Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 1958 2022-07-29 10:27:01 |
2 update references and layout + 1 word(s) 1959 2022-07-29 10:42:12 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Bolneo, E.;  Chau, P.Y.S.;  Noakes, P.G.;  Bellingham, M.C. Importance of GABA in the Nervous System. Encyclopedia. Available online: https://encyclopedia.pub/entry/25649 (accessed on 19 April 2024).
Bolneo E,  Chau PYS,  Noakes PG,  Bellingham MC. Importance of GABA in the Nervous System. Encyclopedia. Available at: https://encyclopedia.pub/entry/25649. Accessed April 19, 2024.
Bolneo, Erika, Pak Yan S. Chau, Peter G. Noakes, Mark C. Bellingham. "Importance of GABA in the Nervous System" Encyclopedia, https://encyclopedia.pub/entry/25649 (accessed April 19, 2024).
Bolneo, E.,  Chau, P.Y.S.,  Noakes, P.G., & Bellingham, M.C. (2022, July 29). Importance of GABA in the Nervous System. In Encyclopedia. https://encyclopedia.pub/entry/25649
Bolneo, Erika, et al. "Importance of GABA in the Nervous System." Encyclopedia. Web. 29 July, 2022.
Importance of GABA in the Nervous System
Edit

Normal development and function of the central nervous system involves a balance between excitatory and inhibitory neurotransmission. Activity of both excitatory and inhibitory neurons is modulated by inhibitory signalling of the GABAergic and glycinergic systems. Mechanisms that regulate formation, maturation, refinement, and maintenance of inhibitory synapses are established in early life. Deviations from ideal excitatory and inhibitory balance, such as down-regulated inhibition, are linked with many neurological diseases, including epilepsy, schizophrenia, anxiety, and autism spectrum disorders. In the mammalian forebrain, gamma-aminobutyric acid (GABA) is the primary inhibitory neurotransmitter, binding to GABA receptors, opening chloride channels and hyperpolarizing the cell.

GABA-receptors GABAergic transmission neural development

1. Inhibition in the Forebrain Is Mediated by Gamma-Aminobutyric Acid (GABA)

Neural excitation by glutamatergic neurons is the primary influence driving central neurons to fire and is constantly counterbalanced by inhibitory synaptic inputs [1]. Inhibition in the central nervous system (CNS) is elicited by two major neurotransmitters: GABA, and glycine [2]. GABA is the primary inhibitory synaptic neurotransmitter of the CNS, with GABAergic synapses found throughout the brain [3] and especially in the forebrain, where they predominate. In addition, since many events of neurogenesis occur before the onset of synapse formation, non-synaptic GABAergic transmission together with endogenously released GABA has been postulated to be involved in early neurodevelopment [4]. Tonic GABA release [5][6] lays down the foundation for normal proliferation of neuronal precursors, neuronal differentiation and migration, and early activity patterns [4]. Clearly, proper formation of neuronal networks relies on the GABAergic system to function optimally in the developing brain.
The distribution of GABAergic circuits has been defined using immunohistochemical and electrophysiological techniques [2]. GABA is synthesised by GABAergic interneurons, and elicits inhibition by binding to GABA receptors on the postsynaptic membrane of other neurons [7]. GABAergic circuits are widely distributed in CNS regions such as the cortex, hippocampus, thalamus, hypothalamus, brainstem, and basal ganglia [3][7][8][9][10]. It is this sculpting of excitatory transmission by the inhibitory GABAergic system that allows normal synapse formation, maturation, and maintenance of neural circuits within the CNS [2][11]. GABAergic transmission is required for modulating circuits involved in complex cognitive behaviours, including personality expression, decision making, and goal-orientation [12].
Several brain regions use both GABAergic and glycinergic inhibition, including the retina, spinal cord, cerebellum, brainstem nuclei, olfactory bulb, and hippocampus [13][14][15][16][17][18]. Within these regions, GABAergic and glycinergic inhibition can act independently or together to modulate excitatory signals [19]. Mixed inhibitory signalling is also evidenced by the co-release of GABA and glycine from the axon terminal of brainstem and spinal cord interneurons, allowing a wider dynamic range of inhibitory control [20][21].

2. The Role of GABA in the Mature Central Nervous System

The effects of GABA are elicited by the binding of GABA to the ionotropic GABA-A and GABA-C receptor subtypes and the metabotropic GABA-B receptor subtype on the postsynaptic membrane of neurons [22][23]. The focuses are on the role of GABA-A receptors, which form the majority of GABA receptors within the CNS. GABA-A receptors incorporate an ion channel that allows the passage of chloride anions, and exert primary control over inhibitory signalling [24]. Not only do GABA-A receptors predominantly control inhibitory signalling within the CNS, they also play a vital modulatory role in neurodevelopmental events leading to the establishment of complex neuronal networks and to behaviours they regulate [24][25]. GABA-A receptors also exhibit binding sites for several modulatory molecules, including ethanol, benzodiazepines, anaesthetics, neurosteroids, barbiturates, and picrotoxin [24]. After GABA binding to their receptors, GABA is removed from the synaptic cleft into glial cells via GAT 2/3 for breakdown into glutamine, or transported into the presynaptic terminal via GAT-1 for recycling into synaptic vesicles, the latter accounting for ~80% of GABA uptake [26]. Additionally, Neuroligin-2 (NL2) is a postsynaptic cell adhesion protein exclusively localised at GABAergic synapses [27] that mediate a bidirectional signalling between pre- and postsynaptic neurons by forming a trans-synaptic signal transduction complex [28]. In the process of forming the complex, NL2 is accountable for recruiting required additional proteins, which is essential for the stabilisation/destabilisation of GABAergic synapses [29][30].
Gamma aminobutyric acid (GABA) is synthesised by glutamic acid decarboxylase (GAD) [31]. GAD comes in two isoforms—67 kDa (GAD67) and 65 kDa (GAD65), which are derived from different genes, rather than alternatively spliced proteins that come from the same gene [32]. GAD67- and GAD65-mediated synthesis of GABA differ in a temporospatial manner [33]. GAD67 contributes to over 90% of basal GABA synthesis and is distributed throughout the cell, while GAD65 remains localised to presynaptic nerve terminals [33]. During development, synthesis of GABA by the two enzymes GAD65 and GAD67 also changes significantly [34]. GAD67 deficient (KO; GAD67−/−) mouse brains contain 70–95% less GABA at birth, compared to wild type (WT; GAD67+/+) mice, and die at birth due to respiratory failure [34]. By contrast, the brains of GAD67+/− mice are 30–50% deficient in GABA at birth and pups survive to adulthood [34]. GAD65−/− mice have normal levels of GABA at birth but show a progressive decrease in GABA levels in the hippocampus and cortex from 60 days postnatal [35]. While both heterozygous (+/−) GAD67 and GAD65 mice exhibit spontaneous seizures and behavioural deficiencies, only the latter are predisposed to premature death [35][36]. GAD67 activity thus provides most GABA synthesis during prenatal and early postnatal (P) development to day 60 (P60), while GAD65 is the predominant source of GABA in the adult CNS [34].

3. The Role of VGAT in GABA and Glycine Signalling

The vesicular GABA transporter (VGAT), also known as vesicular inhibitory amino acid transporter (VIAAT), is a common vesicle transporter for GABA and glycine, and is essential for normal GABAergic and glycinergic neurotransmission [37][38]. It facilitates GABA and glycine signalling by transporting them into their synaptic vesicles prior to exocytosis [39]. The uptake of GABA and glycine by VGAT is heavily dependent on pH and electrochemical gradients across the vesicular membrane that are regulated by Mg2+-activated ATPase [39]. VGAT is highly expressed in the nerve terminals of both glycinergic and GABAergic neurons [37][38]. In regions where GABA and glycine are co-released, GABA has a higher affinity towards VGAT than glycine [39].

4. GABA in Neurodevelopment

A growing focus of research is dedicated towards revealing the critical role of inhibitory neurotransmitters in refining many aspects of neurodevelopment [5][6][40]. In addition to inhibitory signalling in the mature CNS, GABA has been demonstrated to provide significant excitatory activity in the developing CNS [41][42]. Immature neurons use ionotropic GABA transmission by transporting chloride (Cl) anions across the cell membrane, where elevated intracellular Cl concentration causes the cell membrane to depolarise and elicit functionally excitatory actions [43][44]. During two weeks postnatal in rodents and approximately full-term (forty weeks) birth in humans, the efflux of Cl from immature CNS neurons is increased, causing the neurons to shift from depolarising to hyperpolarising [44]. This electrochemical shift is necessary for establishing the subsequent inhibitory action of GABAergic neurons, playing a vital role in the developing CNS. Evidence suggests a combination of excitation from GABAergic, glycinergic, and glutamatergic neurons contribute to activity-dependent remodelling during neurodevelopment [11]. Each of these neurotransmitter systems work together to delicately balance neural activity within a normal physiological range, thus playing a key role in establishing neural circuits of neural pathways necessary for vision, hearing, breathing, and pain [40][45].
GABAergic synaptic inputs mould connectivity and plasticity of the developing mammalian brain [46]. While GABA is the major inhibitory neurotransmitter in the mature mammalian CNS, GABA-A receptors evoke excitation before birth and this period of excitation can even extend into early postnatal life, evoking depolarisation in postsynaptic neurons [44][47][48][49][50][51][52]. Through this signalling, postsynaptic neurons, which are new projection neurons derived from neurogenesis of proliferative zones such as ventricular and subventricular zones [53], are guided to migrate into target brain areas [5][46][54]. This process of migration is dependent on the receptor subtype expressed on the postsynaptic neuron [48]. Neural progenitors in the proliferative zone express GABA-A receptors [55][56][57]. Here, GABAergic signalling can promote proliferation, cell migration, and cell cycle exit [57]. Once cells reach the cortical plate, GABA-A receptor activation results in migration cessation [57][58]. Once foundational GABAergic circuits are in place, continued excitatory signalling through GABA-A activation causes projection neurons to extend dendrite length and form new synaptic contacts [59]. Along with promoting normal development of neural processes, GABA-A receptor activity also regulates the maturation of inhibitory synapses required for differentiation [48][60]. Dynamic expression of GABA receptor subtypes and their respective activity constitute a careful orchestration directed by the GABAergic system in normal CNS development [48].
The shift from an excitatory to an inhibitory action of GABA may also be modulated in an activity-dependent manner [47]. Following excitatory GABAergic activity and near the time of birth, increased expression of the potassium chloride transporter 2 (KCC2) switches the transmembrane chloride gradient from a depolarising to a hyperpolarising direction, thereby changing GABAergic signalling to inhibitory [48]. Through hyperpolarisation, the postsynaptic neuron’s ability to fire action potentials is down regulated [61], allowing for fine tuning of neural information processing [62].
In light of this, genetic mutation of any of the components of the GABAergic system have been shown to cause a wide range of neurological disorders [1]. Abnormal development of the GABAergic system could result either in disorganised neural circuits caused by improper GABA function during development, and/or by alteration in GABAergic synaptic function resulting in atypical action potential firing of neural circuits, underlying the pathophysiology of many synaptic inhibition disorders [48].

5. Excitatory–Inhibitory Balance Is Crucial in Normal Neurodevelopment

Excitatory–inhibitory (E-I) balance in neuronal circuits has become of increasing research interest, in most part due to its potential roles in the aetiology of a wide range of neurological diseases [1]. Neuronal circuits coordinate E-I signalling that varies depending on the individual’s developmental period and interactions with the environment [11][63]. During development, the quantity and distribution of excitatory versus inhibitory synapses across neurons and their individual dendrites serve as the hardware of the brain [11][64]. These cellular computations allow us to perceive our environment and interact accordingly. There is a strong relationship between altered synaptic activity onto neurons (as measured by patch-clamp electrophysiology) and concomitant neuronal dendritic architecture and spine density changes (as measured by single neuron morphology) [11][65][66][67]. This association between abnormal synaptic physiology and dendritic structures occurs in both pyramidal cells and other neuronal types [11][65]. These relationships between alterations in neuronal structure and function have been confirmed using other techniques, such as Golgi impregnation [65][68].
Strong evidence suggests an increased ratio of excitatory over inhibitory (E/I) transmission contributes to the pathogenesis of neurological disorders, with a large number of mutations and variants of different genes culminating to a disturbed E-I balance [48]. Mutations in genes leading to dysfunctional inhibitory synapses and inhibitory neurotransmission thus place an individual at higher risk of developing neurological disorders [48][69]. A substantial proportion of clinically identified neurodevelopmental disorder risk genes encode GABAergic transcription factors, GABA receptors, chloride transporters, and inhibitory synaptic proteins [46]. An increased E-I ratio has been found in animal models for epilepsy, schizophrenia, ASDs, and anxiety [33][69]. Additionally, mutations in genes coding for the enzymes responsible for glycine synthesis have been seen to cause hyperglycinemia—a disease characterized by mental retardation, lethargy, and seizures [70].
More importantly, in the instance of inhibitory signalling, an altered GABAergic system, including GABAergic interneuronal loss, disrupted cell maturation, imbalance in GABA-mediated inhibitory synaptic transmission, or reduced GABA, GAD, parvalbumin (PV), and somatostatin (Sst) expression, have been reported in a variety of neurodevelopmental and neuropsychiatric disorders [71][72][73][74][75].
To understand the vital role GABAergic signalling plays in establishing normal neurodevelopment, and how mutations within genes of the GABAergic system can have devastating impacts in development and the organism at whole, two rodent models will be examined for down-regulated GABAergic influence—the GAD67+/− and VGAT−/− mouse. The GAD67+/− mouse allows us to investigate the effects of a 50% global reduction of GABA in the developing CNS and how a mutation within the gene that codes for GAD67, can manifest in neurological disease [34][76]. Another mouse model worthy of importance when examining GABAergic signalling in normal neurodevelopment and neurological diseases is the vesicular GABA transporter (VGAT) deficient mouse model (i.e., the VGAT−/− mouse). The use of these two mouse models when investigating potential therapeutics for epilepsy, schizophrenia, ASDs, and anxiety are the focuses.

References

  1. He, H.-Y.; Cline, H.T. What Is Excitation/Inhibition and How Is It Regulated? A Case of the Elephant and the Wisemen. J. Exp. Neurosci. 2019, 13, 1179069519859371.
  2. Gamlin, C.R.; Yu, W.-Q.; Wong, R.O.L.; Hoon, M. Assembly and maintenance of GABAergic and Glycinergic circuits in the mammalian nervous system. Neural Dev. 2018, 13, 12.
  3. Jewett, B.E.; Sharma, S. Physiology, GABA. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2021.
  4. Kilb, W.; Kirischuk, S.; Luhmann, H.J. Role of tonic GABAergic currents during pre- and early postnatal rodent development. Front. Neural Circuits 2013, 7, 139.
  5. Cellot, G.; Cherubini, E. Functional role of ambient GABA in refining neuronal circuits early in postnatal development. Front. Neural Circuits 2013, 7, 136.
  6. Manent, J.-B.; Demarque, M.; Jorquera, I.; Pellegrino, C.; Ben-Ari, Y.; Aniksztejn, L.; Represa, A. A Noncanonical Release of GABA and Glutamate Modulates Neuronal Migration. J. Neurosci. 2005, 25, 4755–4765.
  7. Xiang, Z.; Huguenard, J.R.; Prince, D.A. GABAA receptor-mediated currents in interneurons and pyramidal cells of rat visual cortex. J. Physiol. 1998, 506, 715–730.
  8. Huang, Z.J.; Di Cristo, G.; Ango, F. Development of GABA innervation in the cerebral and cerebellar cortices. Nat. Rev. Neurosci. 2007, 8, 673–686.
  9. Decavel, C.; Pol, A.N.V.D. GABA: A dominant neurotransmitter in the hypothalamus. J. Comp. Neurol. 1990, 302, 1019–1037.
  10. Halassa, M.M.; Acsády, L. Thalamic Inhibition: Diverse Sources, Diverse Scales. Trends Neurosci. 2016, 39, 680–693.
  11. Fogarty, M.J.; Kanjhan, R.; Bellingham, M.C.; Noakes, P.G. Glycinergic Neurotransmission: A Potent Regulator of Embryonic Motor Neuron Dendritic Morphology and Synaptic Plasticity. J. Neurosci. 2016, 36, 80–87.
  12. Moriguchi, Y.; Hiraki, K. Prefrontal cortex and executive function in young children: A review of NIRS studies. Front. Hum. Neurosci. 2013, 7, 867.
  13. Dumoulin, A.; Triller, A.; Dieudonné, S. IPSC Kinetics at Identified GABAergic and Mixed GABAergic and Glycinergic Synapses onto Cerebellar Golgi Cells. J. Neurosci. 2001, 21, 6045–6057.
  14. Wässle, H.; Koulen, P.; Brandstätter, J.H.; Fletcher, E.; Becker, C.-M. Glycine and GABA receptors in the mammalian retina. Vis. Res. 1998, 38, 1411–1430.
  15. Pol, A.N.V.D.; Görcs, T. Glycine and glycine receptor immunoreactivity in brain and spinal cord. J. Neurosci. 1988, 8, 472–492.
  16. Sanes, D.H.; Geary, W.A.; Wooten, G.F.; Rubel, E.W. Quantitative distribution of the glycine receptor in the auditory brain stem of the gerbil. J. Neurosci. 1987, 7, 3793–3802.
  17. Trombley, P.Q.; Shepherd, G.M. Glycine exerts potent inhibitory actions on mammalian olfactory bulb neurons. J. Neurophysiol. 1994, 71, 761–767.
  18. Song, W.; Chattipakorn, S.C.; McMahon, L.L. Glycine-Gated Chloride Channels Depress Synaptic Transmission in Rat Hippo-campus. J. Neurophysiol. 2006, 95, 2366–2379.
  19. Diamond, J.S. Inhibitory Interneurons in the Retina: Types, Circuitry, and Function. Annu. Rev. Vis. Sci. 2017, 3, 1–24.
  20. Jonas, P.; Bischofberger, J.; Sandkuühler, J. Corelease of Two Fast Neurotransmitters at a Central Synapse. Science 1998, 281, 419–424.
  21. O’Brien, J.A.; Berger, A.J. Cotransmission of GABA and Glycine to Brain Stem Motoneurons. J. Neurophysiol. 1999, 82, 1638–1641.
  22. Kelsom, C.; Lu, W. Development and specification of GABAergic cortical interneurons. Cell Biosci. 2013, 3, 19.
  23. Sigel, E.; Steinmann, M.E. Structure, Function, and Modulation of GABA(A) Receptors. J. Biol. Chem. 2012, 287, 40224–40231.
  24. Olsen, R.W.; DeLorey, T.M. GABA Receptor Physiology and Pharmacology. In Basic Neurochemistry: Molecular, Cellular and Medical Aspects, 6th ed.; Siegel, G.J., Agranoff, B.W., Albers, R.W., Eds.; Lippincott-Raven: Philadelphia, PA, USA, 1999.
  25. DeLorey, T.M.; Sahbaie, P.; Hashemi, E.; Homanics, G.E.; Clark, J.D. Gabrb3 gene deficient mice exhibit impaired social and exploratory behaviors, deficits in non-selective attention and hypoplasia of cerebellar vermal lobules: A potential model of autism spectrum disorder. Behav. Brain Res. 2008, 187, 207–220.
  26. Zafar, S.; Jabeen, I. Structure, Function, and Modulation of γ-Aminobutyric Acid Transporter 1 (GAT1) in Neurological Disorders: A Pharmacoinformatic Prospective. Front. Chem. 2018, 11, 397.
  27. Varoqueaux, F.; Jamain, S.; Brose, N. Neuroligin 2 is exclusively localized to inhibitory synapses. Eur. J. Cell Biol. 2004, 83, 449–456.
  28. Siddiqui, T.J.; Craig, A.M. Synaptic organizing complexes. Curr. Opin. Neurobiol. 2011, 21, 132–143.
  29. Nam, C.I.; Chen, L. Postsynaptic assembly induced by neurexin-neuroligin interaction and neurotransmitter. Proc. Natl. Acad. Sci. USA 2005, 102, 6137–6142.
  30. Scheiffele, P.; Fan, J.; Choih, J.; Fetter, R.; Serafini, T. Neuroligin Expressed in Nonneuronal Cells Triggers Presynaptic Development in Contacting Axons. Cell 2000, 101, 657–669.
  31. Modi, J.; Prentice, H.; Wu, J.-Y. Regulation of GABA Neurotransmission by Glutamic Acid Decarboxylase (GAD). Curr. Pharm. Des. 2015, 21, 4939–4942.
  32. Erlander, M.G.; Tillakaratne, N.J.; Feldblum, S.; Patel, N.; Tobin, A.J. Two genes encode distinct glutamate decarboxylases. Neuron 1991, 7, 91–100.
  33. Gaetz, W.; Bloy, L.; Wang, D.; Port, R.; Blaskey, L.; Levy, S.; Roberts, T. GABA estimation in the brains of children on the autism spectrum: Measurement precision and regional cortical variation. NeuroImage 2013, 86, 1.
  34. Kuwana, S.; Okada, Y.; Sugawara, Y.; Tsunekawa, N.; Obata, K. Disturbance of neural respiratory control in neonatal mice lacking gaba synthesizing enzyme 67-kda iso-form of glutamic acid decarboxylase. Neuroscience 2003, 120, 861–870.
  35. Stork, O.; Ji, F.-Y.; Kaneko, K.; Stork, S.; Yoshinobu, Y.; Moriya, T.; Shibata, S.; Obata, K. Postnatal development of a GABA deficit and disturbance of neural functions in mice lacking GAD65. Brain Res. 2000, 865, 45–58.
  36. Sandhu, K.V.; Lang, D.; Muller, B.; Nullmeier, S.; Yanagawa, Y.; Schwegler, H.; Stork, O. Glutamic acid decarboxylase 67 haplodeficiency impairs social behavior in mice. Genes Brain Behav. 2014, 13, 439–450.
  37. Sagné, C.; El Mestikawy, S.; Isambert, M.-F.; Hamon, M.; Henry, J.-P.; Giros, B.; Gasnier, B. Cloning of a functional vesicular GABA and glycine transporter by screening of genome databases. FEBS Lett. 1997, 417, 177–183.
  38. Dumoulin, A.; Rostaing, P.; Bedet, C.; Lévi, S.; Isambert, M.F.; Henry, J.P.; Gasnier, B. Presence of the vesicular inhibitory amino acid transporter in GABAergic and glycinergic synaptic ter-minal boutons. J. Cell Sci. 1999, 112, 811–823.
  39. Chaudhry, F.A.; Reimer, R.J.; Bellocchio, E.E.; Danbolt, N.C.; Osen, K.K.; Edwards, R.H.; Storm-Mathisen, J. The Vesicular GABA Transporter, VGAT, Localizes to Synaptic Vesicles in Sets of Glycinergic as Well as GABAergic Neurons. J. Neurosci. 1998, 18, 9733–9750.
  40. Bruining, H.; Hardstone, R.; Juarez-Martinez, E.L.; Sprengers, J.; Avramiea, A.-E.; Simpraga, S.; Houtman, S.J.; Poil, S.-S.; Dallares, E.; Palva, S.; et al. Measurement of excitation-inhibition ratio in autism spectrum disorder using critical brain dynamics. Sci. Rep. 2020, 10, 9195.
  41. Ito, S. GABA and glycine in the developing brain. J. Physiol. Sci. 2016, 66, 375–379.
  42. Wang, W.; Xu, T.L. Chloride homeostasis differentially affects GABA(A) receptor- and glycine receptor-mediated effects on spontaneous circuit activity in hippocampal cell culture. Neurosci. Lett. 2006, 406, 11–16.
  43. Blaesse, P.; Airaksinen, M.S.; Rivera, C.; Kaila, K. Cation-Chloride Cotransporters and Neuronal Function. Neuron 2009, 61, 820–838.
  44. Virtanen, M.A.; Uvarov, P.; Hübner, C.A.; Kaila, K. NKCC1, an Elusive Molecular Target in Brain Development: Making Sense of the Existing Data. Cells 2020, 9, 2607.
  45. Gatto, C.L.; Broadie, K. Genetic controls balancing excitatory and inhibitory synaptogenesis in neurodevelopmental dis-order models. Front. Synaptic Neurosci. 2010, 2, 4.
  46. Tang, X.; Jaenisch, R.; Sur, M. The role of GABAergic signalling in neurodevelopmental disorders. Nat. Rev. Neurosci. 2021, 22, 290–307.
  47. Ben-Ari, Y. Excitatory actions of gaba during development: The nature of the nurture. Nat. Rev. Neurosci. 2002, 3, 728–739.
  48. Schmidt, M.J.; Mirnics, K. Neurodevelopment, GABA System Dysfunction, and Schizophrenia. Neuropsychopharmacology 2014, 40, 190–206.
  49. Murata, Y.; Colonnese, M.T. GABAergic interneurons excite neonatal hippocampus in vivo. Sci. Adv. 2020, 6, eaba1430.
  50. Payne, J.A.; Rivera, C.; Voipio, J.; Kaila, K. Cation-chloride co-transporters in neuronal communication, development and trauma. Trends Neurosci. 2003, 26, 199–206.
  51. Kirmse, K.; Hübner, C.A.; Isbrandt, D.; Witte, O.W.; Holthoff, K. GABAergic Transmission during Brain Development: Multiple Effects at Multiple Stages. Neuroscientist 2017, 24, 36–53.
  52. Watanabe, M.; Fukuda, A. Development and regulation of chloride homeostasis in the central nervous system. Front. Cell. Neurosci. 2015, 9, 371.
  53. Kendler, A.; Golden, J.A. Progenitor Cell Proliferation Outside the Ventricular and Subventricular Zones during Human Brain Development. J. Neuropathol. Exp. Neurol. 1996, 55, 1253–1258.
  54. Owens, D.F.; Kriegstein, A.R. Is there more to GABA than synaptic inhibition? Nat. Rev. Neurosci. 2002, 3, 715–727.
  55. Haydar, T.F.; Wang, F.; Schwartz, M.L.; Rakic, P. Differential modulation of proliferation in the neocortical ventricular and subventricular zones. J. Neurosci. 2000, 20, 5764–5774.
  56. Bortone, D.; Polleux, F. KCC2 Expression Promotes the Termination of Cortical Interneuron Migration in a Voltage-Sensitive Calcium-Dependent Manner. Neuron 2009, 62, 53–71.
  57. LoTurco, J.J.; Owens, D.F.; Heath, M.J.; Davis, M.B.; Kriegstein, A.R. GABA and glutamate depolarize cortical progenitor cells and inhibit DNA synthesis. Neuron 1995, 15, 1287–1298.
  58. Wang, Y.; Kakizaki, T.; Sakagami, H.; Saito, K.; Ebihara, S.; Kato, M.; Hirabayashi, M.; Saito, Y.; Furuya, N.; Yanagawa, Y. Fluorescent labeling of both GABAergic and glycinergic neurons in vesicular GABA transporter (VGAT)–Venus transgenic mouse. Neuroscience 2009, 164, 1031–1043.
  59. Marty, S.; Berninger, B.; Carroll, P.; Thoenen, H. GABAergic Stimulation Regulates the Phenotype of Hippocampal Interneurons through the Regulation of Brain-Derived Neurotrophic Factor. Neuron 1996, 16, 565–570.
  60. Wu, C.; Sun, D. GABA receptors in brain development, function, and injury. Metab. Brain Dis. 2015, 30, 367–379.
  61. Ganguly, K.; Schinder, A.F.; Wong, S.T.; Poo, M.-M. GABA Itself Promotes the Developmental Switch of Neuronal GABAergic Responses from Excitation to Inhibition. Cell 2001, 105, 521–532.
  62. Carment, L.; Dupin, L.; Guedj, L.; Térémetz, M.; Cuenca, M.; Krebs, M.-O.; Amado, I.; Maier, M.A.; Lindberg, P.G. Neural noise and cortical inhibition in schizophrenia. Brain Stimul. 2020, 13, 1298–1304.
  63. Alwis, D.S.; Rajan, R. Environmental enrichment causes a global potentiation of neuronal responses across stimulus com-plexity and lamina of sensory cortex. Front. Cell. Neurosci. 2013, 7, 124.
  64. Villa, K.L.; Nedivi, E. Dendrites; Springer: Tokyo, Japan, 2016; pp. 467–487.
  65. Fogarty, M.J.; Kanjhan, R.; Yanagawa, Y.; Noakes, P.G.; Bellingham, M. Alterations in hypoglossal motor neurons due to GAD67 and VGAT deficiency in mice. Exp. Neurol. 2017, 289, 117–127.
  66. Forrest, M.P.; Hill, M.J.; Kavanagh, D.H.; Tansey, K.E.; Waite, A.J.; Blake, D.J. The Psychiatric Risk Gene Transcription Factor 4 (TCF4) Regulates Neurodevelopmental Pathways Asso-ciated With Schizophrenia, Autism, and Intellectual Disability. Schizophr Bull. 2018, 44, 1100–1110.
  67. Penzes, P.; Woolfrey, K.M.; Srivastava, D.P. Epac2-mediated dendritic spine remodeling: Implications for disease. Mol. Cell. Neurosci. 2011, 46, 368–380.
  68. Klenowski, P.M.; Shariff, M.; Belmer, A.; Fogarty, M.J.; Mu, E.W.H.; Bellingham, M.; Bartlett, S.E. Prolonged Consumption of Sucrose in a Binge-Like Manner, Alters the Morphology of Medium Spiny Neurons in the Nucleus Accumbens Shell. Front. Behav. Neurosci. 2016, 10, 54.
  69. Moy, S.S.; Nadler, J.J.; Young, N.B.; Perez, A.; Holloway, L.P.; Barbaro, R.P.; Barbaro, J.R.; Wilson, L.M.; Threadgill, D.W.; Lauder, J.M.; et al. Mouse behavioral tasks relevant to autism: Phenotypes of 10 inbred strains. Behav. Brain Res. 2007, 176, 4–20.
  70. Purves, D.; Augustine, G.J.; Fitzpatrick, D.; Katz, L.C.; LaMantia, A.S.; McNamara, J.O.; Williams, S.M. GABA and Glycine. In Neuroscience, 2nd ed.; Sinauer Associates: Sunderland, MA, USA, 2001.
  71. Fujihara, K.; Miwa, H.; Kakizaki, T.; Kaneko, R.; Mikuni, M.; Tanahira, C.; Yanagawa, Y. Glutamate Decarboxylase 67 Deficiency in a Subset of GABAergic Neurons Induces Schizophrenia-Related Pheno-types. Neuropsychopharmacology 2015, 40, 2475–2486.
  72. Thompson, M.; Weickert, C.S.; Wyatt, E.; Webster, M.J. Decreased glutamic acid decarboxylase67 mRNA expression in multiple brain areas of patients with schizophrenia and mood disorders. J. Psychiatr. Res. 2009, 43, 970–977.
  73. Whitney, E.R.; Kemper, T.L.; Bauman, M.L.; Rosene, D.; Blatt, G.J. Cerebellar Purkinje Cells are Reduced in a Subpopulation of Autistic Brains: A Stereological Experiment Using Calbindin-D28k. Cerebellum 2008, 7, 406–416.
  74. Fatemi, S.H.; Reutiman, T.J.; Folsom, T.D.; Thuras, P.D. GABA(A) receptor downregulation in brains of subjects with autism. J. Autism. Dev. Disord. 2009, 39, 223–230.
  75. Maglóczky, Z.; Freund, T.F. Impaired and repaired inhibitory circuits in the epileptic human hippocampus. Trends Neurosci. 2005, 28, 334–340.
  76. Lin, C.C.; Huang, T.L. Chapter 3—Epigenetic biomarkers in neuropsychiatric disorders. In Neuropsychiatric Disorders and Epi-genetics; Yasui, D.H., Peedicayil, J., Grayson, D.R., Eds.; Academic Press: Boston, MA, USA, 2017; pp. 35–66.
More
Information
Subjects: Neurosciences
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 463
Revisions: 2 times (View History)
Update Date: 29 Jul 2022
1000/1000