Open access

Introductory Chapter: GABA/Glutamate Balance: A Key for Normal Brain Functioning

Written By

Janko Samardzic, Dragana Jadzic, Boris Hencic, Jasna Jancic and Dubravka Svob Strac

Submitted: 27 June 2017 Published: 21 March 2018

DOI: 10.5772/intechopen.74023

From the Edited Volume

GABA And Glutamate - New Developments In Neurotransmission Research

Edited by Janko Samardzic

Chapter metrics overview

1,856 Chapter Downloads

View Full Metrics

1. Introduction

The basis of information transfer in the mammalian central nervous system (CNS) consists of excitation and inhibition of neuronal networks. The messengers responsible for propagating these excitatory and inhibitory actions are amino acid neurotransmitters [1]. The principal excitatory neurotransmitter is glutamate, while the principle inhibitory neurotransmitter is gamma-aminobutyric acid (GABA). Coordination between these two principal neurotransmitters ensures adequate rhythmic activity, which may involve either a single neuron or multiple neuronal groups, thus altering synaptic plasticity and ensuring a normal functioning of CNS [2]. As this spatiotemporal framework of different patterns in neural oscillations is essential for information processing throughout the brain [3], the deviations in normal activity of either system or their interactions are associated with a number of neurological and psychiatric diseases [4].

The GABA/glutamate functional balance could be achieved by homeostatic control of presynaptic elements such as glutamate and GABA release, which could be the result of changes in their metabolism (synthesis or degradation involving various enzymes), compartmentation, and recycling (involving plasma transporters) and in the amounts of transmitters available for release from synaptic vesicles (involving vesicular transporters). However, it is generally considered that homeostatic plasticity mechanisms in the brain are mediated primarily by regulation of expression and function of glutamate and GABA receptors [5].

Advertisement

2. GABA and its receptors

Every third chemical synapse in the brain uses neurotransmitter GABA as an integral part of the neurotransmission process. GABA mediates its effects via two types of receptors: ionotropic GABAA and metabotropic GABAB receptors [6]. Although a third type of GABA receptor with pharmacological specificities has been identified, the term GABAС has not received broad consensus among experts. Additionally, the International Union of Basic and Clinical Pharmacology (IUPHAR) has classified GABAC as a type of GABAA receptor [7].

GABAA receptors generally contain chloride ion channels but can, in varying degrees, also contain calcium, sodium, and potassium channels. GABAA receptors mediate the majority of GABA inhibitory actions in the CNS [4]. They are pentameric transmembrane receptors made up of 5 subunit proteins that form an ion channel selectively permeable to chloride anions. Although mainly localized on postsynaptic membranes, they can also be found extrasynaptically, especially GABAA receptors containing α4, α5, or α6 subunits [8]. Unexpectedly, GABAA receptors have also been found on glial cells, potentially providing adaptational support for adjacent neurons [9]. Activation of GABAA receptors leads to a change in the conformational state of associated ion channels, resulting in increased permeability to chloride ions. GABAergic mechanisms are also involved in metabolic processes [10], and a negative correlation between the intensity of GABAergic neurotransmission and metabolic processes in cerebral tissue has been established. So far, 19 subunits of GABAA receptors have been cloned and classified into several structurally related subfamilies (α 1–6, β 1–3, γ 1–3, δ, ε, θ, π, ρ 1–3). The most frequently found GABAA receptor composition is an aggregate composed of two α, two β, and one γ subunit [4]. Receptors that, in addition to two α and two β subunits, contain some other non-γ subunit are rare. Receptors composed only of α and β isoforms also exist. The subunit composition determines the functional and pharmacological properties of GABAA receptors. For example, α1 GABAA receptors mediate sedative and anticonvulsant actions, whereas the α2 subunit is responsible for anxiolytic action of benzodiazepines. Zolpidem, a commonly prescribed sedative for sleep initiation, has a high binding affinity for GABAA receptors containing the α1 subunit [11].

GABA action through GABAA receptors results in chloride channel opening and increased postsynaptic membrane permeability. In addition to the well-determined benzodiazepine binding site, at least 13 different and structurally specific sites on the GABAA receptors have been identified: (1) GABA and other agonist-binding sites, as well as competitive antagonists; (2) picrotoxin site near ion channel; (3) barbiturates binding site; (4) neuroactive steroids binding site; (5) ethanol binding site; (6) inhalation anesthetics stereoselective binding sites; (7) furosemide diuretic binding site; (8) Zn2+ ion binding site; (9) other divalent cation binding sites; (10) La3+ ions site; (11) sites for phosphorylation of specific protein kinases; (12) phospholipid-binding sites; and (13) sites involved in interaction of GABAA receptor and microtubules, which promote receptor grouping on postsynaptic membranes [12]. Modulators of GABAA receptor complex interact with these binding sites in three possible ways: positive allosteric modulators that potentiate chloride ion flux (agonists), negative modulators that reduce GABA-induced chloride ion flux (inverse agonists), and neutral allosteric modulators that competitively block the effects of these two types of agonists-antagonists.

On the other hand, GABAB metabotropic receptors, characterized by stereoselective ligand (−)baclofen, belong to the seven transmembrane G-protein-coupled receptor superfamily. They are pre- and postsynaptic G-protein-coupled receptors that negatively modulate adenylyl cyclase and inositol triphosphate synthesis. Heterodimeric structure as a result of GABAB1 and GABAB2 subunit assembly is necessary for appropriate GABAB receptor function. The extracellular domain of the GABAB1 subunit contains GABA-binding site, whereas GABAB2 subunit is important for the interaction with the G-proteins. GABAB receptor activation produces a cascade of signals that result in activation and/or inhibition of voltage-dependent calcium channels. GABAB receptor is located both centrally and peripherally, particularly in the thalamus, brain stem nuclei, and spinal cord. Depending on the localization of GABAB receptors, GABA-mediated inhibitory influences can be potentiated (postsynaptic receptors, presynaptic heteroreceptors on glutamatergic endings) or reduced (autoreceptors) [13, 14]. GABAB receptor function affects behavior, learning, and memory, and therefore their pharmacological targeting may be beneficial in various neuropsychiatric disorders [15, 16].

Advertisement

3. Glutamate and its receptors

Glutamate, the most abundant neurotransmitter in vertebrates and precursor of GABA, is present in over 90% of all synaptic connections in the human brain and is essential for a wide variety of functions [17]. Over 20 types of mammalian glutamate receptors exist, generally classified into two main categories: voltage-sensitive (ionotropic) and ligand-sensitive (metabotropic) receptors. Ionotropic receptor channels are formed from various protein subunits assembled in heterotetrameric or homotetrameric receptors. The three types of ionotropic receptors are N-methyl-d-aspartate (NMDA), α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA), and kainic acid receptors. The discovery of these agonists, after which the receptors were originally named, led to the detection of other receptor agonists and antagonists [18].

The binding of glutamate to NMDA receptors (NMDARs) results in the opening of a nonselective cation channel. The opening and closing of the channel are primarily gated by ligand binding but are also voltage-dependent. Extracellular magnesium and zinc ions can bind to specific sites on the receptor, blocking the passage of other cations through the open ion channel. However, depolarization of the neuronal cell dislodges and repels these ions from the pore, therefore allowing a voltage-dependent influx of sodium and calcium ions and efflux of potassium ions [19]. The NMDA receptor is primarily a ligand-gated channel, but it does display weaker voltage-dependent modulation of the ligand-dependent gating. NMDA requires co-activation by two ligands: glutamate and either d-serine or glycine [20]. Furthermore, NMDA receptors are divided into subtypes, depending on their intracellular protein structure, NR1, NR2, and NR3. NR1 consists of eight different subunits originating from a single gene via alternative splicing. NR2 has four subunits (A–D), and NR3 has two subunits (A and B). NMDA receptors are highly expressed on both neurons and astrocytes [21]. NMDA signaling is crucial for learning, memory, recovery from injury, and brain plasticity. It is especially important for proper functioning of the hippocampus [22]. In pathological circumstances, overactivation of NMDA receptors can lead to excitotoxicity, involved in some neurodegenerative disorders such as Alzheimer’s, Parkinson’s, and Huntington’s disease [23, 24, 25].

AMPA receptors (AMPAR) are composed of four types of subunits, designated as GluA1, GluA2, GluA3, and GluA4 [26]. These receptors are heterotetrameric, containing GluA2 and either GluA1, GluA3, or GluA4 subunits in a “dimer of dimers” structure [27, 28]. Each AMPAR consists of four subunits which make up four binding sites to which an agonist (such as glutamate) can bind. The channel opens when two binding sites are simultaneously occupied, and the current increases as more binding sites become occupied [29]. Once opened, the channel may undergo rapid desensitization and current termination. Since AMPARs open and close quickly (1 ms), they are responsible for fast excitatory synaptic transmission in the CNS [30]. The GluA2 subunit regulates whether the AMPAR is permeable to calcium and other cations, such as sodium and potassium. If receptor does not contain a GluA2, the AMPAR will be permeable to calcium, sodium, and potassium. Both NMDA and AMPA ion channels are important for plasticity and synaptic transmission at many postsynaptic membranes.

Kainate receptors (KAR) are heteromeric receptors assembled from four subunits, formerly referred to as GluR5, GluR6, GluR7, KA1, and KA2 but now named GluK1, GluK2, GluK3, GluK4, and GluK5, and grouped into low affinity (GluK1–3) and high affinity (GluK4–5) receptors. Each subunit has a large extracellular N-terminal domain, four helical transmembrane domains (M1–M4), and an intracellular C-terminal domain. GluK1–3 subunits can form both homomeric and heteromeric receptors, but GluK4 and GluK5 subunits can form only heteromeric functional ion channels together with GluK1–3 subunits. Despite their ion channel structure, KAR can also activate metabotropic signaling through noncanonical G-protein-coupled cascade. They are widely distributed in the brain and can be localized at pre-, post-, and/or extrasynaptic sites. Although KAR are less studied than AMPAR or NMDAR, it is not known that they are multifunctional neuronal modulators which play significant roles in health and disease [31].

Metabotropic glutamate receptors (mGluR) have a G-protein-linked receptor structure consisting of seven transmembrane domains with an extracellular N-terminal and an intracellular COOH terminal. When glutamate binds to a metabotropic receptor, it activates a postsynaptic intracellular G-protein, which eventually results in the opening of a membrane channel for signal transmission. Furthermore, G protein activation also triggers functional changes in the cytoplasm, resulting in gene expression and protein synthesis. For this reason, mGluR is generally considered slower acting channels than the ionotropic glutamate receptors. To date, three groups of mGluR exist. Group I receptors are coupled with phospholipase C, producing diacylglycerol and inositol triphosphate as second messengers. They are mainly expressed on the postsynaptic membrane. Group I receptors are involved in learning and memory, addiction, motor regulation, and Fragile X syndrome [32]. Groups II and III are negatively coupled to adenylyl cyclase. Impaired functioning of group II metabotropic receptors has been linked to anxiety, schizophrenia, and Alzheimer’s disease. Group III metabotropic receptors also inhibit neurotransmitter release but are positioned presynaptically. They are found within the hippocampus and hypothalamus and may play a role in Parkinson’s disease and anxiety disorders [33].

Advertisement

4. Conclusion and clinical implications

The adequate coordination of GABA and glutamate is essential to the normal functioning for the most complex brain processes. Decreased or increased GABA activity is associated with a number of neurological and psychiatric diseases. The GABAergic synapse is the site of action of several different classes of drugs that modulate inhibitory neurotransmission and are used in the pharmacotherapy of anxiety and sleep disorders, epilepsy, alcohol withdrawal, and induction and maintenance of anesthesia [34]. Moreover, glutamate dysfunction is also correlated with a wide range of nervous system disorders, such as Alzheimer’s disease, and neuropsychiatric disorders, including schizophrenia, pain disorders, drug addiction, and traumatic brain and spinal cord injuries [35]. Given the importance of equilibrium of these two systems for neuronal excitability, synaptic plasticity, and cognitive functions such as learning and memory, as well as its involvement in the mood, feeding behavior, reproductive functions, pain sensitivity, aging, etc. [36], it is not surprising that the development of current and prospective pharmaceuticals, including anxiolytics, antidepressants, antipsychotics, antiepileptics, antidementia, and many other drugs, relies increasingly on GABA/glutamate balance.

Advertisement

Acknowledgments

Janko Samardzic’s and Jasna Jancic’s work has been supported by the Ministry of Education, Science and Technological Development of the Republic of Serbia (Grant nos. 175076 and 175031). The authors declare that there are no competing interests regarding the publication of this chapter.

References

  1. 1. Petroff OA. GABA and glutamate in the human brain. The Neuroscientist. 2002;8(6):562-573
  2. 2. Foster AC, Kemp JA. Glutamate- and GABA-based CNS therapeutics. Current Opinion in Pharmacology. 2006;6(1):7-17
  3. 3. Buzsaki G, Draguhn A. Neuronal oscillations in cortical networks. Science. 2004;304(5679):1926-1929
  4. 4. Jacob TC, Moss SJ, Jurd R. GABA(A) receptor trafficking and its role in the dynamic modulation of neuronal inhibition. Nature Reviews. Neuroscience. 2008;9(5):331-343
  5. 5. Erickson JD, De Gois S, Varoqui H, Schafer MK, Weihe E. Activity-dependent regulation of vesicular glutamate and GABA transporters: A means to scale quantal size. Neurochemistry International. 2006;48(6-7):643-649
  6. 6. Chebib M, Johnston GA. The ‘ABC’ of GABA receptors: A brief review. Clinical and Experimental Pharmacology & Physiology. 1999;26(11):937-940
  7. 7. Olsen RW, Sieghart W. International Union of Pharmacology. LXX. Subtypes of gamma-aminobutyric acid(A) receptors: Classification on the basis of subunit composition, pharmacology, and function. Update. Pharmacological Reviews. 2008;60(3):243-260
  8. 8. Fritschy JM, Brunig I. Formation and plasticity of GABAergic synapses: Physiological mechanisms and pathophysiological implications. Pharmacology & Therapeutics. 2003;98(3):299-323
  9. 9. Lin SC, Bergles DE. Synaptic signaling between neurons and glia. Glia. 2004;47(3):290-298
  10. 10. Korpi ER, Grunder G, Luddens H. Drug interactions at GABA(A) receptors. Progress in Neurobiology. 2002;67(2):113-159
  11. 11. Vlainić J, Jembrek MJ, Strac DS, Pericić D. The effects of zolpidem treatment and withdrawal on the in vitro expression of recombinant alpha1beta2gamma2s GABA(A) receptors expressed in HEK 293 cells. Naunyn-Schmiedeberg's Archives of Pharmacology. 2010;382(3):201-212
  12. 12. Chebib M, Johnston GA. GABA-activated ligand gated ion channels: Medicinal chemistry and molecular biology. Journal of Medicinal Chemistry. 2000;43(8):1427-1447
  13. 13. Bettler B, Kaupmann K, Mosbacher J, Gassmann M. Molecular structure and physiological functions of GABA(B) receptors. Physiological Reviews. 2004;84(3):835-867
  14. 14. Heaney CF, Kinney JW. Role of GABA(B) receptors in learning and memory and neurological disorders. Neuroscience and Biobehavioral Reviews. 2016;63:1-28
  15. 15. Cryan JF, Kaupman K. Don't worry ‘B’ happy!: A role for GABAB receptors in anxiety and depression. Trends in Pharmacological Sciences. 2005;26(1):36-43
  16. 16. Bowery NG. GABAB receptor: A site of therapeutic benefit. Current Opinion in Pharmacology. 2006;6(1):37-43
  17. 17. Meldrum BS. Glutamate as a neurotransmitter in the brain: Review of physiology and pathology. The Journal of Nutrition. 2000;130(4S):1007S-10015S
  18. 18. Lesage A, Steckler T. Metabotropic glutamate mGlu1 receptor stimulation and blockade: Therapeutic opportunities in psychiatric illness. European Journal of Pharmacology. 2010;639(1-3):2-16
  19. 19. Paoletti P, Neyton J. NMDA receptor subunits: Function and pharmacology. Current Opinion in Pharmacology. 2007;7(1):39-47
  20. 20. Kleckner NW, Dingledine R. Requirement for glycine in activation of NMDA-receptors expressed in Xenopus oocytes. Science. 1988;241(4867):835-837
  21. 21. Lee MC, Ting KK, Adams S, Brew BJ, Chung R, Guillemin GJ. Characterisation of the expression of NMDA receptors in human astrocytes. PLoS One. 2010;5(11):e14123
  22. 22. Barco A, Bailey CH, Kandel ER. Common molecular mechanisms in explicit and implicit memory. Journal of Neurochemistry. 2006;97(6):1520-1533
  23. 23. Chen HS, Lipton SA. The chemical biology of clinically tolerated NMDA receptor antagonists. Journal of Neurochemistry. 2006;97(6):1611-1626
  24. 24. Lipton SA. Paradigm shift in neuroprotection by NMDA receptor blockade: Memantine and beyond. Nature Reviews. Drug Discovery. 2006;5(2):160-170
  25. 25. Koch HJ, Szecsey A, Haen E. NMDA-antagonism (memantine): An alternative pharmacological therapeutic principle in Alzheimer’s and vascular dementia. Current Pharmaceutical Design. 2004;10(3):253-259
  26. 26. Song I, Huganir RL. Regulation of AMPA receptors during synaptic plasticity. Trends in Neurosciences. 2002;25(11):578-588
  27. 27. Mayer ML. Glutamate receptor ion channels. Current Opinion in Neurobiology. 2005;15(3):282-288
  28. 28. Greger IH, Ziff EB, Penn AC. Molecular determinants of AMPA receptor subunit assembly. Trends in Neurosciences. 2007;30(8):407-416
  29. 29. Rosenmund C, Stern-Bach Y, Stevens CF. The tetrameric structure of a glutamate receptor channel. Science. 1998;280(5369):1596-1599
  30. 30. Platt SR. The role of glutamate in central nervous system health and disease—A review. Veterinary Journal. 2007;173(2):278-286
  31. 31. Evans AJ, Gurung S, Henley JM, Nakamura Y, Wilkinson KA. Exciting times: New advances towards understanding the regulation and roles of kainate receptors. Neurochemical Research. 2017 [Epub ahead of print]
  32. 32. Niswender CM, Conn PJ. Metabotropic glutamate receptors: Physiology, pharmacology, and disease. Annual Review of Pharmacology and Toxicology. 2010;50:295-322
  33. 33. Swanson CJ, Bures M, Johnson MP, Linden AM, Monn JA, Schoepp DD. Metabotropic glutamate receptors as novel targets for anxiety and stress disorders. Nature Reviews. Drug Discovery. 2005;4(2):131-144
  34. 34. Samardzic J, Svob Strac D. Benzodiazepines and Anxiety disorders: From laboratory to clinic. In: Durbano F, editor. New Developments in Anxiety Disorders. Rijeka: InTech; 2016. pp. 23-45
  35. 35. Jadzic D. Role of the Bed Nucleus of Stria Terminalis (BNST) in addiction and depression: a microdialysis study [PhD thesis]. University of Cagliari; 2018
  36. 36. Car H, Wiśniewski K. Similarities and interactions between GABAergic and glutaminergic systems. Roczniki Akademii Medycznej w Białymstoku. 1998;43:5-26

Written By

Janko Samardzic, Dragana Jadzic, Boris Hencic, Jasna Jancic and Dubravka Svob Strac

Submitted: 27 June 2017 Published: 21 March 2018