Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-24T07:54:11.652Z Has data issue: false hasContentIssue false

The tetrapartite synapse: a key concept in the pathophysiology of schizophrenia

Published online by Cambridge University Press:  01 January 2020

Gabriele Chelini
Affiliation:
aTranslational Neuroscience Laboratory, Mclean Hospital, 115 Mill StreetBelmontMA, 02478USA bDept. of Psychiatry, Harvard Medical School, 25 Shattuck St, Boston, MA, 02115USA
Harry Pantazopoulos
Affiliation:
aTranslational Neuroscience Laboratory, Mclean Hospital, 115 Mill StreetBelmontMA, 02478USA bDept. of Psychiatry, Harvard Medical School, 25 Shattuck St, Boston, MA, 02115USA
Peter Durning
Affiliation:
aTranslational Neuroscience Laboratory, Mclean Hospital, 115 Mill StreetBelmontMA, 02478USA
Sabina Berretta
Affiliation:
aTranslational Neuroscience Laboratory, Mclean Hospital, 115 Mill StreetBelmontMA, 02478USA bDept. of Psychiatry, Harvard Medical School, 25 Shattuck St, Boston, MA, 02115USA cProgram in Neuroscience, Harvard Medical School, 220 Longwood Ave., Boston, MA, 02115USA

Abstract

Growing evidence points to synaptic pathology as a core component of the pathophysiology of schizophrenia (SZ). Significant reductions of dendritic spine density and altered expression of their structural and molecular components have been reported in several brain regions, suggesting a deficit of synaptic plasticity. Regulation of synaptic plasticity is a complex process, one that requires not only interactions between pre- and post-synaptic terminals, but also glial cells and the extracellular matrix (ECM). Together, these elements are referred to as the ‘tetrapartite synapse’, an emerging concept supported by accumulating evidence for a role of glial cells and the extracellular matrix in regulating structural and functional aspects of synaptic plasticity. In particular, chondroitin sulfate proteoglycans (CSPGs), one of the main components of the ECM, have been shown to be synthesized predominantly by glial cells, to form organized perisynaptic aggregates known as perineuronal nets (PNNs), and to modulate synaptic signaling and plasticity during postnatal development and adulthood. Notably, recent findings from our group and others have shown marked CSPG abnormalities in several brain regions of people with SZ. These abnormalities were found to affect specialized ECM structures, including PNNs, as well as glial cells expressing the corresponding CSPGs. The purpose of this review is to bring forth the hypothesis that synaptic pathology in SZ arises from a disruption of the interactions between elements of the tetrapartite synapse.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - ND
This is an open access article under the CC BY-NC-ND license
Copyright
Copyright © European Psychiatric Association 2018

1. Introduction

Growing evidence points to synaptic pathology across several brain disorders, including schizophrenia (SZ), bipolar disorder, major depression, autism spectrum disorder and Alzheimer’s disease. Research on the underlying mechanisms for this pathology has only very recently begun to make headway, and important questions arise with regard to the potential common denominators of synaptic pathology among these disorders, and their timeframe across lifespan. With regard to the latter, it is important to consider that synaptic remodeling occurs constantly throughout life. During postnatal development, excessive synaptic formation is followed by elimination of less active synapses, a process named synapse pruning [Reference Huttenlocher1, Reference Huttenlocher and Dabholkar2]. During adult life, synapses are highly dynamic, with constant, experience-driven synaptic growth and elimination. It is plausible to postulate that timeframe-, mechanism- and brain region- specificity underlying synaptic pathology across a spectrum of brain disorders may at least in part contribute to their diverse clinical and pathophysiological manifestations. Within this context, we suggest that the concept of ‘tetrapartite synapse’ may be a useful starting point for investigating synaptic pathology. We focus on schizophrenia as a notable example.

1.1. The tetrapartite synapse

The chemical synapse has classically been considered as composed of two main elements, i.e. the presynaptic and postsynaptic elements. This concept evolved over the past two decades to include a third element, i.e. the astrocyte, as processes from these cells envelope the synapse and play a key role in regulating its functions [Reference Ventura and Harris3, Reference Genoud, Quairiaux, Steiner, Hirling, Welker and Knott4]. The ensemble of pre- and post-synaptic elements and astrocytes has been proposed to form a functional complex referred to as the ‘tripartite synapse’ [Reference Araque, Parpura, Sanzgiri and Haydon5]. Growing evidence indicates that other populations of glial cells, including NG2 glia and microglia, also play critical roles in regulating synaptic functions and plasticity. Thus, we suggest that distinct populations of glial cells with specific functions may be considered together to represent the third element of tripartite synapse. Yet more recently, the extracellular matrix (ECM) has come to the forefront of neuroscience as an active component of neural functions and, in particular, synaptic regulation. On the basis of this evidence, Dityatev et al., proposed the elegant concept of the ‘tetrapartite synapse’, composed of pre- and post-synaptic elements, glial processes and ECM, and elegantly documented the interactions between these components [Reference Dityatev and Rusakov6Reference Song and Dityatev8] (Fig. 1). Here, we review evidence supporting the idea that synaptic functions and plasticity result from interactions between all elements of the tetrapartite synapse, and focus on evidence that these interactions are disrupted in SZ.

Fig. 1 Diagramatic representation of the tetrapartite synapse. Elements composing it are the pre- and post-synaptic terminals, astrocytic processes surrounding them and perisynaptic extracellular matrix condensations interposed between these elements.

1.1.1. The tetrapartite synapse: pre- and post-synaptic elements

The brain possesses the extraordinary ability to continuously reshape itself throughout the entire lifespan. This property, defined as plasticity, is based on the highly dynamic properties of synaptic contacts, i.e. the ability to generate new synapses, eliminate them, and alter the electrophysiological, molecular and structural properties of existing ones in response to experience. The mechanisms underlying synaptic plasticity have been the object of intense work and exciting discoveries over the past few decades, focused initially on the interplay between the presynaptic and postsynaptic elements. For example, the discovery that trains of presynaptic action potentials induce a long–lasting increase in synaptic strength during long term potentiation (LTP) generated intense debate over whether the predominant underlying changes may be related to postsynaptic modification in α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors (AMPARs) or altered presynaptic transmitter release [Reference Nicoll, Kauer and Malenka9, Reference Malenka and Bear10]. Ensuing work demonstrated that these mechanisms include a series of steps, from receptor phosphorylation, to protein synthesis and, eventually, structural changes including growth of new dendritic spines and increased size of pre-existing spines, mediated by changes dendritic spine molecular cytoskeleton, including long-lasting increases in F-actin content within spines and condensation of the post-synaptic density (PSD), a dense aggregate of scaffolding proteins implicated in structural maintenance and signal transduction [Reference Banker, Churchill and Cotman11Reference Kennedy13]. Similarly, long-term depression (LTD) of synaptic strength results in spine shrinkage and elimination [Reference Okamoto, Nagai, Miyawaki and Hayashi14Reference Zhou, Homma and Poo17]. A review of current knowledge on the role of pre- and post-synaptic elements in plasticity is beyond the scope of this manuscript; arguably plastic modifications of these elements may be considered to be the final result of complex, experience-driven mechanisms, and the underlying substrate of learning and memory.

1.1.2. The tetrapartite synapse: role of glial cells

1.1.2.1. Astrocytes

Astrocytic processes envelop the pre- and post-synaptic elements and closely approach the synaptic cleft, thus representing a key component of the synapse [Reference Ventura and Harris3, Reference Witcher, Kirov and Harris18] (Fig. 1). Their robust expression of the high-affinity glutamate transporters EAAT1–EAAT2 allows them to rapidly re-uptake excess of glutamate released from the presynaptic terminals, thus restricting excitatory transmission [Reference Danbolt19]. In turn, astrocytes actively contribute to synaptic transmission and plasticity. Although electrically silent, they respond to presynaptic activation with G-protein-mediated Ca2 signals, triggering the release of ‘gliotransmitters’, including glutamate, ATP, and GABA, which modulate local synaptic transmission [Reference Halassa and Haydon20Reference Rivera, Vanzulli and Butt25]. Astrocyte-derived glutamate facilitates NMDA receptor activation on the postsynaptic sites, enhancing the probability of triggering LTP [Reference Perea and Araque26]. Moreover, stimuli inducing synaptic LTP rapidly induce structural remodeling of astrocytic processes enwrapping synapses, resulting in changes in their ability to modulate synaptic transmission [Reference Perez-Alvarez, Navarrete, Covelo, Martin and Araque27]. Together, these considerations compellingly point to astrocytes as key active mediators of synaptic plasticity.

1.1.2.2. NG2 cells

In the early nineties, a novel population of cells was identified by their expression of the CSPG NG2, platelet-derived growth factor α receptors (PDGFαR), and O4 sulfatide (O4) [Reference Nishiyama, Lin, Giese, Heldin and Stallcup28Reference Stallcup and Beasley30]. These cells are abundant in the white and gray matter, and have been shown to represent the main source of mature oligodendrocytes in the mature brain, earning the name of oligodendrocyte precursor cells [Reference Dawson, Levine and Reynolds31]. However, morphological features of NG2 cells closely resemble those of mature astrocytes, while their molecular signature is quite distinct [Reference Reynolds and Hardy29, Reference Levine and Card32, Reference Butt, Hamilton, Hubbard, Pugh and Ibrahim33]. Further studies showed that NG2 cells represent a distinct mature glial type with unique properties. Indeed, a growing body of evidence indicates that NG2 cell express voltage-gated ion channels and ligand-gated ionotropic neurotransmitter receptors, typically found in neurons but not in glial cells; this peculiar pattern of molecular expression endows them with a unique electrophysiological profile [Reference Larson, Zhang and Bergles34]. They are the only glial cell type capable of receiving synaptic GABAergic and glutamatergic contacts from neurons, and to respond to excitatory inputs with excitatory postsynaptic currents and activity-dependent modifications analogous to LTP at excitatory synapses, although with some notable differences [Reference Ge, Yang, Zhang, Wang, Shen and Deng35Reference Lin and Bergles37]. Electron microscopy studies demonstrated that NG2 cell processes often encapsulate neuronal cell bodies and contact synapses [Reference Butt, Kiff, Hubbard and Berry38]. Potential effects of NG2 cells on synaptic transmission are suggested by experiments showing that downregulation of NG2 expression results in altered subunit composition of AMPA receptors and marked reduction of NMDA-dependent LTP [Reference Sakry, Neitz, Singh, Frischknecht, Marongiu and Biname39]. The full range of NG2 cell functions is only partially understood, and is likely to be age-, stage- and brain region-specific; however, plastic responses of these cells to excitatory neurotransmission and their intimate contacts with synaptic complexes strongly suggest that they play a significant role in the regulation of synaptic functions and plasticity [Reference Larson, Zhang and Bergles34, Reference Eugenin-von Bernhardi and Dimou40Reference Hill and Nishiyama43].

1.1.2.3. Microglia

Microglial cells serve diverse roles during brain development and adulthood, from regulation of synaptic pruning and plasticity to removal of apoptotic cells and debris, and immune responses, representing primary sources of immune response factors such as cytokines (see44). Notably, these seemingly unrelated functions may in fact share similar mechanisms, as growing evidence indicates that immune factors play important roles in the regulation of synaptic plasticity [Reference Wu, Dissing-Olesen, MacVicar and Stevens45]. Microglial cells express an impressive variety of receptors, including not only immune receptors, such as pattern-recognition receptors that allow them to recognize pathogen-associated molecular patterns and tissue damage–associated molecular patterns and chemokine receptors, but also a surprisingly large number of receptors for neurotransmitters and neuropeptides [Reference Colonna and Butovsky44]. These two latter categories include ionotropic and metabotropic glutamate receptors, and receptors for GABA, dopamine, catecholamines and acetylcholine, which mediate neural-glia communications. These receptors modulate the release of cytokines and guide microglial processes toward active synapses, where they regulate synaptic structural plasticity [Reference Colonna and Butovsky44Reference Pocock and Kettenmann46]. In particular, microglia represent a key player in the remodeling, particularly removal, of inactive synapses during brain development and adulthood. Microglia processes are highly dynamic, continually extending, retracting and interacting with synapses, thus as acting like sentinels to assess surrounding synapses and contribute to their remodeling when needed [Reference Halassa and Haydon20Reference Araque, Carmignoto, Haydon, Oliet, Robitaille and Volterra23]. Among the immune-related molecules involved in these mechanisms are several complement factors, which act as signal to microglia to find and engulf pre and post-synaptic elements [Reference Bialas, Presumey, Das, van der Poel, Lapchak and Mesin47Reference Schafer, Lehrman and Stevens49]. Complement signaling pathways also mediate microglia-induced long term depression (LTD), involved in brain circuitry optimization, and potentially in memory impairments and synaptic disruptions in neuroinflammation-related brain disorders [Reference Chai, Fan, Shao, Lu, Zhang and Li50]. Notably, the strongest genetic association observed in SZ involves variation in the Major Histocompatibility Complex locus, shown to arise predominantly from alleles of the complement component 4 (C4) genes [Reference Ripke51, Reference Sekar, Bialas, de Rivera, Davis, Hammond and Kamitaki52]. Elegant work by Sekar et al. recently showed that these alleles do affect C4 expression and that this factor mediates synapse elimination during postnatal development [Reference Sekar, Bialas, de Rivera, Davis, Hammond and Kamitaki52]. In summary, microglial cells play a key role in shaping synaptic connectivity in an activity-dependent manner – the underlying mechanisms involve molecular factors with strong relevant to the pathophysiology of SZ.

1.1.3. The tetrapartite synapse: role of the extracellular matrix

The brain ECM is a complex molecular network that surrounds all cells, occupying approximately a 20% volume fraction of the adult brain [Reference Sykova and Nicholson53]. It’s main components include hyaluronan, proteoglycans, glycoproteins and a variety of posttranslational remodeling proteases, such as matrix metalloproinases (MMPs), ‘a disintegrin and matrix metalloproteases’ (ADAMS), and ‘ADAMS with a thrombospondin domain’ (ADAMTS), which cleave ECM molecules, allowing for highly dynamic functional adaptations [Reference Rivera, Khrestchatisky, Kaczmarek, Rosenberg and Jaworski54Reference Hobohm, Gunther, Grosche, Rossner, Schneider and Bruckner58]. As discussed below, organized forms of ECM surround the synapse, fill the synaptic cleft and interact with cell surface receptors (Fig. 1). Converging evidence points to peri-synaptic ECM aggregates as a critical player contributing to synaptic signaling and plasticity.

1.1.3.1. ECM factors regulating synaptic plasticity

Chondroitin sulfate proteoglycans (CSPGs) have been described as the organizers of the ECM, of which they represent a main component [Reference Yamaguchi59Reference Lendvai, Morawski, Negyessy, Gati, Jager and Baksa64]. These macromolecules consist of core proteins linked to varying numbers of chondroitin sulfate (CS) glycosaminoglycan (GAG) chains. The number and length of GAG chains, and particularly their sulfation patterns (e.g. CS-6, CS-4), are key factors in determining their functions, resulting in highly dynamic structural and functional diversity to these molecules [Reference Karus, Samtleben, Busse, Tsai, Dietzel and Faissner65Reference Maeda68]. While their functions in the developing and mature brain are highly diversified, mounting evidence indicates that CSPGs play a complex role in developmental and adult regulation of synaptic plasticity. For example, enzymatic CSPG removal in vitro mouse hippocampal slices causes a two-fold decrease in long-term potentiation (LTP) [Reference Bukalo, Schachner and Dityatev69]. Overexpression of CS-6 sulfation in mice leads to failure to instate an adult form of restricted plasticity [Reference Miyata, Komatsu, Yoshimura, Taya and Kitagawa66]. Altered expression of several CSPGs, such as PTPRZ1, neurocan and brevican, was found to be associated with synaptic remodeling LTP abnormalities and learning impairment [Reference Niisato, Fujikawa, Komai, Shintani, Watanabe and Sakaguchi70Reference Brakebusch, Seidenbecher, Asztely, Rauch, Matthies and Meyer75]. Notably, several CSPGs have been shown to actively stabilize dendritic spines, while their removal by enzymatic digestion results in increased spine motility [Reference Pizzorusso, Medini, Landi, Baldini, Berardi and Maffei76Reference Orlando, Ster, Gerber, Fawcett and Raineteau79].

Several other ECM molecules have been found to be involved in the regulation of synaptic plasticity. For instance, genetic or pharmacological removal of the ECM component tenascin-C and thrombospondins 1 and 2 resulted in reduced calcium signaling and impaired LTP in rodents (Evers et al.; Dityatev et al.). Hyaluronan, considered to be the backbone of the ECM and enriched in PNNs and other forms of ECM perisynaptic aggregates, was found to regulate hippocampal synaptic plasticity by modulating postsynaptic L-type Ca2+ channels [Reference Kochlamazashvili, Henneberger, Bukalo, Dvoretskova, Senkov and Lievens80]. Other ECM components have been shown to modulate chemical transmission by acting on glutamate NMDA and AMPA receptors and impacting adaptive synapse modifications. Particularly relevant to several brain disorders, including SZ, is the ECM glycoprotein Reelin. Reelin’s effects are mediated through its main lipoprotein receptors, apolipoprotein E receptor 2 and very-low-density lipoprotein receptor [Reference D'Arcangelo, Homayouni, Keshvara, Rice, Sheldon and Curran81, Reference Herz and Chen82], as well as through the integrin family and the Src family kinases [Reference Herz and Chen82Reference Qiu, Zhao, Korwek and Weeber85]. Reelin is secreted into the ECM, where it regulates the composition of NMDA receptors, controlling the predominance and/or phosphorylation of the NR2 NMDA receptor subunits, augments AMPA responses by increasing the number of AMPA receptors on the postsynaptic membrane, and robustly enhances LTP [Reference Qiu, Zhao, Korwek and Weeber85Reference Hellwig, Hack, Kowalski, Brunne, Jarowyj and Unger87]. Reelin powerfully promotes spine remodeling, regulating spine size and stability, and number of synaptic contacts per spine [Reference Pujadas, Gruart, Bosch, Delgado, Teixeira and Rossi88Reference Eastwood and Harrison93]. Integrins, known to interact with Reelin and other ECM molecules, regulate AMPA receptor internalization, surface mobility of NMDA receptor subunits, and synaptic dwell time of glycine receptors and their scaffolding molecule gephyrin [Reference Dityatev, Schachner and Sonderegger94Reference McGeachie, Cingolani and Goda97]. These mechanisms have been postulated to allow integrins to play complex roles in synaptic plasticity, including carrying out structural and functional changes that accompany LTP [Reference McGeachie, Cingolani and Goda97Reference Chan, Chen, Bradley, Dragatsis, Rosenmund and Davis99]. Secreted ECM proteases, such as MMPs, affect excitatory transmission and have extensively been investigated as mediators of synaptic plasticity [Reference Huntley100Reference Verslegers, Lemmens, Van Hove and Moons103]. During development, MMPs play a key role in spine formation and maturation [Reference Michaluk, Wawrzyniak, Alot, Szczot, Wyrembek and Mercik104Reference Levy, Omar and Koleske106]. In mature neurons, MMPs and their interactions with integrins, are required for spine volume changes induced by LTP and LTD [Reference Nagy, Bozdagi, Matynia, Balcerzyk, Okulski and Dzwonek107, Reference Wang, Bozdagi, Nikitczuk, Zhai, Zhou and Huntley108]. MMP-9 is transiently released in response to enhanced neuronal activity and impacts both synaptic potentiation and dendritic spine enlargement in a dependent manner [Reference Wang, Bozdagi, Nikitczuk, Zhai, Zhou and Huntley108, Reference Lepeta and Matrix109]. Notably, several MMPs have been shown to be represented in WFA-labeled PNNs [Reference Rossier, Bernard, Cabungcal, Perrenoud, Savoye and Gallopin110], suggesting a role in regulating their functions. Semaphorins, key components of the ECM, have also been shown to regulate synaptogenesis (Pasterkamp & Giger 2009). For instance, semaphorin 3A, a key component of at least a subpopulation of PNNs, exerts a powerful effect on synapses, possibly through its plexin and neuropilin receptors [Reference Levy, Omar and Koleske106, Reference Fawcett111Reference Dick, Tan, Alves, Ehlert, Miller and Hsieh-Wilson114].

1.1.3.2. ECM perisynaptic aggregates
1.1.3.2.1. Perineuronal nets

In addition to a loosely organized molecular lattice, the ECM forms organized, structured aggregates with distinct molecular composition. PNNs are arguably the most extensively investigated. They tightly surround synaptic contacts on distinct populations of neurons, including GABAergic interneuron populations and GABAergic projection neurons, such as those in the reticular nucleus of the thalamus, central nucleus of the amygdala and Purkinje cells in the cerebellum, as well as subpopulations of cortico-cortical pyramidal cells and spinal cord motor neurons [Reference Gati and Lendvai62, Reference Wegner, Hartig, Bringmann, Grosche, Wohlfarth and Zuschratter115Reference Berretta, Pantazopoulos, Markota, Brown and Batzianouli121]. PNNs represent key players in the regulation of synaptic connectivity and plasticity [Reference Fawcett111, Reference Sorg, Berretta, Blacktop, Fawcett, Kitagawa and Kwok122Reference Gogolla, Caroni, Luthi and Herry129]. They mature late in postnatal development, in an activity-dependent manner [Reference Koppe, Bruckner, Brauer, Hartig and Bigl130Reference Favuzzi, Marques-Smith, Deogracias, Winterflood, Sanchez-Aguilera and Mantoan133]. Their maturation brings to a closure critical periods of development, inducing a profound shift from juvenile forms of plasticity to more restricted mature forms, consolidating successfully established synaptic connectivity and controlling formation of new synapses [Reference Pizzorusso, Medini, Landi, Baldini, Berardi and Maffei76, Reference Bernard and Prochiantz124, Reference Gogolla, Caroni, Luthi and Herry129, Reference Pizzorusso, Medini, Berardi, Chierzi, Fawcett and Maffei134]. Enzymatic CSPG digestion dramatically disrupts PNN integrity, reverting local circuits and learning modalities, from visual perception to emotional learning, to a juvenile state [Reference Gogolla, Caroni, Luthi and Herry129, Reference Pizzorusso, Medini, Berardi, Chierzi, Fawcett and Maffei134]. The molecular composition of mature PNNs is thought to be species-, neuron- and brain region-specific, but to include CSPGs, hyaluronan and a variety of glycoproteins described above in relationship to synaptic plasticity regulation. Thus, functions such as modulation of glutamatergic transmission, LTP and LTD and synaptic motility and structural plasticity, demonstrated for these molecules, are inherent to PNNs. In addition, recent evidence shows that PNNs are in themselves dynamically regulated [Reference Banerjee, Gutzeit, Baman, Aoued, Doshi and Liu123]. Fear learning and consolidation in response to pure tones was shown to induce marked PNN changes in the adult auditory cortex [Reference Banerjee, Gutzeit, Baman, Aoued, Doshi and Liu123]. Expression CSPG mRNA and numbers of PNN were increased within hours following fear conditioning and returned to baseline 24 h later. CSPG enzymatic digestion in the auditory cortex impaired fear learning and consolidation, demonstrating that PNNs are necessary for fear learning [Reference Banerjee, Gutzeit, Baman, Aoued, Doshi and Liu123]. Notably, in the amygdala, CSPG digestion also affected fear learning, reinstating juvenile forms of extinction-vulnerable conditioning [Reference Gogolla, Caroni, Luthi and Herry129].

1.1.3.2.2. CS-6/Glia clusters

Contrary to what initially thought, PNNs are not unique as forms of organized ECM in the brain. Among other forms, CS-6/Glia clusters, or DAndelion-like Clock Structure, may be equally relevant to synaptic plasticity [Reference Hayashi, Tatsumi, Okuda, Yoshikawa, Ishizaka and Miyata135, Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136]. CS-6/Glia clusters are detectable in human and rodent brain using antibodies raised against the CS-6 sulfation patterns (Fig. 2). Their morphology may vary across brain regions, but in general they present as round rosettes of diffuse immunolabeling, with an overall diameter of 100–200 μm, often organized in short, dense segments. Several dendrites, and occasionally some neuronal and glial cell bodies are embedded in these clusters, while several glial cells surround them (Chelini et al; unpublished observations) [Reference Hayashi, Tatsumi, Okuda, Yoshikawa, Ishizaka and Miyata135, Reference Horii-Hayashi, Tatsumi, Matsusue, Okuda, Okuda and Hayashi137]. In the mouse brain, CS-6/Glia clusters were found to develop in late postnatal development, suggesting a role in regulation of synaptic connectivity similar to that shown for PNNs [Reference Hayashi, Tatsumi, Okuda, Yoshikawa, Ishizaka and Miyata135, Reference Horii-Hayashi, Tatsumi, Matsusue, Okuda, Okuda and Hayashi137Reference Takesian and Hensch141]. Increases of CS-6 clusters in response to ketamine treatment suggest that these structures may be responsive to changes of glutamatergic transmission [Reference Matuszko, Curreli, Kaushik, Becker and Dityatev142]. Although still preliminary, information on CS-6/Glia clusters is consistent with their involvement in synaptic functions, potentially representing segregated microenvironments regulated by predominant expression of CS-6 sulfation. The functional effects of these sulfation patterns are currently poorly understood, but evidence suggests a role for CS-6 sulfation patterns in facilitating plasticity. A switch from CS-6 to CS-4 sulfation patterns during postnatal development, and persistent cortical plasticity induced by CS-6 upregulation, suggest that the former may be more permissive, facilitating plasticity, while CS-4 may represent the mature, less permissive, CSPG form [Reference Miyata, Komatsu, Yoshimura, Taya and Kitagawa66] (Fig. 3).

Fig. 2 CS-6/Glia clusters in the healthy human amygdala, immunolabeled with CS-6 antibody CS56. Scale bar 100 μm.

Fig. 3 (A) Rodent CS-6 cluster (red; immunolabeled with CS56) surrounded by astrocytes (green; immunoreactive for glial fibrillary acidic protein (GFAP). (B) Immunolabeled CS-6/Glia clusters (blue) in the mouse hippocampus. These clusters are crossed by several dendrites arising from projection neurons (green, immunolabeled for Thy1) and are often surrounded by interneurons expressing parvalbumin (red). Scale bar 100 μm. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

1.2. Synaptic pathology in SZ

Growing evidence overwhelmingly points to synaptic abnormalities as core component of the pathology of SZ. Significant reductions of dendritic spines have been reported in several cortical areas, including prefrontal and auditory cortical areas and the hippocampus [Reference Glantz and Lewis143Reference Kolomeets, Orlovskaya, Rachmanova and Uranova147]. Anomalous dendritic spine morphology, expression of PSD proteins, including PSD95 and Homer-1, and associated glutamate signaling pathway proteins have also been reported [Reference de Bartolomeis, Latte, Tomasetti and Iasevoli148, Reference MacDonald, Ding, Newman, Hemby, Penzes and Lewis149]. Altered expression of molecules involved in the actin cytoskeleton strongly suggests that dendritic spines in SZ may reflect structural deficits [Reference Datta, Arion, Corradi and Lewis150Reference Konopaske, Subburaju, Coyle and Benes152]. A recent study shows dramatic changes affecting genes involved in synaptic functions in the amygdala [Reference Beach, Huck, Zhu and Bozoki153], notably one of the main regions shown to have significant ECM abnormalities in this disorder. Decreases of dendritic spines in SZ have been proposed to represent the consequence of overpruning during adolescence [Reference MacDonald, Alhassan, Newman, Richard, Gu and Kelly154, Reference Feinberg155]. This hypothesis predicts a loss of large, mature spines in this disorder. Contrary to this expectation, a recent report shows that, at least in the primary auditory cortex, loss of dendritic spines in SZ may predominantly reflect decreases of smaller spines, and potentially be related to the SZ risk gene CACNB4 [Reference MacDonald, Alhassan, Newman, Richard, Gu and Kelly154]. These findings are prompting novel hypotheses on the potential causes of synaptic pathology in SZ, and its links to genetic vulnerabilities. Importantly, GWAS data and de novo CNV analyses strongly support the involvement of genes involved in synaptic plasticity, and specifically, encoding for elements of the postsynaptic density, as risk factors for SZ [Reference Ripke51, Reference Fromer, Pocklington, Kavanagh, Williams, Dwyer and Gormley156]. Among these are alleles of the complement component 4 (C4) genes which, as mentioned above, were recently shown to play a key role in synapse elimination during postnatal development [Reference Ripke51, Reference Sekar, Bialas, de Rivera, Davis, Hammond and Kamitaki52]. Additional genetic loci associated with genetic vulnerability to SZ also include genes encoding for ECM molecules and particularly ECM remodeling proteases [Reference Ripke51]. Together, these studies support the possibility that risk genes affecting synaptic functions and ECM may interact with each other and with secondary or environmental factors to affect elements of the tetrapartite synapse, ultimately resulting in disruption of synaptic functions and plasticity.

2. Methods

Methods briefly described here refer to investigations on postmortem human tissue carried out by our group.

2.1. Human subjects

Tissue blocks containing the regions of interest (e.g. the amygdala) from cohorts of normal control and SZ donors (n = 12–25/group) were used for histochemical, immunocytochemical and qRT-PCR studies. All tissue blocks were obtained from the Harvard Brain Tissue Resource Center (HBTRC), McLean Hospital, Belmont, MA, USA. Diagnoses of SZ and BD were made by two psychiatrists on the basis of retrospective review of medical records, extensive questionnaires concerning social and medical history provided by family members and neuropathological report. Cohorts did not include subjects with evidence for gross and/or macroscopic brain changes, or clinical history, consistent with cerebrovascular accident or other neurological disorders. Subjects with Braak stages III or higher [Reference Braak and Braak157] (modified Bielschowsky stain) were not included.

2.2. Tissue processing and data collection

Tissue blocks were dissected from fresh brains, lightly fixed and cryoprotected (for immunohystochemistry (IHC) only), or quickly frozen in liquid nitrogen vapor (for Western blotting and qRT-PCR), then sectioned using a freezing microtome or a cryostat. Specificity of primary antibodies for IHC was tested by immunoblotting and pre-absorption with the corresponding antigen. WFA, and a variety of antibodies raised against CSPG protein cores (e.g. aggrecan), or CSPG CS-6 sulfation patterns (CS56, 3B3) were used to label PNNs and CS-6 clusters. Procedures for protein and mRNA detection and data collection were carried out as reported previously [e.g. Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136, Reference Pantazopoulos, Woo, Lim, Lange and Berretta158].

2.3. Statistical analysis

Differences between groups relative to the main outcome measures in each of the regions examined were assessed for statistical significance using an ANCOVA stepwise linear regression process. Effect sizes were calculated according Hedges’ g. A logarithmic transformation was uniformly applied to all original values because the data were not normally distributed. Age, gender, postmortem time interval, inflammation (classified as positive or negative for inflammatory condition at time of death), hemisphere, cause of death, brain weight, exposure to alcohol, nicotine, electroconvulsive therapy, and lifetime, as well as final six months’, exposure to antipsychotic drugs, exposure to selective serotonin reuptake inhibitors classified as positive or negative for exposure, and lithium treatment were tested systematically for their effects on the main outcome measures, and included in the model if they significantly improved the model goodness of-fit (see also Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136, Reference Pantazopoulos, Woo, Lim, Lange and Berretta158]

3. Results

3.1. ECM pathology in SZ

Human postmortem studies from our group consistently show marked decreases of PNNs in people with SZ [Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136, Reference Mauney, Athanas, Pantazopoulos, Shaskan, Passeri and Berretta138, Reference Pantazopoulos, Woo, Lim, Lange and Berretta158Reference Pantazopoulos, Boyer-Boiteau, Holbrook, Jang, Hahn and Arnold160]. Our first findings showed significant reduction of PNNs labeled with the lectin Wisteria floribunda agglutinin (WFA) in the amygdala and entorhinal cortex of subjects with SZ [Reference Pantazopoulos, Woo, Lim, Lange and Berretta158]. Similar decreases of WFA-labeled PNNs were also detected in the prefrontal cortex and hippocampus, but not in the visual cortex, consistent with specific involvement of brain regions impacted in SZ [Reference Mauney, Athanas, Pantazopoulos, Shaskan, Passeri and Berretta138, Reference Pantazopolous, Sawyer, Heckers, Berretta and Markota159]. Notably, numbers of neurons predominantly associated with WFA-labeled PNNs, i.e. interneurons expressing parvalbumin [Reference Pantazopoulos, Lange, Hassinger and Berretta161Reference Hartig, Brauer and Bruckner164], were not decreased in these regions [Reference Pantazopoulos, Lange, Baldessarini and Berretta165Reference Gonzalez-Burgos, Cho and Lewis169]. This latter finding supports the idea that PNN decreases do not depend on a corresponding reduction of the neurons they envelope, while at the same time suggest that they may contribute to functional abnormalities affecting these neurons in SZ [Reference Gonzalez-Burgos, Cho and Lewis169, Reference Gonzalez-Burgos and Lewis170]. WFA labels a distal N-acetylgalactosamine on the CS chains of a group of CSPGs. Thus, decreases of WFA-labeled PNNs suggest CSPG involvement in SZ. To test the hypothesis that PNN decreases in SZ may not be restricted to WFA-positive PNNs, we focused on aggrecan, one of the main CSPGs in the brain and a major component of a population of PNNs, and on CS-6 sulfation patterns. Our results showed marked decreases of PNNs containing aggrecan and CS-6 sulfation in the amygdala of people with SZ [Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136]. Decreases of aggrecan-positive and WFA-positive PNNs were detected exclusively in the lateral nucleus of the amygdala; however, only one-third of WFA-positive PNNs also expressed aggrecan, raising the possibility that neuronal populations affected by WFA- and aggrecan-positive PNN loss only partially overlap. Decreases of CS-6 immunoreactive PNNs were much broader, impacting several amygdala nuclei, including the lateral, basal, accessory basal, cortical, and medial nuclei. Together, these findings show that PNN abnormalities in the amygdala of people with SZ include several distinct PNN phenotypes encompassing multiple neuronal populations. Such heterogeneity in PNN phenotypes is consistent with previous findings in several other brain regions [Reference Matthews, Kelly, Zerillo, Gray, Tiemeyer and Hockfield171Reference Yamada and Jinno173].

Results from these postmortem studies also show complex interactions between PNN decreases and glial cells abnormalities. In the amygdala and entorhinal cortex, WFA-positive PNN decreases were accompanied by a robust increase of WFA-positive astrocytes affecting all amygdala nuclei and entorhinal cortex subregions tested, in contrast to the more restricted PNN changes. We speculate that impaired CSPG secretion in glial cells may be causally linked to WFA-positive PNN decreases. In contrast to results with WFA, aggrecan-positive PNN decreases were accompanied by marked reductions of glial cells expressing this CSPG [Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136], suggesting decreased aggrecan supply from glial cells. Together, these findings are consistent with the hypothesis that interactions between glial cells and ECM may be disrupted in SZ. This hypothesis is also supported by our findings that CS-6/Glia clusters, labeled with two different CS-6 antibodies (CS56 and 3B3), are markedly decreased in the amygdala of SZ [Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136]. As mentioned above, CS-6/Glia clusters putatively represent a novel form of structured ECM, postulated to affect dendritic spines. Their decreases may represent a clear example of disruption of the interactions between elements of the tetrapartite synapse, i.e. glial cells, ECM and pre- and post-synaptic elements. Additional evidence for the involvement of the ECM in SZ comes from findings in the olfactory epithelium, a peripheral sensory organ where neurogenesis and axon growth occur throughout adult life [Reference Cascella, Takaki, Lin and Sawa174Reference Monti Graziadei, Karlan, Bernstein and Graziadei176]. ECM components, and CSPGs in particular, are suspected to play a key role in these functions. Postmortem findings from our group show that cytoplasmic CSPG expression is altered in olfactory receptor neurons, potentially contributing to a disruption of olfactory functions observed in people with SZ [Reference Turetsky and Moberg177Reference Atanasova, Graux, El Hage, Hommet, Camus and Belzung185].

4. Discussion

In summary, our findings show consistent ECM abnormalities in people with SZ, including reductions of PNNs, altered CSPG expression in glial cells and reduced numbers of CS-6/Glia clusters. These abnormalities are large in magnitude, robust to confounding factors and shared by several brain regions involved in SZ {[Reference Pantazopoulos, Markota, Jaquet, Ghosh, Wallin and Santos136] Pantazopoulos #1181; [Reference Mauney, Athanas, Pantazopoulos, Shaskan, Passeri and Berretta138] Mauney #7280; [Reference Pantazopoulos, Woo, Lim, Lange and Berretta158] Pantazopoulos #17815}. Yet, they are restricted in their distribution to specific cortical layers and amygdala nuclei, suggesting specialized mechanisms affecting distinct neuronal populations.

The causative mechanisms underlying ECM abnormalities are not yet well understood. However, support for the contribution of genetic factors can be found in a growing number of studies. A common variation of NCAN, encoding neurocan, a brain-specific CSPG highly represented in the brain, has been reported and replicated in several studies, including a large GWAS with more than 36000 SZ cases [Reference Ripke51, Reference Muhleisen, Mattheisen, Strohmaier, Degenhardt, Priebe and Schultz186, Reference Cichon, Muhleisen, Degenhardt, Mattheisen, Miro and Strohmaier187]. This latter study also discovered SZ risk factors from several other ECM molecules, including MMP-16 [Reference Ripke51]. Several additional studies reported the involvement of other MMPs, of which MMP-9 may be the most notable. For instance, polymorphisms of MMP-9 have been shown to be associated with SZ, although some negative findings have also been reported, and plasma levels of MMP-9 were found to be altered in this disorder [Reference Lepeta and Matrix109, Reference Yamamori, Hashimoto, Ishima, Kishi, Yasuda and Ohi188Reference Vafadari, Salamian and Kaczmarek190]. Together, these findings led to suggestions that MMPs may represent a novel therapeutic target for SZ [Reference Vafadari, Salamian and Kaczmarek190, Reference Chopra, Baveja and Kuhad191]. Notably, dose-dependent MMP-9 activity is crucially modulated by mir-132, a microRNA identified as pivotal for synaptic plasticity and found to be decreased in blood and brain tissue of people with SZ [Reference Miller, Zeier, Xi, Lanz, Deng and Strathmann192Reference Tognini and Pizzorusso195]. Together, these findings support the hypothesis of aberrant CSPGs processing during activity dependent plasticity, potentially linking CSPGs abnormalities to synaptic pathology in SZ.

4.1. The tetrapartite synapse: role in synaptic pathology in SZ

Converging evidence reviewed here and elsewhere compellingly support the idea that all elements of the tetrapartite synapse may be altered in SZ. Briefly, several molecular factors contributing to presynaptic functions, including synaptic vesicle trafficking, have been found to be disrupted in SZ [Reference Egbujo, Sinclair and Hahn196Reference Landen, Davidsson, Gottfries, Grenfeldt, Stridsberg and Blennow199]. Similarly, altered expression of PSD proteins, such as PSD95 and Homer-1, and glutamate signaling pathway, as well as proteins molecules involved in the actin cytoskeleton has been reported in people with this disorder [Reference de Bartolomeis, Latte, Tomasetti and Iasevoli148Reference Konopaske, Subburaju, Coyle and Benes152]. Emerging evidence for glial involvement in SZ indicates that all glial populations shown to affect synaptic functions, i.e. astrocytes, NG2 cells and microglia, contribute to the pathophysiology of this disorder [Reference Catts, Wong, Fillman, Fung and Shannon Weickert200Reference Duncan, Holmans, Lee, O'Dushlaine, Kirby and Smoller209]. Finally, ECM abnormalities in SZ have been reviewed above, including evidence for altered CSPG expression in glial cells [see also Reference Berretta, Pantazopoulos, Markota, Brown and Batzianouli121, Reference Pantazopoulos and Berretta210].

It is plausible to postulate that abnormalities affecting distinct elements of the tetrapartite synapse may be causally related, and/or interact with each and result in synaptic dysfunction. Although the underlying mechanisms are not yet understood, current data raises a number of possible hypotheses. For instance, altered CSPG expression in glial cells and decreased PNNs in the same subjects suggests that decreased availability of PNN molecular components may contribute to PNN abnormalities. ECM molecules bind to neuronal surface receptors (e.g. integrins), which in turn link the postsynaptic density to the actin cytoskeleton on one side, and to the ECM and pre-synaptic terminal of the other side. Through this arrangement, cell adhesion molecules (CAMs)-mediate ECM and PSD signaling may impact the dendritic spine actin network, and thus the spine shape [Reference Huntley, Gil and Bozdagi211Reference Koleske217]. This possibility is supported by evidence that several ECM molecules, including CSPGs and Reelin, known to modulate spine formation, size and stability through ECM receptors, are also implicated in the pathology of SZ [Reference Pizzorusso, Medini, Landi, Baldini, Berardi and Maffei76Reference Orlando, Ster, Gerber, Fawcett and Raineteau79, Reference Satz, Ostendorf, Hou, Turner, Kusano and Lee88Reference Trotter, Klein, Jinwal, Abisambra, Dickey and Tharkur92, Reference Beffert, Weeber, Durudas, Qiu, Masiulis and Sweatt218]. The potential contribution of decreased Reelin expression to dendritic spine decreases in SZ has long been postulated [Reference Eastwood and Harrison93, Reference Costa, Davis, Grayson, Guidotti, Pappas and Pesold219]. As reviewed above, ECM proteases such as MMPs, secreted into the ECM by astrocytes and microglia in addition to neurons, have been shown to robustly affect dendritic spine stability [Reference Levy, Omar and Koleske106].

5. Conclusions

We reviewed evidence that each element of the tetrapartite synapse plays a role in synaptic plasticity and is involved in SZ. Several glial populations, including astrocytes, NG2 cells and microglia have been shown to regulate synaptic functions and plasticity. Each of these glial cell populations has been shown to have complex, intimate relationships with the ECM. Several ECM molecules have been shown to have powerful effects on synaptic plasticity, and to be represented in ECM organized perisynaptic structures such as PNNs. Genetic and human postmortem evidence supports the involvement of the ECM in SZ, including loss of PNNs and CS-6/Glia clusters. Synaptic pathology, including loss of dendritic spines and molecular factors involved in spine structural stability and enriched in the PSD, is well established in SZ. We put forth the hypothesis that this pathology results from a disruption of interactions between elements of the tetrapartite synapse. Given the clinical, genetic and pathological heterogeneity of SZ, it is possible that synaptic pathology in specific brain regions may represent a point of convergence, potentially caused by a number of distinct molecular mechanisms in different individuals.

Conflict of interest

The authors declare that there is no conflict of interest regarding the publication of this paper.

Funding sources

This work was supported by the National Institutes of Health: R01 MH086522, R01 MH104488, R01 MH105608.

Acknowledgements

The authors thank the Harvard Brain Tissue Resource Center, funded through NIH NeuroBiobank HHSN-271-2013-00030C (The National Institute of Mental Health (NIMH), National Institute of Neurological Diseases and Stroke (NINDS) and Eunice Kennedy Shriver National Institute of Child Health and Human Development (NICHD), and brain donors and their families for the tissue samples used in these studies.

References

Huttenlocher, PRSynaptic density in human frontal cortex – developmental changes and effects of aging. Brain Res 1979;163(2):195205.Google ScholarPubMed
Huttenlocher, PRDabholkar, ASRegional differences in synaptogenesis in human cerebral cortex. J Comp Neurol 1997;387(2):167–78.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Ventura, RHarris, KMThree-dimensional relationships between hippocampal synapses and astrocytes. J Neurosci 1999;19(16):6897–906.CrossRefGoogle ScholarPubMed
Genoud, CQuairiaux, CSteiner, PHirling, HWelker, EKnott, GWPlasticity of astrocytic coverage and glutamate transporter expression in adult mouse cortex. PLoS Biol 2006;4(11):e343.CrossRefGoogle ScholarPubMed
Araque, AParpura, VSanzgiri, RPHaydon, PGTripartite synapses: glia, the unacknowledged partner. Trends Neurosci 1999;22(5):208–15.CrossRefGoogle ScholarPubMed
Dityatev, ARusakov, DAMolecular signals of plasticity at the tetrapartite synapse. Curr Opin Neurobiol 2011;21(2):353–9.CrossRefGoogle ScholarPubMed
Dityatev, AFrischknecht, RSeidenbecher, CIExtracellular matrix and synaptic functions. Results Probl Cell Differ 2006; 43:6997.CrossRefGoogle ScholarPubMed
Song, IDityatev, ACrosstalk between glia, extracellular matrix and neurons. Brain Res Bull 2018; 136:101–8.CrossRefGoogle ScholarPubMed
Nicoll, RAKauer, JAMalenka, RCThe current excitement on long-term potentiation. Neuron 1988; 1:97103.CrossRefGoogle ScholarPubMed
Malenka, RCBear, MFLTP and LTD: an embarrassment of riches. Neuron 2004;44(1):521.CrossRefGoogle ScholarPubMed
Banker, GChurchill, LCotman, CWProteins of the postsynaptic density. J Cell Biol 63(2 Pt 1)1974; 456–65.CrossRefGoogle ScholarPubMed
Granger, AJNicoll, RAExpression mechanisms underlying long-term potentiation: a postsynaptic view, 10 years on. Philos Trans R Soc Lond B Biol Sci 2014;369(1633):20130136.CrossRefGoogle Scholar
Kennedy, MBThe postsynaptic density. Curr Opin Neurobiol 1993;3(5):732–37.CrossRefGoogle ScholarPubMed
Okamoto, KNagai, TMiyawaki, AHayashi, YRapid and persistent modulation of actin dynamics regulates postsynaptic reorganization underlying bidirectional plasticity. Nat Neurosci 2004;7(10):1104–12.CrossRefGoogle ScholarPubMed
Wiegert, JSOertner, TGLong-term depression triggers the selective elimination of weakly integrated synapses. Proc Natl Acad Sci U S A 2013;110(47):E45109.CrossRefGoogle ScholarPubMed
Nagerl, UVEberhorn, NCambridge, SBBonhoeffer, TBidirectional activity-dependent morphological plasticity in hippocampal neurons. Neuron 2004;44(5):759–67.CrossRefGoogle ScholarPubMed
Zhou, QHomma, KJPoo, MMShrinkage of dendritic spines associated with long-term depression of hippocampal synapses. Neuron 2004;44(5):749–57.CrossRefGoogle ScholarPubMed
Witcher, MRKirov, SAHarris, KMPlasticity of perisynaptic astroglia during synaptogenesis in the mature rat hippocampus. Glia 2007;55(1):1323.CrossRefGoogle ScholarPubMed
Danbolt, NCGlutamate uptake. Prog Neurobiol 2001;65(1):1105.CrossRefGoogle ScholarPubMed
Halassa, MMHaydon, PGIntegrated brain circuits: astrocytic networks modulate neuronal activity and behavior. Annu Rev Physiol 2010; 72:335–55.CrossRefGoogle ScholarPubMed
Charles, ACMerrill, JEDirksen, ERSanderson, MJIntercellular signaling in glial cells: calcium waves and oscillations in response to mechanical stimulation and glutamate. Neuron 1991;6(6):983–92.CrossRefGoogle ScholarPubMed
Parpura, VZorec, RGliotransmission: exocytotic release from astrocytes. Brain Res Rev 2009.Google ScholarPubMed
Araque, ACarmignoto, GHaydon, PGOliet, SHRobitaille, RVolterra, AGliotransmitters travel in time and space. Neuron 2014;81(4):728–39.CrossRefGoogle ScholarPubMed
Harada, KKamiya, TTsuboi, TGliotransmitter release from astrocytes: functional, developmental, and pathological implications in the brain. Front Neurosci 2015; 9:499.Google Scholar
Rivera, AVanzulli, IButt, AMA central role for ATP signalling in glial interactions in the CNS. Curr Drug Targets 2016;17(16):1829–33.CrossRefGoogle ScholarPubMed
Perea, GAraque, AAstrocytes potentiate transmitter release at single hippocampal synapses. Science 2007;317(5841):1083–86.CrossRefGoogle ScholarPubMed
Perez-Alvarez, ANavarrete, MCovelo, AMartin, EDAraque, AStructural and functional plasticity of astrocyte processes and dendritic spine interactions. J Neurosci 2014;34(38):12738–44.CrossRefGoogle ScholarPubMed
Nishiyama, ALin, XHGiese, NHeldin, CHStallcup, WBInteraction between NG2 proteoglycan and PDGF alpha-receptor on O2A progenitor cells is required for optimal response to PDGF. J Neurosci Res 1996;43(3):315–30.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
Reynolds, RHardy, ROligodendroglial progenitors labeled with the O4 antibody persist in the adult rat cerebral cortex in vivo. J Neurosci Res 1997;47(5):455–70.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Stallcup, WBBeasley, LBipotential glial precursor cells of the optic nerve express the NG2 proteoglycan. J Neurosci 1987;7(9):2737–44.CrossRefGoogle ScholarPubMed
Dawson, MRLevine, JMReynolds, RNG2-expressing cells in the central nervous system: are they oligodendroglial progenitors?. J Neurosci Res 2000;61(5):471–79.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Levine, JMCard, JPLight and electron microscopic localization of a cell surface antigen (NG2) in the rat cerebellum: association with smooth protoplasmic astrocytes. J Neurosci 1987;7(9):2711–20.CrossRefGoogle ScholarPubMed
Butt, AMHamilton, NHubbard, PPugh, MIbrahim, MSynantocytes: the fifth element. J Anat 2005;207(6):695706.CrossRefGoogle ScholarPubMed
Larson, VAZhang, YBergles, DEElectrophysiological properties of NG2(+) cells: matching physiological studies with gene expression profiles. Brain Res 1638(Pt B)2016; 138–60.CrossRefGoogle ScholarPubMed
Ge, WPYang, XJZhang, ZWang, HKShen, WDeng, QD et al. Long-term potentiation of neuron-glia synapses mediated by Ca2+-permeable AMPA receptors. Science 2006;312(5779):1533–37.CrossRefGoogle ScholarPubMed
Bergles, DERoberts, JDSomogyi, PJahr, CEGlutamatergic synapses on oligodendrocyte precursor cells in the hippocampus. Nature 2000;405(6783):187–91.CrossRefGoogle ScholarPubMed
Lin, SCBergles, DESynaptic signaling between neurons and glia. Glia 2004;47(3):290–98.CrossRefGoogle ScholarPubMed
Butt, AMKiff, JHubbard, PBerry, MSynantocytes: new functions for novel NG2 expressing glia. J Neurocytol 31(6–7)2002; 551–65.CrossRefGoogle ScholarPubMed
Sakry, DNeitz, ASingh, JFrischknecht, RMarongiu, DBiname, F et al. Oligodendrocyte precursor cells modulate the neuronal network by activity-dependent ectodomain cleavage of glial NG2. PLoS Biol 2014;12(11):e1001993.CrossRefGoogle ScholarPubMed
Eugenin-von Bernhardi, JDimou, LMore than progenitor cells. Adv Exp Med Biol 2016; 949:2745.CrossRefGoogle ScholarPubMed
Sakry, DKarram, KTrotter, JSynapses between NG2 glia and neurons. J Anat 2011;219(1):27.CrossRefGoogle ScholarPubMed
Dimou, LGallo, VNG2-glia and their functions in the central nervous system. Glia 2015;63(8):1429–51.CrossRefGoogle ScholarPubMed
Hill, RANishiyama, ANG2 cells (polydendrocytes): listeners to the neural network with diverse properties. Glia 2014;62(8):1195–210.CrossRefGoogle Scholar
Colonna, MButovsky, OMicroglia function in the central nervous system during health and neurodegeneration. Annu Rev Immunol 2017; 35:441–68.CrossRefGoogle ScholarPubMed
Wu, YDissing-Olesen, LMacVicar, BAStevens, BMicroglia dynamic mediators of synapse development and plasticity. Trends Immunol 2015;36(10):605–13.CrossRefGoogle ScholarPubMed
Pocock, JMKettenmann, HNeurotransmitter receptors on microglia. Trends Neurosci 2007;30(10):527–35.CrossRefGoogle ScholarPubMed
Bialas, ARPresumey, JDas, Avan der Poel, CELapchak, PHMesin, L et al. Microglia-dependent synapse loss in type I interferon-mediated lupus. Nature 2017;546(7659):539–43.CrossRefGoogle ScholarPubMed
Paolicelli, RCBolasco, GPagani, FMaggi, LScianni, MPanzanelli, P et al. Synaptic pruning by microglia is necessary for normal brain development. Science 2011;333(6048):1456–58.CrossRefGoogle ScholarPubMed
Schafer, DPLehrman, EKStevens, BThe quad-partite synapse: microglia-synapse interactions in the developing and mature CNS. Glia 2013;61(1):2436.CrossRefGoogle ScholarPubMed
Chai, XFan, LShao, HLu, XZhang, WLi, J et al. Reelin induces branching of neurons and radial glial cells during corticogenesis. Cereb Cortex 2014.Google ScholarPubMed
Ripke, SSchizophrenia working group of the psychiatric genomics C: biological insights from 108 schizophrenia-associated genetic loci. Nature 2014;511(7510):421–27.Google Scholar
Sekar, ABialas, ARde Rivera, HDavis, AHammond, TRKamitaki, N et al. Schizophrenia risk from complex variation of complement component 4. Nature 2016;530(7589):177–83.CrossRefGoogle ScholarPubMed
Sykova, ENicholson, CDiffusion in brain extracellular space. Physiol Rev 2008;88(4):1277–340.CrossRefGoogle ScholarPubMed
Rivera, SKhrestchatisky, MKaczmarek, LRosenberg, GAJaworski, DMMetzincin proteases and their inhibitors: foes or friends in nervous system physiology?. J Neurosci 2010;30(46):15337–57.CrossRefGoogle ScholarPubMed
Muir, EMAdcock, KHMorgenstern, DAClayton, Rvon Stillfried, NRhodes, K et al. Matrix metalloproteases and their inhibitors are produced by overlapping populations of activated astrocytes. Brain Res Mol Brain Res 100(1–2)2002; 103–17.CrossRefGoogle ScholarPubMed
Abdolmaleky, HMThiagalingam, SWilcox, MGenetics and epigenetics in major psychiatric disorders: dilemmas, achievements, applications, and future scope. Am J Pharmacogenom 2005;5(3):149–60.CrossRefGoogle ScholarPubMed
Medina-Flores, RWang, GBissel, SJMurphey-Corb, MWiley, CADestruction of extracellular matrix proteoglycans is pervasive in simian retroviral neuroinfection. Neurobiol Dis 2004;16(3):604–16.CrossRefGoogle ScholarPubMed
Hobohm, CGunther, AGrosche, JRossner, SSchneider, DBruckner, GDecomposition and long-lasting downregulation of extracellular matrix in perineuronal nets induced by focal cerebral ischemia in rats. J Neurosci Res 2005;80(4):539–48.CrossRefGoogle ScholarPubMed
Yamaguchi, YLecticans: organizers of the brain extracellular matrix. Cell Mol Life Sci 2000;57(2):276–89.CrossRefGoogle ScholarPubMed
Giamanco, KAMatthews, RTDeconstructing the perineuronal net: cellular contributions and molecular composition of the neuronal extracellular matrix. Neuroscience 2012; 218:367–84.CrossRefGoogle ScholarPubMed
Deepa, SSCarulli, DGaltrey, CRhodes, KFukuda, JMikami, T et al. Composition of perineuronal net extracellular matrix in rat brain: a different disaccharide composition for the net-associated proteoglycans. J Biol Chem 2006;281(26):17789–800.CrossRefGoogle ScholarPubMed
Gati, GLendvai, D[The dress makes the neuron – different forms of the extracellular matrix in the central nervous system of vertebrates]. Orv Hetil 2013;154(27):1067–73 quiz 78-9.CrossRefGoogle Scholar
Jager, CLendvai, DSeeger, GBruckner, GMatthews, RTArendt, T et al. Perineuronal and perisynaptic extracellular matrix in the human spinal cord. Neuroscience 2013; 238:168–84.CrossRefGoogle ScholarPubMed
Lendvai, DMorawski, MNegyessy, LGati, GJager, CBaksa, G et al. Neurochemical mapping of the human hippocampus reveals perisynaptic matrix around functional synapses in Alzheimer's disease. Acta Neuropathol 2013;125(2):215–29.CrossRefGoogle ScholarPubMed
Karus, MSamtleben, SBusse, CTsai, TDietzel, IDFaissner, A et al. Normal Sulphation levels regulate spinal cord neural precursor cell proliferation and differentiation. Neural Dev 2012;7(1):20.CrossRefGoogle Scholar
Miyata, SKomatsu, YYoshimura, YTaya, CKitagawa, HPersistent cortical plasticity by upregulation of chondroitin 6-sulfation. Nat Neurosci 2012;15(3):414–22 S1-2.CrossRefGoogle ScholarPubMed
Wang, HKatagiri, YMcCann, TEUnsworth, EGoldsmith, PYu, ZX et al. Chondroitin-4-sulfation negatively regulates axonal guidance and growth. J Cell Sci 121(Pt 18)2008; 3083–91.CrossRefGoogle ScholarPubMed
Maeda, NStructural variation of chondroitin sulfate and its roles in the central nervous system. Cent Nerv Syst Agents Med Chem 2010;10(1):2231.CrossRefGoogle ScholarPubMed
Bukalo, OSchachner, MDityatev, AModification of extracellular matrix by enzymatic removal of chondroitin sulfate and by lack of tenascin-R differentially affects several forms of synaptic plasticity in the hippocampus. Neuroscience 2001;104(2):359–69.CrossRefGoogle ScholarPubMed
Niisato, KFujikawa, AKomai, SShintani, TWatanabe, ESakaguchi, G et al. Age-dependent enhancement of hippocampal long-term potentiation and impairment of spatial learning through the Rho-associated kinase pathway in protein tyrosine phosphatase receptor type Z-deficient mice. J Neurosci 2005;25(5):1081–88.CrossRefGoogle ScholarPubMed
Kawachi, HTamura, HWatakabe, IShintani, TMaeda, NNoda, MProtein tyrosine phosphatase zeta/RPTPbeta interacts with PSD-95/SAP90 family. Brain Res Mol Brain Res 1999;72(1):4754.CrossRefGoogle ScholarPubMed
Snyder, SELi, JSchauwecker, PEMcNeill, THSalton, SRComparison of RPTP zeta/beta, phosphacan, and trkB mRNA expression in the developing and adult rat nervous system and induction of RPTP zeta/beta and phosphacan mRNA following brain injury. Brain Res Mol Brain Res 1996;40(1):7996.CrossRefGoogle ScholarPubMed
Nishiwaki, TMaeda, NNoda, MCharacterization and developmental regulation of proteoglycan-type protein tyrosine phosphatase zeta/RPTPbeta isoforms. J Biochem 1998;123(3):458–67.CrossRefGoogle ScholarPubMed
Shintani, TWatanabe, EMaeda, NNoda, MNeurons as well as astrocytes express proteoglycan-type protein tyrosine phosphatase zeta/RPTPbeta: analysis of mice in which the PTPzeta/RPTPbeta gene was replaced with the LacZ gene. Neurosci Lett 247(2–3)1998; 135–38.CrossRefGoogle ScholarPubMed
Brakebusch, CSeidenbecher, CIAsztely, FRauch, UMatthies, HMeyer, H et al. Brevican-deficient mice display impaired hippocampal CA1 long-term potentiation but show no obvious deficits in learning and memory. Mol Cell Biol 2002;22(21):7417–27.CrossRefGoogle ScholarPubMed
Pizzorusso, TMedini, PLandi, SBaldini, SBerardi, NMaffei, LStructural and functional recovery from early monocular deprivation in adult rats. Proc Natl Acad Sci U S A 2006;103(22):8517–22.CrossRefGoogle ScholarPubMed
Majewska, ASur, MMotility of dendritic spines in visual cortex in vivo: changes during the critical period and effects of visual deprivation. Proc Natl Acad Sci U S A 2003;100(26):16024–29.CrossRefGoogle ScholarPubMed
de Vivo, LLandi, SPanniello, MBaroncelli, LChierzi, SMariotti, L et al. Extracellular matrix inhibits structural and functional plasticity of dendritic spines in the adult visual cortex. Nat Commun 2013; 4:1484.CrossRefGoogle ScholarPubMed
Orlando, CSter, JGerber, UFawcett, JWRaineteau, OPerisynaptic chondroitin sulfate proteoglycans restrict structural plasticity in an integrin-dependent manner. J Neurosci 2012;32(50):18009–17 17a.CrossRefGoogle Scholar
Kochlamazashvili, GHenneberger, CBukalo, ODvoretskova, ESenkov, OLievens, PM et al. The extracellular matrix molecule hyaluronic acid regulates hippocampal synaptic plasticity by modulating postsynaptic L-type Ca(2+) channels. Neuron 2010;67(1):116–28.CrossRefGoogle ScholarPubMed
D'Arcangelo, GHomayouni, RKeshvara, LRice, DSSheldon, MCurran, TReelin is a ligand for lipoprotein receptors. Neuron 1999;24(2):471–79.CrossRefGoogle ScholarPubMed
Herz, JChen, YReelin, lipoprotein receptors and synaptic plasticity. Nat Rev Neurosci 2006;7(11):850–59.CrossRefGoogle ScholarPubMed
Dulabon, LOlson, ECTaglienti, MGEisenhuth, SMcGrath, BWalsh, CA et al. Reelin binds alpha3beta1 integrin and inhibits neuronal migration. Neuron 2000;27(1):3344.CrossRefGoogle ScholarPubMed
Chen, YBeffert, UErtunc, MTang, TSKavalali, ETBezprozvanny, I et al. Reelin modulates NMDA receptor activity in cortical neurons. J Neurosci 2005;25(36):8209–16.CrossRefGoogle ScholarPubMed
Qiu, SZhao, LFKorwek, KMWeeber, EJDifferential reelin-induced enhancement of NMDA and AMPA receptor activity in the adult hippocampus. J Neurosci 2006;26(50):12943–55.CrossRefGoogle ScholarPubMed
Campo, CGSinagra, MVerrier, DManzoni, OJChavis, PReelin secreted by GABAergic neurons regulates glutamate receptor homeostasis. PLoS One 2009;4(5):e5505.CrossRefGoogle ScholarPubMed
Hellwig, SHack, IKowalski, JBrunne, BJarowyj, JUnger, A et al. Role for Reelin in neurotransmitter release. J Neurosci 2011;31(7):2352–60.CrossRefGoogle ScholarPubMed
Pujadas, LGruart, ABosch, CDelgado, LTeixeira, CMRossi, D et al. Reelin regulates postnatal neurogenesis and enhances spine hypertrophy and long-term potentiation. J Neurosci 2010;30(13):4636–49.CrossRefGoogle ScholarPubMed
Pribiag, HPeng, HShah, WAStellwagen, DCarbonetto, SDystroglycan mediates homeostatic synaptic plasticity at GABAergic synapses. Proc Natl Acad Sci U S A 2014;111(18):6810–15.CrossRefGoogle ScholarPubMed
Satz, JSOstendorf, APHou, STurner, AKusano, HLee, JC et al. Distinct functions of glial and neuronal dystroglycan in the developing and adult mouse brain. J Neurosci 2010;30(43):14560–72.CrossRefGoogle ScholarPubMed
Beffert, UDurudas, AWeeber, EJStolt, PCGiehl, KMSweatt, JD et al. Functional dissection of Reelin signaling by site-directed disruption of Disabled-1 adaptor binding to apolipoprotein E receptor 2: distinct roles in development and synaptic plasticity. J Neurosci 2006;26(7):2041–52.CrossRefGoogle ScholarPubMed
Trotter, JHKlein, MJinwal, UKAbisambra, JFDickey, CATharkur, J et al. ApoER2 function in the establishment and maintenance of retinal synaptic connectivity. J Neurosci 2011;31(40):14413–23.CrossRefGoogle ScholarPubMed
Eastwood, SLHarrison, PJCellular basis of reduced cortical reelin expression in schizophrenia. Am J Psychiatry 2006;163(3):540–42.CrossRefGoogle Scholar
Dityatev, ASchachner, MSonderegger, PThe dual role of the extracellular matrix in synaptic plasticity and homeostasis. Nat Rev Neurosci 2010;11(11):735–46.CrossRefGoogle ScholarPubMed
Pozo, KGoda, YUnraveling mechanisms of homeostatic synaptic plasticity. Neuron 2010;66(3):337–51.CrossRefGoogle ScholarPubMed
Charrier, CMachado, PTweedie-Cullen, RYRutishauser, DMansuy, IMTriller, AA crosstalk between beta1 and beta3 integrins controls glycine receptor and gephyrin trafficking at synapses. Nat Neurosci 2010;13(11):1388–95.CrossRefGoogle ScholarPubMed
McGeachie, ABCingolani, LAGoda, YStabilising influence: integrins in regulation of synaptic plasticity. Neurosci Res 2011;70(1):2429.CrossRefGoogle ScholarPubMed
Babayan, AHKramar, EABarrett, RMJafari, MHaettig, JChen, LY et al. Integrin dynamics produce a delayed stage of long-term potentiation and memory consolidation. J Neurosci 2012;32(37):12854–61.CrossRefGoogle ScholarPubMed
Chan, CSChen, HBradley, ADragatsis, IRosenmund, CDavis, RLalpha8-integrins are required for hippocampal long-term potentiation but not for hippocampal-dependent learning. Genes Brain Behav 2010;9(4):402–10.CrossRefGoogle Scholar
Huntley, GWSynaptic circuit remodelling by matrix metalloproteinases in health and disease. Nat Rev Neurosci 2012;13(11):743–57.CrossRefGoogle ScholarPubMed
Iryna, MEthell, DWEMatrix metalloproteinases in brain development and remodeling: synaptic functions and targets. J Neurosci Res 2007;9999(9999):NA.Google Scholar
Smith, ACScofield, MDKalivas, PWThe tetrapartite synapse: extracellular matrix remodeling contributes to corticoaccumbens plasticity underlying drug addiction. Brain Res 2015.CrossRefGoogle ScholarPubMed
Verslegers, MLemmens, KVan Hove, IMoons, LMatrix metalloproteinase-2 and −9 as promising benefactors in development, plasticity and repair of the nervous system. Prog Neurobiol 2013; 105:6078.CrossRefGoogle Scholar
Michaluk, PWawrzyniak, MAlot, PSzczot, MWyrembek, PMercik, K et al. Influence of matrix metalloproteinase MMP-9 on dendritic spine morphology. J Cell Sci 124(Pt 19)2011; 3369–80.CrossRefGoogle ScholarPubMed
Bilousova, TVRusakov, DAEthell, DWEthell, IMMatrix metalloproteinase-7 disrupts dendritic spines in hippocampal neurons through NMDA receptor activation. J Neurochem 2006;97(1):4456.CrossRefGoogle ScholarPubMed
Levy, ADOmar, MHKoleske, AJExtracellular matrix control of dendritic spine and synapse structure and plasticity in adulthood. Front Neuroanat 2014; 8:116.CrossRefGoogle ScholarPubMed
Nagy, VBozdagi, OMatynia, ABalcerzyk, MOkulski, PDzwonek, J et al. Matrix metalloproteinase-9 is required for hippocampal late-phase long-term potentiation and memory. J Neurosci 2006;26(7):1923–34.CrossRefGoogle ScholarPubMed
Wang, XBBozdagi, ONikitczuk, JSZhai, ZWZhou, QHuntley, GWExtracellular proteolysis by matrix metalloproteinase-9 drives dendritic spine enlargement and long-term potentiation coordinately. Proc Natl Acad Sci U S A 2008;105(49):19520–25.CrossRefGoogle ScholarPubMed
Lepeta, KMatrix, Kaczmarek L.Metalloproteinase-9 as a novel player in synaptic plasticity and schizophrenia. Schizophr Bull 2015.CrossRefGoogle Scholar
Rossier, JBernard, ACabungcal, JHPerrenoud, QSavoye, AGallopin, T et al. Cortical fast-spiking parvalbumin interneurons enwrapped in the perineuronal net express the metallopeptidases Adamts8, Adamts15 and Neprilysin. Mol Psychiatry 2015;20(2):154–61.CrossRefGoogle ScholarPubMed
Fawcett, JWThe extracellular matrix in plasticity and regeneration after CNS injury and neurodegenerative disease. Prog Brain Res 2015; 218:213–26.CrossRefGoogle ScholarPubMed
Uesaka, NUchigashima, MMikuni, TNakazawa, TNakao, HHirai, H et al. Retrograde semaphorin signaling regulates synapse elimination in the developing mouse brain. Science 2014;344(6187):1020–23.CrossRefGoogle ScholarPubMed
Carulli, DFoscarin, SFaralli, APajaj, ERossi, FModulation of semaphorin3A in perineuronal nets during structural plasticity in the adult cerebellum. Mol Cell Neurosci 2013; 57:1022.CrossRefGoogle ScholarPubMed
Dick, GTan, CLAlves, JNEhlert, EMMiller, GMHsieh-Wilson, LC et al. Semaphorin 3A binds to the perineuronal nets via chondroitin sulfate type E motifs in rodent brains. J Biol Chem 2013;288(38):27384–95.CrossRefGoogle Scholar
Wegner, FHartig, WBringmann, AGrosche, JWohlfarth, KZuschratter, W et al. Diffuse perineuronal nets and modified pyramidal cells immunoreactive for glutamate and the GABA(A) receptor alpha1 subunit form a unique entity in rat cerebral cortex. Exp Neurol 2003;184(2):705–14.CrossRefGoogle Scholar
Ajmo, JMEakin, AKHamel, MGGottschall, PEDiscordant localization of WFA reactivity and brevican/ADAMTS-derived fragment in rodent brain. BMC Neurosci 2008; 9:14.CrossRefGoogle ScholarPubMed
Bruckner, GBrauer, KHartig, WWolff, JRRickmann, MJDerouiche, A et al. Perineuronal nets provide a polyanionic, glia-associated form of microenvironment around certain neurons in many parts of the rat brain. Glia 1993;8(3):183200.CrossRefGoogle ScholarPubMed
Hartig, WBruckner, GBrauer, KSchmidt, CBigl, VAllocation of perineuronal nets and parvalbumin-, calbindin- D28k- and glutamic acid decarboxylase-immunoreactivity in the amygdala of the rhesus monkey. Brain Res 698(1–2)1995; 265–9.CrossRefGoogle ScholarPubMed
Pantazopoulos, HLange, NHassinger, LBerretta, SSubpopulations of neurons expressing parvalbumin in the human amygdala. J Comp Neurol 2006;496(5):706–22.CrossRefGoogle ScholarPubMed
Seeger, GBrauer, KHartig, WBruckner, GMapping of perineuronal nets in the rat brain stained by colloidal iron hydroxide histochemistry and lectin cytochemistry. Neuroscience 1994;58(2):371–88.CrossRefGoogle ScholarPubMed
Berretta, SPantazopoulos, HMarkota, MBrown, CBatzianouli, ETLosing the sugar coating: potential impact of perineuronal net abnormalities on interneurons in schizophrenia. Schizophr Res 2015.CrossRefGoogle Scholar
Sorg, BABerretta, SBlacktop, JMFawcett, JWKitagawa, HKwok, JC et al. Casting a wide net: role of perineuronal nets in neural plasticity. J Neurosci 2016;36(45):11459–68.CrossRefGoogle ScholarPubMed
Banerjee, SBGutzeit, VABaman, JAoued, HSDoshi, NKLiu, RC et al. Perineuronal nets in the adult sensory cortex are necessary for fear learning. Neuron 2017;95(1):169–79 e3.CrossRefGoogle ScholarPubMed
Bernard, CProchiantz, AOtx2-PNN interaction to regulate cortical plasticity. Neural Plast 2016; 2016:7931693.CrossRefGoogle ScholarPubMed
Frischknecht, RChang, KJRasband, MNSeidenbecher, CINeural ECM molecules in axonal and synaptic homeostatic plasticity. Prog Brain Res 2014; 214:81100.CrossRefGoogle ScholarPubMed
Romberg, CYang, SMelani, RAndrews, MRHorner, AESpillantini, MG et al. Depletion of perineuronal nets enhances recognition memory and long-term depression in the perirhinal cortex. J Neurosci 2013;33(16):7057–65.CrossRefGoogle ScholarPubMed
Wang, DFawcett, JThe perineuronal net and the control of CNS plasticity. Cell Tissue Res 2012;349(1):147–60.CrossRefGoogle ScholarPubMed
Maeda, NFukazawa, NIshii, MChondroitin sulfate proteoglycans in neural development and plasticity. Front Biosci 2010; 15:626–44.CrossRefGoogle ScholarPubMed
Gogolla, NCaroni, PLuthi, AHerry, CPerineuronal nets protect fear memories from erasure. Science 2009;325(5945):1258–61.CrossRefGoogle ScholarPubMed
Koppe, GBruckner, GBrauer, KHartig, WBigl, VDevelopmental patterns of proteoglycan-containing extracellular matrix in perineuronal nets and neuropil of the postnatal rat brain. Cell Tissue Res 1997;288(1):3341.Google ScholarPubMed
Ye, QMiao, QLExperience-dependent development of perineuronal nets and chondroitin sulfate proteoglycan receptors in mouse visual cortex. Matrix Biol 2013;32(6):352–63.CrossRefGoogle ScholarPubMed
Mix, AHoppenrath, KFunke, KReduction in cortical parvalbumin expression due to intermittent theta-burst stimulation correlates with maturation of the perineuronal nets in young rats. Dev Neurobiol 2015;75(1):111.CrossRefGoogle ScholarPubMed
Favuzzi, EMarques-Smith, ADeogracias, RWinterflood, CMSanchez-Aguilera, AMantoan, L et al. Activity-dependent gating of parvalbumin interneuron function by the perineuronal net protein Brevican. Neuron 2017;95(3):639–55 e10.CrossRefGoogle ScholarPubMed
Pizzorusso, TMedini, PBerardi, NChierzi, SFawcett, JWMaffei, LReactivation of ocular dominance plasticity in the adult visual cortex. Science 2002;298(5596):1248–51.CrossRefGoogle ScholarPubMed
Hayashi, NTatsumi, KOkuda, HYoshikawa, MIshizaka, SMiyata, S et al. DACS, novel matrix structure composed of chondroitin sulfate proteoglycan in the brain. Biochem Biophys Res Commun 2007;364(2):410–15.CrossRefGoogle Scholar
Pantazopoulos, HMarkota, MJaquet, FGhosh, DWallin, ASantos, A et al. Aggrecan and chondroitin-6-sulfate abnormalities in schizophrenia and bipolar disorder: a postmortem study on the amygdala. Transl. Psychiatry 2015; 5:e496.CrossRefGoogle Scholar
Horii-Hayashi, NTatsumi, KMatsusue, YOkuda, HOkuda, AHayashi, M et al. Chondroitin sulfate demarcates astrocytic territories in the mammalian cerebral cortex. Neurosci Lett 2010;483(1):6772.CrossRefGoogle ScholarPubMed
Mauney, SAAthanas, KMPantazopoulos, HShaskan, NPasseri, EBerretta, S et al. Developmental pattern of perineuronal nets in the human prefrontal cortex and their deficit in schizophrenia. Biol Psychiatry 2013;74(6):427–35.CrossRefGoogle Scholar
Yamada, JJinno, SSpatio-temporal differences in perineuronal net expression in the mouse hippocampus, with reference to parvalbumin. Neuroscience 2013; 253:368–79.CrossRefGoogle ScholarPubMed
Bavelier, DLevi, DMLi, RWDan, YHensch, TKRemoving brakes on adult brain plasticity: from molecular to behavioral interventions. J Neurosci 2010;30(45):14964–71.CrossRefGoogle ScholarPubMed
Takesian, AEHensch, TKBalancing plasticity/stability across brain development. Prog Brain Res 2013; 207:334.CrossRefGoogle ScholarPubMed
Matuszko, GCurreli, SKaushik, RBecker, ADityatev, AExtracellular matrix alterations in the ketamine model of schizophrenia. Neuroscience 2017; 350:1322.CrossRefGoogle ScholarPubMed
Glantz, LALewis, DADecreased dendritic spine density on prefrontal cortical pyramidal neurons in schizophrenia. Arch Gen Psychiatry 2000;57(1):6573.CrossRefGoogle Scholar
Konopaske, GTLange, NCoyle, JTBenes, FMPrefrontal cortical dendritic spine pathology in schizophrenia and bipolar disorder. JAMA Psychiatry 2014;71(12):1323–31.CrossRefGoogle ScholarPubMed
Sweet, RAHenteleff, RAZhang, WSampson, ARLewis, DAReduced dendritic spine density in auditory cortex of subjects with schizophrenia. Neuropsychopharmacology 2009;34(2):374–89.CrossRefGoogle ScholarPubMed
Law, AJWeickert, CSHyde, TMKleinman, JEHarrison, PJReduced spinophilin but not microtubule-associated protein 2 expression in the hippocampal formation in schizophrenia and mood disorders: molecular evidence for a pathology of dendritic spines. Am J Psychiatry 2004;161(10):1848–55.CrossRefGoogle Scholar
Kolomeets, NSOrlovskaya, DDRachmanova, VIUranova, NAUltrastructural alterations in hippocampal mossy fiber synapses in schizophrenia: a postmortem morphometric study. Synapse 2005;57(1):4755.CrossRefGoogle ScholarPubMed
de Bartolomeis, ALatte, GTomasetti, CIasevoli, FGlutamatergic postsynaptic density protein dysfunctions in synaptic plasticity and dendritic spines morphology: relevance to schizophrenia and other behavioral disorders pathophysiology, and implications for novel therapeutic approaches. Mol Neurobiol 2014;49(1):484511.CrossRefGoogle ScholarPubMed
MacDonald, MLDing, YNewman, JHemby, SPenzes, PLewis, DA et al. Altered glutamate protein co-expression network topology linked to spine loss in the auditory cortex of schizophrenia. Biol Psychiatry 2015;77(11):959–68.CrossRefGoogle ScholarPubMed
Datta, DArion, DCorradi, JPLewis, DAAltered expression of CDC42 signaling pathway components in cortical layer 3 pyramidal cells in schizophrenia. Biol Psychiatry 2015;78(11):775–85.CrossRefGoogle Scholar
Shelton, MANewman, JTGu, HSampson, ARFish, KNMacDonald, ML et al. Loss of microtubule-associated protein 2 immunoreactivity linked to dendritic spine loss in schizophrenia. Biol Psychiatry 2015.CrossRefGoogle Scholar
Konopaske, GTSubburaju, SCoyle, JTBenes, FMAltered prefrontal cortical MARCKS and PPP1R9A mRNA expression in schizophrenia and bipolar disorder. Schizophr Res 164(1–3)2015; 100–8.CrossRefGoogle ScholarPubMed
Beach, PAHuck, JTZhu, DCBozoki, ACAltered behavioral and autonomic pain responses in Alzheimer's disease are associated with dysfunctional affective, self-reflective and salience network resting-state connectivity. Front Aging Neurosci 2017; 9:297.CrossRefGoogle ScholarPubMed
MacDonald, MLAlhassan, JNewman, JTRichard, MGu, HKelly, RM et al. Selective loss of smaller spines in schizophrenia. Am J Psychiatry 2017;174(6):586–94.CrossRefGoogle Scholar
Feinberg, ISchizophrenia: caused by a fault in programmed synaptic elimination during adolescence?. J Psychiatr Res 1982;17(4):319–34.CrossRefGoogle ScholarPubMed
Fromer, MPocklington, AJKavanagh, DHWilliams, HJDwyer, SGormley, P et al. De novo mutations in schizophrenia implicate synaptic networks. Nature 2014;506(7487):179–84.CrossRefGoogle ScholarPubMed
Braak, HBraak, EDiagnostic criteria for neuropathologic assessment of Alzheimer's disease. Neurobiol Aging 18(4 Suppl)1997; S858.CrossRefGoogle ScholarPubMed
Pantazopoulos, HWoo, T-UWLim, MPLange, NBerretta, SExtracellular matrix-glial abnormalities in the amygdala and entorhinal cortex of subjects diagnosed with schizophrenia. Arch Gen Psychiatry 2010;67(2):155–66.CrossRefGoogle ScholarPubMed
Pantazopolous, HSawyer, CHeckers, SBerretta, SMarkota, MChondroitin sulfate proteoglycan abnormalities in the hippocampus of subjects with schizophrenia. Neuropsychopharmacology 2014; 39:S2989.Google Scholar
Pantazopoulos, HBoyer-Boiteau, AHolbrook, EHJang, WHahn, CGArnold, SE et al. Proteoglycan abnormalities in olfactory epithelium tissue from subjects diagnosed with schizophrenia. Schizophr Res 150(2–3)2013; 366–72.CrossRefGoogle ScholarPubMed
Pantazopoulos, HLange, NHassinger, LBerretta, SSubpopulations of neurons expressing parvalbumin in the human amygdala. J Comp Neurol 2006;496(5):706–22.CrossRefGoogle ScholarPubMed
Hartig, WBrauer, KBigl, VBruckner, GChondroitin sulfate proteoglycan-immunoreactivity of lectin-labeled perineuronal nets around parvalbumin-containing neurons. Brain Res 635(1–2)1994; 307–11.CrossRefGoogle ScholarPubMed
Brauer, KHartig, WBigl, VBruckner, GDistribution of parvalbumin-containing neurons and lectin-binding perineuronal nets in the rat basal forebrain. Brain Res 1993;631(1):167–70.CrossRefGoogle ScholarPubMed
Hartig, WBrauer, KBruckner, GWisteria floribunda agglutinin-labelled nets surround parvalbumin-containing neurons. Neuroreport 1992;3(10):869–72.CrossRefGoogle ScholarPubMed
Pantazopoulos, HLange, NBaldessarini, RJBerretta, SParvalbumin neurons in the entorhinal cortex of subjects diagnosed with bipolar disorder or schizophrenia. Biol Psychiatry 2007;61(5):640–52.CrossRefGoogle ScholarPubMed
Konradi, CZimmerman, EIYang, CKLohmann, KMGresch, PPantazopoulos, H et al. Hippocampal interneurons in bipolar disorder. Arch Gen Psychiatry 2011;68(4):340–50.CrossRefGoogle ScholarPubMed
Wang, AYLohmann, KMYang, CKZimmerman, EIPantazopoulos, HHerring, N et al. 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(5):615–26.CrossRefGoogle ScholarPubMed
Chung, DWFish, KNLewis, DAPathological basis for deficient excitatory drive to cortical parvalbumin interneurons in schizophrenia. Am J Psychiatry 2016;173(11):1131–39.CrossRefGoogle Scholar
Gonzalez-Burgos, GCho, RYLewis, DAAlterations in cortical network oscillations and parvalbumin neurons in schizophrenia. Biol Psychiatry 2015;77(12):1031–40.CrossRefGoogle Scholar
Gonzalez-Burgos, GLewis, DANMDA receptor hypofunction, parvalbumin-positive neurons, and cortical gamma oscillations in schizophrenia. Schizophr Bull 2012;38(5):950–57.CrossRefGoogle Scholar
Matthews, RTKelly, GMZerillo, CAGray, GTiemeyer, MHockfield, SAggrecan glycoforms contribute to the molecular heterogeneity of perineuronal nets. J Neurosci 2002;22(17):7536–47.CrossRefGoogle Scholar
Racz, EGaal, BMatesz, CHeterogeneous expression of extracellular matrix molecules in the red nucleus of the rat. Neuroscience 2016; 322:117.CrossRefGoogle ScholarPubMed
Yamada, JJinno, SMolecular heterogeneity of aggrecan-based perineuronal nets around five subclasses of parvalbumin-expressing neurons in the mouse hippocampus. J Comp Neurol 2017;525(5):1234–49.CrossRefGoogle ScholarPubMed
Cascella, NGTakaki, MLin, SSawa, ANeurodevelopmental involvement in schizophrenia: the olfactory epithelium as an alternative model for research. J Neurochem 2007;102(3):587–94.CrossRefGoogle ScholarPubMed
Graziadei, PPMonti Graziadei, GANeurogenesis and neuron regeneration in the olfactory system of mammals. III. Deafferentation and reinnervation of the olfactory bulb following section of the fila olfactoria in rat. J Neurocytol 1980;9(2):145–62.CrossRefGoogle ScholarPubMed
Monti Graziadei, GAKarlan, MSBernstein, JJGraziadei, PPReinnervation of the olfactory bulb after section of the olfactory nerve in monkey (Saimiri sciureus). Brain Res 1980;189(2):343–54.CrossRefGoogle Scholar
Turetsky, BIMoberg, PJAn odor-specific threshold deficit implicates abnormal intracellular cyclic AMP signaling in schizophrenia. Am J Psychiatry 2009;166(2):226–33.CrossRefGoogle Scholar
Brewer, WJPantelis, CAnderson, VVelakoulis, DSingh, BCopolov, DL et al. Stability of olfactory identification deficits in neuroleptic-naive patients with first-episode psychosis. Am J Psychiatry 2001;158(1):107–15.CrossRefGoogle ScholarPubMed
Coleman, EGoetz, RRLeitman, DYale, SStanford, AGorman, JM et al. Odor identification impairments in schizophrenia: relationship with demographic measures, clinical variables, and diagnostic subtypes. CNS Spectr 2002;7(1):4348.CrossRefGoogle ScholarPubMed
Good, KPWhitehorn, DRui, QMilliken, HKopala, LCOlfactory identification deficits in first-episode psychosis may predict patients at risk for persistent negative and disorganized or cognitive symptoms. Am J Psychiatry 2006;163(5):932–33.CrossRefGoogle ScholarPubMed
Kohler, CGMoberg, PJGur, REO'Connor, MJSperling, MRDoty, RLOlfactory dysfunction in schizophrenia and temporal lobe epilepsy. Neuropsychiatry Neuropsychol Behav Neurol 2001;14(2):8388.Google ScholarPubMed
Kopala, LCClark, CHurwitz, TOlfactory deficits in neuroleptic naive patients with schizophrenia. Schizophr Res 1993;8(3):245–50.CrossRefGoogle ScholarPubMed
Moberg, PJArnold, SEDoty, RLGur, REBalderston, CCRoalf, DR et al. Olfactory functioning in schizophrenia: relationship to clinical, neuropsychological, and volumetric MRI measures. J Clin Exp Neuropsychol 2006;28(8):1444–61.CrossRefGoogle ScholarPubMed
Seidman, LJTalbot, NLKalinowski, AGMcCarley, RWFaraone, SVKremen, WS et al. Neuropsychological probes of fronto-limbic system dysfunction in schizophrenia: olfactory identification and Wisconsin Card Sorting performance. Schizophr Res 1991;6(1):5565.CrossRefGoogle ScholarPubMed
Atanasova, BGraux, JEl Hage, WHommet, CCamus, VBelzung, COlfaction: a potential cognitive marker of psychiatric disorders. Neurosci Biobehav Rev 2008;32(7):1315–25.CrossRefGoogle ScholarPubMed
Muhleisen, TWMattheisen, MStrohmaier, JDegenhardt, FPriebe, LSchultz, CC et al. Association between schizophrenia and common variation in neurocan (NCAN), a genetic risk factor for bipolar disorder. Schizophr Res 2012;138(1):6973.CrossRefGoogle Scholar
Cichon, SMuhleisen, TWDegenhardt, FAMattheisen, MMiro, XStrohmaier, J et al. Genome-wide association study identifies genetic variation in neurocan as a susceptibility factor for bipolar disorder. Am J Hum Genet 2011;88(3):372–81.CrossRefGoogle ScholarPubMed
Yamamori, HHashimoto, RIshima, TKishi, FYasuda, YOhi, K et al. Plasma levels of mature brain-derived neurotrophic factor (BDNF) and matrix metalloproteinase-9 (MMP-9) in treatment-resistant schizophrenia treated with clozapine. Neurosci Lett 2013; 556:3741.CrossRefGoogle ScholarPubMed
Bienkowski, PSamochowiec, JPelka-Wysiecka, JGrzywacz, ASkibinska, MJasiewicz, A et al. Functional polymorphism of matrix metalloproteinase-9 (MMP9) gene is not associated with schizophrenia and with its deficit subtype. Pharmacol Rep 2015;67(3):442–5.CrossRefGoogle Scholar
Vafadari, BSalamian, AKaczmarek, LMMP-9 in translation: from molecule to brain physiology, pathology, and therapy. J Neurochem 139(Suppl. 2)2016; 91114.CrossRefGoogle Scholar
Chopra, KBaveja, AKuhad, AMMPs: a novel drug target for schizophrenia. Expert Opin Ther Targets 2015;19(1):7785.CrossRefGoogle Scholar
Miller, BHZeier, ZXi, LLanz, TADeng, SStrathmann, J et al. MicroRNA-132 dysregulation in schizophrenia has implications for both neurodevelopment and adult brain function. Proc Natl Acad Sci U S A 2012;109(8):3125–30.CrossRefGoogle ScholarPubMed
Jasinska, MMilek, JCymerman, IALeski, SKaczmarek, LDziembowska, MmiR-132 regulates dendritic spine structure by direct targeting of matrix metalloproteinase 9 mRNA. Mol Neurobiol 2016;53(7):4701–12.CrossRefGoogle ScholarPubMed
Mazziotti, RBaroncelli, LCeglia, NChelini, GSala, GDMagnan, C et al. Mir-132/212 is required for maturation of binocular matching of orientation preference and depth perception. Nat Commun 2017; 8:15488.CrossRefGoogle ScholarPubMed
Tognini, PPizzorusso, TMicroRNA212/132 family: molecular transducer of neuronal function and plasticity. Int J Biochem Cell Biol 2012;44(1):610.CrossRefGoogle Scholar
Egbujo, CNSinclair, DHahn, CGDysregulations of synaptic vesicle trafficking in schizophrenia. Curr Psychiatry Rep 2016;18(8):77.CrossRefGoogle Scholar
Halim, NDWeickert, CSMcClintock, BWHyde, TMWeinberger, DRKleinman, JE et al. Presynaptic proteins in the prefrontal cortex of patients with schizophrenia and rats with abnormal prefrontal development. Mol Psychiatry 2003;8(9):797810.CrossRefGoogle ScholarPubMed
Siegert, SSeo, JKwon, EJRudenko, ACho, SWang, W et al. The schizophrenia risk gene product miR-137 alters presynaptic plasticity. Nat Neurosci 2015;18(7):1008–16.CrossRefGoogle ScholarPubMed
Landen, MDavidsson, PGottfries, CGGrenfeldt, BStridsberg, MBlennow, KReduction of the small synaptic vesicle protein synaptophysin but not the large dense core chromogranins in the left thalamus of subjects with schizophrenia. Biol Psychiatry 1999;46(12):1698–702.CrossRefGoogle Scholar
Catts, VSWong, JFillman, SGFung, SJShannon Weickert, CIncreased expression of astrocyte markers in schizophrenia: association with neuroinflammation. Aust N Z J Psychiatry 2014;48(8):722–34.CrossRefGoogle ScholarPubMed
Katsel, PByne, WRoussos, PTan, WSiever, LHaroutunian, VAstrocyte and glutamate markers in the superficial, deep, and white matter layers of the anterior cingulate gyrus in schizophrenia. Neuropsychopharmacology 2011;36(6):1171–77.CrossRefGoogle Scholar
Mauney, SAPietersen, CYSonntag, KCWoo, TWDifferentiation of oligodendrocyte precursors is impaired in the prefrontal cortex in schizophrenia. Schizophr Res 169(1–3)2015; 374–80.CrossRefGoogle Scholar
Monji, AKato, TAMizoguchi, YHorikawa, HSeki, YKasai, M et al. Neuroinflammation in schizophrenia especially focused on the role of microglia. Prog Neuropsychopharmacol Biol Psychiatry 2013; 42:115–21.CrossRefGoogle ScholarPubMed
Trepanier, MOHopperton, KEMizrahi, RMechawar, NBazinet, RPPostmortem evidence of cerebral inflammation in schizophrenia: a systematic review. Mol Psychiatry 2016.CrossRefGoogle ScholarPubMed
Weickert, CSWeickert, TWWhat's hot in schizophrenia research?. Psychiatr Clin North Am 2016;39(2):343–51.CrossRefGoogle ScholarPubMed
Williams, AJUmemori, HThe best-laid plans go oft awry: synaptogenic growth factor signaling in neuropsychiatric disease. Front Synaptic Neurosci 2014; 6:4.CrossRefGoogle ScholarPubMed
McCullumsmith, REO'Donovan, SMDrummond, JBBenesh, FSSimmons, MRoberts, R et al. Shaping plasticity: alterations in glutamate transporter localization as a pathophysiological mechanism in severe mental illness. Mol Psychiatry 2016;21(6):723.CrossRefGoogle ScholarPubMed
Bernstein, HGSteiner, JGuest, PCDobrowolny, HBogerts, BGlial cells as key players in schizophrenia pathology: recent insights and concepts of therapy. Schizophr Res 2014.Google ScholarPubMed
Duncan, LEHolmans, PALee, PHO'Dushlaine, CTKirby, AWSmoller, JW et al. Pathway analyses implicate glial cells in schizophrenia. PLoS One 2014;9(2):e89441.CrossRefGoogle Scholar
Pantazopoulos, HBerretta, SIn sickness and in health: perineuronal nets and synaptic plasticity in psychiatric disorders. Neural Plast 2016; 2016:9847696.CrossRefGoogle ScholarPubMed
Huntley, GWGil, OBozdagi, OThe cadherin family of cell adhesion molecules: multiple roles in synaptic plasticity. Neuroscientist 2002;8(3):221–33.CrossRefGoogle ScholarPubMed
Washbourne, PDityatev, AScheiffele, PBiederer, TWeiner, JAChristopherson, KS et al. Cell adhesion molecules in synapse formation. J Neurosci 2004;24(42):9244–49.CrossRefGoogle ScholarPubMed
Lin, YCKoleske, AJMechanisms of synapse and dendrite maintenance and their disruption in psychiatric and neurodegenerative disorders. Annu Rev Neurosci 2010; 33:349–78.CrossRefGoogle ScholarPubMed
Benson, DLHuntley, GWSynapse adhesion: a dynamic equilibrium conferring stability and flexibility. Curr Opin Neurobiol 2012;22(3):397404.CrossRefGoogle ScholarPubMed
Cheadle, LBiederer, TThe novel synaptogenic protein Farp1 links postsynaptic cytoskeletal dynamics and transsynaptic organization. J Cell Biol 2012;199(6):9851001.CrossRefGoogle ScholarPubMed
Sloniowski, SEthell, IMLooking forward to EphB signaling in synapses. Semin Cell Dev Biol 2012;23(1):7582.CrossRefGoogle ScholarPubMed
Koleske, AJMolecular mechanisms of dendrite stability. Nat Rev Neurosci 2013;14(8):536–50.CrossRefGoogle ScholarPubMed
Beffert, UWeeber, EJDurudas, AQiu, SMasiulis, ISweatt, JD et al. Modulation of synaptic plasticity and memory by Reelin involves differential splicing of the lipoprotein receptor Apoer2. Neuron 2005;47(4):567–79.CrossRefGoogle ScholarPubMed
Costa, EDavis, JGrayson, DRGuidotti, APappas, GDPesold, CDendritic spine hypoplasticity and downregulation of reelin and gabaergic tone in schizophrenia vulnerability. Neurobiol Dis 2001;8(5):723–42.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1 Diagramatic representation of the tetrapartite synapse. Elements composing it are the pre- and post-synaptic terminals, astrocytic processes surrounding them and perisynaptic extracellular matrix condensations interposed between these elements.

Figure 1

Fig. 2 CS-6/Glia clusters in the healthy human amygdala, immunolabeled with CS-6 antibody CS56. Scale bar 100 μm.

Figure 2

Fig. 3 (A) Rodent CS-6 cluster (red; immunolabeled with CS56) surrounded by astrocytes (green; immunoreactive for glial fibrillary acidic protein (GFAP). (B) Immunolabeled CS-6/Glia clusters (blue) in the mouse hippocampus. These clusters are crossed by several dendrites arising from projection neurons (green, immunolabeled for Thy1) and are often surrounded by interneurons expressing parvalbumin (red). Scale bar 100 μm. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Submit a response

Comments

No Comments have been published for this article.