GABA Structure, Function, Physiology, and Dysfunction

From brainmatrix

This Page is being created.

GABA[edit]

Gamma-aminobutyric acid (GABA) is the principal inhibitory and abundant neurotransmitter in the mammalian nervous system (Figure 1).[1] Although GABAergic interneurons, each with distinct morphological, physiological, and expression patterns account for less than 30% of cells in the cortex, approximately 60 – 70% of synapses in the central nervous system are proposed to be GABAergic, and are considered crucial for information processing.[2][3][4] GABA signaling is essential for the maturation of the nervous system, synaptic plasticity, fine-tuning of neural circuits, sensory processing, and much more.[4][5][6] Significantly, disruptions in GABA neurotransmission are linked to multiple neurologic and psychiatric disorders those in turn fundamentally inform the development of pharmacological interventions to treat such disorders.

Figure 1. GABA Chemical Structure.(Public domain image accessed on November 25, 2022)

The GABA interneurons are widely distributed in the nervous system and, as mentioned mediate most of the inhibitory input to neurons, whereas glutamate, another ubiquitous neurotransmitter in the nervous system is responsible for majority of the excitatory input.[4][7] The maintenance of the optimal balance of excitatory/inhibitory (E/I) neurotransmission is important for proper cortical function and alterations in E/I equilibrium have been linked to various psychiatric diseases.[8] Of late it has become clearer that apart from dampening neuronal excitability, GABA signals serves to shape the response of the post-synaptic neurons and ensemble of neurons forming a coherent neural circuit.[9][10][11] In this vain, the GABAergic interneurons have been identified by notions of graph theory as unique 'operational hubs', dynamically controlling cortical network activity in their sphere and refining the flow of information beyond.[12] Mapping techniques have advanced this notion by constructing atlases of the anatomical and functional relationships of local and long-range GABAergic connections established between cortical and sub-cortical regions.[13][14]

The cortex has an abundance of phenotypically unique, non-overlapping populations of GABA interneurons. These cells maintain distinctive electrophysiological and biochemical features and are suited to encode pre- and post-synaptic events underlying different functional correlates of behavior. Evidence links behavior with the properties of distinct cellular populations, as in the anterior cingulate cortex (ACC), the somatostatin (STT)- and parvalbumin (PV)-expressing interneurons exhibit spatiotemporally distinct kinetics in response to different aspects of reward stimuli.[15] At the cellular level, dendrite-targeting GABA interneurons dynamically impact post-synaptic neurons by shaping their membrane oscillations and integrating their numerous synaptic inputs.[16][17] Furthermore, GABA interneurons with distinct identities, alter calcium signals in dendrites of prefrontal cortex pyramidal cells, with profound implications for cortical information processing.[18] GABA leverages multiple electrophysiological and biochemical modes of action in its arsenal to crucially contribute to the precise timing of neural signals, serving to engender, initiate and shape rhythmic activity in the brain. These interneurons dramatically and subtly influence activity by engaging in different modes of feedback and feedforward inhibition on different spatial and temporal scales.[19] - enabling them to have a dynamic influence on principal (pyramidal) neurons and on other interneurons and networks of cellular circuits. This endows GABA interneurons with the capacity to make meaningful contributions to pre- and post-synaptic events that underlie synaptic plasticity and learning.[15]

Advances combining noninvasive multimodal techniques have enabled examination of neural activity on broader, macroscopic levels. These techniques allow for changes in neurotransmitter levels to be studied in concert with cognate neurophysiological events that accompany functional states related to behavior or emotion. A recent meta-analysis of studies on healthy individuals, using proton magnetic resonance spectroscopy (1H-MRS) and task-related functional magnetic resonance imaging (fMRI) aimed to examine exactly such relationships. The study showed elevated GABA levels associated with reduced local neural activity in the anterior cingulate and medial prefrontal cortex during emotional processing, and in the occipital cortex during visual processing.[20] A similar local result was not observed in the case of glutamate, which exhibited a positive correlation with more distal brain regions. Although there is a long way to go, such neurochemical-neurofunctional investigations could augment other diagnostic tests, and prove significant in evaluating neurological and psychiatric disorders featuring changing metabolite levels.

GABA signaling undergoes a series of developmental shifts that correspond to the state of maturity of the developing nervous system, and the integral role played by GABA in numerous aspects of its growth. GABA signaling initially entails excitatory, depolarizing currents in the immature nervous system before transitioning to the classic inhibitory, hyperpolarizing currents later on in development.[21] The shift in conductance is a consequence of the altered expression patterns of chloride/cation transporters that induce the outflow of chloride ions from the neurons during early stages of development.[22] These shifts underlie the events of early neurogenesis, namely, proliferation, migration, differentiation, neurite extension, and synaptic plasticity. The developmental shift in signaling from excitation to inhibition also coincides with the advent of post-natal sensory input processing. Deviations in this developmental program have been thought to play a role in certain neurodevelopmental disorders.[21][23]

As GABA neurotransmission is crucial for normal function and its signaling is an essential component of diverse brain activities, GABA malfunction has been implicated in the etiology and pathophysiology of several neurological and psychiatric disorders, including epilepsy, anxiety disorders, stress, autism, schizophrenia, Parkinson’s disease, Huntington's disease, and amyotrophic lateral sclerosis (ALS) - some of which are elaborated on below.[8][24][25][26][27][28]

Moreover, GABA acts on peripheral non-neuronal tissue, and impaired functioning is associated with various disorders including, and not limited to hypertension and diabetes. GABA balance is considered beneficial for health, and it is touted for its anti-oxidant and antimicrobial properties. It is found naturally in foods such as green tea, soybean, and fermented foods such as kimchi and yogurt. Consequently, there is great interest in offering alternative, GABA based therapies in the form of GABA enriched foods or synthetic GABA compounds for various neurological and non-neurological conditions.[29]

Synthesis and Metabolism of GABA[edit]

The neurotransmitter GABA is not synthesized de novo, but is derived from its precursor glutamate by the action of the rate-limiting enzyme, glutamic acid decarboxylase (GAD), and involves what is described as the 'GABA shunt'[30] This shunt is part of the glutamate/GABA-glutamine cycle and connects the tricyclic acid (TCA) cycle to glutamate and GABA metabolism.[31] The cycle entails the release of the neurotransmitters’ glutamate and GABA from pre-synaptic neurons, and the subsequent uptake and processing in neurons and astroglial cells.[32] In the glial cells, the metabolic pathway invokes the mitochondrial TCA cycle and forms the shunt for synthesis of glutamine that is returned to the neurons for conversion to glutamate or GABA, depending on the identity of the neuron.[33]

GABA released from the terminals into the synaptic cleft is not subject to enzymatic degradation and is predominantly cleared by diffusion or by uptake via specialized transporters. These transporters that belong to a neurotransmitter:sodium symporter family, recycle the neurotransmitter back into pre-synaptic neurons or transport it into astroglial cells, the latter representing a minor pathway for GABA replenishment.[34] In humans, these transporters belong to what is referred to as the solute carrier 6 (SLC6) family, and encompass 20 (A1-A20) transporter proteins, that are differentially targeted to neurons or glial cells. Four groups of GABA transporters (GAT), that belong to the SLC6 family have been identified, GAT1-3 (SLC6A1, SLC6A13, SLC6A11), and betaine/GABA transporter 1 (BGT1/SLC6A12), the latter a focus of epilepsy. Although the transporters overlap in their distribution, GAT1 is considered predominantly neuronal, whereas GAT3 is principally glial.[35]

Following the reuptake of GABA into the presynaptic neuron terminals, the neurotransmitter is repackaged into synaptic vesicles by vesicular GABA transporter (VGAT), ready to be rereleased.[36] Another mode of enriching GABA in synaptic vesicles includes the uptake of glutamate into the terminals by excitatory amino acid transporters (EAAT1), the enzymatic conversion by GAD65 to GABA, and loading of the newly synthesized transmitter into synaptic vesicles, ready for release.[37]

The astrocyte mode of recycling GABA involves the glutamate-glutamine cycle and is significantly associated with the subsequent inhibitory events of GABA interneurons, especially at active synapses.[38] This pathway is ultimately purported to provide a substantial amount of GABA required at the synapse for maintaining sustained inhibitory activity. The astrocytic-neuronal - glutamate-glutamine-GABA cycle, dynamically links the excitatory glutamate pathway directly with the activity of the GABA neurosystem, ensuring the availability of loaded synaptic vesicles for release.[38]

The recycled GABA, both in neurons and astrocytes, is catalytically converted to succinic semialdehyde by the enzyme GABA-transaminase (GABA-T), which is further processed into succinic acid by succinic semialdehyde dehydrogenase (SSADH). Succinic acid then enters the TCA cycle to form α-ketoglutarate, that is ultimately converted into the GABA precursor, glutamate.[39] The transamination of GABA by GABA-T to subsequently form glutamate requires α-ketoglutarate to be in a position to accept the amino group generated from GABA processing. In this way the GABA supply is sustained, and adequate levels of the neurotransmitter are maintained for optimal function. The astroglial glutamate is catalyzed by glutamine synthetase (GS) to form glutamine that is released for uptake into the presynaptic terminals of GABAergic neurons. In GABA neurons, glutamine is converted to glutamate via the enzyme, phosphate-activated glutaminase (PAG), and subsequently glutamate is catalytically decarboxylated by GAD, in an irreversible reaction that generates GABA.[40]

Glutamic acid decarboxylase, is an enzyme unique to GABAergic neurons and requires the cofactor pyridoxil phosphate (vitamin B6) for its catalytic activity. Two distinct isoforms of the enzyme, GAD65 and GAD67 have been discovered and are encoded by separate genes, exhibiting discrete functional properties, subcellular localization, and cofactor requirements. GAD65 is targeted to synaptic vesicles, being primarily deployed in neurotransmission, whereas cytoplasmic GAD67 is active and supplies GABA as a trophic factor during synaptogenesis, and injury protection.[40][41][42]

Pathological states associated with auto-antibodies against GAD65 lower GABA levels, and have been linked with certain disorders, including diabetes mellitus type 1 (DM1) and certain neurological diseases. Patients with high anti-GAD65 antibodies suffering from motor and cognitive disorders manifest an incomplete and variable response to immuotherapy.[43] These lines of investigation into anti-GAD65 syndromes and other pathologies linked to GABA metabolism need to be extended in order to gain a firm grip on the basis of these neurologic conditions, alleviate symptoms and provide relief to those suffering from these GABA-associated disorders.

The synthesis and metabolism of GABA regulates the levels of the transmitter available for signaling and GABA concentration has a dynamic impact on network activity and maintenance of excitatory and inhibitory homeostasis. The various enzymes, transporters, and proteins involved in these mechanisms have a powerful impact on the effectiveness of GABA function, and ultimately the functioning of the entire nervous system. In this regard, inadequacies in processing are implicated in various physiological, neurological, and psychiatric diseases, and are the targets of intense investigation for their therapeutic potential.

GABA Interneurons[edit]

Interneurons, especially inhibitory interneurons make up approximately 20-30% of neurons in the adult neocortex, and their numbers vary with respect to particular cortical layers, regions, and species. Cortical GABAergic interneurons are critical for development, plasticity and normal function, and their complexity is increasingly compounded by their diversity. These inhibitory interneurons differ in many respects and are remarkable for their unique morphological and physiological features, intrinsic membrane properties, their engagement at the synaptic level, and their molecular identity.[13][44][45][46]

Majority of these neurons have characteristic aspiny dendrites and smooth appearance attributable to inhibitory neurons. Decades of research along with the refinement of technical and analytical tools have elucidated the diversity of GABAergic interneurons.[47][48] They are distinguished into particular types based on differences at the level of cell bodies, dendrites, and axons.[4][49] Their cell bodies receive inhibitory and excitatory inputs and are integral components of neural circuits, serving to meter the excitatory drive and contribute to the appropriate synaptic output.[48]

Arguably, it is the axon extent and arborization that is a more reliable identifier of a particular interneuron type than the pattern of its dendritic branching.[4] Neocortical GABAergic neurons are ‘local circuit neurons’ that maintain a distinct columnar and laminar organization, with dendrites extending within and across columns, and with axons arborizing within the neocortex. They remain confined within the neocortical architecture, remaining true to their local aspect and refrain from elaborating neurites to sub-cortical regions.[3] The different subtypes of GABA interneurons specialize in targeting different neuronal compartments to influence cell function. They form synapses at proximal and other aspects of dendrites, cell bodies, or on the axon initial segment.[3][4][44][50] This targeting affords them the capacity to determinedly sculpt the spatiotemporal response generated by the target neuron and by extension, the whole assembly of neurons in a circuit.[10][51] At present there are estimated to be approximately 50 types of GABAergic neurons identified in the cerebral cortex.[52]

GABA interneurons exert their inhibition on principal pyramidal neurons, and in turn these neurons remain connected with other circuit principal neurons and interneurons in order to maintain E/I balance.[19] These interactions and the activity contained in these circuits drives the local response and the communication between different brain regions. Disturbances of these finely tuned physiological patterns of brain activity has been linked to various disorders, including epilepsy and schizophrenia.

The inhibitory medium spiny neurons of the basal ganglia, are a GABAergic population of neurons that are uniquely positioned to influence motor function.[53] The direct pathway from the cortex that loops through the basal ganglia serves to enhance motor activity by releasing GABA inhibition of the thalamocortical motor circuit. In contrast the indirect pathway serves to suppress motor activity by amplifying GABA inhibition of the thalamocortical motor circuit. Dysfunction in this system is implicated in the motor impairments observed in and not limited to Parkinson's and Huntington's diseases. In Huntington's disease, the GABAergic striatal medium spiny neurons are 'selectively' vulnerable to degeneration early on and contribute to the progressive motor and cognitive dysfunction.[28]

Phenotypically the interneurons are differentiated into several subtypes that each concentrate different neuropeptides, calcium binding proteins, and receptors, including distinct parvalbumin (PV), somatostatin (STT), and ionotropic serotonin receptors (5-HT3A), vasoactive intestinal peptide (VIP) populations, that altogether account for 85 % - 100% of all interneurons in the cortex.[44][45][54][55] Different interneuron types differentially target and release GABA and neuropeptides at specific post-synaptic sites and times in accordance with the requirements of behavior - fine tuning and tailoring their signals to the specific activity related to movement or alternatively, maintaining homeostasis [56] Recently the PV+, STT+, and VIP+ interneuronal populations in the mouse somatosensory cortex were mapped to determine the extent and type of cortical input received. These kinds of investigations reveal that different interneuronal populations are determined to leverage feedback and feedforward mechanisms in order to modulate the dynamics of their embedded local networks.[13]

Based on decades of accumulated data on GABA interneurons the Petilla terminology was adopted in 2008 to enumerate these cell types owing to their morphological, physiological, and molecular characteristics.[57] The organizing principle of this nomenclature takes into account the targeting by these interneurons of pyramidal cells or other interneurons. GABA interneurons are further distinguished by their terminations on different aspects of the pyramidal cells. Specifically, interneurons that form synapses at the axon initial segment and comprise the axo-axonic chandelier cells; those that connect with the soma, such as basket cells; and interneurons that target dendrites, either the dendritic shaft of neurons that go on to elaborate axons vertically, descending to the white matter, or those like the Martinotti cells whose axons ascend to the pia, and lastly the neurogliaform cells that preferentially target the dendritic spine.[58]

The classification scheme further differentiates GABAergic interneurons based on their expression of certain molecular markers. These consist of the expression of PV, STT, 5-HT3A receptor/VIP, neuropeptide Y (NPY), and cholecystokinin (CCK). They are further categorized depending on whether each type colocalizes other peptides, and their complement of ion channels and receptors, structural and synapse related proteins, calcium binding proteins, transcription factors, among other distinguishing features.[46] It is now evident that there is a distinct class of ionotropic serotonin 5-HT3A receptor expressing neocortical interneurons that based on their independent lineage do not overlap with the PV- or STT- interneuron populations. A significant subset, ~ 40%, of these interneurons accumulate VIP.[45][54]

Recent studies conducted on transgenic mice coupled with optogenetic investigation of distinct, non-overlapping populations of PV, STT, and VIP expressing interneurons, have shed light on the contributions of these cells to neural circuits. Notably, the fast-spiking, PV-expressing interneurons such as the basket and chandelier cells, target the cell body and axon initial segment of other neurons, respectively. Additionally, their fast and sustained firing characteristics puts them in a dynamic position to impact the electrical activity of target neurons.[59] Moreover, the interconnected cohort of PV-expressing interneurons have an equally strong impact on driving the synchronous activity of neurons.[3][9][58] The signals from this select , PV+ subset of GABAergic interneurons influence the ensuing events considered consequential for cognitive function and behavior, and dysfunction in this process could lead to impairments associated with neurological and psychiatric diseases.[60]

STT-expressing interneurons, including Martinotti cells project axons superficially in the cortex to influence the output related to sensory information processing.[61] The axons of these interneurons target the apical dendrites of deeper, cortical layer 5, pyramidal neurons. Using fiber-optic recording methods, these layer 1, STT+ interneurons have been observed to inhibit the calcium spikes and bursting in the dendrites of the pyramidal neurons, a result also recorded in the hippocampus.[51] The STT+ interneurons di-synaptically inhibit adjacent neurons, enabling them to powerfully control the activity of the micro-circuit that is responsible for the bottom-up encoding of sensory information.[61] In the hippocampus, STT containing interneurons target the dendrites of CA1 pyramidal cells and their slow time course of action mediated by the α-5 GABAA receptors is suited to and influences the post-synaptic NMDA receptor activation kinetics. In this way the STT-expressing GABA interneurons control the synaptic integration and excitability of the post-synaptic pyramidal cells, proving to be an attractive site for synaptic plasticity.[17] In contrast to their inhibitory effects, STT+ interneurons exert their influence on cortical microcircuitry and cortical information processing by disinhibiting the thalamic afferents to pyramidal neurons in layer 4 of the cortex, the critical input layer in sensory cortices that is important for information processing. The STT-expressing GABA interneurons have been classified into 5 types with distinct morphological and molecular features and laminar location in the human cerebral cortex.[62] Of note is that in Alzheimer's disease, there is a selective loss of STT+ versus PVV+interneurons in the temporal lobe of the cortex, an area crucial for cognitive function and a predominant deficit seen in people with this disorder.[63] This just underscores the significance of well balanced interneuronal circuits in maintaining optimal function.

The third class of GABAergic cortical interneurons with multiple morphologic features are superficially located, and express 5-HT3A receptor and VIP. They have a disinhibitory role of principally inhibiting STT- and a segment of PV-expressing interneurons. In sensory cortices, such as the visual cortex, the mutual, reciprocal antagonism exerted by VIP and STT interneurons contributes to the appropriate behavioral output, especially in response to weak stimuli.[64] The VIP-expressing interneurons influence on local circuit activity, i.e., input and output of cortical pyramidal cells, is principally brought about by inhibiting the inhibitory STT and PV interneurons and is observed across many other regions of the cerebral cortex. This conserved feature has been demonstrated to be relevant for pyramidal cells response to reward-punishment, which constitutes a reinforcement signal and could be crucial for certain forms of learning.[65] This subset of GABAergic-VIP interneurons plays an important role in different behavioral states, as it exerts inhibition and disinhibition on GABAergic and other neurons to modulate the activity of distinct circuits.[66] STT+ and VIP+ interneurons in the somatosensory cortex have been found to be active early in the development of sensory processing, individually contributing to the activity that continues to mature into adulthood.[67]

It is evident that the diversity of GABA interneurons is a salient and indispensable feature of the nervous system, making important contributions to cortical functioning and resultant behavioral states. At the individual level, these interneurons serve to temper the excitability and shape the response patterns of the pyramidal cell that can be tied to behavioral states. Zooming out to the network level, interneurons influence the formation of neural circuits that are crucial for propagating local activity and spreading excitation in the form of membrane oscillations. This architecture enables local interneurons to exert their actions at the network level, constantly refining cortical circuits, and providing a mechanism for synaptic plasticity and sensory processing.[5] The above-described classification scheme lays bare the intriguingly complex connectivity and heterogeneity of these cell types, especially as there are about 50 identified GABA interneurons types as of now. Furthermore, investigations in humans of different GABA receptor subunit gene expression, combined with unique interneuron subtypes biomarkers, and receptor binding assays, demonstrate a correlation between certain subsets of inhibitory GABA interneurons and their receptors.[68] Comprehensive, noninvasive studies, including but not limited to PET, optogenetic studies, (1H) MRS, of these kinds on humans shine a light on the identity of unique markers of diverse cortical GABA circuits, and how dysfunction at this level could have ramifications for neurological and psychiatric diseases.

GABA Receptors[edit]

GABA is critical for coordinating the activity of neurons at the individual and circuit level, and it exerts its inhibitory effects via two subtypes of receptors, namely GABAA and GABAB receptors.

Ionotropic GABAA Receptors[edit]

Figure 2. 3D structure of GABAA Receptors. The upper image depicts the extracellular domain of the receptor with the transmembrane domains shown traversing the plasma membrane. The bottom image shows the different subunits (individually colored), surrounding the central pore for conducting chloride ions, and the binding sites for GABA and benzodiazepines. (Public domain image accessed December 20, 2022)

GABAA receptors are ligand-gated ion channels.[69] The receptors are integral membrane proteins composed of 5 subunits that form a pore for regulating the passage of chloride into the cell (Figure 2). Alternative splicing of 20 gene products generates different types of subunits, i.e., six alpha (1-6), three beta (1-3), three gamma (1-3), three rho (1-3), and single delta, epsilon, pi, and theta subunits.[25][70] Several classes of pharmacological agents bind to the receptor, and the combination of subunits forming the receptor define its functional properties.[70][71][72] Majority of mature receptor isoforms in the mammalian nervous system are formed by three of these subtypes and are proposed to assume a combined stoichiometry of α1β2γ2 subunits.[73][74] Furthermore, the GABAC ionotropic receptor isoform is composed of rho subunits and is considered to be part of GABAA family of receptors.[75]

GABAA receptor subunits bear structural homology with other ligand-gated ion channels. Each subunit is composed of 4 transmembrane domains (TMD1-4), with GABA and psychoactive drugs such as benzodiazepine binding to sites on the extracellular N-terminal domain. The second TMD forms the pore for chloride ions, and the intracellular domain (ICD) represents sites for protein and other modulatory interactions. The long intracellular loop between TMD 3 and 4 binds to the scaffold protein gephyrin, that is the chief architect for organizing and maintaining the inhibitory post-synaptic density at the post-synaptic membrane.[76] The receptor has distinct sites for binding neurosteroids and sedatives, such as barbiturates.[70][77][78] The cell surface expression and turnover of GABAA receptors at synaptic sites is integral to robust inhibitory signals. Disruption of the processes that underlie the appropriate distribution of the receptors and intracellular recycling and turnover leads to perturbations of inhibitory neurotransmission. This perturbation in GABA inhibition alters the E/I balance and the resultant increase in excitatory tone is commonly linked with various disorders of the nervous system.[79]

The diverse subunit composition of GABAA receptors confers differential sensitivity to distinct endogenous neurochemicals and pharmacological agents for modulating neuronal excitability. The subunit differences and distinct regional distribution patterns and cellular localization of the receptor enables it to have a profound influence on brain function. It further defines the post- or extra-synaptic location of the receptor that mediate phasic or tonic inhibition, respectively.[72][78][80][81][82] Importantly, transient (< 1ms), phasic inhibition is a feature of classic, point-to-point neurotransmission at post-synaptic receptor sites that activate at higher (0.3-1mM) concentrations of the GABA in the synaptic cleft. This activity answers to benzodiazepines. Whereas the persistent, tonic form results from longer lasting modulation of inhibitory post-synaptic currents at extra-synaptic sites, requiring lower (μM) levels of the neurotransmitter to activate.[81] The latter underlie more than just post-synaptic excitability but extend to oscillations at the network level, and are important for synaptic plasticity, neurogenesis, and cognitive functioning.[83] Tonic GABAA receptor neurotransmission is the predominant form of GABA signaling in the brain and neurosteroids are known to be potent modulators of these high affinity extra-synaptic receptors.[84] Neurosteroids are positive allosteric modulators (PAM) of extra-synaptic delta (δ)- subunit containing GABAA receptors. They enhance the tonic inhibition mediated by these receptors and serve as powerful anticonvulsants.[78] Currently, research on surgically resected human cortical tissue from patients with various pathologies has confirmed the presence of GABAA mediated tonic inhibition.[85]

The disruption in either phasic or tonic forms of inhibition is consequential for proper development and functioning and has been implicated in many neurological and psychiatric diseases. Pathological disease states and disorders such as epilepsy, stress, anxiety, schizophrenia, Parkinson’s disease, recovery following a stroke are all associated with dysfunction of GABAA receptor.[70][78][83][86]

GABAA receptors have been shown to modulate the activity of excitatory cortical circuits by providing inhibitory inputs to influence the timing of spikes and the persistent activity of the network.[87] There is considerable interest in further developing selective therapeutic agents based on the diversity of GABAA receptors and their impact for neuroprotection, maintenance of E/I balance, and ramifications for diseases.[70][71]

Metabotropic GABAB Receptors[edit]

Figure 3. 3D Structure of the GABAB G-protein coupled receptor. Represented is the inactive receptor dimers. (Public domain image accessed December 20, 2022)

The metabotropic GABAB receptors belongs to the 7 transmembrane domain class C G protein-coupled receptor (GPCRs) superfamily.[88] They play an important role in modulating neuronal and circuit excitability in the nervous system. These receptors are obligate heterodimers of GABAB1 (GB1) and GABAB2 (GB2) subunits, and heteromeric interactions between the two subunits is essential for the formation of a functional receptor (Figure 3).[89][90][91][92][93] Each subunit has the extracellular Venus Flytrap (VFT) domain - characteristic of this class of receptors - wherein lies the ligand binding site.[94][88] Whilst the GB1 subunits have been shown by radioligand binding and site-directed mutagenesis studies to bind with high affinity to GABA; allosteric compounds act on the GB2 subunit, which is crucial to intracellular coupling with G-proteins.[95][96]

These receptors influence neuronal signaling by either inhibiting the release of neurotransmitters at presynaptic terminals, including as auto- or heteroreceptors on GABA or glutamate nerve terminals, respectively, or activating K+ channels at post-synaptic sites. GABAB receptors mediate slow and prolonged inhibition via intracellular activation of Gai/o protein cascades which include the inhibition of adenylate cyclase, activation of inwardly-rectifying K+ channels (GIRK), and inactivation of voltage-gated Ca2+ channels (Cav).[97] The GABAB receptor inactivation of Cav channels at presynaptic terminals serves to interfere with neurotransmitter release, which is significantly consequential for neurotransmission.[98] It follows that pre-synaptically located GABAB receptors inhibit glutamate release.[99][100] Post-synaptic membrane excitability is altered by the enhanced conductivity of GIRK channels induced by GABAB receptor activation. The subsequent K+ efflux through these channels, results in membrane hyperpolarization that underlies the slow and prolonged inhibitory post-synaptic potentials (IPSP) attributed to GABA signaling.[101] Tonic activation of extra-synaptic GABAB receptors by other interneurons in the vicinity occurs in response to GABA spillover, i.e., escaping of the the neurotransmitter from the immediate post-synaptic site. This activation of the GABAB receptors is thought to contribute the important inhibitory component of hippocampal rhythmic activity.[102] In contrast to others GPCRs, these receptors remain stable at the plasma membrane, resisting internalization even after agonist exposure.[103] The absence of endocytosis of GABAB receptors until their subsequent degradation in the natural course of events, correlates with lack of arrestin recruitment and with enhanced stability after cAMP- mediated PKA phosphorylation.[103] Their membrane stability is consistent with the importance role they have in moderating neuronal excitability, and contributing to rhythmic activity. Interestingly, it is glutamate and not GABA that has been tied to the turnover of GABAB receptors, as it decreases the number of receptors at the plasma membrane and initiates trafficking through the lysosomal pathway.[104]

GABAB receptors signal transduction is augmented by interactions with other proteins, such as scaffold, trafficking proteins, enzymes, ion channels, GPCRs, including glutamate receptors.[97][105] The GABAB receptors regulates the ionotropic glutamate NMDA receptor (NMDAR) function indirectly by controlling glutamate release from pre-synaptic terminals and directly by influencing NMDAR calcium signaling at post-synaptic sites. The post-synaptic NMDAR calcium influx, important for membrane excitability, is inhibited by GABAB receptors, however the glutamate receptor retains the ability to respond to NMDA with synaptic currents.[106] The hyperpolarization and diminished current attributed to the activity of GIRK channels, is conducive to maintaining the Mg2+ block of NMDAR channels. The GABAB receptors also inhibit the activity of PKA, which is thought to be important for NMDAR calcium signaling.[106] Furthermore, as mentioned above, GABAB receptors also function to attenuate the presynaptic release of glutamate by suppressing calcium entry directly through voltage-gated calcium channels.[99] The activity of and calcium signaling by NMDARs is essential for neuronal excitability, neuroplasticity, and dendritic spine remodeling. Therefore, widespread prevalence of GABAB receptors has broad implications for calcium signaling in the nervous system.

Dysfunction of NMDARs is implicated in neuropathologies and neuropsychiatric diseases, such as pain, autism, and schizophrenia. GABAB receptor agents have been shown to reverse certain deficits of the glutamate receptor, and thereby could be suitable therapeutic targets for treating NMDAR-dysfunction related disorders.[107] Additionally, GABAB receptors interact with ionotropic AMPA receptors, providing another suitable target for these agents.[97] The activation of GABAB receptors serves to inhibit the persistent activity of cortical circuits.[87] The excitatory-inhibitory E/I balance in the nervous system is mediated by both these neurotransmitters, and imbalances contributing to diseases in humans, like schizophrenia, are being studied in post-mortem tissue for the development of biomarkers.[26]

The GABARs form complexs with metabotropic glutamate receptors (mGluRs) at the plasma membrane, and this higher-order GPCR complex is functionally regulated in a manner distinct from the regulation of the individual receptor unit.[108] The two receptors are known to be colocalized at peri-synaptic sites of dendritic spines, which enables them to interact, and is conducive to cross-talk between them. These effects have a significant impact on neurotransmitter release, neuronal and circuit excitability, synaptic plasticity, neurogenesis. GABAB receptors interact with the glutamate system in the hippocampus to influence spatial learning and cognition, functioning that is impaired in certain psychiatric disorders.[97]

The various mechanisms at play alter the pharmacology and functioning of GABAB receptors and offer up a unique palette of sites for receptor modulation and drug development for therapeutic purposes.[96] Currently, the only GABAB receptor compound in use is the agonist baclofen, which acts at both pre- and post-synaptic sites in the central nervous system.[109] It is approved for use as an antispastic agent and in treatment of multiple sclerosis and spinal cord injury, and is used off-label as a muscle relaxant.[110]

GABA Disorders[edit]

Its is evident that the proper functioning of the brain and nervous system on the whole relies on balanced E/I signaling. Dysregulation at the levels of GABA and glutamate is considered central to the etiology of various neurological and neuropsychiatric disorders. Integrating the the realm of imaging, which is continually evolving, with more refined technologies such as proton magnetic resonance spectrography (1H MRS) that samples neurotransmitter concentrations in vivo, are proving to be exciting and invaluable resources for providing insights into the neuroanatomical and neurochemical correlates of brain dysfunction. These applications combined with behavioral, neurophysiological, and molecular approaches is adding to our understanding of the very reasons for these complex disorders, and potentially opening up interventional approaches that have previously not been identified. Additionally, more potent and selective drugs are being recruited for their novel and strategic therapeutic potential to treat disorders related to GABA dysfunction.[71] Impairment in GABA signaling is crucially implicated in many disorders, as discussed below.

Depression[edit]

GABA neurotransmission is critical for cortical synaptic plasticity and maintainance of E/I balance as it impacts the internal cortical circuitry and output of its principal, pyramidal neurons. Depression has been associated with a deficit of GABA functioning in the brain at multiple levels, from reduced concentrations of the neurotransmitter to structural and functional alterations of GABAA receptors, and the ability of available anti-depressants acting via the GABAergic system to mitigate depression.[111] The rapid effects of anti-depressants that target either monoamines, glutamate or acetylcholine signaling in order to ameliorate depressive symptoms are though to act by altering localized GABA activity.[112][113] In this regard, the NMDA receptor subunit GluN2B on GABA interneurons in the medial prefrontal cortex, appears to be the preliminary site of action of the rapid-acting antidepressant ketamine.[114] GABAA receptors are of particular interest in the treatment for depression, and the neurosteroid allopregnanolone along with other positive allosteric modulators (PAM) of the receptor are being deployed for treatment.[115] The extra-synaptic GABAA receptors located on pyramidal cells in cortical and subcortical hippocampal sites - regions implicated in depressive and anxious states, respectively - are targets of GABAA-receptor selective positive allosteric modulator antidepressants (GASPAMA).[116][117]

There have been reports from multimodal studies contradicting the GABA-deficit hypothesis, however these studies have different paradigms, cohorts and drug regimes that preclude any definite synthesis of results. GABA concentrations have been found to be unchanged in patients at risk of developing major depressive disorder (MDD) suggesting they cannot be considered a trait marker for the eventual development of the disorder.[118] The regions evaluated in patients at risk of developing major depressive disorder (MDD) using combined investigational approaches, including 1H MRS, have comprised the prefrontal and parito-occipital cortex, preoptic area, and other regions of the brain.

It is clear that despite the challenges, the novel GABA agents currently in use or in the pipeline to treat depression, are based on a better understanding of GABA and its crucial role in neurotransmission, and are proving to be beneficial in management of depression.

Stress, Fear, and Anxiety[edit]

Low levels of GABA are associated with stress, fear, and anxiety. The amygdala in the temporal lobe is a structure central to the processing of unpleasant, stressful, and fearful stimuli.[119] GABA neurons in the amygdala are key components in maintaining the stability of excitatory and inhibitory neurotransmission.[16] The amygdala is composed of numerous nuclei, and amongst its subdivisions is the basolateral amygdala (BLA), a structure crucial for processing emotional and motivational states, and central amygdala (CeA), that is entirely gabaergic and critical for processing of stressful and fearful stimuli. A comparatively modest set of GABA interneurons in the BLA balances out the glutamatergic excitation, and this group provides input to the GABA interneurons of the CeA that exit the amygdala.[120] GABA neurons in the amygdala are influenced by a diversity of afferent neurotransmitter systems, arriving from both cortical and subcortical regions that serve to facilitate GABA signaling.[121] GABA interneurons are uniquely placed to efficiently influence the circuitry and function of the amygdala, and importantly to effectively contribute to the behavioral responses in stressful circumstances.

Disruption of BLA GABAergic interneurons, including cellular loss, reduction in the activity of the GABA synthesizing enzyme (GAD65/67), or dysfunction related to the activity of GABAA receptors, undermines the E/I balance and underlies states of hyperexcitability.[121] This excitability presents as anxiety and other behavioral and emotional disturbances, including seizures. Dysregulation of these neurons and the accompanying anxiety can have longer term consequences for the development of psychiatric post-traumatic stress disorders (PTSD), and neurodegenerative diseases such as Alzheimer’s disease.

Furthermore, GABA has an inhibitory effect on the hypothalamic release of corticotropin-releasing hormone (CRF) and vasopressin that are involved in the activation of the stress-related, hypothalamo-pituitary adrenal (HPA) axis. Reduced GABA in pathological states removes the inhibition that keeps the HPA system in check. Subsequently, the over-activity associated with the HPA axis is associated with, and not limited to stress and anxiety.[122]

Treatment for stress disorders include benzodiazepines, which are limited in their utility based on the delineation of the underlying physiological determinant of the dysfunctional response, a limitation which would obviate their effectiveness. Impairments such an underlying loss of BLA GABAergic interneurons or GAD activity responsible for bringing about the dysfunctional stress response, would prove refractory to treatment with benzodiazepines.

Neuropsychiatric Diseases[edit]

Schizophrenia[edit]

The neural substrates of working memory, are underpinned by activity of both glutamate and GABA acting in concert, and presents as gamma oscillations in the dorsolateral prefrontal cortex (DLPFC). Working memory and other cognitive deficits are a defining, negative feature of schizophrenia.[123][124] Alterations in GABA transmission in the supragranular cortical layers, especially the circuits involving the deep layer 3/4 PV+ neurons, and the more superficial layer 2/3 and SST+ neurons that form synapses with layer 3 pyramidal cells, are responsible for these cognitive deficits.[125] Oscillatory activity of DLPFC pyramidal neurons is synchronized by the inhibitory synaptic input of GABA interneurons - pyramidal interneuron network gamma (PING) model.[26] These gamma oscillations (30-80Hz) that underpin working memory are observed to be compromised in patients with schizophrenia.[126][127] Thus, the paucity of cortical GABA signaling contributes to this negative symptom of schizophrenia, and noninvasive tests could identify aberrant cortical gamma activity and serve as a biomarker for schizophrenia.

Certain subpopulations of GABA interneurons, like PV+ basket cells, may be selectively vulnerable in schizophrenia.[128] Postmortem tissue analysis has consistently shown a decrease in the mRNA levels of the enzyme GAD67 in the presynaptic terminals of especially PV+ and SST+ GABA interneurons, pointing to a reduction in the synthesis of GABA and robustness of neurotransmission.[126][128] In accordance with this, the activity of the post-synaptic GABAA receptors targeted by these interneurons is also weakened due to lowered alpha subunit levels. Apart from this, lower levels of SST+ neurons in the DLPFC are also observed and could contribute to the microcircuits that underlie the oscillatory activity underpinning the deficits in the working memory of schizophrenia patients. Agents that selectively and positively modulate the activity of these receptors have been shown to positively impact cortical activity and task performance in patients with schizophrenia.[127][129]

Multimodal studies, combining 1H MRS to determine the concentration of GABA, have revealed regional reductions in GABA in the frontal and occipital cortex, and basal ganglia.[118] These studies on their own do not indicate a causal role of GABA in schizophrenia, but suggest a perturbation in E/I balance - a subject of many current investigations.

Epilepsy[edit]

Epilepsy bestows an enormous burden of disease to the world, afflicting 50 million people globally and 3.4 million people in the US alone.[130][131] Temporal lobe epilepsy is the most common form of partial epilepsy, accounting for 60% of adult cases, and is characterized by recurrent and unprovoked focal seizures of the temporal lobe.[132] GABA neurotransmission is held as a counterbalance to the excitation brought about by glutamate signaling, and both are tasked with maintaining E/I homeostasis.[133][134] Valproate, a broad spectrum drug in wide use to treat epilepsy is believed to act mechanistically by potentiating GABA inhibition and reducing glutamate/NMDA receptor mediated excitation.[135] The consequences of GABA signaling in maintaining E/I balance is more significant and richer than just its actions assigned to inhibition, and dysregulation of the system has implications for generation of epileptic seizures.[136][137] The diversity and dysfunction of GABA interneurons at the network, electrophysiological, subcellular, and molecular levels contributes to the complex nature of epileptogenesis.[138] However, GABA signaling has recently been revealed to both promote and suppress epileptic seizures.[137][139] Additionally, astroglial cells play a critical role in the metabolic changes associated with the E/I balance of GABA and glutamate signaling. Long-lasting impairment of these cells as a consequence of epileptogenic changes could lead to alterations in the astrocyte-neuron network and could perturb the E/I homeostasis.[140]

Disrupted GABA inhibition perpetuates and intensifies the self-sustaining epileptic seizures, with attendant behavioral and neurological disturbances. GABAA receptors are considered to be mainly responsible for these seizures, and positive allosteric modulators (PAM), benzodiazepines, that target the γ2 subunits of these receptors are deployed as the first line of treatment.[141] These receptors mediate the phasic, fast inhibition in response to high GABA concentrations attributed to synaptic GABAA receptors. Cationic ion channel transporters such as the K+-Cl- cotransporter isoform 2, KCC2, and Na+-K+-2Cl- cotransporter isoform 1, NKCC1 are critical for maintaining the ionic gradients that define the flow of ion currents through the receptor ion channels. These ion channels and receptors are subject to immediate and pathological seizure-related post-translational changes, and this has an impact on their availability and function, and consequentially this 'ion plasticity' is responsible for changes in membrane excitability.[139][142] These rapid changes at the membrane level are also consequential for the time-limited responsiveness to benzodiazepines, as they are rendered ineffective in treatment of prolonged seizures.

Following precipitating events of epileptogeneisis, such as stroke, traumatic brain injury (TBI), seizures, hypoxia, transcriptional activity alters the levels of receptor subunits expressed. These transcriptional changes brought about by growth factors such as BDNF and other transcription factors like CREB, alters the targeting of GABAA receptors and changes the levels of post- and extra-synaptic receptors, resulting in a shift toward tonic signaling.[143] These changes in subunit composition renders the receptor less sensitive to benzodiazepines, and may be a factor that should be considered in treating epilepsy. Tonic inhibition is preserved with the changes brought about by epileptic seizures, and it serves to constrains excitatory network activity. This mode of GABAA receptor signaling may well represent targets for drug interventions in the treatment of epilepsy.[83] Of note is the neurosteroid sensitivity of extra-synaptic GABAA receptors. Eplileptogenesis has numerous causes, symptoms, prognosis and treatment, and as such has certain common, clearly defining and more individually unique features.

Pain[edit]

GABA signaling and plasticity are relevant to the development of neuropathic and other painful conditions.[144] GABA neurons and GABA receptors are found in all regions of the brain and spinal cord essential for mediating the perceiving pain. Under noxious circumstances disturbances in GABA neurotransmission disrupts the glutamate/GABA mediated E/I homeostasis.[145] It has long be considered that central sensitization of excitatory signals in the dorsal horn of the spinal cord, that receives inputs from the peripheral sensory receptors of the primary afferent fibers, is important for pain processing.[146] However, it is abundantly clear that deficits in inhibitory GABA signaling is crucial for the generation of neuropathic and other pain, and that agents that activate GABAA receptors can be leveraged for treatment of these pathological conditions.[147] Pain- and damage-related alterations in GABA function, extend from paucity of neurotransmitter synthesis to neuronal apoptosis in the dorsal horn, some portion of the latter presumably from excitotoxic insult.[148][149][150] Plasticity of GABA interneurons is observed along the neuraxis, in the descending nociceptive pathways, from the receptors in the periphery to the dorsal horn of the spinal cord, and higher, central pain processing centers. The plasticity attributed to GABA interneurons expands its repertoire of contributions to neural circuits, making it amenable to incorporate an effective and adaptive response to healthy physiological demands and those placed by noxious stimuli.

References[edit]

  1. Roberts, Eugene; Frankel, Sam (1950). "γ-Aminobutryic acid in Brain: Its formation from glutamic acid". Journal of Biological Chemistry. 187 (1): 55–63. doi:10.1016/s0021-9258(19)50929-2. ISSN 0021-9258.
  2. Schwartz, Rochelle D. (1988). "The GABAa receptor-gated ion channel: Biochemical and pharmacological studies of structure and function". Biochemical Pharmacology. 37 (18): 3369–3375. doi:10.1016/0006-2952(88)90684-3. ISSN 0006-2952.
  3. 3.0 3.1 3.2 3.3 Somogyi, Peter; Tamás, Gábor; Lujan, Rafael; Buhl, Eberhard H. (1998). "Salient features of synaptic organisation in the cerebral cortex1Published on the World Wide Web on 3 March 1998.1". Brain Research Reviews. 26 (2–3): 113–135. doi:10.1016/s0165-0173(97)00061-1. ISSN 0165-0173.
  4. 4.0 4.1 4.2 4.3 4.4 4.5 Markram, Henry; Toledo-Rodriguez, Maria; Wang, Yun; Gupta, Anirudh; Silberberg, Gilad; Wu, Caizhi (2004). "Interneurons of the neocortical inhibitory system". Nature Reviews Neuroscience. 5 (10): 793–807. doi:10.1038/nrn1519. ISSN 1471-003X.
  5. 5.0 5.1 Griffen, Trevor C; Maffei, A. (2014). "GABAergic synapses: their plasticity and role in sensory cortex". Frontiers in Cellular Neuroscience. 8. doi:10.3389/fncel.2014.00091. ISSN 1662-5102. PMC 3972456. PMID 24723851.
  6. Chiu, Chiayu Q.; Barberis, Andrea; Higley, Michael J. (2019). "Preserving the balance: diverse forms of long-term GABAergic synaptic plasticity". Nature Reviews Neuroscience. 20 (5): 272–281. doi:10.1038/s41583-019-0141-5. ISSN 1471-003X.
  7. Hollmann, Michael; Heinemann, Stephen (1994). "Cloned Glutamate Receptors". Annual Review of Neuroscience. 17 (1): 31–108. doi:10.1146/annurev.ne.17.030194.000335. ISSN 0147-006X.
  8. 8.0 8.1 Allen, Paul; Sommer, Iris E.; Jardri, Renaud; Eysenck, Michael W.; Hugdahl, Kenneth (2019). "Extrinsic and default mode networks in psychiatric conditions: Relationship to excitatory-inhibitory transmitter balance and early trauma". Neuroscience & Biobehavioral Reviews. 99: 90–100. doi:10.1016/j.neubiorev.2019.02.004. ISSN 0149-7634.
  9. 9.0 9.1 Kepecs, Adam; Fishell, Gordon (2014). "Interneuron cell types are fit to function". Nature. 505 (7483): 318–326. doi:10.1038/nature12983. ISSN 0028-0836. PMC 4349583. PMID 24429630.
  10. 10.0 10.1 Cardin, Jessica A. (2018). "Inhibitory Interneurons Regulate Temporal Precision and Correlations in Cortical Circuits". Trends in Neurosciences. 41 (10): 689–700. doi:10.1016/j.tins.2018.07.015. ISSN 0166-2236. PMC 6173199. PMID 30274604.
  11. Hangya, Balázs; Pi, Hyun-Jae; Kvitsiani, Duda; Ranade, Sachin P; Kepecs, Adam (2014). "From circuit motifs to computations: mapping the behavioral repertoire of cortical interneurons". Current Opinion in Neurobiology. 26: 117–124. doi:10.1016/j.conb.2014.01.007. ISSN 0959-4388. PMC 4090079. PMID 24508565.
  12. Cossart, Rosa (2014). "Operational hub cells: a morpho-physiologically diverse class of GABAergic neurons united by a common function". Current Opinion in Neurobiology. 26: 51–56. doi:10.1016/j.conb.2013.12.002. ISSN 0959-4388.
  13. 13.0 13.1 13.2 Wall, Nicholas R.; De La Parra, Mauricio; Sorokin, Jordan M.; Taniguchi, Hiroki; Huang, Z. Josh; Callaway, Edward M. (2016). "Brain-Wide Maps of Synaptic Input to Cortical Interneurons". The Journal of Neuroscience. 36 (14): 4000–4009. doi:10.1523/jneurosci.3967-15.2016. ISSN 0270-6474. PMC 4821911. PMID 27053207.
  14. Sun, Qingtao; Li, Xiangning; Ren, Miao; Zhao, Mengting; Zhong, Qiuyuan; Ren, Yuqi; Luo, Pan; Ni, Hong; Zhang, Xiaoyu; Zhang, Chen; Yuan, Jing (2019). "A whole-brain map of long-range inputs to GABAergic interneurons in the mouse medial prefrontal cortex". Nature Neuroscience. 22 (8): 1357–1370. doi:10.1038/s41593-019-0429-9. ISSN 1097-6256.
  15. 15.0 15.1 Kvitsiani, D.; Ranade, S.; Hangya, B.; Taniguchi, H.; Huang, J. Z.; Kepecs, A. (2013). "Distinct behavioural and network correlates of two interneuron types in prefrontal cortex". Nature. 498 (7454): 363–366. doi:10.1038/nature12176. ISSN 0028-0836. PMC 4349584. PMID 23708967.
  16. 16.0 16.1 Klausberger, Thomas; Somogyi, Peter (2008). "Neuronal Diversity and Temporal Dynamics: The Unity of Hippocampal Circuit Operations". Science. 321 (5885): 53–57. doi:10.1126/science.1149381. ISSN 0036-8075. PMC 4487503. PMID 18599766.
  17. 17.0 17.1 Schulz, Jan M.; Knoflach, Frederic; Hernandez, Maria-Clemencia; Bischofberger, Josef (2018). "Dendrite-targeting interneurons control synaptic NMDA-receptor activation via nonlinear α5-GABAA receptors". Nature Communications. 9 (1): 1–16. doi:10.1038/s41467-018-06004-8. ISSN 2041-1723. PMC 6120902. PMID 30177704.
  18. Marlin, J. J.; Carter, A. G. (2014). "GABA-A Receptor Inhibition of Local Calcium Signaling in Spines and Dendrites". Journal of Neuroscience. 34 (48): 15898–15911. doi:10.1523/jneurosci.0869-13.2014. ISSN 0270-6474. PMC 4244464. PMID 25429132.
  19. 19.0 19.1 Roux, Lisa; Buzsáki, György (2015). "Tasks for inhibitory interneurons in intact brain circuits". Neuropharmacology. 88: 10–23. doi:10.1016/j.neuropharm.2014.09.011. ISSN 0028-3908. PMC 4254329. PMID 25239808.
  20. Kiemes, Amanda; Davies, Cathy; Kempton, Matthew J.; Lukow, Paulina B.; Bennallick, Carly; Stone, James M.; Modinos, Gemma (2021). "GABA, Glutamate and Neural Activity: A Systematic Review With Meta-Analysis of Multimodal 1H-MRS-fMRI Studies". Frontiers in Psychiatry. 12: 255. doi:10.3389/fpsyt.2021.644315. ISSN 1664-0640. PMC 7982484 Check |pmc= value (help). PMID 33762983 Check |pmid= value (help).
  21. 21.0 21.1 Peerboom, Carlijn; Wierenga, Corette J. (2021). "The postnatal GABA shift: A developmental perspective". Neuroscience & Biobehavioral Reviews. 124: 179–192. doi:10.1016/j.neubiorev.2021.01.024. ISSN 0149-7634.
  22. Watanabe, Miho; Fukuda, Atsuo (2015). "Development and regulation of chloride homeostasis in the central nervous system". Frontiers in Cellular Neuroscience. 9. doi:10.3389/fncel.2015.00371. ISSN 1662-5102. PMC 4585146. PMID 26441542.CS1 maint: PMC format (link)
  23. Wu, Connie; Sun, Dandan (2015). "GABA receptors in brain development, function, and injury". Metabolic Brain Disease. 30 (2): 367–379. doi:10.1007/s11011-014-9560-1. ISSN 0885-7490. PMC 4231020. PMID 24820774.
  24. Ramírez-Jarquín, Uri Nimrod; Lazo-Gómez, Rafael; Tovar-y-Romo, Luis B.; Tapia, Ricardo (2014). "Spinal inhibitory circuits and their role in motor neuron degeneration". Neuropharmacology. 82: 101–107. doi:10.1016/j.neuropharm.2013.10.003. ISSN 0028-3908.
  25. 25.0 25.1 Ghit, Amr; Assal, Dina; Al-Shami, Ahmed S.; Hussein, Diaa Eldin E. (2021). "GABAA receptors: structure, function, pharmacology, and related disorders". Journal of Genetic Engineering and Biotechnology. 19 (1): 1–15. doi:10.1186/s43141-021-00224-0. ISSN 2090-5920. PMC 8380214 Check |pmc= value (help). PMID 34417930 Check |pmid= value (help).
  26. 26.0 26.1 26.2 Schoonover, Kirsten E.; Dienel, Samuel J.; Lewis, David A. (2020). "Prefrontal cortical alterations of glutamate and GABA neurotransmission in schizophrenia: Insights for rational biomarker development". Biomarkers in Neuropsychiatry. 3: 100015. doi:10.1016/j.bionps.2020.100015. ISSN 2666-1446. PMC 7351254. PMID 32656540.
  27. Błaszczyk, Janusz W. (2016). "Parkinson's Disease and Neurodegeneration: GABA-Collapse Hypothesis". Frontiers in Neuroscience. 10. doi:10.3389/fnins.2016.00269. ISSN 1662-453X. PMC 4899466. PMID 27375426.
  28. 28.0 28.1 A. Rikani, Azadeh; Choudhry, Zia; Choudhry, Adnan Maqsood; Rizvi, Nasir; Ikram, Huma; Mobassarah, Nusrat Jahan (2014). "The mechanism of degeneration of striatal neuronal subtypes in Huntington disease". Annals of Neurosciences. 21 (3): 112. doi:10.5214/ans.0972.7531.210308. ISSN 0972-7531. PMC 4158784. PMID 25206077.
  29. Ngo, Dai-Hung; Vo, Thanh Sang (2019). "An Updated Review on Pharmaceutical Properties of Gamma-Aminobutyric Acid". Molecules. 24 (15): 2678. doi:10.3390/molecules24152678. ISSN 1420-3049. PMC 6696076. PMID 31344785.
  30. Watanabe, M., Maemura, K., Kanbara, K., Tamayama, T., & Hayasaki, H. (2002). GABA and GABA receptors in the central nervous system and other organs. International review of cytology, 213, 1-47. https://doi.org/10.1016/S0074-7696(02)13011-7
  31. Schousboe, Arne; Bak, Lasse K.; Waagepetersen, Helle S. (2013). "Astrocytic Control of Biosynthesis and Turnover of the Neurotransmitters Glutamate and GABA". Frontiers in Endocrinology. 4: 102. doi:10.3389/fendo.2013.00102. ISSN 1664-2392. PMC 3744088. PMID 23966981.
  32. Hertz, Leif (2013). "The Glutamate–Glutamine (GABA) Cycle: Importance of Late Postnatal Development and Potential Reciprocal Interactions between Biosynthesis and Degradation". Frontiers in Endocrinology. 4. doi:10.3389/fendo.2013.00059. ISSN 1664-2392. PMC 3664331. PMID 23750153.
  33. Bak, Lasse K.; Schousboe, Arne; Waagepetersen, Helle S. (2006). "The glutamate/GABA-glutamine cycle: aspects of transport, neurotransmitter homeostasis and ammonia transfer". Journal of Neurochemistry. 98 (3): 641–653. doi:10.1111/j.1471-4159.2006.03913.x. ISSN 0022-3042.
  34. Scimemi, Annalisa (2014). "Structure, function, and plasticity of GABA transporters". Frontiers in Cellular Neuroscience. 8: 161. doi:10.3389/fncel.2014.00161. ISSN 1662-5102. PMC 4060055. PMID 24987330.
  35. Zhou, Yun; Danbolt, Niels Christian (2013). "GABA and Glutamate Transporters in Brain". Frontiers in Endocrinology. 4: 165. doi:10.3389/fendo.2013.00165. ISSN 1664-2392. PMC 3822327. PMID 24273530.
  36. Yamashita, Manami; Kawaguchi, Shin-ya; Hori, Tetsuya; Takahashi, Tomoyuki (2018). "Vesicular GABA Uptake Can Be Rate Limiting for Recovery of IPSCs from Synaptic Depression". Cell Reports. 22 (12): 3134–3141. doi:10.1016/j.celrep.2018.02.080. ISSN 2211-1247.
  37. Jin, Hong; Wu, Heng; Osterhaus, Gregory; Wei, Jianning; Davis, Kathleen; Sha, Di; Floor, Eric; Hsu, Che-Chang; Kopke, Richard D.; Wu, Jang-Yen (2003). "Demonstration of functional coupling between γ-aminobutyric acid (GABA) synthesis and vesicular GABA transport into synaptic vesicles". Proceedings of the National Academy of Sciences. 100 (7): 4293–4298. doi:10.1073/pnas.0730698100. ISSN 0027-8424. PMC 153086. PMID 12634427.
  38. 38.0 38.1 Liang, S-L.; Carlson, G. C.; Coulter, D. A. (2006). "Dynamic Regulation of Synaptic GABA Release by the Glutamate-Glutamine Cycle in Hippocampal Area CA1". Journal of Neuroscience. 26 (33): 8537–8548. doi:10.1523/jneurosci.0329-06.2006. ISSN 0270-6474. PMC 2471868. PMID 16914680.
  39. Silverman, Richard B. (2018). "Design and Mechanism of GABA Aminotransferase Inactivators. Treatments for Epilepsies and Addictions". Chemical Reviews. 118 (7): 4037–4070. doi:10.1021/acs.chemrev.8b00009. ISSN 0009-2665. PMC 8459698 Check |pmc= value (help). PMID 29569907.
  40. 40.0 40.1 Martin, David L.; Rimvall, Karin (1993). "Regulation of γ-Aminobutyric Acid Synthesis in the Brain". Journal of Neurochemistry. 60 (2): 395–407. doi:10.1111/j.1471-4159.1993.tb03165.x. ISSN 0022-3042.
  41. Lamigeon, C.; Bellier, J. P.; Sacchettoni, S.; Rujano, M.; Jacquemont, B. (2001). "Enhanced neuronal protection from oxidative stress by coculture with glutamic acid decarboxylase-expressing astrocytes". Journal of Neurochemistry. 77 (2): 598–606. doi:10.1046/j.1471-4159.2001.00278.x. ISSN 0022-3042.
  42. Waagepetersen, Helle S.; Sonnewald, Ursula; Gegelashvili, Georgi; Larsson, Orla M.; Schousboe, Arne (2001). <347::aid-jnr1029>3.0.co;2-g "Metabolic distinction between vesicular and cytosolic GABA in cultured GABAergic neurons using13C magnetic resonance spectroscopy". Journal of Neuroscience Research. 63 (4): 347–355. doi:10.1002/1097-4547(20010215)63:4<347::aid-jnr1029>3.0.co;2-g. ISSN 0360-4012.
  43. Muñoz-Lopetegi, Amaia; de Bruijn, Marienke A.A.M.; Boukhrissi, Sanae; Bastiaansen, Anna E.M.; Nagtzaam, Mariska M.P.; Hulsenboom, Esther S.P.; Boon, Agnita J.W.; Neuteboom, Rinze F.; de Vries, Juna M.; Sillevis Smitt, Peter A.E.; Schreurs, Marco W.J. (2020). "Neurologic syndromes related to anti-GAD65". Neurology - Neuroimmunology Neuroinflammation. 7 (3): e696. doi:10.1212/nxi.0000000000000696. ISSN 2332-7812. PMC 7136051. PMID 32123047.
  44. 44.0 44.1 44.2 Gupta, Anirudh; Wang, Yun; Markram, Henry (2000). "Organizing Principles for a Diversity of GABAergic Interneurons and Synapses in the Neocortex". Science. 287 (5451): 273–278. doi:10.1126/science.287.5451.273. ISSN 0036-8075.
  45. 45.0 45.1 45.2 Rudy, Bernardo; Fishell, Gordon; Lee, SooHyun; Hjerling-Leffler, Jens (2011). "Three groups of interneurons account for nearly 100% of neocortical GABAergic neurons". Developmental Neurobiology. 71 (1): 45–61. doi:10.1002/dneu.20853. ISSN 1932-8451. PMC 3556905. PMID 21154909.
  46. 46.0 46.1 Que, Lin; Winterer, Jochen; Földy, Csaba (2019). "Deep Survey of GABAergic Interneurons: Emerging Insights From Gene-Isoform Transcriptomics". Frontiers in Molecular Neuroscience. 12. doi:10.3389/fnmol.2019.00115. ISSN 1662-5099. PMC 6514532. PMID 31133800.
  47. DeFelipe, J. (2002). Cortical interneurons: from Cajal to 2001. Progress in brain research, 136, 215-238.https://doi.org/10.1016/S0079-6123(02)36019-9
  48. 48.0 48.1 Tremblay, Robin; Lee, Soohyun; Rudy, Bernardo (2016). "GABAergic Interneurons in the Neocortex: From Cellular Properties to Circuits". Neuron. 91 (2): 260–292. doi:10.1016/j.neuron.2016.06.033. ISSN 0896-6273. PMC 4980915. PMID 27477017.
  49. Anton-Sanchez, Laura; Bielza, Concha; Benavides-Piccione, Ruth; DeFelipe, Javier; Larrañaga, Pedro (2016). "Dendritic and Axonal Wiring Optimization of Cortical GABAergic Interneurons". Neuroinformatics. 14 (4): 453–464. doi:10.1007/s12021-016-9309-6. ISSN 1539-2791. PMC 5010609. PMID 27345531.
  50. Kawaguchi, Y; Kubota, Y. (1997). "GABAergic cell subtypes and their synaptic connections in rat frontal cortex". Cerebral Cortex. 7 (6): 476–486. doi:10.1093/cercor/7.6.476. ISSN 1460-2199.
  51. 51.0 51.1 Royer, Sébastien; Zemelman, Boris V; Losonczy, Attila; Kim, Jinhyun; Chance, Frances; Magee, Jeffrey C; Buzsáki, György (2012). "Control of timing, rate and bursts of hippocampal place cells by dendritic and somatic inhibition". Nature Neuroscience. 15 (5): 769–775. doi:10.1038/nn.3077. ISSN 1097-6256. PMC 4919905. PMID 22446878.
  52. Lim, Lynette; Mi, Da; Llorca, Alfredo; Marín, Oscar (2018). "Development and Functional Diversification of Cortical Interneurons". Neuron. 100 (2): 294–313. doi:10.1016/j.neuron.2018.10.009. ISSN 0896-6273. PMC 6290988. PMID 30359598.
  53. Lee, Hyun Joo; Weitz, Andrew J.; Bernal-Casas, David; Duffy, Ben A.; Choy, ManKin; Kravitz, Alexxai V.; Kreitzer, Anatol C.; Lee, Jin Hyung (2016). "Activation of Direct and Indirect Pathway Medium Spiny Neurons Drives Distinct Brain-wide Responses". Neuron. 91 (2): 412–424. doi:10.1016/j.neuron.2016.06.010. ISSN 0896-6273. PMC 5528162. PMID 27373834.
  54. 54.0 54.1 Lee, S.; Hjerling-Leffler, J.; Zagha, E.; Fishell, G.; Rudy, B. (2010). "The Largest Group of Superficial Neocortical GABAergic Interneurons Expresses Ionotropic Serotonin Receptors". Journal of Neuroscience. 30 (50): 16796–16808. doi:10.1523/jneurosci.1869-10.2010. ISSN 0270-6474. PMC 3025500. PMID 21159951.
  55. Liguz-Lecznar, Monika; Urban-Ciecko, Joanna; Kossut, Malgorzata (2016). "Somatostatin and Somatostatin-Containing Neurons in Shaping Neuronal Activity and Plasticity". Frontiers in Neural Circuits. 10. doi:10.3389/fncir.2016.00048. ISSN 1662-5110. PMC 4927943. PMID 27445703.
  56. Lapray, Damien; Lasztoczi, Balint; Lagler, Michael; Viney, Tim James; Katona, Linda; Valenti, Ornella; Hartwich, Katja; Borhegyi, Zsolt; Somogyi, Peter; Klausberger, Thomas (2012). "Behavior-dependent specialization of identified hippocampal interneurons". Nature Neuroscience. 15 (9): 1265–1271. doi:10.1038/nn.3176. ISSN 1097-6256. PMC 3433735. PMID 22864613.
  57. "Petilla terminology: nomenclature of features of GABAergic interneurons of the cerebral cortex. (2008)". Nature Reviews Neuroscience. 9 (7): 557–568. doi:10.1038/nrn2402. ISSN 1471-003X. PMC 2868386. PMID 18568015.
  58. 58.0 58.1 DeFelipe, Javier; López-Cruz, Pedro L.; Benavides-Piccione, Ruth; Bielza, Concha; Larrañaga, Pedro; Anderson, Stewart; Burkhalter, Andreas; Cauli, Bruno; Fairén, Alfonso; Feldmeyer, Dirk; Fishell, Gord (2013). "New insights into the classification and nomenclature of cortical GABAergic interneurons". Nature Reviews Neuroscience. 14 (3): 202–216. doi:10.1038/nrn3444. ISSN 1471-003X. PMC 3619199. PMID 23385869.CS1 maint: PMC format (link)
  59. Bartholome, Odile; de la Brassinne Bonardeaux, Orianne; Neirinckx, Virginie; Rogister, Bernard (2020). "A Composite Sketch of Fast-Spiking Parvalbumin-Positive Neurons". Cerebral Cortex Communications. 1 (1). doi:10.1093/texcom/tgaa026. ISSN 2632-7376. PMC 8153048 Check |pmc= value (help). PMID 34296100 Check |pmid= value (help).
  60. Nahar, Lailun; Delacroix, Blake M.; Nam, Hyung W. (2021). "The Role of Parvalbumin Interneurons in Neurotransmitter Balance and Neurological Disease". Frontiers in Psychiatry. 12. doi:10.3389/fpsyt.2021.679960. ISSN 1664-0640. PMC 8249927 Check |pmc= value (help). PMID 34220586 Check |pmid= value (help).
  61. 61.0 61.1 Murayama, Masanori; Pérez-Garci, Enrique; Nevian, Thomas; Bock, Tobias; Senn, Walter; Larkum, Matthew E. (2009). "Dendritic encoding of sensory stimuli controlled by deep cortical interneurons". Nature. 457 (7233): 1137–1141. doi:10.1038/nature07663. ISSN 0028-0836.
  62. Banovac, Ivan; Sedmak, Dora; Esclapez, Monique; Petanjek, Zdravko (2022). "The Distinct Characteristics of Somatostatin Neurons in the Human Brain". Molecular Neurobiology. 59 (8): 4953–4965. doi:10.1007/s12035-022-02892-6. ISSN 0893-7648.
  63. Waller, Rachel; Mandeya, Memory; Viney, Edward; Simpson, Julie E.; Wharton, Stephen B. (2020). "Histological characterization of interneurons in Alzheimer's disease reveals a loss of somatostatin interneurons in the temporal cortex". Neuropathology. 40 (4): 336–346. doi:10.1111/neup.12649. ISSN 0919-6544.
  64. Millman, Daniel J; Ocker, Gabriel Koch; Caldejon, Shiella; Kato, India; Larkin, Josh D; Lee, Eric Kenji; Luviano, Jennifer; Nayan, Chelsea; Nguyen, Thuyanh V; North, Kat; Seid, Sam (2020). "Author response: VIP interneurons in mouse primary visual cortex selectively enhance responses to weak but specific stimuli". doi:10.7554/elife.55130.sa2. Cite journal requires |journal= (help)
  65. Pi, Hyun-Jae; Hangya, Balázs; Kvitsiani, Duda; Sanders, Joshua I.; Huang, Z. Josh; Kepecs, Adam (2013). "Cortical interneurons that specialize in disinhibitory control". Nature. 503 (7477): 521–524. doi:10.1038/nature12676. ISSN 0028-0836. PMC 4017628. PMID 24097352.
  66. Apicella, Alfonso junior; Marchionni, Ivan (2022). "VIP-Expressing GABAergic Neurons: Disinhibitory vs. Inhibitory Motif and Its Role in Communication Across Neocortical Areas". Frontiers in Cellular Neuroscience. 16. doi:10.3389/fncel.2022.811484. ISSN 1662-5102. PMC 8867699 Check |pmc= value (help). PMID 35221922 Check |pmid= value (help).
  67. Baruchin, Liad J; Ghezzi, Filippo; Kohl, Michael M; Butt, Simon J B (2022). "Contribution of Interneuron Subtype-Specific GABAergic Signaling to Emergent Sensory Processing in Mouse Somatosensory Whisker Barrel Cortex". Cerebral Cortex. 32 (12): 2538–2554. doi:10.1093/cercor/bhab363. ISSN 1047-3211. PMC 9201598 Check |pmc= value (help). PMID 34613375 Check |pmid= value (help).
  68. Lukow, Paulina Barbara; Martins, Daniel; Veronese, Mattia; Vernon, Anthony Christopher; McGuire, Philip; Turkheimer, Federico Edoardo; Modinos, Gemma (2022). "Cellular and molecular signatures of in vivo imaging measures of GABAergic neurotransmission in the human brain". Communications Biology. 5 (1): 372. doi:10.1038/s42003-022-03268-1. ISSN 2399-3642. PMC PMC9018713 Check |pmc= value (help). PMID 35440709 Check |pmid= value (help).CS1 maint: PMC format (link)
  69. Schwartz, R. D. (1988). The GABAA receptor-gated ion channel: Biochemical and pharmacological studies of structure and function. Biochemical Pharmacology. 37 (18): 3369-3375.
  70. 70.0 70.1 70.2 70.3 70.4 Sieghart, Werner; Savić, Miroslav M. (2018). "International Union of Basic and Clinical Pharmacology. CVI: GABAA Receptor Subtype- and Function-selective Ligands: Key Issues in Translation to Humans". Pharmacological Reviews. 70 (4): 836–878. doi:10.1124/pr.117.014449. ISSN 0031-6997.
  71. 71.0 71.1 71.2 Etherington, Lori-An; Mihalik, Balázs; Pálvölgyi, Adrienn; Ling, István; Pallagi, Katalin; Kertész, Szabolcs; Varga, Péter; Gunn, Ben G.; Brown, Adam R.; Livesey, Matthew R.; Monteiro, Olivia (2017). "Selective inhibition of extra-synaptic α5-GABA A receptors by S44819, a new therapeutic agent". Neuropharmacology. 125: 353–364. doi:10.1016/j.neuropharm.2017.08.012. ISSN 0028-3908.
  72. 72.0 72.1 Sallard, Erwan; Letourneur, Diane; Legendre, Pascal (2021). "Electrophysiology of ionotropic GABA receptors". Cellular and Molecular Life Sciences. 78 (13): 5341–5370. doi:10.1007/s00018-021-03846-2. ISSN 1420-682X. PMC 8257536 Check |pmc= value (help). PMID 34061215 Check |pmid= value (help).
  73. Baur, Roland; Minier, Frédéric; Sigel, Erwin (2006). "A GABAAreceptor of defined subunit composition and positioning: Concatenation of five subunits". FEBS Letters. 580 (6): 1616–1620. doi:10.1016/j.febslet.2006.02.002. ISSN 0014-5793.
  74. Rudolph, Uwe; Knoflach, Frédéric (2011). "Beyond classical benzodiazepines: novel therapeutic potential of GABAA receptor subtypes". Nature Reviews Drug Discovery. 10 (9): 685–697. doi:10.1038/nrd3502. ISSN 1474-1776. PMC 3375401. PMID 21799515.
  75. Olsen, Richard W.; Sieghart, Werner (2008). "International Union of Pharmacology. LXX. Subtypes of γ-Aminobutyric AcidAReceptors: Classification on the Basis of Subunit Composition, Pharmacology, and Function. Update". Pharmacological Reviews. 60 (3): 243–260. doi:10.1124/pr.108.00505. ISSN 0031-6997. PMC 2847512. PMID 18790874.
  76. Kasaragod, Vikram Babu; Schindelin, Hermann (2019). "Structure of Heteropentameric GABAA Receptors and Receptor-Anchoring Properties of Gephyrin". Frontiers in Molecular Neuroscience. 12. doi:10.3389/fnmol.2019.00191. ISSN 1662-5099. PMC 6693554. PMID 31440140.
  77. Lambert, Jeremy J.; Belelli, Delia; Peden, Dianne R.; Vardy, Audrey W.; Peters, John A. (2003). "Neurosteroid modulation of GABAA receptors". Progress in Neurobiology. 71 (1): 67–80. doi:10.1016/j.pneurobio.2003.09.001. ISSN 0301-0082.
  78. 78.0 78.1 78.2 78.3 Reddy, D. S. (2018). GABA-A receptors mediate tonic inhibition and neurosteroid sensitivity in the brain. Vitamins and hormones, 107, 177-191. https://doi.org/10.1016/bs.vh.2017.12.001
  79. Mele, Miranda; Costa, Rui O.; Duarte, Carlos B. (2019). "Alterations in GABAA-Receptor Trafficking and Synaptic Dysfunction in Brain Disorders". Frontiers in Cellular Neuroscience. 13. doi:10.3389/fncel.2019.00077. ISSN 1662-5102. PMC 6416223. PMID 30899215.
  80. Nusser, Zoltan; Sieghart, Werner; Somogyi, Peter (1998). "Segregation of Different GABAAReceptors to Synaptic and Extrasynaptic Membranes of Cerebellar Granule Cells". The Journal of Neuroscience. 18 (5): 1693–1703. doi:10.1523/jneurosci.18-05-01693.1998. ISSN 0270-6474. PMC 6792611. PMID 9464994.
  81. 81.0 81.1 Mody, Istvan; Pearce, Robert A. (2004). "Diversity of inhibitory neurotransmission through GABAA receptors". Trends in Neurosciences. 27 (9): 569–575. doi:10.1016/j.tins.2004.07.002. ISSN 0166-2236.
  82. Halonen, Lauri M.; Sinkkonen, Saku T.; Chandra, Dev; Homanics, Gregg E.; Korpi, Esa R. (2009). "Brain regional distribution of GABAA receptors exhibiting atypical GABA agonism: Roles of receptor subunits". Neurochemistry International. 55 (6): 389–396. doi:10.1016/j.neuint.2009.04.008. ISSN 0197-0186. PMC 2760098. PMID 19397945.
  83. 83.0 83.1 83.2 Schipper, S.; Aalbers, M. W.; Rijkers, K.; Swijsen, A.; Rigo, J. M.; Hoogland, G.; Vles, J. S. H. (2016). "Tonic GABAA Receptors as Potential Target for the Treatment of Temporal Lobe Epilepsy". Molecular Neurobiology. 53 (8): 5252–5265. doi:10.1007/s12035-015-9423-8. ISSN 0893-7648. PMC 5012145. PMID 26409480.
  84. Semyanov, Alexey; Walker, Matthew C.; Kullmann, Dimitri M.; Silver, R.Angus (2004). "Tonically active GABAA receptors: modulating gain and maintaining the tone". Trends in Neurosciences. 27 (5): 262–269. doi:10.1016/j.tins.2004.03.005. ISSN 0166-2236.
  85. Field, Martin; Lukacs, Istvan P.; Hunter, Emily; Stacey, Richard; Plaha, Puneet; Livermore, Laurent; Ansorge, Olaf; Somogyi, Peter (2021). "Tonic GABAA Receptor-Mediated Currents of Human Cortical GABAergic Interneurons Vary Amongst Cell Types". The Journal of Neuroscience. 41 (47): 9702–9719. doi:10.1523/jneurosci.0175-21.2021. ISSN 0270-6474. PMC 8612645 Check |pmc= value (help). PMID 34667071 Check |pmid= value (help).
  86. Brickley, Stephen G.; Mody, Istvan (2012). "Extrasynaptic GABAA Receptors: Their Function in the CNS and Implications for Disease". Neuron. 73 (1): 23–34. doi:10.1016/j.neuron.2011.12.012. ISSN 0896-6273. PMC 3399243. PMID 22243744. no-break space character in |first= at position 8 (help)
  87. 87.0 87.1 Mann, E. O.; Kohl, M. M.; Paulsen, O. (2009). "Distinct Roles of GABAA and GABAB Receptors in Balancing and Terminating Persistent Cortical Activity". Journal of Neuroscience. 29 (23): 7513–7518. doi:10.1523/jneurosci.6162-08.2009. ISSN 0270-6474. PMC 4326656. PMID 19515919.
  88. 88.0 88.1 Yang, Moon Young; Kim, Soo-Kyung; Goddard, William A. (2022). "G protein coupling and activation of the metabotropic GABAB heterodimer". Nature Communications. 13 (1). doi:10.1038/s41467-022-32213-3. ISSN 2041-1723. PMC 9360005 Check |pmc= value (help). PMID 35941188 Check |pmid= value (help).
  89. White, Julia H.; Wise, Alan; Main, Martin J.; Green, Andrew; Fraser, Neil J.; Disney, Graham H.; Barnes, Ashley A.; Emson, Piers; Foord, Steven M.; Marshall, Fiona H. (1998). "Heterodimerization is required for the formation of a functional GABAB receptor". Nature. 396 (6712): 679–682. doi:10.1038/25354. ISSN 0028-0836.
  90. Marshall, Fiona H; Jones, Kenneth A; Kaupmann, Klemens; Bettler, Bernhard (1999). "GABAB receptors – the first 7TM heterodimers". Trends in Pharmacological Sciences. 20 (10): 396–399. doi:10.1016/s0165-6147(99)01383-8. ISSN 0165-6147.
  91. Darlison, Mark G., ed. (2008). "Inhibitory Regulation of Excitatory Neurotransmission". Results and Problems in Cell Differentiation. doi:10.1007/978-3-540-72602-9.
  92. Monnier, Carine; Tu, Haijun; Bourrier, Emmanuel; Vol, Claire; Lamarque, Laurent; Trinquet, Eric; Pin, Jean-Philippe; Rondard, Philippe (2011). "Trans-activation between 7TM domains: implication in heterodimeric GABABreceptor activation". The EMBO Journal. 30 (1): 32–42. doi:10.1038/emboj.2010.270. ISSN 0261-4189. PMC 3020105. PMID 21063387.
  93. Xue, Li; Sun, Qian; Zhao, Han; Rovira, Xavier; Gai, Siyu; He, Qianwen; Pin, Jean-Philippe; Liu, Jianfeng; Rondard, Philippe (2019). "Rearrangement of the transmembrane domain interfaces associated with the activation of a GPCR hetero-oligomer". Nature Communications. 10 (1): 2765. doi:10.1038/s41467-019-10834-5. ISSN 2041-1723. PMC 6591306. PMID 31235691.CS1 maint: PMC format (link)
  94. Shen, Cangsong; Mao, Chunyou; Xu, Chanjuan; Jin, Nan; Zhang, Huibing; Shen, Dan-Dan; Shen, Qingya; Wang, Xiaomei; Hou, Tingjun; Chen, Zhong; Rondard, Philippe (2021). "Structural basis of GABAB receptor–Gi protein coupling". Nature. 594 (7864): 594–598. doi:10.1038/s41586-021-03507-1. ISSN 0028-0836. PMC 8222003 Check |pmc= value (help). PMID 33911284 Check |pmid= value (help).
  95. Margeta-Mitrovic, Marta; Jan, Yuh Nung; Jan, Lily Yeh (2001). "Function of GB1 and GB2 subunits in G protein coupling of GABA B receptors". Proceedings of the National Academy of Sciences. 98 (25): 14649–14654. doi:10.1073/pnas.251554498. ISSN 0027-8424. PMC 64736. PMID 11724956. line feed character in |title= at position 63 (help)
  96. 96.0 96.1 Evenseth, Linn Samira Mari; Gabrielsen, Mari; Sylte, Ingebrigt (2020). "The GABAB Receptor—Structure, Ligand Binding and Drug Development". Molecules. 25 (13): 3093. doi:10.3390/molecules25133093. ISSN 1420-3049. PMC 7411975. PMID 32646032.
  97. 97.0 97.1 97.2 97.3 Terunuma, Miho (2018). "Diversity of structure and function of GABAB receptors: a complexity of GABAB-mediated signaling". Proceedings of the Japan Academy, Series B. 94 (10): 390–411. doi:10.2183/pjab.94.026. ISSN 0386-2208. PMC 6374141. PMID 30541966.
  98. Bettler, Bernhard; Tiao, Jim Yu-Hsiang (2006). "Molecular diversity, trafficking and subcellular localization of GABAB receptors". Pharmacology & Therapeutics. 110 (3): 533–543. doi:10.1016/j.pharmthera.2006.03.006. ISSN 0163-7258.
  99. 99.0 99.1 Perkinton, Michael S.; Sihra, Talvinder S. (1998). "Presynaptic GABAB Receptor Modulation of Glutamate Exocytosis from Rat Cerebrocortical Nerve Terminals: Receptor Decoupling by Protein Kinase C". Journal of Neurochemistry. 70 (4): 1513–1522. doi:10.1046/j.1471-4159.1998.70041513.x. ISSN 0022-3042.
  100. Ladera, Carolina; del Carmen Godino, María; Cabañero, María José; Torres, Magdalena; Watanabe, Masahiko; Luján, Rafael; Sánchez-Prieto, José (2008). "Pre-synaptic GABABreceptors inhibit glutamate release through GIRK channels in rat cerebral cortex". Journal of Neurochemistry. 107 (6): 1506–1517. doi:10.1111/j.1471-4159.2008.05712.x. ISSN 0022-3042.
  101. Lüscher, Christian; Jan, Lily Y; Stoffel, Markus; Malenka, Robert C; Nicoll, Roger A (1997). "G Protein-Coupled Inwardly Rectifying K+ Channels (GIRKs) Mediate Postsynaptic but Not Presynaptic Transmitter Actions in Hippocampal Neurons". Neuron. 19 (3): 687–695. doi:10.1016/s0896-6273(00)80381-5. ISSN 0896-6273.
  102. Scanziani, Massimo (2000). "GABA Spillover Activates Postsynaptic GABAB Receptors to Control Rhythmic Hippocampal Activity". Neuron. 25 (3): 673–681. doi:10.1016/s0896-6273(00)81069-7. ISSN 0896-6273.
  103. 103.0 103.1 Fairfax, Benjamin P.; Pitcher, Julie A.; Scott, Mark G.H.; Calver, Andrew R.; Pangalos, Menelas N.; Moss, Stephen J.; Couve, Andrés (2004). "Phosphorylation and Chronic Agonist Treatment Atypically Modulate GABAB Receptor Cell Surface Stability". Journal of Biological Chemistry. 279 (13): 12565–12573. doi:10.1074/jbc.m311389200. ISSN 0021-9258.
  104. Vargas, Karina J.; Terunuma, Miho; Tello, Judith A.; Pangalos, Menelas N.; Moss, Stephen J.; Couve, Andrés (2008). "The Availability of Surface GABAB Receptors Is Independent of γ-Aminobutyric Acid but Controlled by Glutamate in Central Neurons". Journal of Biological Chemistry. 283 (36): 24641–24648. doi:10.1074/jbc.m802419200. ISSN 0021-9258. PMC 3259848. PMID 18579521.
  105. Magalhaes, Ana C; Dunn, Henry; Ferguson, Stephen SG (2012). "Regulation of GPCR activity, trafficking and localization by GPCR-interacting proteins". British Journal of Pharmacology. 165 (6): 1717–1736. doi:10.1111/j.1476-5381.2011.01552.x. ISSN 0007-1188. PMC 3372825. PMID 21699508.
  106. 106.0 106.1 Chalifoux, Jason R.; Carter, Adam G. (2010). "GABAB Receptors Modulate NMDA Receptor Calcium Signals in Dendritic Spines". Neuron. 66 (1): 101–113. doi:10.1016/j.neuron.2010.03.012. ISSN 0896-6273. PMC 2861500. PMID 20399732.
  107. Gandal, M J; Sisti, J; Klook, K; Ortinski, P I; Leitman, V; Liang, Y; Thieu, T; Anderson, R; Pierce, R C; Jonak, G; Gur, R E (2012). "GABAB-mediated rescue of altered excitatory–inhibitory balance, gamma synchrony and behavioral deficits following constitutive NMDAR-hypofunction". Translational Psychiatry. 2 (7): e142–e142. doi:10.1038/tp.2012.69. ISSN 2158-3188. PMC 3410621. PMID 22806213.
  108. Sakairi, Hakushun; Kamikubo, Yuji; Abe, Masayoshi; Ikeda, Keisuke; Ichiki, Arata; Tabata, Toshihide; Kano, Masanobu; Sakurai, Takashi (2020). "G Protein-Coupled Glutamate and GABA Receptors Form Complexes and Mutually Modulate Their Signals". ACS Chemical Neuroscience. 11 (4): 567–578. doi:10.1021/acschemneuro.9b00599. ISSN 1948-7193.
  109. Froestl, W. (2010). Chemistry and pharmacology of GABAB receptor ligands. In Advances in pharmacology (Vol. 58, pp. 19-62). Academic Press. https://doi.org/10.1016/S1054-3589(10)58002-5
  110. Romito, Jia W; Turner, Emily R; Rosener, John A; Coldiron, Landon; Udipi, Ashutosh; Nohrn, Linsey; Tausiani, Jacob; Romito, Bryan T (2021). "Baclofen therapeutics, toxicity, and withdrawal: A narrative review". SAGE Open Medicine. 9: 205031212110221. doi:10.1177/20503121211022197. ISSN 2050-3121. PMC 8182184 Check |pmc= value (help). PMID 34158937 Check |pmid= value (help).
  111. Luscher, B; Shen, Q; Sahir, N (2011). "The GABAergic deficit hypothesis of major depressive disorder". Molecular Psychiatry. 16 (4): 383–406. doi:10.1038/mp.2010.120. ISSN 1359-4184. PMC 3412149. PMID 21079608.
  112. Fogaça, Manoela V.; Wu, Min; Li, Chan; Li, Xiao-Yuan; Picciotto, Marina R.; Duman, Ronald S. (2021). "Inhibition of GABA interneurons in the mPFC is sufficient and necessary for rapid antidepressant responses". Molecular Psychiatry. 26 (7): 3277–3291. doi:10.1038/s41380-020-00916-y. ISSN 1359-4184. PMC 8052382 Check |pmc= value (help). PMID 33070149 Check |pmid= value (help).
  113. Luscher, B., & Fuchs, T. (2015). GABAergic control of depression-related brain states. In Advances in pharmacology. Vol. 73, pp. 97-144. Academic Press. https://doi.org/10.1016/bs.apha.2014.11.003
  114. Gerhard, Danielle M.; Pothula, Santosh; Liu, Rong-Jian; Wu, Min; Li, Xiao-Yuan; Girgenti, Matthew J.; Taylor, Seth R.; Duman, Catharine H.; Delpire, Eric; Picciotto, Marina; Wohleb, Eric S. (2020). "GABA interneurons are the cellular trigger for ketamine's rapid antidepressant actions". Journal of Clinical Investigation. 130 (3): 1336–1349. doi:10.1172/jci130808. ISSN 0021-9738. PMC 7269589. PMID 31743111.
  115. Zorumski, Charles F.; Paul, Steven M.; Covey, Douglas F.; Mennerick, Steven (2019). "Neurosteroids as novel antidepressants and anxiolytics: GABA-A receptors and beyond". Neurobiology of Stress. 11: 100196. doi:10.1016/j.ynstr.2019.100196. ISSN 2352-2895. PMC 6804800. PMID 31649968.
  116. Fasipe, Olumuyiwa John; Agede, Olalekan Ayodele; Enikuomehin, Adenike Christiana (2020). "Announcing the novel class of GABA–A receptor selective positive allosteric modulator antidepressants". Future Science OA. 7 (2). doi:10.2144/fsoa-2020-0108. ISSN 2056-5623. PMC 7787135 Check |pmc= value (help). PMID 33437518 Check |pmid= value (help).
  117. Wilkinson, Samuel T.; Sanacora, Gerard (2019). "A new generation of antidepressants: an update on the pharmaceutical pipeline for novel and rapid-acting therapeutics in mood disorders based on glutamate/GABA neurotransmitter systems". Drug Discovery Today. 24 (2): 606–615. doi:10.1016/j.drudis.2018.11.007. ISSN 1359-6446. PMC 6397075. PMID 30447328.
  118. 118.0 118.1 Chiapponi, Chiara; Piras, Federica; Piras, Fabrizio; Caltagirone, Carlo; Spalletta, Gianfranco (2016). "GABA System in Schizophrenia and Mood Disorders: A Mini Review on Third-Generation Imaging Studies". Frontiers in Psychiatry. 7: 61. doi:10.3389/fpsyt.2016.00061. ISSN 1664-0640. PMC 4835487. PMID 27148090.
  119. Jie, Fan; Yin, Guanghao; Yang, Wei; Yang, Modi; Gao, Shuohui; Lv, Jiayin; Li, Bingjin (2018). "Stress in Regulation of GABA Amygdala System and Relevance to Neuropsychiatric Diseases". Frontiers in Neuroscience. 12. doi:10.3389/fnins.2018.00562. ISSN 1662-453X. PMC 6103381. PMID 30154693.
  120. Li, Bingjin; Ge, Tongtong; Cui, Ranji (2018). "Long-Term Plasticity in Amygdala Circuits: Implication of CB1-Dependent LTD in Stress". Molecular Neurobiology. 55: 4107–4114. doi:10.1007/s12035-017-0643-y. ISSN 0893-7648.
  121. 121.0 121.1 Prager, Eric M.; Bergstrom, Hadley C.; Wynn, Gary H.; Braga, Maria F.M. (2016). "The basolateral amygdala γ-aminobutyric acidergic system in health and disease". Journal of Neuroscience Research. 94 (6): 548–567. doi:10.1002/jnr.23690. ISSN 0360-4012. PMC 4837071. PMID 26586374.
  122. Lydiard RB. (2003). The role of GABA in anxiety disorders. J Clin Psychiatry. 64 Suppl 3:21-7. PMID: 12662130.
  123. Lewis, D. A., & Hashimoto, T. (2007). Deciphering the disease process of schizophrenia: the contribution of cortical GABA neurons. International review of neurobiology, 78, 109-131. https://doi.org/10.1016/S0074-7742(06)78004-7
  124. Carter, Cameron S.; Barch, Deanna M.; Buchanan, Robert W.; Bullmore, Ed; Krystal, John H.; Cohen, Jonathan; Geyer, Mark; Green, Michael; Nuechterlein, Keith H.; Robbins, Trevor; Silverstein, Steven (2008). "Identifying Cognitive Mechanisms Targeted for Treatment Development in Schizophrenia: An Overview of the First Meeting of the Cognitive Neuroscience Treatment Research to Improve Cognition in Schizophrenia Initiative". Biological Psychiatry. 64 (1): 4–10. doi:10.1016/j.biopsych.2008.03.020. ISSN 0006-3223. PMC 2577821. PMID 18466880.
  125. Gonzalez-Burgos, Guillermo; Cho, Raymond Y.; Lewis, David A. (2015). "Alterations in Cortical Network Oscillations and Parvalbumin Neurons in Schizophrenia". Biological Psychiatry. 77 (12): 1031–1040. doi:10.1016/j.biopsych.2015.03.010. ISSN 0006-3223. PMC 4444373. PMID 25863358.
  126. 126.0 126.1 Dienel, Samuel J.; Lewis, David A. (2019). "Alterations in cortical interneurons and cognitive function in schizophrenia". Neurobiology of Disease. 131: 104208. doi:10.1016/j.nbd.2018.06.020. ISSN 0969-9961. PMC 6309598. PMID 29936230.
  127. 127.0 127.1 Chen, Chi-Ming A.; Stanford, Arielle D.; Mao, Xiangling; Abi-Dargham, Anissa; Shungu, Dikoma C.; Lisanby, Sarah H.; Schroeder, Charles E.; Kegeles, Lawrence S. (2014). "GABA level, gamma oscillation, and working memory performance in schizophrenia". NeuroImage: Clinical. 4: 531–539. doi:10.1016/j.nicl.2014.03.007. ISSN 2213-1582. PMC 3989525. PMID 24749063.
  128. 128.0 128.1 Nakazawa, Kazu; Zsiros, Veronika; Jiang, Zhihong; Nakao, Kazuhito; Kolata, Stefan; Zhang, Shuqin; Belforte, Juan E. (2012). "GABAergic interneuron origin of schizophrenia pathophysiology". Neuropharmacology. 62 (3): 1574–1583. doi:10.1016/j.neuropharm.2011.01.022. ISSN 0028-3908. PMC 3090452. PMID 21277876.
  129. Lewis, David A.; Cho, Raymond Y.; Carter, Cameron S.; Eklund, Kevin; Forster, Sarah; Kelly, Mary Ann; Montrose, Debra (2008). "Subunit-Selective Modulation of GABA Type A Receptor Neurotransmission and Cognition in Schizophrenia". American Journal of Psychiatry. 165 (12): 1585–1593. doi:10.1176/appi.ajp.2008.08030395. ISSN 0002-953X. PMC 2876339. PMID 18923067.
  130. Beghi, E; Giussani, G; Nichols, E; Abd-Allah, F; Abdela, J; Abdelalim, A; et al. (2019). "Global, regional, and national burden of epilepsy, 1990-2016: a systematic analysis for the Global Burden of Disease Study 2016". The Lancet. Neurology. 18 (4): 357–375. doi:10.1016/S1474-4422(18)30454-X. ISSN 1474-4465. PMC 6416168. PMID 30773428. Explicit use of et al. in: |first= (help)
  131. Zack, Matthew M.; Kobau, Rosemarie (2017). "National and State Estimates of the Numbers of Adults and Children with Active Epilepsy — United States, 2015". MMWR. Morbidity and Mortality Weekly Report. 66 (31): 821–825. doi:10.15585/mmwr.mm6631a1. ISSN 0149-2195. PMC 5687788. PMID 28796763.
  132. Téllez-Zenteno, Jose F.; Hernández-Ronquillo, Lizbeth (2012). "A Review of the Epidemiology of Temporal Lobe Epilepsy". Epilepsy Research and Treatment. 2012: 1–5. doi:10.1155/2012/630853. ISSN 2090-1348. PMC 3420432. PMID 22957234.
  133. Dehghani, Nima; Peyrache, Adrien; Telenczuk, Bartosz; Le Van Quyen, Michel; Halgren, Eric; Cash, Sydney S.; Hatsopoulos, Nicholas G.; Destexhe, Alain (2016). "Dynamic Balance of Excitation and Inhibition in Human and Monkey Neocortex". Scientific Reports. 6 (1): 23176. doi:10.1038/srep23176. ISSN 2045-2322. PMC 4793223. PMID 26980663.CS1 maint: PMC format (link)
  134. Fritschy, Jean-Marc (2008). "E/I balance and GABAA receptor plasticity". Frontiers in Molecular Neuroscience. 1. doi:10.3389/neuro.02.005.2008. ISSN 1662-5099. PMC 2525999. PMID 18946538.
  135. Löscher, Wolfgang (2012). "Basic Pharmacology of Valproate". CNS Drugs. 16 (10): 669–694. doi:10.2165/00023210-200216100-00003. ISSN 1172-7047.
  136. Treiman, David M. (2001). "GABAergic Mechanisms in Epilepsy". Epilepsia. 42: 8–12. doi:10.1046/j.1528-1157.2001.042suppl.3008.x. ISSN 0013-9580.
  137. 137.0 137.1 Cossart, Rosa; Bernard, Christophe; Ben-Ari, Yehezkel (2005). "Multiple facets of GABAergic neurons and synapses: multiple fates of GABA signalling in epilepsies". Trends in Neurosciences. 28 (2): 108–115. doi:10.1016/j.tins.2004.11.011. ISSN 0166-2236.
  138. Righes Marafiga, Joseane; Vendramin Pasquetti, Mayara; Calcagnotto, Maria Elisa (2021). "GABAergic interneurons in epilepsy: More than a simple change in inhibition". Epilepsy & Behavior. 121: 106935. doi:10.1016/j.yebeh.2020.106935. ISSN 1525-5050.
  139. 139.0 139.1 Kaila, Kai; Ruusuvuori, Eva; Seja, Patricia; Voipio, Juha; Puskarjov, Martin (2014). "GABA actions and ionic plasticity in epilepsy". Current Opinion in Neurobiology. 26: 34–41. doi:10.1016/j.conb.2013.11.004. ISSN 0959-4388.
  140. Bonansco, Christian; Fuenzalida, Marco (2016). "Plasticity of Hippocampal Excitatory-Inhibitory Balance: Missing the Synaptic Control in the Epileptic Brain". Neural Plasticity. 2016: 1–13. doi:10.1155/2016/8607038. ISSN 2090-5904. PMC 4783563. PMID 27006834.
  141. Greenfield, L. John (2013). "Molecular mechanisms of antiseizure drug activity at GABAA receptors". Seizure. 22 (8): 589–600. doi:10.1016/j.seizure.2013.04.015. ISSN 1059-1311. PMC 3766376. PMID 23683707.
  142. Loup, F.; Picard, F.; Andre, V. M.; Kehrli, P.; Yonekawa, Y.; Wieser, H.-G.; Fritschy, J.-M. (2006). "Altered expression of  3-containing GABAA receptors in the neocortex of patients with focal epilepsy". Brain. 129 (12): 3277–3289. doi:10.1093/brain/awl287. ISSN 0006-8950. no-break space character in |title= at position 23 (help)
  143. Grabenstatter, Heidi L.; Russek, Shelley J.; Brooks-Kayal, Amy R. (2012). "Molecular pathways controlling inhibitory receptor expression". Epilepsia. 53: 71–78. doi:10.1111/epi.12036. ISSN 0013-9580. PMC 3776022. PMID 23216580.
  144. Enna, S. J., & McCarson, K. E. (2006). The role of GABA in the mediation and perception of pain. Advances in pharmacology, 54, 1-27. https://doi.org/10.1016/S1054-3589(06)54001-3
  145. Peek, Aimie Laura; Rebbeck, Trudy; Puts, Nicolaas AJ.; Watson, Julia; Aguila, Maria-Eliza R.; Leaver, Andrew M. (2020). "Brain GABA and glutamate levels across pain conditions: A systematic literature review and meta-analysis of 1H-MRS studies using the MRS-Q quality assessment tool". NeuroImage. 210: 116532. doi:10.1016/j.neuroimage.2020.116532. ISSN 1053-8119.
  146. Harding, Erika K.; Fung, Samuel Wanchi; Bonin, Robert P. (2020). "Insights Into Spinal Dorsal Horn Circuit Function and Dysfunction Using Optical Approaches". Frontiers in Neural Circuits. 14: 31. doi:10.3389/fncir.2020.00031. ISSN 1662-5110. PMC 7303281. PMID 32595458.
  147. Moon, Hyeong Cheol; Park, Young Seok (2017). "Reduced GABAergic neuronal activity in zona incerta causes neuropathic pain in a rat sciatic nerve chronic constriction injury model". Journal of Pain Research. Volume 10: 1125–1134. doi:10.2147/jpr.s131104. ISSN 1178-7090. PMC 5436785. PMID 28546770.
  148. Scholz, J. (2005). "Blocking Caspase Activity Prevents Transsynaptic Neuronal Apoptosis and the Loss of Inhibition in Lamina II of the Dorsal Horn after Peripheral Nerve Injury". Journal of Neuroscience. 25 (32): 7317–7323. doi:10.1523/jneurosci.1526-05.2005. ISSN 0270-6474. PMC 6725303. PMID 16093381.
  149. Fu, Huiqun; Li, Fenghua; Thomas, Sebastian; Yang, Zhongjin (2017). "Hyperbaric oxygenation alleviates chronic constriction injury (CCI)-induced neuropathic pain and inhibits GABAergic neuron apoptosis in the spinal cord". Scandinavian Journal of Pain. 17 (1): 330–338. doi:10.1016/j.sjpain.2017.08.014. ISSN 1877-8860.
  150. Li, Caijuan; Lei, Yanying; Tian, Yi; Xu, Shiqin; Shen, Xiaofeng; Wu, Haibo; Bao, Senzhu; Wang, Fuzhou (2019). "The etiological contribution of GABAergic plasticity to the pathogenesis of neuropathic pain". Molecular Pain. 15: 174480691984736. doi:10.1177/1744806919847366. ISSN 1744-8069. PMC 6509976. PMID 30977423.