  34. Kayarian, F.B.; Jannati, A.; Rotenberg, A.; Santarnecchi, E. Targeting Gamma-Related Pathophysiology in Autism Spectrum Disorder Using Transcranial Electrical Stimulation: Opportunities and Challenges. Autism Res. 2020, 13, 1051–1071. [Google Scholar] [CrossRef] [PubMed]
  35. Ramamoorthi, K.; Lin, Y. The Contribution of GABAergic Dysfunction to Neurodevelopmental Disorders. Trends Mol. Med. 2011, 17, 452–462. [Google Scholar] [CrossRef] [PubMed]
  36. Sohal, V.S.; Rubenstein, J.L.R. Excitation-Inhibition Balance as a Framework for Investigating Mechanisms in Neuropsychiatric Disorders. Mol. Psychiatry 2019, 24, 1248–1257. [Google Scholar] [CrossRef] [PubMed]
  37. Verret, L.; Mann, E.O.; Hang, G.B.; Barth, A.M.I.; Cobos, I.; Ho, K.; Devidze, N.; Masliah, E.; Kreitzer, A.C.; Mody, I.; et al. Inhibitory Interneuron Deficit Links Altered Network Activity and Cognitive Dysfunction in Alzheimer Model. Cell 2012, 149, 708–721. [Google Scholar] [CrossRef] [PubMed]
  38. Gonzalez-Burgos, G.; Lewis, D.A. GABA Neurons and the Mechanisms of Network Oscillations: Implications for Understanding Cortical Dysfunction in Schizophrenia. Schizophr. Bull. 2008, 34, 944–961. [Google Scholar] [CrossRef] [PubMed]
  39. Hu, H.; Jonas, P. A Supercritical Density of Na+ Channels Ensures Fast Signaling in GABAergic Interneuron Axons. Nat. Neurosci. 2014, 17, 686–693. [Google Scholar] [CrossRef] [PubMed]
  40. Stedehouder, J.; Couey, J.J.; Brizee, D.; Hosseini, B.; Slotman, J.A.; Dirven, C.M.F.; Shpak, G.; Houtsmuller, A.B.; Kushner, S.A. Fast-Spiking Parvalbumin Interneurons Are Frequently Myelinated in the Cerebral Cortex of Mice and Humans. Cereb. Cortex 2017, 27, 5001–5013. [Google Scholar] [CrossRef] [PubMed]
  41. Bartos, M.; Vida, I.; Frotscher, M.; Meyer, A.; Monyer, H.; Geiger, J.R.P.; Jonas, P. Fast Synaptic Inhibition Promotes Synchronized Gamma Oscillations in Hippocampal Interneuron Networks. Proc. Natl. Acad. Sci. USA 2002, 99, 13222. [Google Scholar] [CrossRef] [PubMed]
  42. Hájos, N.; Pálhalmi, J.; Mann, E.O.; Németh, B.; Paulsen, O.; Freund, T.F. Spike Timing of Distinct Types of GABAergic Interneuron during Hippocampal Gamma Oscillations In Vitro. J. Neurosci. 2004, 24, 9127. [Google Scholar] [CrossRef] [PubMed]
  43. Oren, I.; Mann, E.O.; Paulsen, O.; Hájos, N. Synaptic Currents in Anatomically Identified CA3 Neurons during Hippocampal Gamma Oscillations In Vitro. J. Neurosci. 2006, 26, 9923. [Google Scholar] [CrossRef] [PubMed]
  44. Penttonen, M.; Kamondi, A.; Acsády, L.; Buzsáki, G. Gamma Frequency Oscillation in the Hippocampus of the Rat: Intracellular Analysis in Vivo. Eur. J. Neurosci. 1998, 10, 718–728. [Google Scholar] [CrossRef] [PubMed]
  45. Klausberger, T.; Roberts, J.D.B.; Somogyi, P. Cell Type- and Input-Specific Differences in the Number and Subtypes of Synaptic GABAA Receptors in the Hippocampus. J. Neurosci. 2002, 22, 2513. [Google Scholar] [CrossRef]
  46. Möhler, H. GABAA Receptor Diversity and Pharmacology. Cell Tissue Res. 2006, 326, 505–516. [Google Scholar] [CrossRef]
  47. Hensch, T.K. Critical Period Plasticity in Local Cortical Circuits. Nat. Rev. Neurosci. 2005, 6, 877–888. [Google Scholar] [CrossRef] [PubMed]
  48. Do, K.Q.; Cuenod, M.; Hensch, T.K. Targeting Oxidative Stress and Aberrant Critical Period Plasticity in the Developmental Trajectory to Schizophrenia. Schizophr. Bull. 2015, 41, 835–846. [Google Scholar] [CrossRef] [PubMed]
  49. Berardi, N.; Pizzorusso, T.; Maffei, L. Extracellular Matrix and Visual Cortical Plasticity: Freeing the Synapse. Neuron 2004, 44, 905–908. [Google Scholar] [CrossRef] [PubMed]
  50. Celio, M.R.; Spreafico, R.; De Biasi, S.; Vitellaro-Zuccarello, L. Perineuronal Nets: Past and Present. Trends Neurosci. 1998, 21, 510–515. [Google Scholar] [CrossRef] [PubMed]
  51. Dityatev, A.; Schachner, M.; Sonderegger, P. The Dual Role of the Extracellular Matrix in Synaptic Plasticity and Homeostasis. Nat. Rev. Neurosci. 2010, 11, 735–746. [Google Scholar] [CrossRef] [PubMed]
  52. Gundelfinger, E.D.; Frischknecht, R.; Choquet, D.; Heine, M. Converting Juvenile into Adult Plasticity: A Role for the Brain’s Extracellular Matrix. Eur. J. Neurosci. 2010, 31, 2156–2165. [Google Scholar] [CrossRef]
  53. Pizzorusso, T.; Medini, P.; Berardi, N.; Chierzi, S.; Fawcett, J.W.; Maffei, L. Reactivation of Ocular Dominance Plasticity in the Adult Visual Cortex. Science 2002, 298, 1248. [Google Scholar] [CrossRef] [PubMed]
  54. Härtig, W.; Derouiche, A.; Welt, K.; Brauer, K.; Grosche, J.; Mäder, M.; Reichenbach, A.; Brückner, G. Cortical Neurons Immunoreactive for the Potassium Channel Kv3.1b Subunit Are Predominantly Surrounded by Perineuronal Nets Presumed as a Buffering System for Cations. Brain Res. 1999, 842, 15–29. [Google Scholar] [CrossRef] [PubMed]
  55. Lensjø, K.K.; Lepperød, M.E.; Dick, G.; Hafting, T.; Fyhn, M. Removal of Perineuronal Nets Unlocks Juvenile Plasticity Through Network Mechanisms of Decreased Inhibition and Increased Gamma Activity. J. Neurosci. 2017, 37, 1269–1283. [Google Scholar] [CrossRef] [PubMed]
  56. Morris, N.P.; Henderson, Z. Perineuronal Nets Ensheath Fast Spiking, Parvalbumin-Immunoreactive Neurons in the Medial Septum/Diagonal Band Complex. Eur. J. Neurosci. 2000, 12, 828–838. [Google Scholar] [CrossRef] [PubMed]
  57. Wen, T.H.; Binder, D.K.; Ethell, I.M.; Razak, K.A. The Perineuronal ‘Safety’ Net? Perineuronal Net Abnormalities in Neurological Disorders. Front. Mol. Neurosci. 2018, 11, 270. [Google Scholar] [CrossRef] [PubMed]
  58. Cabungcal, J.H.; Counotte, D.S.; Lewis, E.; Tejeda, H.A.; Piantadosi, P.; Pollock, C.; Calhoon, G.G.; Sullivan, E.; Presgraves, E.; Kil, J.; et al. Juvenile Antioxidant Treatment Prevents Adult Deficits in a Developmental Model of Schizophrenia. Neuron 2014, 83, 1073–1084. [Google Scholar] [CrossRef] [PubMed]
  59. Morawski, M.; Brückner, M.K.; Riederer, P.; Brückner, G.; Arendt, T. Perineuronal Nets Potentially Protect against Oxidative Stress. Exp. Neurol. 2004, 188, 309–315. [Google Scholar] [CrossRef] [PubMed]
  60. McRae, P.A.; Rocco, M.M.; Kelly, G.; Brumberg, J.C.; Matthews, R.T. Sensory Deprivation Alters Aggrecan and Perineuronal Net Expression in the Mouse Barrel Cortex. J. Neurosci. 2007, 27, 5405. [Google Scholar] [CrossRef] [PubMed]
  61. Reimers, S.; Hartlage-Rübsamen, M.; Brückner, G.; Roßner, S. Formation of Perineuronal Nets in Organotypic Mouse Brain Slice Cultures Is Independent of Neuronal Glutamatergic Activity. Eur. J. Neurosci. 2007, 25, 2640–2648. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, D.; Fawcett, J. The Perineuronal Net and the Control of CNS Plasticity. Cell Tissue Res. 2012, 349, 147–160. [Google Scholar] [CrossRef] [PubMed]
  63. Pirbhoy, P.S.; Rais, M.; Lovelace, J.W.; Woodard, W.; Razak, K.A.; Binder, D.K.; Ethell, I.M. Acute Pharmacological Inhibition of Matrix Metalloproteinase-9 Activity during Development Restores Perineuronal Net Formation and Normalizes Auditory Processing in Fmr1 KO Mice. J. Neurochem. 2020, 155, 538–558. [Google Scholar] [CrossRef] [PubMed]
  64. Steullet, P.; Cabungcal, J.-H.; Coyle, J.; Didriksen, M.; Gill, K.; Grace, A.A.; Hensch, T.K.; LaMantia, A.-S.; Lindemann, L.; Maynard, T.M.; et al. Oxidative Stress-Driven Parvalbumin Interneuron Impairment as a Common Mechanism in Models of Schizophrenia. Mol. Psychiatry 2017, 22, 936–943. [Google Scholar] [CrossRef]
  65. Dwir, D.; Giangreco, B.; Xin, L.; Tenenbaum, L.; Cabungcal, J.-H.; Steullet, P.; Goupil, A.; Cleusix, M.; Jenni, R.; Chtarto, A.; et al. MMP9/RAGE Pathway Overactivation Mediates Redox Dysregulation and Neuroinflammation, Leading to Inhibitory/Excitatory Imbalance: A Reverse Translation Study in Schizophrenia Patients. Mol. Psychiatry 2020, 25, 2889–2904. [Google Scholar] [CrossRef] [PubMed]
  66. Guidotti, A.; Auta, J.; Davis, J.M.; Gerevini, V.D.; Dwivedi, Y.; Grayson, D.R.; Impagnatiello, F.; Pandey, G.; Pesold, C.; Sharma, R.; et al. Decrease in Reelin and Glutamic Acid Decarboxylase67 (GAD67) Expression in Schizophrenia and Bipolar Disorder: A Postmortem Brain Study. Arch. Gen. Psychiatry 2000, 57, 1061–1069. [Google Scholar] [CrossRef] [PubMed]
  67. Pesold, C.; Liu, W.S.; Guidotti, A.; Costa, E.; Caruncho, H.J. Cortical Bitufted, Horizontal, and Martinotti Cells Preferentially Express and Secrete Reelin into Perineuronal Nets, Nonsynaptically Modulating Gene Expression. Proc. Natl. Acad. Sci. USA 1999, 96, 3217–3222. [Google Scholar] [CrossRef] [PubMed]
  68. Berretta, S.; Pantazopoulos, H.; Markota, M.; Brown, C.; Batzianouli, E.T. Losing the Sugar Coating: Potential Impact of Perineuronal Net Abnormalities on Interneurons in Schizophrenia. Schizophr. Res. 2015, 167, 18–27. [Google Scholar] [CrossRef]
  69. Fatemi, S.H.; Earle, J.A.; McMenomy, T. Reduction in Reelin Immunoreactivity in Hippocampus of Subjects with Schizophrenia, Bipolar Disorder and Major Depression. Mol. Psychiatry 2000, 5, 654–663. [Google Scholar] [CrossRef] [PubMed]
  70. Gursoy-Ozdemir, Y.; Qiu, J.; Matsuoka, N.; Bolay, H.; Bermpohl, D.; Jin, H.; Wang, X.; Rosenberg, G.A.; Lo, E.H.; Moskowitz, M.A. Cortical Spreading Depression Activates and Upregulates MMP-9. J. Clin. Investig. 2004, 113, 1447–1455. [Google Scholar] [CrossRef] [PubMed]
  71. Rybakowski, J.K.; Remlinger-Molenda, A.; Czech-Kucharska, A.; Wojcicka, M.; Michalak, M.; Losy, J. Increased Serum Matrix Metalloproteinase-9 (MMP-9) Levels in Young Patients during Bipolar Depression. J. Affect. Disord. 2013, 146, 286–289. [Google Scholar] [CrossRef] [PubMed]
  72. Sorg, B.A.; Berretta, S.; Blacktop, J.M.; Fawcett, J.W.; Kitagawa, H.; Kwok, J.C.F.; Miquel, M. Casting a Wide Net: Role of Perineuronal Nets in Neural Plasticity. J. Neurosci. 2016, 36, 11459. [Google Scholar] [CrossRef] [PubMed]
  73. Bordt, E.A.; Polster, B.M. NADPH Oxidase- and Mitochondria-Derived Reactive Oxygen Species in Proinflammatory Microglial Activation: A Bipartisan Affair? Free Radic. Biol. Med. 2014, 76, 34–46. [Google Scholar] [CrossRef]
  74. Ganzinelli, S.; Borda, E.; Sterin-Borda, L. Autoantibodies from Schizophrenia Patients Induce Cerebral Cox-1/iNOS mRNA Expression with NO/PGE2/MMP-3 Production. Int. J. Neuropsychopharmacol. 2010, 13, 293–303. [Google Scholar] [CrossRef] [PubMed]
  75. Tendilla-Beltrán, H.; Sanchez-Islas, N.D.C.; Marina-Ramos, M.; Leza, J.C.; Flores, G. The Prefrontal Cortex as a Target for Atypical Antipsychotics in Schizophrenia, Lessons of Neurodevelopmental Animal Models. Prog. Neurobiol. 2021, 199, 101967. [Google Scholar] [CrossRef] [PubMed]
  76. François, J.; Ferrandon, A.; Koning, E.; Angst, M.-J.; Sandner, G.; Nehlig, A. Selective Reorganization of GABAergic Transmission in Neonatal Ventral Hippocampal-Lesioned Rats. Int. J. Neuropsychopharmacol. 2009, 12, 1097–1110. [Google Scholar] [CrossRef] [PubMed]
  77. Tendilla-Beltrán, H.; Meneses-Prado, S.; Vázquez-Roque, R.A.; Tapia-Rodríguez, M.; Vázquez-Hernández, A.J.; Coatl-Cuaya, H.; Martín-Hernández, D.; MacDowell, K.S.; Garcés-Ramírez, L.; Leza, J.C.; et al. Risperidone Ameliorates Prefrontal Cortex Neural Atrophy and Oxidative/Nitrosative Stress in Brain and Peripheral Blood of Rats with Neonatal Ventral Hippocampus Lesion. J. Neurosci. Off. J. Soc. Neurosci. 2019, 39, 8584–8599. [Google Scholar] [CrossRef] [PubMed]
  78. Casey, C.; Fullard, J.F.; Sleator, R.D. Unravelling the Genetic Basis of Schizophrenia. Gene 2024, 902, 148198. [Google Scholar] [CrossRef]
  79. Zhao, M.-Z.; Song, X.-S.; Ma, J.-S. Gene × Environment Interaction in Major Depressive Disorder. World J. Clin. Cases 2021, 9, 9368–9375. [Google Scholar] [CrossRef] [PubMed]
  80. Brown, A.S. The Environment and Susceptibility to Schizophrenia. Prog. Neurobiol. 2011, 93, 23–58. [Google Scholar] [CrossRef]
  81. Stapp, E.K.; Mendelson, T.; Merikangas, K.R.; Wilcox, H.C. Parental Bipolar Disorder, Family Environment, and Offspring Psychiatric Disorders: A Systematic Review. J. Affect. Disord. 2020, 268, 69–81. [Google Scholar] [CrossRef]
  82. Grace, A.A. Dysregulation of the Dopamine System in the Pathophysiology of Schizophrenia and Depression. Nat. Rev. Neurosci. 2016, 17, 524–532. [Google Scholar] [CrossRef] [PubMed]
  83. Berk, M.; Dodd, S.; Kauer-Sant’anna, M.; Malhi, G.S.; Bourin, M.; Kapczinski, F.; Norman, T. Dopamine Dysregulation Syndrome: Implications for a Dopamine Hypothesis of Bipolar Disorder. Acta Psychiatr. Scand. Suppl. 2007, 116, 41–49. [Google Scholar] [CrossRef] [PubMed]
  84. Szczypiński, J.J.; Gola, M. Dopamine Dysregulation Hypothesis: The Common Basis for Motivational Anhedonia in Major Depressive Disorder and Schizophrenia? Rev. Neurosci. 2018, 29, 727–744. [Google Scholar] [CrossRef] [PubMed]
  85. Kraguljac, N.V.; McDonald, W.M.; Widge, A.S.; Rodriguez, C.I.; Tohen, M.; Nemeroff, C.B. Neuroimaging Biomarkers in Schizophrenia. Am. J. Psychiatry 2021, 178, 509–521. [Google Scholar] [CrossRef] [PubMed]
  86. Qiu, H.; Li, J. Major Depressive Disorder and Magnetic Resonance Imaging: A Mini-Review of Recent Progress. Curr. Pharm. Des. 2018, 24, 2524–2529. [Google Scholar] [CrossRef]
  87. Abé, C.; Liberg, B.; Klahn, A.L.; Petrovic, P.; Landén, M. Mania-Related Effects on Structural Brain Changes in Bipolar Disorder-A Narrative Review of the Evidence. Mol. Psychiatry 2023, 28, 2674–2682. [Google Scholar] [CrossRef] [PubMed]
  88. Charlson, F.J.; Ferrari, A.J.; Santomauro, D.F.; Diminic, S.; Stockings, E.; Scott, J.G.; McGrath, J.J.; Whiteford, H.A. Global Epidemiology and Burden of Schizophrenia: Findings From the Global Burden of Disease Study 2016. Schizophr. Bull. 2018, 44, 1195–1203. [Google Scholar] [CrossRef] [PubMed]
  89. Freedman, R.; Waldo, M.; Bickford-Wimer, P.; Nagamoto, H. Elementary Neuronal Dysfunctions in Schizophrenia. Schizophr. Res. 1991, 4, 233–243. [Google Scholar] [CrossRef] [PubMed]
  90. Pearlson, G.D. Neurobiology of Schizophrenia. Ann. Neurol. 2000, 48, 556–566. [Google Scholar] [CrossRef] [PubMed]
  91. Ross, C.A.; Margolis, R.L.; Reading, S.A.J.; Pletnikov, M.; Coyle, J.T. Neurobiology of Schizophrenia. Neuron 2006, 52, 139–153. [Google Scholar] [CrossRef]
  92. Ross, C.A.; Margolis, R.L. Neurogenetics: Insights into Degenerative Diseases and Approaches to Schizophrenia. Clin. Neurosci. Res. 2005, 5, 3–14. [Google Scholar] [CrossRef]
  93. Bilbo, S.D.; Schwarz, J.M. Early-Life Programming of Later-Life Brain and Behavior: A Critical Role for the Immune System. Front. Behav. Neurosci. 2009, 3, 14. [Google Scholar] [CrossRef] [PubMed]
  94. Buka, S.L.; Tsuang, M.T.; Torrey, E.F.; Klebanoff, M.A.; Wagner, R.L.; Yolken, R.H. Maternal Cytokine Levels during Pregnancy and Adult Psychosis. Brain. Behav. Immun. 2001, 15, 411–420. [Google Scholar] [CrossRef] [PubMed]
  95. Dickerson, F.; Jones-Brando, L.; Ford, G.; Genovese, G.; Stallings, C.; Origoni, A.; O’Dushlaine, C.; Katsafanas, E.; Sweeney, K.; Khushalani, S.; et al. Schizophrenia Is Associated With an Aberrant Immune Response to Epstein-Barr Virus. Schizophr. Bull. 2019, 45, 1112–1119. [Google Scholar] [CrossRef] [PubMed]
  96. Trovão, N.; Prata, J.; VonDoellinger, O.; Santos, S.; Barbosa, M.; Coelho, R. Peripheral Biomarkers for First-Episode Psychosis—Opportunities from the Neuroinflammatory Hypothesis of Schizophrenia. Psychiatry Investig. 2019, 16, 177–184. [Google Scholar] [CrossRef] [PubMed]
  97. Yolken, R.H.; Torrey, E.F. Are Some Cases of Psychosis Caused by Microbial Agents? A Review of the Evidence. Mol. Psychiatry 2008, 13, 470–479. [Google Scholar] [CrossRef]
  98. Buckley, P.F. Neuroinflammation and Schizophrenia. Curr. Psychiatry Rep. 2019, 21, 72. [Google Scholar] [CrossRef] [PubMed]
  99. Monji, A.; Kato, T.A.; Mizoguchi, Y.; Horikawa, H.; Seki, Y.; Kasai, M.; Yamauchi, Y.; Yamada, S.; Kanba, S. Neuroinflammation in Schizophrenia Especially Focused on the Role of Microglia. Prog. Neuropsychopharmacol. Biol. Psychiatry 2013, 42, 115–121. [Google Scholar] [CrossRef] [PubMed]
  100. Rothermundt, M.; Arolt, V.; Peters, M.; Gutbrodt, H.; Fenker, J.; Kersting, A.; Kirchner, H. Inflammatory Markers in Major Depression and Melancholia. J. Affect. Disord. 2001, 63, 93–102. [Google Scholar] [CrossRef] [PubMed]
  101. Watkins, C.C.; Andrews, S.R. Clinical Studies of Neuroinflammatory Mechanisms in Schizophrenia. Schizophr. Res. 2016, 176, 14–22. [Google Scholar] [CrossRef] [PubMed]
  102. Semiz, M.; Yildirim, O.; Canan, F.; Demir, S.; Hasbek, E.; Tuman, T.C.; Kayka, N.; Tosun, M. Elevated Neutrophil/Lymphocyte Ratio in Patients with Schizophrenia. Psychiatr. Danub. 2014, 26, 220–225. [Google Scholar] [PubMed]
  103. Hof, P.R.; Haroutunian, V.; Friedrich, V.L.; Byne, W.; Buitron, C.; Perl, D.P.; Davis, K.L. Loss and Altered Spatial Distribution of Oligodendrocytes in the Superior Frontal Gyrus in Schizophrenia. Biol. Psychiatry 2003, 53, 1075–1085. [Google Scholar] [CrossRef] [PubMed]
  104. Schwartz, M.; Butovsky, O.; Brück, W.; Hanisch, U.-K. Microglial Phenotype: Is the Commitment Reversible? Trends Neurosci. 2006, 29, 68–74. [Google Scholar] [CrossRef] [PubMed]
  105. Wierzba-Bobrowicz, T.; Lewandowska, E.; Kosno-Kruszewska, E.; Lechowicz, W.; Pasennik, E.; Schmidt-Sidor, B. Degeneration of Microglial Cells in Frontal and Temporal Lobes of Chronic Schizophrenics. Folia Neuropathol. 2004, 42, 157–165. [Google Scholar]
  106. Cuenod, M.; Steullet, P.; Cabungcal, J.-H.; Dwir, D.; Khadimallah, I.; Klauser, P.; Conus, P.; Do, K.Q. Caught in Vicious Circles: A Perspective on Dynamic Feed-Forward Loops Driving Oxidative Stress in Schizophrenia. Mol. Psychiatry 2022, 27, 1886–1897. [Google Scholar] [CrossRef] [PubMed]
  107. Kohen, R.; Nyska, A. Oxidation of Biological Systems: Oxidative Stress Phenomena, Antioxidants, Redox Reactions, and Methods for Their Quantification. Toxicol. Pathol. 2002, 30, 620–650. [Google Scholar] [CrossRef] [PubMed]
  108. Cabungcal, J.-H.; Steullet, P.; Morishita, H.; Kraftsik, R.; Cuenod, M.; Hensch, T.K.; Do, K.Q. Perineuronal Nets Protect Fast-Spiking Interneurons against Oxidative Stress. Proc. Natl. Acad. Sci. USA 2013, 110, 9130–9135. [Google Scholar] [CrossRef] [PubMed]
  109. Sullivan, E.M.; O’Donnell, P. Inhibitory Interneurons, Oxidative Stress, and Schizophrenia. Schizophr. Bull. 2012, 38, 373–376. [Google Scholar] [CrossRef] [PubMed]
  110. Kann, O.; Huchzermeyer, C.; Kovács, R.; Wirtz, S.; Schuelke, M. Gamma Oscillations in the Hippocampus Require High Complex I Gene Expression and Strong Functional Performance of Mitochondria. Brain 2011, 134, 345–358. [Google Scholar] [CrossRef] [PubMed]
  111. Dwir, D.; Khadimallah, I.; Xin, L.; Rahman, M.; Du, F.; Öngür, D.; Do, K.Q. Redox and Immune Signaling in Schizophrenia: New Therapeutic Potential. Int. J. Neuropsychopharmacol. 2023, 26, pyad012. [Google Scholar] [CrossRef]
  112. Steullet, P.; Cabungcal, J.H.; Monin, A.; Dwir, D.; O’Donnell, P.; Cuenod, M.; Do, K.Q. Redox Dysregulation, Neuroinflammation, and NMDA Receptor Hypofunction: A “Central Hub” in Schizophrenia Pathophysiology? Schizophr. Res. 2016, 176, 41–51. [Google Scholar] [CrossRef] [PubMed]
  113. Jenkins, T.A.; Harte, M.K.; Stenson, G.; Reynolds, G.P. Neonatal Lipopolysaccharide Induces Pathological Changes in Parvalbumin Immunoreactivity in the Hippocampus of the Rat. Behav. Brain Res. 2009, 205, 355–359. [Google Scholar] [CrossRef] [PubMed]
  114. Meyer, U.; Nyffeler, M.; Yee, B.K.; Knuesel, I.; Feldon, J. Adult Brain and Behavioral Pathological Markers of Prenatal Immune Challenge during Early/Middle and Late Fetal Development in Mice. Brain. Behav. Immun. 2008, 22, 469–486. [Google Scholar] [CrossRef] [PubMed]
  115. Deverman, B.E.; Patterson, P.H. Cytokines and CNS Development. Neuron 2009, 64, 61–78. [Google Scholar] [CrossRef] [PubMed]
  116. Jonakait, G.M. The Effects of Maternal Inflammation on Neuronal Development: Possible Mechanisms. Int. J. Dev. Neurosci. Off. J. Int. Soc. Dev. Neurosci. 2007, 25, 415–425. [Google Scholar] [CrossRef] [PubMed]
  117. Meyer, U. Developmental Neuroinflammation and Schizophrenia. Prog. Neuropsychopharmacol. Biol. Psychiatry 2013, 42, 20–34. [Google Scholar] [CrossRef] [PubMed]
  118. Clark, D.; Dedova, I.; Cordwell, S.; Matsumoto, I. A Proteome Analysis of the Anterior Cingulate Cortex Gray Matter in Schizophrenia. Mol. Psychiatry 2006, 11, 459–470. [Google Scholar] [CrossRef] [PubMed]
  119. Gawryluk, J.W.; Wang, J.-F.; Andreazza, A.C.; Shao, L.; Young, L.T. Decreased Levels of Glutathione, the Major Brain Antioxidant, in Post-Mortem Prefrontal Cortex from Patients with Psychiatric Disorders. Int. J. Neuropsychopharmacol. 2011, 14, 123–130. [Google Scholar] [CrossRef] [PubMed]
  120. Nishioka, N.; Arnold, S.E. Evidence for Oxidative DNA Damage in the Hippocampus of Elderly Patients with Chronic Schizophrenia. Am. J. Geriatr. Psychiatry 2004, 12, 167–175. [Google Scholar] [CrossRef]
  121. Prabakaran, S.; Swatton, J.E.; Ryan, M.M.; Huffaker, S.J.; Huang, J.-J.; Griffin, J.L.; Wayland, M.; Freeman, T.; Dudbridge, F.; Lilley, K.S.; et al. Mitochondrial Dysfunction in Schizophrenia: Evidence for Compromised Brain Metabolism and Oxidative Stress. Mol. Psychiatry 2004, 9, 684–697. [Google Scholar] [CrossRef] [PubMed]
  122. Yao, J.K.; Leonard, S.; Reddy, R.D. Increased Nitric Oxide Radicals in Postmortem Brain from Patients with Schizophrenia. Schizophr. Bull. 2004, 30, 923–934. [Google Scholar] [CrossRef] [PubMed]
  123. Longson, D.; Deakin, J.F.W.; Benes, F.M. Increased Density of Entorhinal Glutamate-Immunoreactive Vertical Fibers in Schizophrenia. J. Neural Transm. 1996, 103, 503–507. [Google Scholar] [CrossRef] [PubMed]
  124. Pantazopoulos, H.; Markota, M.; Jaquet, F.; Ghosh, D.; Wallin, A.; Santos, A.; Caterson, B.; Berretta, S. Aggrecan and Chondroitin-6-Sulfate Abnormalities in Schizophrenia and Bipolar Disorder: A Postmortem Study on the Amygdala. Transl. Psychiatry 2015, 5, e496. [Google Scholar] [CrossRef] [PubMed]
  125. Reynolds, G.P. Developments in the Drug Treatment of Schizophrenia. Trends Pharmacol. Sci. 1992, 13, 116–121. [Google Scholar] [CrossRef]
  126. Pantazopoulos, H.; Woo, T.-U.W.; Lim, M.P.; Lange, N.; Berretta, S. Extracellular Matrix-Glial Abnormalities in the Amygdala and Entorhinal Cortex of Subjects Diagnosed With Schizophrenia. Arch. Gen. Psychiatry 2010, 67, 155–166. [Google Scholar] [CrossRef] [PubMed]
  127. Pasternak, O.; Westin, C.-F.; Bouix, S.; Seidman, L.J.; Goldstein, J.M.; Woo, T.-U.W.; Petryshen, T.L.; Mesholam-Gately, R.I.; McCarley, R.W.; Kikinis, R.; et al. Excessive Extracellular Volume Reveals a Neurodegenerative Pattern in Schizophrenia Onset. J. Neurosci. Off. J. Soc. Neurosci. 2012, 32, 17365–17372. [Google Scholar] [CrossRef] [PubMed]
  128. Buxbaum, J.D.; Georgieva, L.; Young, J.J.; Plescia, C.; Kajiwara, Y.; Jiang, Y.; Moskvina, V.; Norton, N.; Peirce, T.; Williams, H.; et al. Molecular Dissection of NRG1-ERBB4 Signaling Implicates PTPRZ1 as a Potential Schizophrenia Susceptibility Gene. Mol. Psychiatry 2008, 13, 162–172. [Google Scholar] [CrossRef]
  129. 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] [PubMed]
  130. Mauney, S.A.; Athanas, K.M.; Pantazopoulos, H.; Shaskan, N.; Passeri, E.; Berretta, S.; Woo, T.-U.W. Developmental Pattern of Perineuronal Nets in the Human Prefrontal Cortex and Their Deficit in Schizophrenia. Biol. Psychiatry 2013, 74, 427–435. [Google Scholar] [CrossRef] [PubMed]
  131. Enwright, J.F.; Sanapala, S.; Foglio, A.; Berry, R.; Fish, K.N.; Lewis, D.A. Reduced Labeling of Parvalbumin Neurons and Perineuronal Nets in the Dorsolateral Prefrontal Cortex of Subjects with Schizophrenia. Neuropsychopharmacology 2016, 41, 2206–2214. [Google Scholar] [CrossRef] [PubMed]
  132. Bourgeois, J.-P.; Goldman-Rakic, P.S.; Rakic, P. Synaptogenesis in the Prefrontal Cortex of Rhesus Monkeys. Cereb. Cortex 1994, 4, 78–96. [Google Scholar] [CrossRef] [PubMed]
  133. Goldman-Rakic, P.S.; Selemon, L.D. Functional and Anatomical Aspects of Prefrontal Pathology in Schizophrenia. Schizophr. Bull. 1997, 23, 437–458. [Google Scholar] [CrossRef] [PubMed]
  134. Woo, T.U.; Miller, J.L.; Lewis, D.A. Schizophrenia and the Parvalbumin-Containing Class of Cortical Local Circuit Neurons. Am. J. Psychiatry 1997, 154, 1013–1015. [Google Scholar] [CrossRef] [PubMed]
  135. Cabungcal, J.-H.; Steullet, P.; Kraftsik, R.; Cuenod, M.; Do, K.Q. Early-Life Insults Impair Parvalbumin Interneurons via Oxidative Stress: Reversal by N-Acetylcysteine. Biol. Psychiatry 2013, 73, 574–582. [Google Scholar] [CrossRef] [PubMed]
  136. Pierri, J.N.; Volk, C.L.; Auh, S.; Sampson, A.; Lewis, D.A. Decreased Somal Size of Deep Layer 3 Pyramidal Neurons in the Prefrontal Cortex of Subjects with Schizophrenia. Arch. Gen. Psychiatry 2001, 58, 466–473. [Google Scholar] [CrossRef]
  137. Curley, A.A.; Eggan, S.M.; Lazarus, M.S.; Huang, Z.J.; Volk, D.W.; Lewis, D.A. Role of Glutamic Acid Decarboxylase 67 in Regulating Cortical Parvalbumin and GABA Membrane Transporter 1 Expression: Implications for Schizophrenia. Neurobiol. Dis. 2013, 50, 179–186. [Google Scholar] [CrossRef] [PubMed]
  138. Jaaro-Peled, H.; Ayhan, Y.; Pletnikov, M.V.; Sawa, A. Review of Pathological Hallmarks of Schizophrenia: Comparison of Genetic Models With Patients and Nongenetic Models. Schizophr. Bull. 2010, 36, 301–313. [Google Scholar] [CrossRef] [PubMed]
  139. Reynolds, G.P. The Neurochemical Pathology of Schizophrenia: Post-Mortem Studies from Dopamine to Parvalbumin. J. Neural Transm. 2021, 129, 643–647. [Google Scholar] [CrossRef] [PubMed]
  140. Hashimoto, T.; Volk, D.W.; Eggan, S.M.; Mirnics, K.; Pierri, J.N.; Sun, Z.; Sampson, A.R.; Lewis, D.A. Gene Expression Deficits in a Subclass of GABA Neurons in the Prefrontal Cortex of Subjects with Schizophrenia. J. Neurosci. 2003, 23, 6315–6326. [Google Scholar] [CrossRef]
  141. Glausier, J.R.; Fish, K.N.; Lewis, D.A. Altered Parvalbumin Basket Cell Inputs in the Dorsolateral Prefrontal Cortex of Schizophrenia Subjects. Mol. Psychiatry 2014, 19, 30–36. [Google Scholar] [CrossRef] [PubMed]
  142. Lewis, D.A.; Curley, A.A.; Glausier, J.R.; Volk, D.W. Cortical Parvalbumin Interneurons and Cognitive Dysfunction in Schizophrenia. Trends Neurosci. 2012, 35, 57–67. [Google Scholar] [CrossRef]
  143. Nyíri, G.; Freund, T.F.; Somogyi, P. Input-Dependent Synaptic Targeting of A2-Subunit-Containing GABAA Receptors in Synapses of Hippocampal Pyramidal Cells of the Rat. Eur. J. Neurosci. 2001, 13, 428–442. [Google Scholar] [CrossRef] [PubMed]
  144. Beneyto, M.; Abbott, A.; Hashimoto, T.; Lewis, D.A. Lamina-Specific Alterations in Cortical GABAA Receptor Subunit Expression in Schizophrenia. Cereb. Cortex 2011, 21, 999–1011. [Google Scholar] [CrossRef] [PubMed]
  145. Hashimoto, T.; Bazmi, H.H.; Mirnics, K.; Wu, Q.; Sampson, A.R.; Lewis, D.A. Conserved Regional Patterns of GABA-Related Transcript Expression in the Neocortex of Subjects With Schizophrenia. Am. J. Psychiatry 2008, 165, 479–489. [Google Scholar] [CrossRef] [PubMed]
  146. Glausier, J.R.; Lewis, D.A. Selective Pyramidal Cell Reduction of GABAA Receptor A1 Subunit Messenger RNA Expression in Schizophrenia. Neuropsychopharmacology 2011, 36, 2103–2110. [Google Scholar] [CrossRef] [PubMed]
  147. Volk, D.W.; Pierri, J.N.; Fritschy, J.-M.; Auh, S.; Sampson, A.R.; Lewis, D.A. Reciprocal Alterations in Pre- and Postsynaptic Inhibitory Markers at Chandelier Cell Inputs to Pyramidal Neurons in Schizophrenia. Cereb. Cortex 2002, 12, 1063–1070. [Google Scholar] [CrossRef] [PubMed]
  148. Lewis, D.A. GABAergic Local Circuit Neurons and Prefrontal Cortical Dysfunction in Schizophrenia. Brain Res. Rev. 2000, 31, 270–276. [Google Scholar] [CrossRef] [PubMed]
  149. Jones, E.G. Cortical Development and Thalamic Pathology in Schizophrenia1. Schizophr. Bull. 1997, 23, 483–501. [Google Scholar] [CrossRef] [PubMed]
  150. Pierri, J.N.; Chaudry, A.S.; Woo, T.-U.W.; Lewis, D.A. Alterations in Chandelier Neuron Axon Terminals in the Prefrontal Cortex of Schizophrenic Subjects. Am. J. Psychiatry 1999, 156, 1709–1719. [Google Scholar] [CrossRef] [PubMed]
  151. Fazzari, P.; Paternain, A.V.; Valiente, M.; Pla, R.; Luján, R.; Lloyd, K.; Lerma, J.; Marín, O.; Rico, B. Control of Cortical GABA Circuitry Development by Nrg1 and ErbB4 Signalling. Nature 2010, 464, 1376–1380. [Google Scholar] [CrossRef] [PubMed]
  152. Chen, Y.-J.; Zhang, M.; Yin, D.-M.; Wen, L.; Ting, A.; Wang, P.; Lu, Y.-S.; Zhu, X.-H.; Li, S.-J.; Wu, C.-Y.; et al. ErbB4 in Parvalbumin-Positive Interneurons Is Critical for Neuregulin 1 Regulation of Long-Term Potentiation. Proc. Natl. Acad. Sci. USA 2010, 107, 21818–21823. [Google Scholar] [CrossRef]
  153. Fisahn, A.; Neddens, J.; Yan, L.; Buonanno, A. Neuregulin-1 Modulates Hippocampal Gamma Oscillations: Implications for Schizophrenia. Cereb. Cortex 2009, 19, 612–618. [Google Scholar] [CrossRef] [PubMed]
  154. Flames, N.; Long, J.E.; Garratt, A.N.; Fischer, T.M.; Gassmann, M.; Birchmeier, C.; Lai, C.; Rubenstein, J.L.R.; Marín, O. Short- and Long-Range Attraction of Cortical GABAergic Interneurons by Neuregulin-1. Neuron 2004, 44, 251–261. [Google Scholar] [CrossRef] [PubMed]
  155. Lu, Y.; Sun, X.-D.; Hou, F.-Q.; Bi, L.-L.; Yin, D.-M.; Liu, F.; Chen, Y.-J.; Bean, J.C.; Jiao, H.-F.; Liu, X.; et al. Maintenance of GABAergic Activity by Neuregulin 1-ErbB4 in Amygdala for Fear Memory. Neuron 2014, 84, 835–846. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, Z.J.; Reynolds, G.P. A Selective Decrease in the Relative Density of Parvalbumin-Immunoreactive Neurons in the Hippocampus in Schizophrenia. Schizophr. Res. 2002, 55, 1–10. [Google Scholar] [CrossRef] [PubMed]
  157. Antonoudiou, P.; Tan, Y.L.; Kontou, G.; Upton, A.L.; Mann, E.O. Parvalbumin and Somatostatin Interneurons Contribute to the Generation of Hippocampal Gamma Oscillations. J. Neurosci. 2020, 40, 7668. [Google Scholar] [CrossRef] [PubMed]
  158. Weinberger, D. Cell Biology of the Hippocampal Formation in Schizophrenia. Biol. Psychiatry 1999, 45, 395–402. [Google Scholar] [CrossRef] [PubMed]
  159. Hikida, T.; Jaaro-Peled, H.; Seshadri, S.; Oishi, K.; Hookway, C.; Kong, S.; Wu, D.; Xue, R.; Andradé, M.; Tankou, S.; et al. Dominant-Negative DISC1 Transgenic Mice Display Schizophrenia-Associated Phenotypes Detected by Measures Translatable to Humans. Proc. Natl. Acad. Sci. USA 2007, 104, 14501–14506. [Google Scholar] [CrossRef] [PubMed]
  160. Shen, S.; Lang, B.; Nakamoto, C.; Zhang, F.; Pu, J.; Kuan, S.-L.; Chatzi, C.; He, S.; Mackie, I.; Brandon, N.J.; et al. Schizophrenia-Related Neural and Behavioral Phenotypes in Transgenic Mice Expressing Truncated Disc1. J. Neurosci. 2008, 28, 10893–10904. [Google Scholar] [CrossRef] [PubMed]
  161. Seshadri, S.; Kamiya, A.; Yokota, Y.; Prikulis, I.; Kano, S.; Hayashi-Takagi, A.; Stanco, A.; Eom, T.-Y.; Rao, S.; Ishizuka, K.; et al. Disrupted-in-Schizophrenia-1 Expression Is Regulated by Beta-Site Amyloid Precursor Protein Cleaving Enzyme-1-Neuregulin Cascade. Proc. Natl. Acad. Sci. USA 2010, 107, 5622–5627. [Google Scholar] [CrossRef] [PubMed]
  162. Goate, A.; Chartier-Harlin, M.-C.; Mullan, M.; Brown, J.; Crawford, F.; Fidani, L.; Giuffra, L.; Haynes, A.; Irving, N.; James, L.; et al. Segregation of a Missense Mutation in the Amyloid Precursor Protein Gene with Familial Alzheimer’s Disease. Nature 1991, 349, 704–706. [Google Scholar] [CrossRef]
  163. Young-Pearse, T.L.; Suth, S.; Luth, E.S.; Sawa, A.; Selkoe, D.J. Biochemical and Functional Interaction of Disrupted-in-Schizophrenia 1 and Amyloid Precursor Protein Regulates Neuronal Migration during Mammalian Cortical Development. J. Neurosci. 2010, 30, 10431–10440. [Google Scholar] [CrossRef]
  164. Lesh, T.A.; Niendam, T.A.; Minzenberg, M.J.; Carter, C.S. Cognitive Control Deficits in Schizophrenia: Mechanisms and Meaning. Neuropsychopharmacology 2011, 36, 316–338. [Google Scholar] [CrossRef] [PubMed]
  165. Uhlhaas, P.J.; Singer, W. Abnormal Neural Oscillations and Synchrony in Schizophrenia. Nat. Rev. Neurosci. 2010, 11, 100–113. [Google Scholar] [CrossRef] [PubMed]
  166. Woo, R.-S.; Li, X.-M.; Tao, Y.; Carpenter-Hyland, E.; Huang, Y.Z.; Weber, J.; Neiswender, H.; Dong, X.-P.; Wu, J.; Gassmann, M.; et al. Neuregulin-1 Enhances Depolarization-Induced GABA Release. Neuron 2007, 54, 599–610. [Google Scholar] [CrossRef] [PubMed]
  167. Lodge, D.J.; Behrens, M.M.; Grace, A.A. A Loss of Parvalbumin-Containing Interneurons Is Associated with Diminished Oscillatory Activity in an Animal Model of Schizophrenia. J. Neurosci. Off. J. Soc. Neurosci. 2009, 29, 2344–2354. [Google Scholar] [CrossRef] [PubMed]
  168. Nguyen, R.; Morrissey, M.D.; Mahadevan, V.; Cajanding, J.D.; Woodin, M.A.; Yeomans, J.S.; Takehara-Nishiuchi, K.; Kim, J.C. Parvalbumin and GAD65 Interneuron Inhibition in the Ventral Hippocampus Induces Distinct Behavioral Deficits Relevant to Schizophrenia. J. Neurosci. Off. J. Soc. Neurosci. 2014, 34, 14948–14960. [Google Scholar] [CrossRef] [PubMed]
  169. Kumari, V.; Soni, W.; Mathew, V.M.; Sharma, T. Prepulse Inhibition of the Startle Response in Men With Schizophrenia: Effects of Age of Onset of Illness, Symptoms, and Medication. Arch. Gen. Psychiatry 2000, 57, 609–614. [Google Scholar] [CrossRef]
  170. Cho, K.K.A.; Hoch, R.; Lee, A.T.; Patel, T.; Rubenstein, J.L.R.; Sohal, V.S. Gamma Rhythms Link Prefrontal Interneuron Dysfunction with Cognitive Inflexibility in Dlx5/6+/− Mice. Neuron 2015, 85, 1332–1343. [Google Scholar] [CrossRef] [PubMed]
  171. Kalweit, A.N.; Amanpour-Gharaei, B.; Colitti-Klausnitzer, J.; Manahan-Vaughan, D. Changes in Neuronal Oscillations Accompany the Loss of Hippocampal LTP That Occurs in an Animal Model of Psychosis. Front. Behav. Neurosci. 2017, 11, 36. [Google Scholar] [CrossRef] [PubMed]
  172. Grent-’t-Jong, T.; Rivolta, D.; Gross, J.; Gajwani, R.; Lawrie, S.M.; Schwannauer, M.; Heidegger, T.; Wibral, M.; Singer, W.; Sauer, A.; et al. Acute Ketamine Dysregulates Task-Related Gamma-Band Oscillations in Thalamo-Cortical Circuits in Schizophrenia. Brain J. Neurol. 2018, 141, 2511–2526. [Google Scholar] [CrossRef] [PubMed]
  173. Siegel, M.; Donner, T.H.; Engel, A.K. Spectral Fingerprints of Large-Scale Neuronal Interactions. Nat. Rev. Neurosci. 2012, 13, 121–134. [Google Scholar] [CrossRef] [PubMed]
  174. Rutter, L.; Carver, F.W.; Holroyd, T.; Nadar, S.R.; Mitchell-Francis, J.; Apud, J.; Weinberger, D.R.; Coppola, R. Magnetoencephalographic Gamma Power Reduction in Patients with Schizophrenia during Resting Condition. Hum. Brain Mapp. 2009, 30, 3254–3264. [Google Scholar] [CrossRef] [PubMed]
  175. Gruetzner, C.; Wibral, M.; Sun, L.; Rivolta, D.; Singer, W.; Maurer, K.; Uhlhaas, P. Deficits in High- (>60 Hz) Gamma-Band Oscillations during Visual Processing in Schizophrenia. Front. Hum. Neurosci. 2013, 7, 88. [Google Scholar] [CrossRef]
  176. Rivolta, D.; Heidegger, T.; Scheller, B.; Sauer, A.; Schaum, M.; Birkner, K.; Singer, W.; Wibral, M.; Uhlhaas, P.J. Ketamine Dysregulates the Amplitude and Connectivity of High-Frequency Oscillations in Cortical–Subcortical Networks in Humans: Evidence From Resting-State Magnetoencephalography-Recordings. Schizophr. Bull. 2015, 41, 1105–1114. [Google Scholar] [CrossRef] [PubMed]
  177. Spencer, K.M.; Nestor, P.G.; Perlmutter, R.; Niznikiewicz, M.A.; Klump, M.C.; Frumin, M.; Shenton, M.E.; McCarley, R.W. Neural Synchrony Indexes Disordered Perception and Cognition in Schizophrenia. Proc. Natl. Acad. Sci. USA 2004, 101, 17288–17293. [Google Scholar] [CrossRef] [PubMed]
  178. Leicht, G.; Kirsch, V.; Giegling, I.; Karch, S.; Hantschk, I.; Möller, H.-J.; Pogarell, O.; Hegerl, U.; Rujescu, D.; Mulert, C. Reduced Early Auditory Evoked Gamma-Band Response in Patients with Schizophrenia. Biol. Psychiatry 2010, 67, 224–231. [Google Scholar] [CrossRef] [PubMed]
  179. Leicht, G.; Vauth, S.; Polomac, N.; Andreou, C.; Rauh, J.; Mußmann, M.; Karow, A.; Mulert, C. EEG-Informed fMRI Reveals a Disturbed Gamma-Band-Specific Network in Subjects at High Risk for Psychosis. Schizophr. Bull. 2016, 42, 239–249. [Google Scholar] [CrossRef]
  180. Tada, M.; Kirihara, K.; Koshiyama, D.; Fujioka, M.; Usui, K.; Uka, T.; Komatsu, M.; Kunii, N.; Araki, T.; Kasai, K. Gamma-Band Auditory Steady-State Response as a Neurophysiological Marker for Excitation and Inhibition Balance: A Review for Understanding Schizophrenia and Other Neuropsychiatric Disorders. Clin. EEG Neurosci. 2020, 51, 234–243. [Google Scholar] [CrossRef] [PubMed]
  181. Tsuchimoto, R.; Kanba, S.; Hirano, S.; Oribe, N.; Ueno, T.; Hirano, Y.; Nakamura, I.; Oda, Y.; Miura, T.; Onitsuka, T. Reduced High and Low Frequency Gamma Synchronization in Patients with Chronic Schizophrenia. Schizophr. Res. 2011, 133, 99–105. [Google Scholar] [CrossRef] [PubMed]
  182. Wang, J.; Tang, Y.; Curtin, A.; Chan, R.C.K.; Wang, Y.; Li, H.; Zhang, T.; Qian, Z.; Guo, Q.; Li, Y.; et al. Abnormal Auditory-Evoked Gamma Band Oscillations in First-Episode Schizophrenia during both Eye Open and Eye Close States. Prog. Neuropsychopharmacol. Biol. Psychiatry 2018, 86, 279–286. [Google Scholar] [CrossRef] [PubMed]
  183. Dimitriadis, S.I.; Perry, G.; Foley, S.F.; Tansey, K.E.; Jones, D.K.; Holmans, P.; Zammit, S.; Hall, J.; O’Donovan, M.C.; Owen, M.J.; et al. Genetic Risk for Schizophrenia Is Associated with Altered Visually-Induced Gamma Band Activity: Evidence from a Population Sample Stratified Polygenic Risk. Transl. Psychiatry 2021, 11, 592. [Google Scholar] [CrossRef] [PubMed]
  184. Alamian, G.; Hincapié, A.-S.; Pascarella, A.; Thiery, T.; Combrisson, E.; Saive, A.-L.; Martel, V.; Althukov, D.; Haesebaert, F.; Jerbi, K. Measuring Alterations in Oscillatory Brain Networks in Schizophrenia with Resting-State MEG: State-of-the-Art and Methodological Challenges. Clin. Neurophysiol. 2017, 128, 1719–1736. [Google Scholar] [CrossRef] [PubMed]
  185. Northoff, G.; Duncan, N.W. How Do Abnormalities in the Brain’s Spontaneous Activity Translate into Symptoms in Schizophrenia? From an Overview of Resting State Activity Findings to a Proposed Spatiotemporal Psychopathology. Prog. Neurobiol. 2016, 145–146, 26–45. [Google Scholar] [CrossRef] [PubMed]
  186. Sun, L.; Castellanos, N.; Grützner, C.; Koethe, D.; Rivolta, D.; Wibral, M.; Kranaster, L.; Singer, W.; Leweke, M.F.; Uhlhaas, P.J. Evidence for Dysregulated High-Frequency Oscillations during Sensory Processing in Medication-Naïve, First Episode Schizophrenia. Schizophr. Res. 2013, 150, 519–525. [Google Scholar] [CrossRef]
  187. Garrity, A.G.; Pearlson, G.D.; McKiernan, K.; Lloyd, D.; Kiehl, K.A.; Calhoun, V.D. Aberrant “Default Mode” Functional Connectivity in Schizophrenia. Am. J. Psychiatry 2007, 164, 450–457. [Google Scholar] [CrossRef] [PubMed]
  188. Kim, J.S.; Shin, K.S.; Jung, W.H.; Kim, S.N.; Kwon, J.S.; Chung, C.K. Power Spectral Aspects of the Default Mode Network in Schizophrenia: An MEG Study. BMC Neurosci. 2014, 15, 104. [Google Scholar] [CrossRef] [PubMed]
  189. Levit-Binnun, N.; Litvak, V.; Pratt, H.; Moses, E.; Zaroor, M.; Peled, A. Differences in TMS-Evoked Responses between Schizophrenia Patients and Healthy Controls Can Be Observed without a Dedicated EEG System. Clin. Neurophysiol. 2010, 121, 332–339. [Google Scholar] [CrossRef] [PubMed]
  190. Ferrarelli, F.; Massimini, M.; Peterson, M.J.; Riedner, B.A.; Lazar, M.; Murphy, M.J.; Huber, R.; Rosanova, M.; Alexander, A.L.; Kalin, N.; et al. Reduced Evoked Gamma Oscillations in the Frontal Cortex in Schizophrenia Patients: A TMS/EEG Study. Am. J. Psychiatry 2008, 165, 996–1005. [Google Scholar] [CrossRef] [PubMed]
  191. Hoy, K.E.; Coyle, H.; Gainsford, K.; Hill, A.T.; Bailey, N.W.; Fitzgerald, P.B. Investigating Neurophysiological Markers of Impaired Cognition in Schizophrenia. Schizophr. Res. 2021, 233, 34–43. [Google Scholar] [CrossRef] [PubMed]
  192. Palmisano, A.; Chiarantoni, G.; Bossi, F.; Conti, A.; D’Elia, V.; Tagliente, S.; Nitsche, M.A.; Rivolta, D. Face Pareidolia Is Enhanced by 40 Hz Transcranial Alternating Current Stimulation (tACS) of the Face Perception Network. Sci. Rep. 2023, 13, 2035. [Google Scholar] [CrossRef] [PubMed]
  193. Strüber, D.; Herrmann, C.S. Modulation of Gamma Oscillations as a Possible Therapeutic Tool for Neuropsychiatric Diseases: A Review and Perspective. Int. J. Psychophysiol. 2020, 152, 15–25. [Google Scholar] [CrossRef]
  194. Helfrich, R.F.; Schneider, T.R.; Rach, S.; Trautmann-Lengsfeld, S.A.; Engel, A.K.; Herrmann, C.S. Entrainment of brain oscillations by transcranial alternating current stimulation. Curr. Biol. 2014, 24, 333–339. [Google Scholar] [CrossRef]
  195. Haller, N.; Hasan, A.; Padberg, F.; Brunelin, J.; da Costa Lane Valiengo, L.; Palm, U. Gamma Transcranial Alternating Current Stimulation in Patients with Negative Symptoms in Schizophrenia: A Case Series. Neurophysiol. Clin. 2020, 50, 301–304. [Google Scholar] [CrossRef] [PubMed]
  196. Beasley, C.L.; Zhang, Z.J.; Patten, I.; Reynolds, G.P. Selective Deficits in Prefrontal Cortical GABAergic Neurons in Schizophrenia Defined by the Presence of Calcium-Binding Proteins. Biol. Psychiatry 2002, 52, 708–715. [Google Scholar] [CrossRef] [PubMed]
  197. Basar-Eroglu, C.; Brand, A.; Hildebrandt, H.; Karolina Kedzior, K.; Mathes, B.; Schmiedt, C. Working Memory Related Gamma Oscillations in Schizophrenia Patients. Int. J. Psychophysiol. Off. J. Int. Organ. Psychophysiol. 2007, 64, 39–45. [Google Scholar] [CrossRef]
  198. Light, G.A.; Hsu, J.L.; Hsieh, M.H.; Meyer-Gomes, K.; Sprock, J.; Swerdlow, N.R.; Braff, D.L. Gamma Band Oscillations Reveal Neural Network Cortical Coherence Dysfunction in Schizophrenia Patients. Biol. Psychiatry 2006, 60, 1231–1240. [Google Scholar] [CrossRef] [PubMed]
  199. Yeragani, V.K.; Cashmere, D.; Miewald, J.; Tancer, M.; Keshavan, M.S. Decreased Coherence in Higher Frequency Ranges (Beta and Gamma) between Central and Frontal EEG in Patients with Schizophrenia: A Preliminary Report. Psychiatry Res. 2006, 141, 53–60. [Google Scholar] [CrossRef] [PubMed]
  200. Radhu, N.; Dominguez, L.G.; Greenwood, T.A.; Farzan, F.; Semeralul, M.O.; Richter, M.A.; Kennedy, J.L.; Blumberger, D.M.; Chen, R.; Fitzgerald, P.B.; et al. Investigating Cortical Inhibition in First-Degree Relatives and Probands in Schizophrenia. Sci. Rep. 2017, 7, 43629. [Google Scholar] [CrossRef] [PubMed]
  201. Alonso, J.; Petukhova, M.; Vilagut, G.; Chatterji, S.; Heeringa, S.; Üstün, T.B.; Alhamzawi, A.O.; Viana, M.C.; Angermeyer, M.; Bromet, E.; et al. Days out of Role Due to Common Physical and Mental Conditions: Results from the WHO World Mental Health Surveys. Mol. Psychiatry 2011, 16, 1234–1246. [Google Scholar] [CrossRef] [PubMed]
  202. Solé, E.; Garriga, M.; Valentí, M.; Vieta, E. Mixed Features in Bipolar Disorder. CNS Spectr. 2017, 22, 134–140. [Google Scholar] [CrossRef] [PubMed]
  203. Phillips, M.L.; Kupfer, D.J. Bipolar Disorder Diagnosis: Challenges and Future Directions. Lancet 2013, 381, 1663–1671. [Google Scholar] [CrossRef] [PubMed]
  204. Grande, I.; Berk, M.; Birmaher, B.; Vieta, E. Bipolar Disorder. Lancet 2016, 387, 1561–1572. [Google Scholar] [CrossRef] [PubMed]
  205. Sigitova, E.; Fišar, Z.; Hroudová, J.; Cikánková, T.; Raboch, J. Biological Hypotheses and Biomarkers of Bipolar Disorder. Psychiatry Clin. Neurosci. 2017, 71, 77–103. [Google Scholar] [CrossRef] [PubMed]
  206. Lima, D.D.; Cyrino, L.A.R.; Ferreira, G.K.; Magro, D.D.D.; Calegari, C.R.; Cabral, H.; Cavichioli, N.; Ramos, S.A.; Ullmann, O.M.; Mayer, Y.; et al. Neuroinflammation and Neuroprogression Produced by Oxidative Stress in Euthymic Bipolar Patients with Different Onset Disease Times. Sci. Rep. 2022, 12, 16742. [Google Scholar] [CrossRef] [PubMed]
  207. Canetta, S.; Bolkan, S.; Padilla-Coreano, N.; Song, L.J.; Sahn, R.; Harrison, N.L.; Gordon, J.A.; Brown, A.; Kellendonk, C. Maternal Immune Activation Leads to Selective Functional Deficits in Offspring Parvalbumin Interneurons. Mol. Psychiatry 2016, 21, 956–968. [Google Scholar] [CrossRef] [PubMed]
  208. Weber-Stadlbauer, U.; Richetto, J.; Labouesse, M.A.; Bohacek, J.; Mansuy, I.M.; Meyer, U. Transgenerational Transmission and Modification of Pathological Traits Induced by Prenatal Immune Activation. Mol. Psychiatry 2017, 22, 102–112. [Google Scholar] [CrossRef] [PubMed]
  209. Kulak, A.; Steullet, P.; Cabungcal, J.-H.; Werge, T.; Ingason, A.; Cuenod, M.; Do, K.Q. Redox Dysregulation in the Pathophysiology of Schizophrenia and Bipolar Disorder: Insights from Animal Models. Antioxid. Redox Signal. 2013, 18, 1428–1443. [Google Scholar] [CrossRef] [PubMed]
  210. Valvassori, S.S.; Aguiar-Geraldo, J.M.; Possamai-Della, T.; da-Rosa, D.D.; Peper-Nascimento, J.; Cararo, J.H.; Quevedo, J. Depressive-like Behavior Accompanies Neuroinflammation in an Animal Model of Bipolar Disorder Symptoms Induced by Ouabain. Pharmacol. Biochem. Behav. 2022, 219, 173434. [Google Scholar] [CrossRef] [PubMed]
  211. Konradi, C.; Eaton, M.; MacDonald, M.L.; Walsh, J.; Benes, F.M.; Heckers, S. Molecular Evidence for Mitochondrial Dysfunction in Bipolar Disorder. Arch. Gen. Psychiatry 2004, 61, 300–308. [Google Scholar] [CrossRef]
  212. Sun, X.; Wang, J.-F.; Tseng, M.; Young, L.T. Downregulation in Components of the Mitochondrial Electron Transport Chain in the Postmortem Frontal Cortex of Subjects with Bipolar Disorder. J. Psychiatry Neurosci. 2006, 31, 189–196. [Google Scholar] [PubMed]
  213. Wang, J.-F.; Shao, L.; Sun, X.; Young, L.T. Increased Oxidative Stress in the Anterior Cingulate Cortex of Subjects with Bipolar Disorder and Schizophrenia. Bipolar Disord. 2009, 11, 523–529. [Google Scholar] [CrossRef] [PubMed]
  214. Scaini, G.; Rezin, G.T.; Carvalho, A.F.; Streck, E.L.; Berk, M.; Quevedo, J. Mitochondrial Dysfunction in Bipolar Disorder: Evidence, Pathophysiology and Translational Implications. Neurosci. Biobehav. Rev. 2016, 68, 694–713. [Google Scholar] [CrossRef] [PubMed]
  215. Barbosa, I.G.; Huguet, R.B.; Mendonça, V.A.; Sousa, L.P.; Neves, F.S.; Bauer, M.E.; Teixeira, A.L. Increased Plasma Levels of Soluble TNF Receptor I in Patients with Bipolar Disorder. Eur. Arch. Psychiatry Clin. Neurosci. 2011, 261, 139–143. [Google Scholar] [CrossRef]
  216. Knable, M.B.; Barci, B.M.; Webster, M.J.; Meador-Woodruff, J.; Torrey, E.F. Molecular Abnormalities of the Hippocampus in Severe Psychiatric Illness: Postmortem Findings from the Stanley Neuropathology Consortium. Mol. Psychiatry 2004, 9, 609–620. [Google Scholar] [CrossRef] [PubMed]
  217. Berretta, S.; Pantazopoulos, H.; Lange, N. Neuron Numbers and Volume of the Amygdala in Subjects Diagnosed with Bipolar Disorder or Schizophrenia. Biol. Psychiatry 2007, 62, 884–893. [Google Scholar] [CrossRef] [PubMed]
  218. Steullet, P.; Cabungcal, J.-H.; Bukhari, S.A.; Ardelt, M.I.; Pantazopoulos, H.; Hamati, F.; Salt, T.E.; Cuenod, M.; Do, K.Q.; Berretta, S. The Thalamic Reticular Nucleus in Schizophrenia and Bipolar Disorder: Role of Parvalbumin-Expressing Neuron Networks and Oxidative Stress. Mol. Psychiatry 2018, 23, 2057–2065. [Google Scholar] [CrossRef] [PubMed]
  219. Alcaide, J.; Guirado, R.; Crespo, C.; Blasco-Ibáñez, J.M.; Varea, E.; Sanjuan, J.; Nacher, J. Alterations of Perineuronal Nets in the Dorsolateral Prefrontal Cortex of Neuropsychiatric Patients. Int. J. Bipolar Disord. 2019, 7, 24. [Google Scholar] [CrossRef] [PubMed]
  220. Bernard, C.; Prochiantz, A. Otx2-PNN Interaction to Regulate Cortical Plasticity. Neural Plast. 2016, 2016, e7931693. [Google Scholar] [CrossRef] [PubMed]
  221. Beurdeley, M.; Spatazza, J.; Lee, H.H.C.; Sugiyama, S.; Bernard, C.; Di Nardo, A.A.; Hensch, T.K.; Prochiantz, A. Otx2 Binding to Perineuronal Nets Persistently Regulates Plasticity in the Mature Visual Cortex. J. Neurosci. 2012, 32, 9429. [Google Scholar] [CrossRef]
  222. Sabunciyan, S.; Yolken, R.; Ragan, C.M.; Potash, J.B.; Nimgaonkar, V.L.; Dickerson, F.; Llenos, I.C.; Weis, S. Polymorphisms in the Homeobox Gene OTX2 May Be a Risk Factor for Bipolar Disorder. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2007, 144B, 1083–1086. [Google Scholar] [CrossRef] [PubMed]
  223. Bernard, C.; Vincent, C.; Testa, D.; Bertini, E.; Ribot, J.; Nardo, A.A.D.; Volovitch, M.; Prochiantz, A. A Mouse Model for Conditional Secretion of Specific Single-Chain Antibodies Provides Genetic Evidence for Regulation of Cortical Plasticity by a Non-Cell Autonomous Homeoprotein Transcription Factor. PLoS Genet. 2016, 12, e1006035. [Google Scholar] [CrossRef] [PubMed]
  224. Morishita, H.; Cabungcal, J.-H.; Chen, Y.; Do, K.Q.; Hensch, T.K. Prolonged Period of Cortical Plasticity upon Redox Dysregulation in Fast-Spiking Interneurons. Biol. Psychiatry 2015, 78, 396–402. [Google Scholar] [CrossRef] [PubMed]
  225. Cichon, S.; Mühleisen, T.W.; Degenhardt, F.A.; Mattheisen, M.; Miró, X.; Strohmaier, J.; Steffens, M.; Meesters, C.; Herms, S.; Weingarten, M.; et al. Genome-Wide Association Study Identifies Genetic Variation in Neurocan as a Susceptibility Factor for Bipolar Disorder. Am. J. Hum. Genet. 2011, 88, 372–381. [Google Scholar] [CrossRef] [PubMed]
  226. Ripke, S.; Neale, B.M.; Corvin, A.; Walters, J.T.; Farh, K.-H.; Holmans, P.A.; Lee, P.; Bulik-Sullivan, B.; Collier, D.A.; Huang, H.; et al. Biological Insights From 108 Schizophrenia-Associated Genetic Loci. Nature 2014, 511, 421–427. [Google Scholar] [CrossRef] [PubMed]
  227. Miró, X.; Meier, S.; Dreisow, M.L.; Frank, J.; Strohmaier, J.; Breuer, R.; Schmäl, C.; Albayram, Ö.; Pardo-Olmedilla, M.T.; Mühleisen, T.W.; et al. Studies in Humans and Mice Implicate Neurocan in the Etiology of Mania. Am. J. Psychiatry 2012, 169, 982–990. [Google Scholar] [CrossRef] [PubMed]
  228. Schmidt, S.; Arendt, T.; Morawski, M.; Sonntag, M. Neurocan Contributes to Perineuronal Net Development. Neuroscience 2020, 442, 69–86. [Google Scholar] [CrossRef] [PubMed]
  229. Lopez, A.Y.; Wang, X.; Xu, M.; Maheshwari, A.; Curry, D.; Lam, S.; Adesina, A.M.; Noebels, J.L.; Sun, Q.-Q.; Cooper, E.C. Ankyrin-G Isoform Imbalance and Interneuronopathy Link Epilepsy and Bipolar Disorder. Mol. Psychiatry 2017, 22, 1464–1472. [Google Scholar] [CrossRef] [PubMed]
  230. Fatemi, S.H.; Stary, J.M.; Earle, J.A.; Araghi-Niknam, M.; Eagan, E. GABAergic Dysfunction in Schizophrenia and Mood Disorders as Reflected by Decreased Levels of Glutamic Acid Decarboxylase 65 and 67 kDa and Reelin Proteins in Cerebellum. Schizophr. Res. 2005, 72, 109–122. [Google Scholar] [CrossRef]
  231. Heckers, S.; Stone, D.; Walsh, J.; Shick, J.; Koul, P.; Benes, F.M. Differential Hippocampal Expression of Glutamic Acid Decarboxylase 65 and 67 Messenger RNA in Bipolar Disorder and Schizophrenia. Arch. Gen. Psychiatry 2002, 59, 521–529. [Google Scholar] [CrossRef] [PubMed]
  232. Fatemi, S.H.; Folsom, T.D.; Thuras, P.D. Deficits in GABA(B) Receptor System in Schizophrenia and Mood Disorders: A Postmortem Study. Schizophr. Res. 2011, 128, 37–43. [Google Scholar] [CrossRef] [PubMed]
  233. Booker, S.A.; Gross, A.; Althof, D.; Shigemoto, R.; Bettler, B.; Frotscher, M.; Hearing, M.; Wickman, K.; Watanabe, M.; Kulik, Á.; et al. Differential GABAB-Receptor-Mediated Effects in Perisomatic- and Dendrite-Targeting Parvalbumin Interneurons. J. Neurosci. 2013, 33, 7961–7974. [Google Scholar] [CrossRef] [PubMed]
  234. Ishikawa, M.; Mizukami, K.; Iwakiri, M.; Asada, T. Immunohistochemical and Immunoblot Analysis of γ-Aminobutyric Acid B Receptor in the Prefrontal Cortex of Subjects with Schizophrenia and Bipolar Disorder. Neurosci. Lett. 2005, 383, 272–277. [Google Scholar] [CrossRef]
  235. Bhagwagar, Z.; Wylezinska, M.; Jezzard, P.; Evans, J.; Ashworth, F.; Sule, A.; Matthews, P.M.; Cowen, P.J. Reduction in Occipital Cortex γ-Aminobutyric Acid Concentrations in Medication-Free Recovered Unipolar Depressed and Bipolar Subjects. Biol. Psychiatry 2007, 61, 806–812. [Google Scholar] [CrossRef]
  236. Berrettini, W.H.; Nurnberger, J.I.; Hare, T.A.; Simmons-Alling, S.; Gershon, E.S.; Post, R.M. Reduced Plasma and CSF !G-Aminobutyric Acid in Affective Illness: Effect of Lithium Carbonate. Biol. Psychiatry 1983, 18, 185–194. [Google Scholar] [PubMed]
  237. Kaiya, H.; Namba, M.; Yoshida, H.; Nakamura, S. Plasma Glutamate Decarboxylase Activity in Neuropsychiatry. Psychiatry Res. 1982, 6, 335–343. [Google Scholar] [CrossRef] [PubMed]
  238. Kaufman, R.E.; Ostacher, M.J.; Marks, E.H.; Simon, N.M.; Sachs, G.S.; Jensen, J.E.; Renshaw, P.F.; Pollack, M.H. Brain GABA Levels in Patients with Bipolar Disorder. Prog. Neuropsychopharmacol. Biol. Psychiatry 2009, 33, 427–434. [Google Scholar] [CrossRef] [PubMed]
  239. Benes, F.M.; Berretta, S. GABAergic Interneurons: Implications for Understanding Schizophrenia and Bipolar Disorder. Neuropsychopharmacology 2001, 25, 1–27. [Google Scholar] [CrossRef] [PubMed]
  240. Sakai, T.; Oshima, A.; Nozaki, Y.; Ida, I.; Haga, C.; Akiyama, H.; Nakazato, Y.; Mikuni, M. Changes in Density of Calcium-Binding-Protein-Immunoreactive GABAergic Neurons in Prefrontal Cortex in Schizophrenia and Bipolar Disorder. Neuropathol. Off. J. Jpn. Soc. Neuropathol. 2008, 28, 143–150. [Google Scholar] [CrossRef]
  241. Wang, A.Y.; Lohmann, K.M.; Yang, C.K.; Zimmerman, E.I.; Pantazopoulos, H.; Herring, N.; Berretta, S.; Heckers, S.; Konradi, C. Bipolar Disorder Type 1 and Schizophrenia Are Accompanied by Decreased Density of Parvalbumin- and Somatostatin-Positive Interneurons in the Parahippocampal Region. Acta Neuropathol. 2011, 122, 615–626. [Google Scholar] [CrossRef] [PubMed]
  242. Cotter, D.; Landau, S.; Beasley, C.; Stevenson, R.; Chana, G.; MacMillan, L.; Everall, I. The Density and Spatial Distribution of GABAergic Neurons, Labelled Using Calcium Binding Proteins, in the Anterior Cingulate Cortex in Major Depressive Disorder, Bipolar Disorder, and Schizophrenia. Biol. Psychiatry 2002, 51, 377–386. [Google Scholar] [CrossRef] [PubMed]
  243. Sibille, E.; Morris, H.M.; Kota, R.S.; Lewis, D.A. GABA-Related Transcripts in the Dorsolateral Prefrontal Cortex in Mood Disorders. Int. J. Neuropsychopharmacol. 2011, 14, 721–734. [Google Scholar] [CrossRef]
  244. Pantazopoulos, H.; Lange, N.; Baldessarini, R.J.; Berretta, S. Parvalbumin Neurons in the Entorhinal Cortex of Subjects Diagnosed With Bipolar Disorder or Schizophrenia. Biol. Psychiatry 2007, 61, 640–652. [Google Scholar] [CrossRef] [PubMed]
  245. Reynolds, G.P.; Abdul-Monim, Z.; Neill, J.C.; Zhang, Z.-J. Calcium Binding Protein Markers of GABA Deficits in Schizophrenia--Postmortem Studies and Animal Models. Neurotox. Res. 2004, 6, 57–61. [Google Scholar] [CrossRef] [PubMed]
  246. Chung, D.W.; Chung, Y.; Bazmi, H.H.; Lewis, D.A. Altered ErbB4 Splicing and Cortical Parvalbumin Interneuron Dysfunction in Schizophrenia and Mood Disorders. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2018, 43, 2478–2486. [Google Scholar] [CrossRef]
  247. Parboosing, R.; Bao, Y.; Shen, L.; Schaefer, C.A.; Brown, A.S. Gestational Influenza and Bipolar Disorder in Adult Offspring. JAMA Psychiatry 2013, 70, 677–685. [Google Scholar] [CrossRef] [PubMed]
  248. Brown, A.S.; Derkits, E.J. Prenatal Infection and Schizophrenia: A Review of Epidemiologic and Translational Studies. Am. J. Psychiatry 2010, 167, 261–280. [Google Scholar] [CrossRef] [PubMed]
  249. Grandjean, E.M.; Aubry, J.-M. Lithium: Updated Human Knowledge Using an Evidence-Based Approach. CNS Drugs 2009, 23, 397–418. [Google Scholar] [CrossRef] [PubMed]
  250. Dzirasa, K.; Coque, L.; Sidor, M.M.; Kumar, S.; Dancy, E.A.; Takahashi, J.S.; McClung, C.A.; Nicolelis, M.A.L. Lithium Ameliorates Nucleus Accumbens Phase-Signaling Dysfunction in a Genetic Mouse Model of Mania. J. Neurosci. Off. J. Soc. Neurosci. 2010, 30, 16314–16323. [Google Scholar] [CrossRef] [PubMed]
  251. Kim, D.-J.; Bolbecker, A.R.; Howell, J.; Rass, O.; Sporns, O.; Hetrick, W.P.; Breier, A.; O’Donnell, B.F. Disturbed Resting State EEG Synchronization in Bipolar Disorder: A Graph-Theoretic Analysis. NeuroImage Clin. 2013, 2, 414–423. [Google Scholar] [CrossRef]
  252. Özerdem, A.; Güntekin, B.; Saatçi, E.; Tunca, Z.; Başar, E. Disturbance in Long Distance Gamma Coherence in Bipolar Disorder. Prog. Neuropsychopharmacol. Biol. Psychiatry 2010, 34, 861–865. [Google Scholar] [CrossRef] [PubMed]
  253. Özerdem, A.; Güntekin, B.; Atagün, İ.; Turp, B.; Başar, E. Reduced Long Distance Gamma (28–48Hz) Coherence in Euthymic Patients with Bipolar Disorder. J. Affect. Disord. 2011, 132, 325–332. [Google Scholar] [CrossRef] [PubMed]
  254. Hall, M.-H.; Spencer, K.M.; Schulze, K.; McDonald, C.; Kalidindi, S.; Kravariti, E.; Kane, F.; Murray, R.M.; Bramon, E.; Sham, P.; et al. The Genetic and Environmental Influences of Event-Related Gamma Oscillations on Bipolar Disorder. Bipolar Disord. 2011, 13, 260–271. [Google Scholar] [CrossRef] [PubMed]
  255. Maharajh, K.; Abrams, D.; Rojas, D.C.; Teale, P.; Reite, M.L. Auditory Steady State and Transient Gamma Band Activity in Bipolar Disorder. Int. Congr. Ser. 2007, 1300, 707–710. [Google Scholar] [CrossRef]
  256. Oda, Y.; Onitsuka, T.; Tsuchimoto, R.; Hirano, S.; Oribe, N.; Ueno, T.; Hirano, Y.; Nakamura, I.; Miura, T.; Kanba, S. Gamma Band Neural Synchronization Deficits for Auditory Steady State Responses in Bipolar Disorder Patients. PLoS ONE 2012, 7, e39955. [Google Scholar] [CrossRef] [PubMed]
  257. Rass, O.; Krishnan, G.; Brenner, C.A.; Hetrick, W.P.; Merrill, C.C.; Shekhar, A.; O’Donnell, B.F. Auditory Steady State Response in Bipolar Disorder: Relation to Clinical State, Cognitive Performance, Medication Status, and Substance Disorders. Bipolar Disord. 2010, 12, 793–803. [Google Scholar] [CrossRef] [PubMed]
  258. Brealy, J.A.; Shaw, A.; Richardson, H.; Singh, K.D.; Muthukumaraswamy, S.D.; Keedwell, P.A. Increased Visual Gamma Power in Schizoaffective Bipolar Disorder. Psychol. Med. 2015, 45, 783–794. [Google Scholar] [CrossRef] [PubMed]
  259. Onitsuka, T.; Oribe, N.; Kanba, S. Neurophysiological Findings in Patients with Bipolar Disorder. Suppl. Clin. Neurophysiol. 2013, 62, 197–206. [Google Scholar] [CrossRef] [PubMed]
  260. Levinson, A.J.; Young, L.T.; Fitzgerald, P.B.; Daskalakis, Z.J. Cortical Inhibitory Dysfunction in Bipolar Disorder: A Study Using Transcranial Magnetic Stimulation. J. Clin. Psychopharmacol. 2007, 27, 493–497. [Google Scholar] [CrossRef] [PubMed]
  261. Radhu, N.; Ravindran, L.N.; Levinson, A.J.; Daskalakis, Z.J. Inhibition of the Cortex Using Transcranial Magnetic Stimulation in Psychiatric Populations: Current and Future Directions. J. Psychiatry Neurosci. JPN 2012, 37, 369–378. [Google Scholar] [CrossRef] [PubMed]
  262. Canali, P.; Sarasso, S.; Rosanova, M.; Casarotto, S.; Sferrazza-Papa, G.; Gosseries, O.; Fecchio, M.; Massimini, M.; Mariotti, M.; Cavallaro, R.; et al. Shared Reduction of Oscillatory Natural Frequencies in Bipolar Disorder, Major Depressive Disorder and Schizophrenia. J. Affect. Disord. 2015, 184, 111–115. [Google Scholar] [CrossRef] [PubMed]
  263. Canali, P.; Casarotto, S.; Rosanova, M.; Sferrazza-Papa, G.; Casali, A.G.; Gosseries, O.; Massimini, M.; Smeraldi, E.; Colombo, C.; Benedetti, F. Abnormal Brain Oscillations Persist after Recovery from Bipolar Depression. Eur. Psychiatry 2017, 41, 10–15. [Google Scholar] [CrossRef]
  264. Olejarczyk, E.; Zuchowicz, U.; Wozniak-Kwasniewska, A.; Kaminski, M.; Szekely, D.; David, O. The Impact of Repetitive Transcranial Magnetic Stimulation on Functional Connectivity in Major Depressive Disorder and Bipolar Disorder Evaluated by Directed Transfer Function and Indices Based on Graph Theory. Int. J. Neural Syst. 2020, 30, 2050015. [Google Scholar] [CrossRef] [PubMed]
  265. Hardeveld, F.; Spijker, J.; De Graaf, R.; Nolen, W.A.; Beekman, A.T.F. Prevalence and Predictors of Recurrence of Major Depressive Disorder in the Adult Population. Acta Psychiatr. Scand. 2010, 122, 184–191. [Google Scholar] [CrossRef]
  266. Fekadu, N.; Shibeshi, W.; Engidawork, E. Major Depressive Disorder: Pathophysiology and Clinical Management. J. Depress Anxiety 2016, 6, 1000255. [Google Scholar] [CrossRef]
  267. Pitsillou, E.; Bresnehan, S.M.; Kagarakis, E.A.; Wijoyo, S.J.; Liang, J.; Hung, A.; Karagiannis, T.C. The Cellular and Molecular Basis of Major Depressive Disorder: Towards a Unified Model for Understanding Clinical Depression. Mol. Biol. Rep. 2020, 47, 753–770. [Google Scholar] [CrossRef]
  268. Avramopoulos, D.; Pearce, B.D.; McGrath, J.; Wolyniec, P.; Wang, R.; Eckart, N.; Hatzimanolis, A.; Goes, F.S.; Nestadt, G.; Mulle, J.; et al. Infection and Inflammation in Schizophrenia and Bipolar Disorder: A Genome Wide Study for Interactions with Genetic Variation. PLoS ONE 2015, 10, e0116696. [Google Scholar] [CrossRef] [PubMed]
  269. Li, B.; Yang, W.; Ge, T.; Wang, Y.; Cui, R. Stress Induced Microglial Activation Contributes to Depression. Pharmacol. Res. 2022, 179, 106145. [Google Scholar] [CrossRef] [PubMed]
  270. Li, H.; Sheng, Z.; Khan, S.; Zhang, R.; Liu, Y.; Zhang, Y.; Yong, V.W.; Xue, M. Matrix Metalloproteinase-9 as an Important Contributor to the Pathophysiology of Depression. Front. Neurol. 2022, 13, 861843. [Google Scholar] [CrossRef] [PubMed]
  271. Zhang, B.; Yang, X.; Ye, L.; Liu, R.; Ye, B.; Du, W.; Shen, F.; Li, Q.; Guo, F.; Liu, J.; et al. Ketamine Activated Glutamatergic Neurotransmission by GABAergic Disinhibition in the Medial Prefrontal Cortex. Neuropharmacology 2021, 194, 108382. [Google Scholar] [CrossRef] [PubMed]
  272. Kappelmann, N.; Lewis, G.; Dantzer, R.; Jones, P.B.; Khandaker, G.M. Antidepressant Activity of Anti-Cytokine Treatment: A Systematic Review and Meta-Analysis of Clinical Trials of Chronic Inflammatory Conditions. Mol. Psychiatry 2018, 23, 335–343. [Google Scholar] [CrossRef] [PubMed]
  273. Schiepers, O.J.G.; Wichers, M.C.; Maes, M. Cytokines and Major Depression. Prog. Neuropsychopharmacol. Biol. Psychiatry 2005, 29, 201–217. [Google Scholar] [CrossRef] [PubMed]
  274. Albeely, A.M.; Ryan, S.D.; Perreault, M.L. Pathogenic Feed-Forward Mechanisms in Alzheimer’s and Parkinson’s Disease Converge on GSK-3. Brain Plast. 2018, 4, 151–167. [Google Scholar] [CrossRef] [PubMed]
  275. Manduca, J.D.; Thériault, R.-K.; Perreault, M.L. Glycogen Synthase Kinase-3: The Missing Link to Aberrant Circuit Function in Disorders of Cognitive Dysfunction? Pharmacol. Res. 2020, 157, 104819. [Google Scholar] [CrossRef] [PubMed]
  276. Heppner, F.L.; Ransohoff, R.M.; Becher, B. Immune Attack: The Role of Inflammation in Alzheimer Disease. Nat. Rev. Neurosci. 2015, 16, 358–372. [Google Scholar] [CrossRef] [PubMed]
  277. McCallum, R.T.; Perreault, M.L. Glycogen Synthase Kinase-3: A Focal Point for Advancing Pathogenic Inflammation in Depression. Cells 2021, 10, 2270. [Google Scholar] [CrossRef] [PubMed]
  278. Nguyen, T.; Fan, T.; George, S.R.; Perreault, M.L. Disparate Effects of Lithium and a GSK-3 Inhibitor on Neuronal Oscillatory Activity in Prefrontal Cortex and Hippocampus. Front. Aging Neurosci. 2018, 9, 434. [Google Scholar] [CrossRef] [PubMed]
  279. Nakao, K.; Singh, M.; Sapkota, K.; Hagler, B.C.; Hunter, R.N.; Raman, C.; Hablitz, J.J.; Nakazawa, K. GSK3β Inhibition Restores Cortical Gamma Oscillation and Cognitive Behavior in a Mouse Model of NMDA Receptor Hypofunction Relevant to Schizophrenia. Neuropsychopharmacology 2020, 45, 2207–2218. [Google Scholar] [CrossRef] [PubMed]
  280. Chen, S.; Chen, F.; Amin, N.; Ren, Q.; Ye, S.; Hu, Z.; Tan, X.; Jiang, M.; Fang, M. Defects of Parvalbumin-Positive Interneurons in the Ventral Dentate Gyrus Region Are Implicated Depression-like Behavior in Mice. Brain Behav. Immun. 2022, 99, 27–42. [Google Scholar] [CrossRef] [PubMed]
  281. Peng, Y.-L.; Liu, Y.-N.; Liu, L.; Wang, X.; Jiang, C.-L.; Wang, Y.-X. Inducible Nitric Oxide Synthase Is Involved in the Modulation of Depressive Behaviors Induced by Unpredictable Chronic Mild Stress. J. Neuroinflamm. 2012, 9, 75. [Google Scholar] [CrossRef] [PubMed]
  282. Zhou, X.; Zhang, F.; Ying, C.; Chen, J.; Chen, L.; Dong, J.; Shi, Y.; Tang, M.; Hu, X.; Pan, Z.; et al. Inhibition of iNOS Alleviates Cognitive Deficits and Depression in Diabetic Mice through Downregulating the NO/sGC/cGMP/PKG Signal Pathway. Behav. Brain Res. 2017, 322, 70–82. [Google Scholar] [CrossRef] [PubMed]
  283. Blanco, I.; Conant, K. Extracellular Matrix Remodeling with Stress and Depression: Studies in Human, Rodent and Zebrafish Models. Eur. J. Neurosci. 2021, 53, 3879–3888. [Google Scholar] [CrossRef] [PubMed]
  284. Yu, Z.; Chen, N.; Hu, D.; Chen, W.; Yuan, Y.; Meng, S.; Zhang, W.; Lu, L.; Han, Y.; Shi, J. Decreased Density of Perineuronal Net in Prelimbic Cortex Is Linked to Depressive-Like Behavior in Young-Aged Rats. Front. Mol. Neurosci. 2020, 13, 4. [Google Scholar] [CrossRef] [PubMed]
  285. Alaiyed, S.; Conant, K. A Role for Matrix Metalloproteases in Antidepressant Efficacy. Front. Mol. Neurosci. 2019, 12, 117. [Google Scholar] [CrossRef]
  286. Alaiyed, S.; McCann, M.; Mahajan, G.; Rajkowska, G.; Stockmeier, C.A.; Kellar, K.J.; Wu, J.Y.; Conant, K. Venlafaxine Stimulates an MMP-9-Dependent Increase in Excitatory/Inhibitory Balance in a Stress Model of Depression. J. Neurosci. Off. J. Soc. Neurosci. 2020, 40, 4418–4431. [Google Scholar] [CrossRef]
  287. Wen, T.H.; Afroz, S.; Reinhard, S.M.; Palacios, A.R.; Tapia, K.; Binder, D.K.; Razak, K.A.; Ethell, I.M. Genetic Reduction of Matrix Metalloproteinase-9 Promotes Formation of Perineuronal Nets Around Parvalbumin-Expressing Interneurons and Normalizes Auditory Cortex Responses in Developing Fmr1 Knock-Out Mice. Cereb. Cortex 2018, 28, 3951–3964. [Google Scholar] [CrossRef]
  288. Riga, D.; Kramvis, I.; Koskinen, M.K.; van Bokhoven, P.; van der Harst, J.E.; Heistek, T.S.; Jaap Timmerman, A.; van Nierop, P.; van der Schors, R.C.; Pieneman, A.W.; et al. Hippocampal Extracellular Matrix Alterations Contribute to Cognitive Impairment Associated with a Chronic Depressive-like State in Rats. Sci. Transl. Med. 2017, 9, eaai8753. [Google Scholar] [CrossRef] [PubMed]
  289. Castillo-Gómez, E.; Pérez-Rando, M.; Bellés, M.; Gilabert-Juan, J.; Llorens, J.V.; Carceller, H.; Bueno-Fernández, C.; García-Mompó, C.; Ripoll-Martínez, B.; Curto, Y.; et al. Early Social Isolation Stress and Perinatal NMDA Receptor Antagonist Treatment Induce Changes in the Structure and Neurochemistry of Inhibitory Neurons of the Adult Amygdala and Prefrontal Cortex. eNeuro 2017, 4, ENEURO.0034-17.2017. [Google Scholar] [CrossRef]
  290. Spijker, S.; Koskinen, M.-K.; Riga, D. Incubation of Depression: ECM Assembly and Parvalbumin Interneurons after Stress. Neurosci. Biobehav. Rev. 2020, 118, 65–79. [Google Scholar] [CrossRef] [PubMed]
  291. de Araújo Costa Folha, O.A.; Bahia, C.P.; de Aguiar, G.P.S.; Herculano, A.M.; Coelho, N.L.G.; de Sousa, M.B.C.; Shiramizu, V.K.M.; de Menezes Galvão, A.C.; de Carvalho, W.A.; Pereira, A. Effect of Chronic Stress during Adolescence in Prefrontal Cortex Structure and Function. Behav. Brain Res. 2017, 326, 44–51. [Google Scholar] [CrossRef] [PubMed]
  292. Fatemi, S.H.; Kroll, J.L.; Stary, J.M. Altered Levels of Reelin and Its Isoforms in Schizophrenia and Mood Disorders. NeuroReport 2001, 12, 3209–3215. [Google Scholar] [CrossRef] [PubMed]
  293. Pantazopoulos, H.; Berretta, S. In Sickness and in Health: Perineuronal Nets and Synaptic Plasticity in Psychiatric Disorders. Neural Plast. 2016, 2016, 9847696. [Google Scholar] [CrossRef] [PubMed]
  294. Yoshida, T.; Ishikawa, M.; Niitsu, T.; Nakazato, M.; Watanabe, H.; Shiraishi, T.; Shiina, A.; Hashimoto, T.; Kanahara, N.; Hasegawa, T.; et al. Decreased Serum Levels of Mature Brain-Derived Neurotrophic Factor (BDNF), but Not Its Precursor proBDNF, in Patients with Major Depressive Disorder. PLoS ONE 2012, 7, e42676. [Google Scholar] [CrossRef] [PubMed]
  295. Todorović, N.; Mićić, B.; Schwirtlich, M.; Stevanović, M.; Filipović, D. Subregion-Specific Protective Effects of Fluoxetine and Clozapine on Parvalbumin Expression in Medial Prefrontal Cortex of Chronically Isolated Rats. Neuroscience 2019, 396, 24–35. [Google Scholar] [CrossRef] [PubMed]
  296. Banasr, M.; Lepack, A.; Fee, C.; Duric, V.; Maldonado-Aviles, J.; DiLeone, R.; Sibille, E.; Duman, R.S.; Sanacora, G. Characterization of GABAergic Marker Expression in the Chronic Unpredictable Stress Model of Depression. Chronic Stress Thousand Oaks Calif 2017, 1, 2470547017720459. [Google Scholar] [CrossRef] [PubMed]
  297. Czeh, B.; Simon, M.; van der Hart, M.G.; Schmelting, B.; Hesselink, M.B.; Fuchs, E. Chronic Stress Decreases the Number of Parvalbumin-Immunoreactive Interneurons in the Hippocampus: Prevention by Treatment with a Substance P Receptor (NK1) Antagonist. Neuropsychopharmacology 2005, 30, 67–79. [Google Scholar] [CrossRef] [PubMed]
  298. Czéh, B.; Varga, Z.K.K.; Henningsen, K.; Kovács, G.L.; Miseta, A.; Wiborg, O. Chronic Stress Reduces the Number of GABAergic Interneurons in the Adult Rat Hippocampus, Dorsal-Ventral and Region-Specific Differences. Hippocampus 2015, 25, 393–405. [Google Scholar] [CrossRef] [PubMed]
  299. Czéh, B.; Vardya, I.; Varga, Z.; Febbraro, F.; Csabai, D.; Martis, L.-S.; Højgaard, K.; Henningsen, K.; Bouzinova, E.V.; Miseta, A.; et al. Long-Term Stress Disrupts the Structural and Functional Integrity of GABAergic Neuronal Networks in the Medial Prefrontal Cortex of Rats. Front. Cell. Neurosci. 2018, 12, 148. [Google Scholar] [CrossRef]
  300. Kigawa, Y.; Hashimoto, E.; Ukai, W.; Ishii, T.; Furuse, K.; Tsujino, H.; Shirasaka, T.; Saito, T. Stem Cell Therapy: A New Approach to the Treatment of Refractory Depression. J. Neural Transm. 2014, 121, 1221–1232. [Google Scholar] [CrossRef]
  301. Zhong, H.; Rong, J.; Zhu, C.; Liang, M.; Li, Y.; Zhou, R. Epigenetic Modifications of GABAergic Interneurons Contribute to Deficits in Adult Hippocampus Neurogenesis and Depression-Like Behavior in Prenatally Stressed Mice. Int. J. Neuropsychopharmacol. 2020, 23, 274–285. [Google Scholar] [CrossRef] [PubMed]
  302. Petty, F. GABA and Mood Disorders: A Brief Review and Hypothesis. J. Affect. Disord. 1995, 34, 275–281. [Google Scholar] [CrossRef] [PubMed]
  303. Sanacora, G.; Mason, G.F.; Rothman, D.L.; Behar, K.L.; Hyder, F.; Petroff, O.A.C.; Berman, R.M.; Charney, D.S.; Krystal, J.H. Reduced Cortical γ-Aminobutyric Acid Levels in Depressed Patients Determined by Proton Magnetic Resonance Spectroscopy. Arch. Gen. Psychiatry 1999, 56, 1043–1047. [Google Scholar] [CrossRef] [PubMed]
  304. Sanacora, G.; Mason, G.F.; Krystal, J.H. Impairment of GABAergic Transmission in Depression: New Insights from Neuroimaging Studies. Crit. Rev. Neurobiol. 2000, 14, 23. [Google Scholar] [CrossRef] [PubMed]
  305. Bhagwagar, Z.; Wylezinska, M.; Taylor, M.; Jezzard, P.; Matthews, P.M.; Cowen, P.J. Increased Brain GABA Concentrations Following Acute Administration of a Selective Serotonin Reuptake Inhibitor. Am. J. Psychiatry 2004, 161, 368–370. [Google Scholar] [CrossRef] [PubMed]
  306. Sanacora, G.; Mason, G.F.; Rothman, D.L.; Hyder, F.; Ciarcia, J.J.; Ostroff, R.B.; Berman, R.M.; Krystal, J.H. Increased Cortical GABA Concentrations in Depressed Patients Receiving ECT. Am. J. Psychiatry 2003, 160, 577–579. [Google Scholar] [CrossRef] [PubMed]
  307. Rajkowska, G.; O’Dwyer, G.; Teleki, Z.; Stockmeier, C.A.; Miguel-Hidalgo, J.J. GABAergic Neurons Immunoreactive for Calcium Binding Proteins Are Reduced in the Prefrontal Cortex in Major Depression. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2007, 32, 471–482. [Google Scholar] [CrossRef] [PubMed]
  308. Smiley, J.F.; Hackett, T.A.; Bleiwas, C.; Petkova, E.; Stankov, A.; Mann, J.J.; Rosoklija, G.; Dwork, A.J. Reduced GABA Neuron Density in Auditory Cerebral Cortex of Subjects with Major Depressive Disorder. J. Chem. Neuroanat. 2016, 76, 108–121. [Google Scholar] [CrossRef]
  309. de Sousa, T.R.; Correia, D.T.; Novais, F. Exploring the Hypothesis of a Schizophrenia and Bipolar Disorder Continuum: Biological, Genetic and Pharmacologic Data. Available online: https://www.ingentaconnect.com/content/ben/cnsnddt/pre-prints/content-34477537 (accessed on 29 November 2021).
  310. Fitzgerald, P.J.; Watson, B.O. Gamma Oscillations as a Biomarker for Major Depression: An Emerging Topic. Transl. Psychiatry 2018, 8, 177. [Google Scholar] [CrossRef] [PubMed]
  311. Gazit, T.; Friedman, A.; Lax, E.; Samuel, M.; Zahut, R.; Katz, M.; Abraham, L.; Tischler, H.; Teicher, M.; Yadid, G. Programmed Deep Brain Stimulation Synchronizes VTA Gamma Band Field Potential and Alleviates Depressive-like Behavior in Rats. Neuropharmacology 2015, 91, 135–141. [Google Scholar] [CrossRef] [PubMed]
  312. Khalid, A.; Kim, B.S.; Seo, B.A.; Lee, S.-T.; Jung, K.-H.; Chu, K.; Lee, S.K.; Jeon, D. Gamma Oscillation in Functional Brain Networks Is Involved in the Spontaneous Remission of Depressive Behavior Induced by Chronic Restraint Stress in Mice. BMC Neurosci. 2016, 17, 4. [Google Scholar] [CrossRef] [PubMed]
  313. Riga, M.S.; Sanchez, C.; Celada, P.; Artigas, F. Sub-Chronic Vortioxetine (but Not Escitalopram) Normalizes Brain Rhythm Alterations and Memory Deficits Induced by Serotonin Depletion in Rats. Neuropharmacology 2020, 178, 108238. [Google Scholar] [CrossRef] [PubMed]
  314. Akar, S.A.; Kara, S.; Agambayev, S.; Bilgiç, V. Nonlinear Analysis of EEG in Major Depression with Fractal Dimensions. In Proceedings of the 2015 37th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Milan, Italy, 25–29 August 2015; pp. 7410–7413. [Google Scholar]
  315. Pizzagalli, D.A.; Peccoralo, L.A.; Davidson, R.J.; Cohen, J.D. Resting Anterior Cingulate Activity and Abnormal Responses to Errors in Subjects with Elevated Depressive Symptoms: A 128-Channel EEG Study. Hum. Brain Mapp. 2006, 27, 185–201. [Google Scholar] [CrossRef] [PubMed]
  316. Murphy, O.W.; Hoy, K.E.; Wong, D.; Bailey, N.W.; Fitzgerald, P.B.; Segrave, R.A. Individuals with Depression Display Abnormal Modulation of Neural Oscillatory Activity during Working Memory Encoding and Maintenance. Biol. Psychol. 2019, 148, 107766. [Google Scholar] [CrossRef] [PubMed]
  317. Siegle, G.J.; Condray, R.; Thase, M.E.; Keshavan, M.; Steinhauer, S.R. Sustained Gamma-Band EEG Following Negative Words in Depression and Schizophrenia. Int. J. Psychophysiol. 2010, 75, 107–118. [Google Scholar] [CrossRef] [PubMed]
  318. Lee, C.H.; Hwang, I.K.; Choi, J.H.; Yoo, K.-Y.; Han, T.H.; Park, O.K.; Lee, S.Y.; Ryu, P.D.; Won, M.-H. Calcium Binding Proteins Immunoreactivity in the Rat Basolateral Amygdala Following Myocardial Infarction. Cell. Mol. Neurobiol. 2010, 30, 333–338. [Google Scholar] [CrossRef] [PubMed]
  319. Liu, T.-Y.; Hsieh, J.-C.; Chen, Y.-S.; Tu, P.-C.; Su, T.-P.; Chen, L.-F. Different Patterns of Abnormal Gamma Oscillatory Activity in Unipolar and Bipolar Disorder Patients during an Implicit Emotion Task. Neuropsychologia 2012, 50, 1514–1520. [Google Scholar] [CrossRef]
  320. Corlier, J.; Burnette, E.; Wilson, A.C.; Lou, J.J.; Landeros, A.; Minzenberg, M.J.; Leuchter, A.F. Effect of Repetitive Transcranial Magnetic Stimulation (rTMS) Treatment of Major Depressive Disorder (MDD) on Cognitive Control. J. Affect. Disord. 2020, 265, 272–277. [Google Scholar] [CrossRef] [PubMed]
  321. Chen, F.; Gu, C.; Zhai, N.; Duan, H.; Zhai, A.; Zhang, X. Repetitive Transcranial Magnetic Stimulation Improves Amygdale Functional Connectivity in Major Depressive Disorder. Front. Psychiatry 2020, 11, 732. [Google Scholar] [CrossRef] [PubMed]
  322. Steele, J.D.; Glabus, M.F.; Shajahan, P.M.; Ebmeier, K.P. Increased Cortical Inhibition in Depression: A Prolonged Silent Period with Transcranial Magnetic Stimulation (TMS). Psychol. Med. 2000, 30, 565–570. [Google Scholar] [CrossRef] [PubMed]
  323. Bajbouj, M.; Lisanby, S.H.; Lang, U.E.; Danker-Hopfe, H.; Heuser, I.; Neu, P. Evidence for Impaired Cortical Inhibition in Patients with Unipolar Major Depression. Biol. Psychiatry 2006, 59, 395–400. [Google Scholar] [CrossRef] [PubMed]
  324. Noda, Y.; Zomorrodi, R.; Saeki, T.; Rajji, T.K.; Blumberger, D.M.; Daskalakis, Z.J.; Nakamura, M. Resting-State EEG Gamma Power and Theta-Gamma Coupling Enhancement Following High-Frequency Left Dorsolateral Prefrontal rTMS in Patients with Depression. Clin. Neurophysiol. Off. J. Int. Fed. Clin. Neurophysiol. 2017, 128, 424–432. [Google Scholar] [CrossRef] [PubMed]
  325. Mitoma, R.; Tamura, S.; Tateishi, H.; Mitsudo, T.; Tanabe, I.; Monji, A.; Hirano, Y. Oscillatory Brain Network Changes after Transcranial Magnetic Stimulation Treatment in Patients with Major Depressive Disorder: Oscillatory Changes Following rTMS in Depression. J. Affect. Disord. Rep. 2021, 7, 100277. [Google Scholar] [CrossRef]
  326. Sun, Y.; Giacobbe, P.; Tang, C.W.; Barr, M.S.; Rajji, T.; Kennedy, S.H.; Fitzgerald, P.B.; Lozano, A.M.; Wong, W.; Daskalakis, Z.J. Deep Brain Stimulation Modulates Gamma Oscillations and Theta-Gamma Coupling in Treatment Resistant Depression. Brain Stimulat. 2015, 8, 1033–1042. [Google Scholar] [CrossRef] [PubMed]
  327. Jiang, H.; Tian, S.; Bi, K.; Lu, Q.; Yao, Z. Hyperactive Frontolimbic and Frontocentral Resting-State Gamma Connectivity in Major Depressive Disorder. J. Affect. Disord. 2019, 257, 74–82. [Google Scholar] [CrossRef] [PubMed]
  328. Yener, G.G.; Başar, E. Chapter 16-Biomarkers in Alzheimer’s Disease with a Special Emphasis on Event-Related Oscillatory Responses. In Supplements to Clinical Neurophysiology; Başar, E., Başar-Eroĝlu, C., Özerdem, A., Rossini, P.M., Yener, G.G., Eds.; Application of Brain Oscillations in Neuropsychiatric Diseases; Elsevier: Amsterdam, The Netherlands, 2013; Volume 62, pp. 237–273. [Google Scholar]
  329. Uhlhaas, P.J.; Singer, W. Neuronal Dynamics and Neuropsychiatric Disorders: Toward a Translational Paradigm for Dysfunctional Large-Scale Networks. Neuron 2012, 75, 963–980. [Google Scholar] [CrossRef] [PubMed]
  330. Lisman, J. Excitation, Inhibition, Local Oscillations, or Large-Scale Loops: What Causes the Symptoms of Schizophrenia? Curr. Opin. Neurobiol. 2012, 22, 537–544. [Google Scholar] [CrossRef] [PubMed]
  331. Favuzzi, E.; Marques-Smith, A.; Deogracias, R.; Winterflood, C.M.; Sánchez-Aguilera, A.; Mantoan, L.; Maeso, P.; Fernandes, C.; Ewers, H.; Rico, B. Activity-Dependent Gating of Parvalbumin Interneuron Function by the Perineuronal Net Protein Brevican. Neuron 2017, 95, 639–655.e10. [Google Scholar] [CrossRef] [PubMed]
  332. Park, K.; Lee, J.; Jang, H.J.; Richards, B.A.; Kohl, M.M.; Kwag, J. Optogenetic Activation of Parvalbumin and Somatostatin Interneurons Selectively Restores Theta-Nested Gamma Oscillations and Oscillation-Induced Spike Timing-Dependent Long-Term Potentiation Impaired by Amyloid β Oligomers. BMC Biol. 2020, 18, 7. [Google Scholar] [CrossRef] [PubMed]
  333. Tukker, J.J.; Fuentealba, P.; Hartwich, K.; Somogyi, P.; Klausberger, T. Cell Type-Specific Tuning of Hippocampal Interneuron Firing during Gamma Oscillations in Vivo. J. Neurosci. Off. J. Soc. Neurosci. 2007, 27, 8184–8189. [Google Scholar] [CrossRef] [PubMed]
  334. Chow, A.; Erisir, A.; Farb, C.; Nadal, M.S.; Ozaita, A.; Lau, D.; Welker, E.; Rudy, B. K+ Channel Expression Distinguishes Subpopulations of Parvalbumin- and Somatostatin-Containing Neocortical Interneurons. J. Neurosci. 1999, 19, 9332. [Google Scholar] [CrossRef] [PubMed]
  335. McDonald, A.J.; Mascagni, F. Differential Expression of Kv3.1b and Kv3.2 Potassium Channel Subunits in Interneurons of the Basolateral Amygdala. Neuroscience 2006, 138, 537–547. [Google Scholar] [CrossRef] [PubMed]
  336. Rudy, B.; McBain, C.J. Kv3 Channels: Voltage-Gated K+ Channels Designed for High-Frequency Repetitive Firing. Trends Neurosci. 2001, 24, 517–526. [Google Scholar] [CrossRef]
  337. Steel, A.; Mikkelsen, M.; Edden, R.A.E.; Robertson, C.E. Regional Balance between Glutamate+glutamine and GABA+ in the Resting Human Brain. NeuroImage 2020, 220, 117112. [Google Scholar] [CrossRef] [PubMed]
  338. Ren, Z.; Pribiag, H.; Jefferson, S.J.; Shorey, M.; Fuchs, T.; Stellwagen, D.; Luscher, B. Bidirectional Homeostatic Regulation of a Depression-Related Brain State by Gamma-Aminobutyric Acidergic Deficits and Ketamine Treatment. Biol. Psychiatry 2016, 80, 457–468. [Google Scholar] [CrossRef] [PubMed]
  339. Carlén, M.; Meletis, K.; Siegle, J.H.; Cardin, J.A.; Futai, K.; Vierling-Claassen, D.; Rühlmann, C.; Jones, S.R.; Deisseroth, K.; Sheng, M.; et al. A Critical Role for NMDA Receptors in Parvalbumin Interneurons for Gamma Rhythm Induction and Behavior. Mol. Psychiatry 2012, 17, 537–548. [Google Scholar] [CrossRef]
  340. Glantz, L.A.; Lewis, D.A. Decreased Dendritic Spine Density on Prefrontal Cortical Pyramidal Neurons in Schizophrenia. Arch. Gen. Psychiatry 2000, 57, 65–73. [Google Scholar] [CrossRef]
  341. Théberge, J.; Al-Semaan, Y.; Williamson, P.C.; Menon, R.S.; Neufeld, R.W.J.; Rajakumar, N.; Schaefer, B.; Densmore, M.; Drost, D.J. Glutamate and Glutamine in the Anterior Cingulate and Thalamus of Medicated Patients with Chronic Schizophrenia and Healthy Comparison Subjects Measured with 4.0-T Proton MRS. Am. J. Psychiatry 2003, 160, 2231–2233. [Google Scholar] [CrossRef]
  342. Cohen, S.M.; Tsien, R.W.; Goff, D.C.; Halassa, M.M. The Impact of NMDA Receptor Hypofunction on GABAergic Neurons in the Pathophysiology of Schizophrenia. Schizophr. Res. 2015, 167, 98–107. [Google Scholar] [CrossRef]
  343. Belforte, J.E.; Zsiros, V.; Sklar, E.R.; Jiang, Z.; Yu, G.; Li, Y.; Quinlan, E.M.; Nakazawa, K. Postnatal NMDA Receptor Ablation in Corticolimbic Interneurons Confers Schizophrenia-like Phenotypes. Nat. Neurosci. 2010, 13, 76–83. [Google Scholar] [CrossRef]
  344. Soeiro-de-Souza, M.G.; Henning, A.; Machado-Vieira, R.; Moreno, R.A.; Pastorello, B.F.; da Costa Leite, C.; Vallada, H.; Otaduy, M.C.G. Anterior Cingulate Glutamate–Glutamine Cycle Metabolites Are Altered in Euthymic Bipolar I Disorder. Eur. Neuropsychopharmacol. 2015, 25, 2221–2229. [Google Scholar] [CrossRef]
  345. Woo, T.-U.W.; Walsh, J.P.; Benes, F.M. Density of Glutamic Acid Decarboxylase 67 Messenger RNA-Containing Neurons That Express the N-Methyl-D-Aspartate Receptor Subunit NR2A in the Anterior Cingulate Cortex in Schizophrenia and Bipolar Disorder. Arch. Gen. Psychiatry 2004, 61, 649–657. [Google Scholar] [CrossRef] [PubMed]
  346. Constantinidis, C.; Williams, G.V.; Goldman-Rakic, P.S. A Role for Inhibition in Shaping the Temporal Flow of Information in Prefrontal Cortex. Nat. Neurosci. 2002, 5, 175–180. [Google Scholar] [CrossRef]
  347. Demchenko, I.; Tassone, V.K.; Kennedy, S.H.; Dunlop, K.; Bhat, V. Intrinsic Connectivity Networks of Glutamate-Mediated Antidepressant Response: A Neuroimaging Review. Front. Psychiatry 2022, 13, 864902. [Google Scholar] [CrossRef] [PubMed]
  348. Feyissa, A.M.; Chandran, A.; Stockmeier, C.A.; Karolewicz, B. Reduced Levels of NR2A and NR2B Subunits of NMDA Receptor and PSD-95 in the Prefrontal Cortex in Major Depression. Prog. Neuropsychopharmacol. Biol. Psychiatry 2009, 33, 70–75. [Google Scholar] [CrossRef] [PubMed]
  349. Niciu, M.J.; Ionescu, D.F.; Richards, E.M.; Zarate, C.A. Glutamate and Its Receptors in the Pathophysiology and Treatment of Major Depressive Disorder. J. Neural Transm. 2014, 121, 907–924. [Google Scholar] [CrossRef]
  350. Lewis, D.A.; Moghaddam, B. Cognitive Dysfunction in Schizophrenia: Convergence of γ-Aminobutyric Acid and Glutamate Alterations. Arch. Neurol. 2006, 63, 1372–1376. [Google Scholar] [CrossRef] [PubMed]
  351. Briones, B.A.; Pitcher, M.N.; Fleming, W.T.; Libby, A.; Diethorn, E.J.; Haye, A.E.; MacDowell, C.J.; Zych, A.D.; Waters, R.C.; Buschman, T.J.; et al. Perineuronal Nets in the Dorsomedial Striatum Contribute to Behavioral Dysfunction in Mouse Models of Excessive Repetitive Behavior. Biol. Psychiatry Glob. Open Sci. 2021, 2, 460–469. [Google Scholar] [CrossRef]
  352. Kalueff, A.V.; Wheaton, M.; Murphy, D.L. What’s Wrong with My Mouse Model? Advances and Strategies in Animal Modeling of Anxiety and Depression. Behav. Brain Res. 2007, 179, 1–18. [Google Scholar] [CrossRef]
  353. Li, M.; D’Arcy, C.; Meng, X. Maltreatment in Childhood Substantially Increases the Risk of Adult Depression and Anxiety in Prospective Cohort Studies: Systematic Review, Meta-Analysis, and Proportional Attributable Fractions. Psychol. Med. 2016, 46, 717–730. [Google Scholar] [CrossRef]
  354. Aleksandrova, L.R.; Chen, W.; Wang, Y.T. An Erbin Story: Amygdala Excitation-Inhibition Balance in Anxiety. Biol. Psychiatry 2020, 87, 872–874. [Google Scholar] [CrossRef] [PubMed]
  355. Luo, Z.-Y.; Huang, L.; Lin, S.; Yin, Y.-N.; Jie, W.; Hu, N.-Y.; Hu, Y.-Y.; Guan, Y.-F.; Liu, J.-H.; You, Q.-L.; et al. Erbin in Amygdala Parvalbumin-Positive Neurons Modulates Anxiety-like Behaviors. Biol. Psychiatry 2020, 87, 926–936. [Google Scholar] [CrossRef]
  356. Filice, F.; Vörckel, K.J.; Sungur, A.Ö.; Wöhr, M.; Schwaller, B. Reduction in Parvalbumin Expression Not Loss of the Parvalbumin-Expressing GABA Interneuron Subpopulation in Genetic Parvalbumin and Shank Mouse Models of Autism. Mol. Brain 2016, 9, 10. [Google Scholar] [CrossRef] [PubMed]
  357. Hashemi, E.; Ariza, J.; Rogers, H.; Noctor, S.C.; Martínez-Cerdeño, V. The Number of Parvalbumin-Expressing Interneurons Is Decreased in the Prefrontal Cortex in Autism. Cereb. Cortex 2017, 27, 1931–1943. [Google Scholar] [CrossRef] [PubMed]
  358. Brandenburg, C.; Blatt, G.J. Region-Specific Alterations of Perineuronal Net Expression in Postmortem Autism Brain Tissue. Front. Mol. Neurosci. 2022, 15, 838918. [Google Scholar] [CrossRef] [PubMed]
  359. Xia, D.; Li, L.; Yang, B.; Zhou, Q. Altered Relationship between Parvalbumin and Perineuronal Nets in an Autism Model. Front. Mol. Neurosci. 2021, 14, 597812. [Google Scholar] [CrossRef] [PubMed]
  360. El-Ansary, A.; Al-Ayadhi, L. GABAergic/Glutamatergic Imbalance Relative to Excessive Neuroinflammation in Autism Spectrum Disorders. J. Neuroinflamm. 2014, 11, 189. [Google Scholar] [CrossRef] [PubMed]
  361. Pangrazzi, L.; Balasco, L.; Bozzi, Y. Oxidative Stress and Immune System Dysfunction in Autism Spectrum Disorders. Int. J. Mol. Sci. 2020, 21, 3293. [Google Scholar] [CrossRef] [PubMed]
  362. Burket, J.A.; Webb, J.D.; Deutsch, S.I. Perineuronal Nets and Metal Cation Concentrations in the Microenvironments of Fast-Spiking, Parvalbumin-Expressing GABAergic Interneurons: Relevance to Neurodevelopment and Neurodevelopmental Disorders. Biomolecules 2021, 11, 1235. [Google Scholar] [CrossRef]
  363. Bi, D.; Wen, L.; Wu, Z.; Shen, Y. GABAergic Dysfunction in Excitatory and Inhibitory (E/I) Imbalance Drives the Pathogenesis of Alzheimer’s Disease. Alzheimers Dement. 2020, 16, 1312–1329. [Google Scholar] [CrossRef] [PubMed]
  364. Palop, J.J.; Mucke, L. Amyloid-β–Induced Neuronal Dysfunction in Alzheimer’s Disease: From Synapses toward Neural Networks. Nat. Neurosci. 2010, 13, 812–818. [Google Scholar] [CrossRef]
  365. Ali, A.B.; Islam, A.; Constanti, A. The Fate of Interneurons, GABAA Receptor Sub-Types and Perineuronal Nets in Alzheimer’s Disease. Brain Pathol. 2023, 33, e13129. [Google Scholar] [CrossRef] [PubMed]
  366. Reichelt, A.C. Is Loss of Perineuronal Nets a Critical Pathological Event in Alzheimer’s Disease? eBioMedicine 2020, 59, 102946. [Google Scholar] [CrossRef]
  367. Caccavano, A.; Bozzelli, P.L.; Forcelli, P.A.; Pak, D.T.S.; Wu, J.-Y.; Conant, K.; Vicini, S. Inhibitory Parvalbumin Basket Cell Activity Is Selectively Reduced during Hippocampal Sharp Wave Ripples in a Mouse Model of Familial Alzheimer’s Disease. J. Neurosci. 2020, 40, 5116–5136. [Google Scholar] [CrossRef]
  368. Hijazi, S.; Heistek, T.S.; Scheltens, P.; Neumann, U.; Shimshek, D.R.; Mansvelder, H.D.; Smit, A.B.; van Kesteren, R.E. Early Restoration of Parvalbumin Interneuron Activity Prevents Memory Loss and Network Hyperexcitability in a Mouse Model of Alzheimer’s Disease. Mol. Psychiatry 2020, 25, 3380–3398. [Google Scholar] [CrossRef]
  369. Hijazi, S.; Heistek, T.S.; van der Loo, R.; Mansvelder, H.D.; Smit, A.B.; van Kesteren, R.E. Hyperexcitable Parvalbumin Interneurons Render Hippocampal Circuitry Vulnerable to Amyloid Beta. iScience 2020, 23, 101271. [Google Scholar] [CrossRef]
  370. Etter, G.; van der Veldt, S.; Manseau, F.; Zarrinkoub, I.; Trillaud-Doppia, E.; Williams, S. Optogenetic Gamma Stimulation Rescues Memory Impairments in an Alzheimer’s Disease Mouse Model. Nat. Commun. 2019, 10, 5322. [Google Scholar] [CrossRef] [PubMed]
  371. Iaccarino, H.F.; Singer, A.C.; Martorell, A.J.; Rudenko, A.; Gao, F.; Gillingham, T.Z.; Mathys, H.; Seo, J.; Kritskiy, O.; Abdurrob, F.; et al. Gamma Frequency Entrainment Attenuates Amyloid Load and Modifies Microglia. Nature 2016, 540, 230–235. [Google Scholar] [CrossRef]
  372. Singer, A.C.; Martorell, A.J.; Douglas, J.M.; Abdurrob, F.; Attokaren, M.K.; Tipton, J.; Mathys, H.; Adaikkan, C.; Tsai, L.-H. Noninvasive 40-Hz Light Flicker to Recruit Microglia and Reduce Amyloid Beta Load. Nat. Protoc. 2018, 13, 1850–1868. [Google Scholar] [CrossRef]
  373. Benussi, A.; Cantoni, V.; Cotelli, M.S.; Cotelli, M.; Brattini, C.; Datta, A.; Thomas, C.; Santarnecchi, E.; Pascual-Leone, A.; Borroni, B. Exposure to Gamma tACS in Alzheimer’s Disease: A Randomized, Double-Blind, Sham-Controlled, Crossover, Pilot Study. Brain Stimulat. 2021, 14, 531–540. [Google Scholar] [CrossRef]
  374. Buss, S.S.; Fried, P.J.; Pascual-Leone, A. Therapeutic Noninvasive Brain Stimulation in Alzheimer’s Disease and Related Dementias. Curr. Opin. Neurol. 2019, 32, 292–304. [Google Scholar] [CrossRef] [PubMed]
  375. Dhaynaut, M.; Cappon, D.; Paciorek, R.; Macone, J.; Connor, A.; Guehl, N.; Pascual-Leone, A.; Fakhri, G.E.; Santarnecchi, E. Effects of Modulating Gamma Oscillations via 40Hz Transcranial Alternating Current Stimulation (tACS) on Tau PET Imaging in Mild to Moderate Alzheimer’s Disease. J. Nucl. Med. 2020, 61, 340. [Google Scholar]
  376. Manippa, V.; Palmisano, A.; Filardi, M.; Vilella, D.; Nitsche, M.; Rivolta, D.; Logroscino, G. An Update on the Use of Gamma (Multi)Sensory Stimulation for Alzheimer’s Disease Treatment. Front. Aging Neurosci. 2022, 14, 1095081. [Google Scholar] [CrossRef] [PubMed]
  377. Sprugnoli, G.; Munsch, F.; Cappon, D.; Paciorek, R.; Macone, J.; Connor, A.; El Fakhri, G.; Salvador, R.; Ruffini, G.; Donohoe, K.; et al. Impact of Multisession 40Hz tACS on Hippocampal Perfusion in Patients with Alzheimer’s Disease. Alzheimers Res. Ther. 2021, 13, 203. [Google Scholar] [CrossRef] [PubMed]
  378. Buzsáki, G. Rhythms of the Brain, 2006th ed.; Oxford University Press: Oxford, UK.
  379. Honda, S.; Matsumoto, M.; Tajinda, K.; Mihara, T. Enhancing Clinical Trials Through Synergistic Gamma Power Analysis. Front. Psychiatry 2020, 11, 537. [Google Scholar] [CrossRef] [PubMed]
  380. McLoughlin, G.; Makeig, S.; Tsuang, M.T. In Search of Biomarkers in Psychiatry: EEG-Based Measures of Brain Function. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2014, 165, 111–121. [Google Scholar] [CrossRef]
  381. Hui, J.; Tremblay, S.; Daskalakis, Z.J. The Current and Future Potential of Transcranial Magnetic Stimulation with Electroencephalography in Psychiatry. Clin. Pharmacol. Ther. 2019, 106, 734–746. [Google Scholar] [CrossRef] [PubMed]
  382. Manippa, V.; Palmisano, A.; Nitsche, M.A.; Filardi, M.; Vilella, D.; Logroscino, G.; Rivolta, D. Cognitive and Neuropathophysiological Outcomes of Gamma-tACS in Dementia: A Systematic Review. Neuropsychol. Rev. 2023, 34, 338–361. [Google Scholar] [CrossRef] [PubMed]
  383. Padberg, F.; Bulubas, L.; Mizutani-Tiebel, Y.; Burkhardt, G.; Kranz, G.S.; Koutsouleris, N.; Kambeitz, J.; Hasan, A.; Takahashi, S.; Keeser, D.; et al. The Intervention, the Patient and the Illness–Personalizing Non-Invasive Brain Stimulation in Psychiatry. Exp. Neurol. 2021, 341, 113713. [Google Scholar] [CrossRef] [PubMed]
  384. Brunoni, A.R.; Sampaio-Junior, B.; Moffa, A.H.; Aparício, L.V.; Gordon, P.; Klein, I.; Rios, R.M.; Razza, L.B.; Loo, C.; Padberg, F.; et al. Noninvasive Brain Stimulation in Psychiatric Disorders: A Primer. Braz. J. Psychiatry 2018, 41, 70–81. [Google Scholar] [CrossRef] [PubMed]
  385. Fitzgerald, P.B.; Daskalakis, Z.J. A Review of Repetitive Transcranial Magnetic Stimulation Use in the Treatment of Schizophrenia. Can. J. Psychiatry 2008, 53, 567–576. [Google Scholar] [CrossRef] [PubMed]
  386. Sreeraj, V.S.; Shanbhag, V.; Nawani, H.; Shivakumar, V.; Damodharan, D.; Bose, A.; Narayanaswamy, J.C.; Venkatasubramanian, G. Feasibility of Online Neuromodulation Using Transcranial Alternating Current Stimulation in Schizophrenia. Indian J. Psychol. Med. 2017, 39, 92–95. [Google Scholar] [CrossRef]
  387. Alexander, M.L.; Alagapan, S.; Lugo, C.E.; Mellin, J.M.; Lustenberger, C.; Rubinow, D.R.; Fröhlich, F. Double-Blind, Randomized Pilot Clinical Trial Targeting Alpha Oscillations with Transcranial Alternating Current Stimulation (tACS) for the Treatment of Major Depressive Disorder (MDD). Transl. Psychiatry 2019, 9, 106. [Google Scholar] [CrossRef] [PubMed]
  388. Elyamany, O.; Leicht, G.; Herrmann, C.S.; Mulert, C. Transcranial Alternating Current Stimulation (tACS): From Basic Mechanisms towards First Applications in Psychiatry. Eur. Arch. Psychiatry Clin. Neurosci. 2021, 271, 135–156. [Google Scholar] [CrossRef] [PubMed]
  389. Tăuƫan, A.-M.; Casula, E.P.; Pellicciari, M.C.; Borghi, I.; Maiella, M.; Bonni, S.; Minei, M.; Assogna, M.; Palmisano, A.; Smeralda, C.; et al. TMS-EEG Perturbation Biomarkers for Alzheimer’s Disease Patients Classification. Sci. Rep. 2023, 13, 7667. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of the interaction between GABAergic PV+ interneurons and pyramidal neurons. Pyramidal neurons send excitatory (+) glutamate-mediated inputs to fast-spiking PV+ interneurons, which in turn send inhibitory (−) feedback signals back to the pyramidal cells. Activated interneurons can propagate inhibitory long-range signals to multiple pyramidal cells and synchronize their activity (E-I balance). PV+ = parvalbumin-positive, E = excitatory, I = inhibitory, PNN = perineuronal net. Created in BioRender.com.
Figure 1. Schematic representation of the interaction between GABAergic PV+ interneurons and pyramidal neurons. Pyramidal neurons send excitatory (+) glutamate-mediated inputs to fast-spiking PV+ interneurons, which in turn send inhibitory (−) feedback signals back to the pyramidal cells. Activated interneurons can propagate inhibitory long-range signals to multiple pyramidal cells and synchronize their activity (E-I balance). PV+ = parvalbumin-positive, E = excitatory, I = inhibitory, PNN = perineuronal net. Created in BioRender.com.
Life 14 00578 g001
Figure 2. Overview of the outlined pathophysiological cascade in ScZ (left panel) with relevant sample evidence (right panel). (a) Representative photomicrographs showing reduced density of PNNs in the PFC of patients with ScZ as compared to HC (Reprinted from [130] with permission from Elsevier); (b) Scatterplots showing lower cortical density of PV+ interneurons in the DLPFC of SCZ patients vs. HC (horizontal lines indicate mean values) (Reprinted from [196] with permission from Elsevier); (c) γ rhythms abnormalities during WM tasks in ScZ patients vs. HC: grand average event-related oscillations in HC (left upper panel) and ScZ patients (right upper panel) during N-back tasks with varying WM demands (easy/hard), with T = 0 ms representing the stimulus onset (Reprinted from [197] with permission from Elsevier); (d) Intertrial coherence time/frequency analyses: 40-Hz phase-locking deficits in ScZ patients (lower panel) relative to HC (upper panel) (Reprinted from [198] with permission from Elsevier); (e) Decreased coherence in higher frequency ranges (β and γ) between central and frontal areas in ScZ patients vs. HC (Reprinted from [199] with permission from Elsevier); (f) Gamma-power difference maps between resting state and cognitive task in HCs and ScZ patients (Reprinted from [188] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023)); (g) Reduced frontal inhibition in ScZ patients compared to HCs following DLPFC stimulation (Reprinted from [200] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023)). PNN = Perineuronal Net; PFC = Prefrontal Cortex; HC = Healthy Controls; PV+ = Parvalbumin-positive; DLPFC = Dorsolateral Prefrontal Cortex; WM = Working Memory; EEG = Electroencephalography; DMN = Default Mode Network; TMS = Transcranial Magnetic Stimulation; ERSP = Event-Related Spectral Perturbation.
Figure 2. Overview of the outlined pathophysiological cascade in ScZ (left panel) with relevant sample evidence (right panel). (a) Representative photomicrographs showing reduced density of PNNs in the PFC of patients with ScZ as compared to HC (Reprinted from [130] with permission from Elsevier); (b) Scatterplots showing lower cortical density of PV+ interneurons in the DLPFC of SCZ patients vs. HC (horizontal lines indicate mean values) (Reprinted from [196] with permission from Elsevier); (c) γ rhythms abnormalities during WM tasks in ScZ patients vs. HC: grand average event-related oscillations in HC (left upper panel) and ScZ patients (right upper panel) during N-back tasks with varying WM demands (easy/hard), with T = 0 ms representing the stimulus onset (Reprinted from [197] with permission from Elsevier); (d) Intertrial coherence time/frequency analyses: 40-Hz phase-locking deficits in ScZ patients (lower panel) relative to HC (upper panel) (Reprinted from [198] with permission from Elsevier); (e) Decreased coherence in higher frequency ranges (β and γ) between central and frontal areas in ScZ patients vs. HC (Reprinted from [199] with permission from Elsevier); (f) Gamma-power difference maps between resting state and cognitive task in HCs and ScZ patients (Reprinted from [188] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023)); (g) Reduced frontal inhibition in ScZ patients compared to HCs following DLPFC stimulation (Reprinted from [200] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023)). PNN = Perineuronal Net; PFC = Prefrontal Cortex; HC = Healthy Controls; PV+ = Parvalbumin-positive; DLPFC = Dorsolateral Prefrontal Cortex; WM = Working Memory; EEG = Electroencephalography; DMN = Default Mode Network; TMS = Transcranial Magnetic Stimulation; ERSP = Event-Related Spectral Perturbation.
Life 14 00578 g002
Figure 3. Overview of the outlined pathophysiological cascade in BD (left panel) with relevant sample evidence (right panel). (a) Decrease in PV+ neurons/PNNs number and density in the TRN of patients with BD as compared to HC (* indicate significance) (Reprinted from [218] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (b,1) Panoramic confocal microphotograph showing PNNs (blue) surrounding PV+ somata (red) in the deep layers of the DLPFC in HC vs. BD patients (white arrowheads: PV+ somata surrounded by PNNs, yellow arrowheads: PV+ cells lacking PNNs, white arrows: PNNs surrounding PV+ somata); (b,2) Histograms of significant differences in the density of PNNs in HC vs. BD patients (as well as SCZ and MDD) (* indicates significance) (Reprinted from [219] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (c) Group averaged time-frequency maps of ASSR power for each hemisphere, with color scales signifying ASSR power, showing BD patients’ reduced mean ASSR power and PLF to 40- and 80- Hz stimulation (Reprinted from [256] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023)); (d) Lower post-stimulus EEG coherence between F4 and T6 locations in BD patients (red line) as compared to HC (blue line) (Reprinted from [252] with permission from Elsevier); (e) Representative data from an HC and a BD patient: average EEG responses to TMS (grey traces represent the 60 recording channels) for the channel closest to the stimulation site (black trace) over the premotor area; color-coded: ERSP plots reflecting the significant TMS-related changes in amplitude and their duration (Reprinted from [263] with permission from Cambridge University Press). PV+ = Parvalbumin-positive; PNN = Perineuronal Net; TRN = Thalamic Reticular Nucleus; BPD = Bipolar Disorder; SCZ = Schizophrenia; HC = Healthy Controls; DLPFC = Dorsolateral Prefrontal Cortex; SCZ = Schizophrenia; ASSR = Auditory Steady-State Response; PLF = Phase-Locking Factor; EEG = Electroencephalography; TMS = Transcranial Magnetic Stimulation; ERSP = Event-Related Spectral Perturbation.
Figure 3. Overview of the outlined pathophysiological cascade in BD (left panel) with relevant sample evidence (right panel). (a) Decrease in PV+ neurons/PNNs number and density in the TRN of patients with BD as compared to HC (* indicate significance) (Reprinted from [218] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (b,1) Panoramic confocal microphotograph showing PNNs (blue) surrounding PV+ somata (red) in the deep layers of the DLPFC in HC vs. BD patients (white arrowheads: PV+ somata surrounded by PNNs, yellow arrowheads: PV+ cells lacking PNNs, white arrows: PNNs surrounding PV+ somata); (b,2) Histograms of significant differences in the density of PNNs in HC vs. BD patients (as well as SCZ and MDD) (* indicates significance) (Reprinted from [219] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (c) Group averaged time-frequency maps of ASSR power for each hemisphere, with color scales signifying ASSR power, showing BD patients’ reduced mean ASSR power and PLF to 40- and 80- Hz stimulation (Reprinted from [256] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023)); (d) Lower post-stimulus EEG coherence between F4 and T6 locations in BD patients (red line) as compared to HC (blue line) (Reprinted from [252] with permission from Elsevier); (e) Representative data from an HC and a BD patient: average EEG responses to TMS (grey traces represent the 60 recording channels) for the channel closest to the stimulation site (black trace) over the premotor area; color-coded: ERSP plots reflecting the significant TMS-related changes in amplitude and their duration (Reprinted from [263] with permission from Cambridge University Press). PV+ = Parvalbumin-positive; PNN = Perineuronal Net; TRN = Thalamic Reticular Nucleus; BPD = Bipolar Disorder; SCZ = Schizophrenia; HC = Healthy Controls; DLPFC = Dorsolateral Prefrontal Cortex; SCZ = Schizophrenia; ASSR = Auditory Steady-State Response; PLF = Phase-Locking Factor; EEG = Electroencephalography; TMS = Transcranial Magnetic Stimulation; ERSP = Event-Related Spectral Perturbation.
Life 14 00578 g003
Figure 4. Overview of the outlined pathophysiological cascade in MDD (left panel) with relevant sample evidence (right panel). (a) Representative images of immunofluorescence staining of PNNs and quantification of neurons (neuronal marker: NeuN+) in the PrL of HC compared to vulnerable and resilient rats in response to CUMS (Reprinted from [284] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (b) Decreased number of PV+ cells in coronal sections of the cingulate cortex, amygdala and hippocampus in HC vs. refractory depressed rats (Reprinted from [300] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (c) Graph for modulation of γ oscillatory power calculated as ERS/ERD% (with data pooled across O1 and O2 electrodes): positive values reflect ERS, negative values reflect ERD. Gamma ERS/ERD% during the Sternberg task was lower for MDD patients (orange) compared to HC (blue) (Reprinted from [316] with permission from Elsevier)); (d) Hyperactive frontolimbic and frontocentral resting-state γ connectivity in MDD: γ-band functional connectivity network differences when comparing MDD patients to HC at rest: top 11 most connected brain regions with node sizes representing the number of connections within the network (upper left), grand average of γ functional connectivity strengths within the identified network across MDD patients and HC, respectively (bottom right) (Reprinted from [327] with permission from Elsevier); (e) Increased γ power in MDD patients following rTMS: (e,1) Resting γ power: bar graphs showing mean resting γ power before and after rTMS at the F3 and F4 electrode sites across MDD patients (* indicates significance); (e,2) EEG topographical plots of resting-state γ power; from right to left, the topoplots depict the γ power distribution pre-rTMS, post-rTMS, the difference between pre- and post-rTMS, and the t-value map corresponding to the difference between pre- and post-rTMS; (e,3,4) Scatter plots showing positive correlation between the resting γ power increase at the F3 electrode site and improvements in HAM-D17 and BDI (Reprinted from [324] with permission from Elsevier). PNN = Perineuronal Net; HC = Healthy Controls; PrL = Prelimbic cortex; CUMS = Chronic Unpredictable Mild Stress; PV+ = Parvalbumin-positive; ERS = Event-Related Synchronization; ERD = Event-Related Desynchronization; ERS/ERD% = Event-Related Synchronization/Desynchronization Percentage; MDD = Major Depressive Disorder; rTMS = repetitive Transcranial Magnetic Stimulation; EEG = Electroencephalography; HAM-D17 = Hamilton Depression Rating Scale-17; BDI = Beck Depression Inventory.
Figure 4. Overview of the outlined pathophysiological cascade in MDD (left panel) with relevant sample evidence (right panel). (a) Representative images of immunofluorescence staining of PNNs and quantification of neurons (neuronal marker: NeuN+) in the PrL of HC compared to vulnerable and resilient rats in response to CUMS (Reprinted from [284] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (b) Decreased number of PV+ cells in coronal sections of the cingulate cortex, amygdala and hippocampus in HC vs. refractory depressed rats (Reprinted from [300] (licensed under CCBY 4.0, https://creativecommons.org/licenses/by/4.0/, accessed on 7 December 2023); (c) Graph for modulation of γ oscillatory power calculated as ERS/ERD% (with data pooled across O1 and O2 electrodes): positive values reflect ERS, negative values reflect ERD. Gamma ERS/ERD% during the Sternberg task was lower for MDD patients (orange) compared to HC (blue) (Reprinted from [316] with permission from Elsevier)); (d) Hyperactive frontolimbic and frontocentral resting-state γ connectivity in MDD: γ-band functional connectivity network differences when comparing MDD patients to HC at rest: top 11 most connected brain regions with node sizes representing the number of connections within the network (upper left), grand average of γ functional connectivity strengths within the identified network across MDD patients and HC, respectively (bottom right) (Reprinted from [327] with permission from Elsevier); (e) Increased γ power in MDD patients following rTMS: (e,1) Resting γ power: bar graphs showing mean resting γ power before and after rTMS at the F3 and F4 electrode sites across MDD patients (* indicates significance); (e,2) EEG topographical plots of resting-state γ power; from right to left, the topoplots depict the γ power distribution pre-rTMS, post-rTMS, the difference between pre- and post-rTMS, and the t-value map corresponding to the difference between pre- and post-rTMS; (e,3,4) Scatter plots showing positive correlation between the resting γ power increase at the F3 electrode site and improvements in HAM-D17 and BDI (Reprinted from [324] with permission from Elsevier). PNN = Perineuronal Net; HC = Healthy Controls; PrL = Prelimbic cortex; CUMS = Chronic Unpredictable Mild Stress; PV+ = Parvalbumin-positive; ERS = Event-Related Synchronization; ERD = Event-Related Desynchronization; ERS/ERD% = Event-Related Synchronization/Desynchronization Percentage; MDD = Major Depressive Disorder; rTMS = repetitive Transcranial Magnetic Stimulation; EEG = Electroencephalography; HAM-D17 = Hamilton Depression Rating Scale-17; BDI = Beck Depression Inventory.
Life 14 00578 g004
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

Palmisano, A.; Pandit, S.; Smeralda, C.L.; Demchenko, I.; Rossi, S.; Battelli, L.; Rivolta, D.; Bhat, V.; Santarnecchi, E. The Pathophysiological Underpinnings of Gamma-Band Alterations in Psychiatric Disorders. Life 2024, 14, 578. https://doi.org/10.3390/life14050578

AMA Style

Palmisano A, Pandit S, Smeralda CL, Demchenko I, Rossi S, Battelli L, Rivolta D, Bhat V, Santarnecchi E. The Pathophysiological Underpinnings of Gamma-Band Alterations in Psychiatric Disorders. Life. 2024; 14(5):578. https://doi.org/10.3390/life14050578

Chicago/Turabian Style

Palmisano, Annalisa, Siddhartha Pandit, Carmelo L. Smeralda, Ilya Demchenko, Simone Rossi, Lorella Battelli, Davide Rivolta, Venkat Bhat, and Emiliano Santarnecchi. 2024. "The Pathophysiological Underpinnings of Gamma-Band Alterations in Psychiatric Disorders" Life 14, no. 5: 578. https://doi.org/10.3390/life14050578

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