Skip to main content
Experimental Biology and Medicine logoLink to Experimental Biology and Medicine
. 2021 Feb 7;246(9):1069–1083. doi: 10.1177/1535370221989263

Influence of glutamate and GABA transport on brain excitatory/inhibitory balance

Sheila MS Sears 1, Sandra J Hewett 1,
PMCID: PMC8113735  PMID: 33554649

Abstract

An optimally functional brain requires both excitatory and inhibitory inputs that are regulated and balanced. A perturbation in the excitatory/inhibitory balance—as is the case in some neurological disorders/diseases (e.g. traumatic brain injury Alzheimer’s disease, stroke, epilepsy and substance abuse) and disorders of development (e.g. schizophrenia, Rhett syndrome and autism spectrum disorder)—leads to dysfunctional signaling, which can result in impaired cognitive and motor function, if not frank neuronal injury. At the cellular level, transmission of glutamate and GABA, the principle excitatory and inhibitory neurotransmitters in the central nervous system control excitatory/inhibitory balance. Herein, we review the synthesis, release, and signaling of GABA and glutamate followed by a focused discussion on the importance of their transport systems to the maintenance of excitatory/inhibitory balance.

Keywords: Excitatory/inhibitory balance, glutamate transport, GABA transport, EAATs, GATs, system xc

Impact statement

Excitatory and inhibitory (E/I) balance broadly refers to a stable global neuronal activity that is predominately achieved in brain by a coordinated balance between excitatory and inhibitory inputs. E/I imbalance contributes to the pathobiology of neurodevelopmental disorders, neurodegenerative/neurological disease, as well as, acute neurological disorders. Hence, a deeper understanding of the cellular and molecular mechanisms regulating physiological E/I balance is needed to improve current clinical strategies for managing disorders/diseases associated with E/I perturbations. Herein, we review the synthesis, release, and signaling of the principle CNS excitatory and inhibitory neurotransmitters (namely glutamate and GABA, respectively) and discuss the capacity of glutamate and GABA transporters to modulate release and uptake of neurotransmitter, as well as neural network activity.

Introduction

An optimally functional brain requires both excitatory and inhibitory inputs that are regulated and balanced. Optimal balance is necessary for efficient information processing at both the cellular and network level, each of which ultimately subserves cognition and behavior.1,2 At the cellular level, modulation of the intrinsic excitability and synaptic strength maintain balance, thereby regulating the overall firing probability of a neuron.36 At the network level, optimal balance maintains stable circuitry (reviewed in Gao and Penzes7 and Nelson and Valakh8). A perturbation in E/I balance has been implicated in the etiology and expression of autism spectrum disorders (ASD), schizophrenia and anxiety, cerebral ischemia, traumatic brain injury, epilepsy and substance abuse.913 As such, a deeper understanding of the cellular and molecular mechanisms regulating physiological E/I balance would allow for improvement of current clinical strategies for managing such disorders. In this review, we will focus solely on the neurotransmitters glutamate and GABA, including how they are made, released, and signal, as well as, whether and how specific transport processes influence their activity.

Gabaergic neural transmission

Gamma-aminobutyric acid (GABA) is present in high concentrations in the CNS. Studies in cortex showing that application by ionotophoresis inhibits cell firing paved the way for its classification as an inhibitory neurotransmitter in mature, adult mammalian brain.1417 Nonetheless, when the potassium/chloride cotransporter KCC2 levels are low, as occurs early in development, GABA signaling is excitatory, exerting trophic effects that contribute to normal neuronal growth and expansion.18,19

GABA synthesis and packaging

GABA is formed predominately by the enzymes glutamate decarboxylase 65 (GAD65) or GAD67, both of which use pyroxidine as a co-factor to convert glutamate to GABA in the CNS. GAD65, located in nerve terminals, produces GABA for classic tonic neurotransmission.20,21 GABA produced via GAD 67, expressed principally in the neuronal somata, functions in a non-neurotransmitter, metabolic, capacity contributing to synaptogenesis as well as to oxidation-reduction (redox) regulation.2022 Glutamine, serving as a precursor for glutamate synthesis via phosphate activated glutaminase, is also a substrate of GABA.23,24 Once synthesized, GABA is packaged for release into synaptic vesicles by vesicular transporters specific for GABA (VGATs) in a manner dependent on both the electrochemical and pH gradient.25,26

GABA signaling

GABA signals via membrane bound receptor proteins that either open chloride channels (GABAAR and GABACR) or activate a G protein (GABABR) (for review see Bormann27). Both lead to hyperpolarization of the cell membrane albeit via different mechanisms.28 GABAARs are composed of an obligatory α and β subunit (α1-6, β1-4) and at least one other subunit (γ1-4, δ, ε, π, and θ) and are ubiquitously expressed throughout the vertebrate CNS. GABACR are composed exclusively of ρ subunits (ρ1–3) (see Hedblom and Kirkness29 Bonnert et al.30 and reviewed in Zhang et al. and Macdonald and Olsen)31,32 and are near exclusively found in the retina.27,33 Sedative-hypnotics of the barbiturate and benzodiazepine family increase channel opening frequency,32,34,35 while bicuculline competitively antagonizes GABAARs,36 but these drugs have no effect on GABACRs.33 The GABABR is a G-protein coupled receptor formed by the dimerization of GABAB1 and GABAB2 subunits (reviewed in Heaney and Kinney37).

Bicuculline-resistant, GABABRs are activated by baclofen and inhibited by phaclofen.38 Located both presynaptically and postsynaptically, the primary effects of activation of GABABRs are the inhibition of adenylate cyclase, inhibition of voltage-gated Ca2+ channels, and activation of inwardly rectifying K+ channels, all of which contribute to a gradual and protracted synaptic inhibition.3941

GABA uptake

The length and size of GABA R mediated responses are controlled by four different sodium symporters belonging to the solute carrier 6 (Slc6) family. The high affinity GABA transporters—GABA transporter (GAT) 1–3—and low affinity betaine-GABA transporter 1 (BGT1) are all coupled to Na+ and Cl gradients.4247 In the CNS, GAT1 is expressed on inhibitory interneuron axon terminals, on the somato-dendritic compartment of developing interneurons, in pyramidal neurons, and in astrocytic processes and has a Km value of 8 µM.44,45,48 Selectively expressed on astrocytes,42,4954 the Km of GAT3 is 0.8 µM. GAT2 (Km = 18 µM) is found in high abundance in kidney and liver and is only weakly expressed in brain, most auspiciously in cells forming the pia and arachnoid barrier and in a subset of blood vessels.55,56 Transcript for BGT-1 has been found in both mouse and human brain 45,57 with a demonstrated Km of 80 µM.

Following its synaptic release, GABA uptake by neurons can be recycled/reloaded into synaptic vesicles to sustain subsequent rounds of release.58,59 Alternatively, it can be metabolized in both neurons and astrocytes by GABA-transaminase and succinic semialdehyde dehydrogenase, a process known as the GABA shunt, to replenish the TCA cycle,60,61 thereby constituting an alternative energy substrate.

Contribution of GABA transport to the maintenance of E/I balance

Because GABA released into the extracellular space is not enzymatically broken down, GAT activity is positioned to control the basal extracellular GABA in the extracellular space and E/I balance by acting as the primary mechanism to terminate synaptic inhibitory neurotransmission.

GAT1

In the CNS, GAT1 signaling accounts for ∼75–80% of GABA uptake,62 with high expression in GABAergic neurons of the neocortex, hippocampus, basal ganglia, brain stem, cerebellum, olfactory bulbs, and retina.63 Indeed, GABAAR-mediated currents derived from hippocampal slice recordings from GAT1−/− mice are increased compared to control slices.62 These results were recapitulated by using GAT1 selective inhibitors at wild-type synapses.64,65 Interestingly, a reduction in inhibitory tone at hippocampal presynaptic GABABR has been reported in GAT1−/− mice,62 which is potentially due to receptor desensitization following prolonged agonist exposure, as has been reported with other G-protein coupled receptors.66 This occurs alongside decreased phasic inhibition—manifest as a reduction in miniature inhibitory postsynaptic current (mIPSC) frequency as compared to wild-type.62 Given that presynaptic GABABR activation typically inhibits neurotransmitter release, this, and the finding that spontaneous IPSCs were unaffected by presynaptic GABAB tone,62 suggests that diminished phasic inhibition in GAT1−/− mice likely occurs via a GABAB receptor-independent mechanism.

Loss of GAT1 signaling also has a profound effect on behavior. Mice null for GAT1−/− display decreased depression and anxiety-like behavior,67,68 are less aggressive,69 and display signs of hypoalgesia, when compared to wild-type control mice.70 They also demonstrate impairment in hippocampus-dependent learning and memory.67,71 Moreover, selective GAT1 inhibitors—including tiagabine, NO-711, and DDPM-2571 72,73—have been demonstrated to recapitulate GAT1−/− behavioral phenotypes in wild-type rodents.7376 These findings have paved the way for human clinical trials of GAT1 inhibitors for treatment of behavioral complications of psychiatric disorders including aggression,77,78 anxiety,79,80 cocaine addiction 81 and for the improvement of pain symptoms in individuals suffering from sensory neuropathy.82 Inhibition of GAT1 also reduces the hyperexcitation of GABAergic neurons produced by opiate withdrawal.83,84

Apart from its role in mediating GABA uptake, multiple lines of in vitro evidence demonstrate GAT1 reversal represents a route of non-vesicular GABA release in the brain. Wu et al. found that the reversal potential of GAT1 is close to equilibrium to the cell’s resting membrane potential, such that action potentials and high-frequency firing are sufficient to induce GAT1 reversal.85 GAT1 reversal is observed in response to membrane depolarization in hippocampal cultures.86 Interestingly, an increase in GAT1 immunoreactivity has been observed in the hippocampus of rats following 4-AP and kainic acid-induced epileptiform activity 87,88 and recent studies show that the anti-seizure medications gabapentin and vigabatrin enhance GAT1-mediated GABA release,86,89 with vigabatrin potently increasing ambient [GABA]e and inducing tonic inhibition of neurons.90 Additionally, tiagabine, a selective GAT1 inhibitor—commonly prescribed as an add-on therapeutic option for epileptics with complex partial seizures 91—has been demonstrated to elevate the pentylenetetrazole (PTZ)-evoked seizure threshold and reduce generalized seizures in amygdala kindled rats.92,93 Thus, GAT1 inhibition or reversal—the latter occurring either naturally or in response to drug treatment—may represent a potent neuromodulatory mechanism to terminate ongoing seizure activity (E/I imbalance) by directly increasing GABAergic transmission.

Taken altogether, the bi-directionality of GAT1-mediated GABA transport, controlled by the dynamic driving force equilibrium, underscores its ability to modify brain excitability and behavior by modulating the level of both tonic and/or phasic inhibition.

GAT2

In the adult rodent brain, GAT2-mRNA is found in leptomeningeal cells, in ependymal cells that line the ventricles, and in cells that constitute the pia and arachnoidea.94,95 As might be expected based on these localization studies, GAT2 does not appear to influence network function, at least under physiological conditions.55,56,96 However, GAT2 knockout mice do have a slight elevation in brain taurine levels,55 in agreement with evidence that GAT2 expressed on blood vessels permits taurine efflux from brain to blood.55,56,96

BGT-1

BGT-1 levels are nearly one thousand times lower than those of GAT1.97 This, as well as its low affinity for GABA, suggests it may lack a role in the reuptake of extracellular GABA under physiological conditions. In keeping with this contention, seizure severity of BGT1 deficient mice (both male and female) elicited by acute administration of PTZ did not differ from wild-type littermates.97 Yet, pharmacological inhibition of BGT-1 reduced spontaneous interictal-like bursting activity recorded from brain slices taken from rats who experienced prolonged kainic acid-induced seizures.98 Also of interest, is data demonstrating that BGT-1 is up-regulated in astrocytes of cortical and hippocampal tissue taken from human Alzheimer disease (AD) patients, thus begging the question as to whether BGT-1 might regulate neuronal excitability imbalances demonstrated to occur in AD.99101

GAT3

Evidence indicates that concurrent block of both GAT1 and GAT3 in the hippocampus in vivo results in a synergistic enhancement of extracellular GABA levels and increased GABAA receptor tonic inhibition—suggesting that GAT3 represents an important GABA reuptake mechanism in brain.102 In support of this assertion, rats subjected to juvenile stress have decreased GAT3 mRNA expression in hippocampus, which was experimentally demonstrated to underlie increased inhibition and reduced paired-pulse facilitation, which persisted into adulthood.103 Interestingly, GAT3 expressed in Xenopus oocytes is inhibited by physiological levels of zinc, and immunohistochemistry studies in rat hippocampus indicate GAT3 is expressed at zinc-containing glutamatergic synapses in regions CA1 and CA3.104 These results suggest that zinc co-released with glutamate 105,106 could serve to enhance GABAergic transmission via GAT3 inhibition. However, in at least one study, selective antagonism of GAT3 using SNAP-5114 93 increased the excitability of neocortical interneurons, suggesting that (as discussed for GAT1) a reduction in transporter-mediated GABA release may be responsible for the reduction in GABA levels.107 Interestingly, an increase in GAT3 mRNA is observed in the amygdala and cortex of rats following amygdala-kindling,108 whereas GAT3 mRNA levels are decreased in the amygdala of alcohol-preferring rats as compared to controls, an effect also observed in the amygdala of alcohol-dependent humans.109 Whether these changes reflect dynamic regulation of GABA transport (uptake or release) in efforts to restore E/I balance requires confirmation. Given that GAT3 is predominately localized to astrocyte processes surrounding symmetric (typically inhibitory) and asymmetric (typically excitatory) synapses,50 future studies should address the cell-type specificity of all of these effects using astrocyte conditional knockout mice.

Glutamatergic neural transmission

Glutamate, the most abundant amino acid in the vertebrate nervous system, is found at concentrations three- to four-fold higher than the next three most abundant amino acids, aspartate, glutamine, and taurine.110,111 As mentioned previously, glutamate is a substrate for GABA synthesis as well as a precursor for other intermediates of the TCA cycle. It participates in both osmotic balance and ammonia homeostasis 112114 and is also incorporated into peptides, fatty acids, lipids, and proteins.115 The pivotal discovery that application of glutamate to brains of rats resulted in seizure activity provided evidence for its role as an excitatory neurotransmitter.116118 Later work determined that glutamate fulfills the five criteria for classification as a neurotransmitter: (1) localization to nerve terminals; (2) release upon neuronal stimulation; (3) activation of cognate receptors; (4) a rapid termination mechanism; and (5) application of glutamate mimics neuronal stimulation.110,119

A large proportion of neuronal synapses in the CNS (≈80–90%) release glutamate 111,120 contributing to a myriad of sensory, cognitive, and behavioral processes (for review see Hassel and Dingledine121). Maintenance of low basal extracellular glutamate concentrations as well as efficient release and uptake of the neurotransmitter are necessary to maintain proper balance of synaptic excitation and inhibition with imbalance leading to neurological disorders and disease states. For example, cognitive deficits associated with schizophrenia may result from glutamatergic hypofunction,12 while disproportionate release of glutamate and/or prolonged glutamate receptor activation can lead to over-excitation and excitotoxic neuronal cell death.122124 Hence, it is imperative that the release of glutamate be exquisitely controlled by specific and efficient uptake, as will be discussed below.

Glutamate synthesis and packaging

Because of its inability to cross the blood–brain barrier, glutamate is synthesized primarily from glutamine (glutamate-glutamine cycle) in both neurons and astrocytes by the action of phosphate-dependent mitochondrial glutaminase (for review see McKenna125). An additional source of glutamate results from the transamidation of α-ketoglutarate, a key intermediate in the TCA cycle, by the enzyme glutamate dehydrogenase.126 Once formed, 70–210 mM glutamate can be packaged into synaptic vesicles via one of three vesicular glutamate transporters ((VGLUT1, Slc17a6); (VGLUT2, Slc17a7); and (VGLUT3, Slc17a8)),127129 all of which rely on the vacuolar type H+-ATPase for function.127,130

Vesicular and non-vesicular release of glutamate

The majority of fast synaptic excitatory neurotransmission is facilitated by action potential driven, Ca2+-dependent vesicular release of glutamate from neurons. However, glutamate can be released by both astrocytes and neurons by other cellular mechanisms. For instance, when the gradients of Na+, K+ and H+ are disrupted across the plasma membrane, as occurs during cerebral ischemia, the Na+-dependent excitatory amino acid transporters (EAATS) reverse, dumping glutamate into the extracellular space.131133 Volume-regulated anion channels (VRACs) are glutamate permeable when physiological and/or pathological swelling occurs.134,135 Functional hemichannels in astrocytes can efflux amino acids, including glutamate.136 This mechanism of glutamate release may occur under physiological 137140 as well as pathophysiological 141,142 conditions. Additionally, purinergic P2X7 receptors are responsible for ATP-induced glutamate release.143 Finally, astrocytes express the protein machinery (see literature144147) that would support fusion-related release of neurotransmitter 148150 and, indeed, vesicular glutamate release has been described from astrocytes in response to neuronal activity.151,152 There is a suggestion that this may be a significant source of extracellular glutamate during development, but not in adulthood.153 Overall, the existence and importance of vesicular glutamate release from astrocytes in vitro appear incontrovertible; however, whether this occurs in vivo remains contested (for dual perspective reviews see Savtchouk and Volterra154 and Fiacco and McCarthy155) Finally, the ambient, basal levels of extracellular glutamate that bathe the CNS are maintained by the activity of a heteromeric amino acid transporter known as system xc, found near exclusively on astrocytes, that facilitates entry of cystine in exchange for glutamate in a one-to-one fashion.156

Glutamate signaling

Fast and slow excitatory synaptic transmission in the CNS occurs via ligand-gated ion channels (i.e. ionotropic (iGluRs)) and G-protein coupled (i.e. metabotropic (mGluR)) receptor subtypes, respectively (see reviews Traynelis et al. 157Niswender and Conn 158).

Cognate iGluRs, composed of four subunits that assemble as dimer pairs, were classified over 40 years ago according to the exogenous ligands that activate them, namely α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid receptors (AMPARs), kainic acid receptors (kainate, KARs), and N-methyl-D-aspartate receptors (NMDARs) (for review see Traynelis et al.157Lodge159). Of interest, delta receptors have also been classified as iGluR subtypes, but this is based solely on sequence homology as neither δ1 and δ2 are gated by glutamate.160,161 They do, however, respond to D-serine and glycine.162

AMPAR tetramers arise from a combination of GluA1-4 subunits.163 Glutamate binding to AMPAR facilitates the fast opening of an ion channel pore, fluxing Na+ in and K+ out, which then rapidly desensitizes.164 The editing of the GluA2 subunit (Q/R) is responsible for the impermeability of AMPARs to calcium..165 However, AMPARs can flux calcium when the tetramer contains unedited GluA2 subunits and/or lacks GluA2 altogether.166,167 Enriched at glutamatergic synapses, AMPARs mediate fast synaptic transmission and are a key determinate of the morphology of the dendritic spine.168,169 Membrane trafficking of AMPARs into and out of the synapse also regulates synaptic strength and plasticity.170

Tetraheteromers of GluK1-5 form KARs that show prominent localization to both pre- and post-synaptic sites in the cerebellum (comprised of GluK1, 2, and 5) and hippocampus (comprised of GluK2, 3, 4, and 5).171 Interestingly, KARs are allosterically modulated by monovalent anions and cations, which serve to stabilize the ligand binding core domain.172,173 Similar to AMPARs, Q/R RNA editing, in this case of GluK5 and GluK6 subunits, renders the receptor impermeable to calcium.174,175 Unlike AMPARs, KAR-mediated excitatory postsynaptic currents (EPSCs) are small, with both slow rise and decay times.176,177 Depending on the concentration of agonist, activation of presynaptic KARs results in either synaptic facilitation or depression at excitatory CA3-CA1 or mossy fiber-CA3 synapses.178180 Presynaptic KAR activation can also depress GABA release in the hippocampus, presumably through a novel second messenger metabotropic signaling mechanism.181

Functional NMDARs are composed of a combination of two GluN1 subunits (termed the obligate receptor subunit; glycine/serine binding) and two GluN2 (GluN2A-D, glutamate binding) and/or GluN3 (GluN3A-B, glycine binding) subunits, the unique composition of which renders distinct physiological properties to each receptor combination along with regional specificity (reviewed in Sanz-Clemente et al. 182). For instance, assembly of GluN1 with GluN3, in the absence of GluN2, creates an excitatory glycine receptor located at sites distant from synaptic terminals.183 In contrast, a significant proportion of forebrain NMDARs are triheteromers found post-synaptically and feature two GluN1 subunits together with two different GluN2 subunits (GluN2A-D), the activation of which creates the slow component of EPSCs. These NMDARs have several features that differentiate them from other iGluRs (for more detail see Glasgow et al.184) First, they function as coincidence detectors, requiring both ligand and voltage for channel opening. Second, their deactivation kinetics are slow, creating an opportunity for temporal integration of synaptic activity.185,186 Third, they show remarkable calcium permeability, thus making them essential players in both Hebbian and homeostatic types of plasticity.187,188

Metabotropic glutamate receptors (mGluRs), eight in total, do not flux ions but instead are coupled to G-proteins that possess seven transmembrane spanning regions, the activation of which initiates distinct intracellular signaling cascades that result in diverse cellular and electrophysiological effects (for reviews see literature158,189,190). mGluRs are divided into three groups (Group I, II, and III) based on amino acid sequence homology and the intracellular second messenger cascade that they initiate. Coupled to phospholipase C, Group I receptors (mGluR1 and 5) hydrolyze phosphoinositol, mobilize calcium, and facilitate protein phosphorylation. Negatively coupled to adenylate cyclase, Group II (mGluR 2 and 3) and Group III (mGluR4 and mGluR6-8) receptors decrease cyclic AMP production and ultimately protein phosphorylation. Activation of mGluRs in the CNS has diverse functional outcomes ranging from activation or inhibition of K+ and Ca2+ channels, potentiation and inhibition of AMPA and NMDA receptor-mediated responses, and/or presynaptic facilitation or inhibition of neurotransmitter release.189,190

Glutamate uptake

Glutamate is not broken down in the extracellular space, and as such excitatory amino acid transporters (EAATs) are responsible for termination of glutamate signaling by its removal from the synapse following release.191 EAAT1-5 are members of the Slc1 family of transporters.192 Removal of glutamate by EAATs is said to be electrogenic as one K+ ion is transported out of the cell each time a glutamate anion and three Na+ ions are transported in.131,193197 Uptake via EAATs is also associated with an uncoupled Cl gradient,198200 which may also contribute to a reduction in excitability.

Found throughout the brain, EAAT1 (GLAST) 201,202 is localized exclusively on astrocytes.203 EAAT2 (GLT-1a,b) 204 is also predominately an astrocyte protein although a minor proportion of GLT-1a can be found on certain axon terminals.205208 EAAT3 (EAAC1) 209,210 is localized to neuronal somata and dendrites.211,212 EAAT4 localizes to cerebellar Purkinje cells,213,214 but also astrocytes,215 while EAAT5 expression appears to be limited to photoreceptors and bipolar cells of the retina.216 In the CNS, the majority of extracellular glutamate clearance is performed by EAAT2217219.

Role of glutamate transport in maintenance of E/I balance

Excitatory amino acid transporter 1

In humans, a heterozygous mutation in EAAT1 phenocopies with reductions in glutamate uptake that likely contributes to neuronal hyperexcitability resulting in episodic ataxia and, depending upon the extent of the reduction, seizures.200,220,221 In mice, loss of EAAT1 does not result in spontaneous seizure generation, but the duration of seizures elicited by electrical stimulation of the amygdala is significantly prolonged, whereas the latency to seizure induced via systemic administration of PTZ is shortened and the seizures themselves more severe.222 EAAT1 null mice also show locomotor hyperactivity when placed in a novel open field.223 Additional studies with this mouse demonstrate abnormalities on behavioral symptoms (positive, negative, and attentional/cognitive symptoms) associated with the developmental disorder schizophrenia,224 which arises from alterations in E/I balance.7

Excitatory amino acid transporter 2

Consistent with its outsized role in synaptic glutamate uptake, mice with a genetic deletion of astrocyte (but not neuronal) EAAT2 demonstrate excessive synaptic glutamate, which precipitates spontaneous seizures that are lethal by three–six weeks of age.225,226 Of interest, Amara et al. determined that Eaat2 (Slc1a2) is located on mouse chromosome 2 near quantitative trait loci shown to modulate seizure frequency in mouse models of epilepsy and alcohol withdrawal.227 In humans, glutamate levels are increased in interictal epileptogenic foci,228 leading to the speculation that clearance is impaired. In keeping with this idea, protein expression of EAAT2 (and EAAT1) in the CA1 hippocampus of patients with temporal lobe epilepsy was reduced by 25% (and 40%), respectively.229 Reduction of EAAT2 protein expression at human neocortical epileptic foci has also been described.230 These reductions could be caused by the production of alternative EAAT2 mRNA splice variants.231 Finally, seizure control in both mouse and monkey models of epilepsy was achieved via strategies that upregulate EAAT2 expression.232,233 Upregulation of EAAT2 expression was also shown to attenuate alcohol consumption in male alcohol preferring rats,234 contributing to the idea that glutamate transport might be a target for treatment of alcohol dependence.235

Very recently, it was demonstrated that chemically elicited cortical spreading depression (CSD), a pathological neural depolarization that underlies migraine pathophysiology 236238 as well as secondary neuronal damage and infarct expansion following cerebral ischemia,239,240 occurs with increased frequency and velocity in EAAT2 astrocyte conditional knock-out mice.241 In contrast, the germ‐line EAAT1 and EAAT3 null mutants show no such effect.241

Excitatory amino acid transporter 3

Although approximately 100-fold less abundant than EAAT2 246, EAAT3/EAAC1, primarily found on dendrites and soma of hippocampal neurons,212 is known to regulate the duration of glutamate in the synaptic space that immediately surrounds active terminals. This prevents glutamate spillover to extrasynaptic regions,242,243 which are morphologically defined as receptors lying more than 100 nm from the post-synaptic density.244 Accordingly, the slower component of CA1 pyramidal cell glutamatergic EPSCs is enhanced when EAAT3/EAAC1 242 is absent and extrasynaptic NMDA receptors are activated in EAAT3 knockout mice.245 This latter result could, at best, result in modulation of synaptic activity246 and, at worse, contribute to neurodegeneration.247,248 Interestingly, a reduction in both EAAT3 message and protein was found in human neocortical tissue taken from epileptic foci in comparison to non-epileptic regions of similar neuronal density using quantitative real‐time PCR and immunoblotting, respectively.230 Thus, it may not be surprising that chronic antisense oligonucleotide treatment against EAAT3/EAAC1 in rats resulted in behavioral episodes of staring/freezing that correlated with electroencephalogram changes manifest by short runs of rhythmic spikes.249,250 The physiological basis of EAAT3/EAAC1 antisense oligonucleotide epileptogenesis was attributed, in part, to hippocampal GABA synthesis reduction.250 However, these phenotypes were not recapitulated in an EAAT3/EAAC1 null mice.251 Additionally, EAAT3 mRNA and protein expression are enhanced (not reduced) by approximately 3-fold in granule cells of the dentate gyrus of pilocarpine-treated rats that seize spontaneously as compared to control rats.252 EAAT3 message is also higher in granule cells of the dentate gyrus taken from human patients with temporal lobe epilepsy.252 Whether this increase represents a compensatory change to increase glutamate clearance or enhance GABA production/activity,250,253 or is merely an epiphenomenon, remains to be definitively determined.

With respect to neuropsychiatric disorders, a functionally relevant deletion of Slc1a1, which encodes for EAAT3/EAAC1 has been shown to co-segregate with psychotic disorders (e.g. bipolar disorder and schizophrenia) in an extended 5-generation pedigree.254 Furthermore, mice with EAAT3/EAAC1 haploinsufficiency show biochemical, behavioral, and histological changes that reflect an altered redox state, congruous with changes found in patients with schizophrenia.255 Finally, genetic linkage and association studies performed in obsessive convulsive disorder (OCD), itself linked with cortical excitability abnormalities,256 point to gene variants in Slc1a1 (for review see Escobar et al. 257). While mice null for EAAT3 do not show behaviors consistent with OCD, overexpression of EAAT3 in forebrain neurons alone does result in anxiety-like and repetitive behaviors, which are also often reported in persons diagnosed with OCD.245

Excitatory amino acid transporter 4

Soma and dendrites of cerebellar Purkinje neurons show prominent expression of EAAT4, although it is not restricted to these cells.258260 While EAAT4 shows high affinity for glutamate,258,261 it takes up just 10% or less of released glutamate at climbing fiber (CF) synapses;262 the balance is removed by Bergmann glial cells via EAAT1.263,264 Consistent with this notion, no Purkinje cell death followed 5 min global ischemia in mice null for EAAT4, whereas significant loss occurred in mice lacking EAAT1.265 Supporting the idea that EAAT4 transporters prevent glutamate spillover to adjacent synapses, a pronounced tail current (seconds in length) appears during the decay phase, the initial kinetics of which is not different, of both CF- and parallel fiber (PF)-EPSCs when evoked in slices derived from mice deficient in EAAT4.266

Excitatory amino acid transporter 5

EAAT5 is localized near exclusively to the retina (see Dalet et al.260) with expression in the synaptic terminals of photoreceptors and rod bipolar cells.216,267 EAAT5 exhibits two distinct properties, acting both as a rather ineffective, low-affinity and low-capacity glutamate transporter and as a glutamate-gated inhibitory “receptor”.267269 This EAAT5-mediated anion channel, optimized for conduction in the negative voltage range,270 is postulated to reduce excitability of neurons by maintaining membrane potential at its optimum.

Electrogenecity and anions

Electrophysiological measurement of tonic NMDAR activity in acute brain slice has informed our understanding of the amount of ambient glutamate in the extracellular space, which has been reported to range from 25 to 90nM.271275 Studies using in vivo microdialysis report higher concentrations (0.2–35 µM),276279 which could be due to tissue damage inflicted by the sampling probe.280 These low values are maintained despite the fact that intracellular concentration of glutamate ranges from high μM to mM concentrations in astrocytes and neurons.128,129,281 This is because glutamate transport is electrogenic,131,193197 allowing for efficient uptake of glutamate against this concentration gradient. The ion-coupled substrate transport current generated by each EAAT subtype varies with the bioenergetics tightly controlling the rate and amount of glutamate removed (for detailed review see Divito and Underhill282). Glutamate uptake creates a chloride flux (anion channel) that is thermodynamically uncoupled to transport but is generated when Na+ ions and/or glutamate bind to the transporter.198,199,283 For a detailed review of the molecular transport mechanisms see Grewer et al.284 EAAT1, EAAT2, and EAAT3 produce smaller anion currents as compared to EAAT4 and EAAT5, both of which show large chloride conductance.216,283,285 It has been suggested that this chloride conductance shapes excitatory signaling by counterbalancing the entry of positive charges that occurs along with glutamate influx, thereby preventing depolarization of the cell.286 Additionally, it could serve to clamp the membrane potential at negative values, further inhibiting glutamate release and/or supporting electrogenic glutamate uptake by favoring Na+ entry.285 Most interestingly, the anion current could effectively function as glutamate-dependent inhibitory receptor, thereby directly counteracting glutamate’s excitatory effects.284,287

System xc

System xc (Sxc)—described by Bannai and Kitamura in 1980—is a Na+-independent, Cl -dependent, heteromeric amino acid transporter that functions physiologically to import L-cystine in exchange for L-glutamate in a 1:1 ratio.156,288 Transport is electroneutral and limited to amino acids in their anionic forms. While Sxc is expressed in cultured microglia,289,290 neurons,290293 HT22 neuronal cell line,294 rodent astrocytes, 290,295 and human glioma cell lines,292,295299 astrocytes appear to be the main cell type expressing Sxc in the mature brain in vivo. To wit: Pow et al. demonstrated via immunocytochemical analyses that α-aminoadipate, a substrate inhibitor of Sxc, was absent from neurons and oligodendrocytes but accumulated in astrocytes, radial, and Bergman glia.299 Transcriptome analysis of parenchymal cells from mouse and human cortex, revealed enrichment of the transcript for the substrate specific light chain of Sxc (xCT encoded by Slc7a11) in astrocytes when compared to neurons, microglia, endothelial cells, and other cell types.144,300 Finally, immunohistochemical analysis for xCT in adult mouse brain showed that Sxc is expressed in a subset of astrocytes but not in neurons, microglia, or oligodendrocytes.301 Labeling of xCT was found in most brain regions including the molecular layer and the stratum lacumosum moleculare of the hippocampus, the striatum, the hypothalamus, the thalamus, and the cortex; it was also concentrated a the blood/brain/cerebral spinal fluid barriers.301

Apart from transporting cystine into cells, a process important for cellular redox balance, export of glutamate by Sxc—estimated to be 0.6 µM/s 302—contributes to/maintains basal extracellular glutamate concentrations,276,279,303309 which itself contributes importantly to brain E/I balance. 273,274,303,308,310318 Numerous studies demonstrate that Sxc -derived extracellular glutamate, specifically, is important for maintaining balanced transmission. For example, increased glutamate receptor clustering and excitatory junction potentials occurred in association with reductions in glutamate at the neuromuscular junction of the Drosophila melanogaster mutant for Sxc, an effect phenocopied by bathing larvae in low glutamate concentrations.308 Findings that Sxc -derived glutamate is important for maintenance of synaptic strength were also reported in mouse CA1 hippocampus taken from male mice mutant for xCT−/−. Specifically, AMPAR immunoreactivity was enhanced as was both spontaneous and evoked excitatory currents, effects phenocopied by maintaining slices in glutamate-free bathing solution and/or by incubation with a pharmacological inhibitor of Sxc−.311 Given this, it is somewhat surprising that a higher dose of the chemoconvulsant pilocarpine or kainic acid, provided via intravenous infusion, is needed to precipitate behavioral seizures in transgenic xCT−/− mice when compared to wild-type control mice.276 Also surprising is that following a single intraperitoneal (i.p.) dose of NMDA, latency to convulsive seizure is increased and incidence of mortality is reduced in a cohort of xCT−/− mice.276 This contrasts with our own findings, which are in keeping with the electrophysiological results described above, that demonstrate a reduction in convulsive seizure threshold of Slc7a11sut/sut, as compared to Slc7a11+/+ littermates, in response to a single dose of the chemoconvulsant PTZ or kainic acid delivered i.p. (manuscript in review).

A deleterious role for aberrant Sxc expression is demonstrated by the following few studies. A rapid increase in Sxc activity, demonstrated by positron emission tomography, in rat brain followed a focal cerebral ischemic insult induced by transient occlusion of the middle cerebral artery.319 In this same study, cell death and neuronal currents induced by oxygen–glucose deprivation in slice and slice culture—so called anoxic depolarizations—were both reduced by block of Sxc−.319 Previously, we found that when astrocyte activity of Sxc is enhanced, glutamate-mediated excitotoxic neuronal death during simulated ischemia is also increased.290,320,321 Finally, both glutamate concentrations and Sxc levels are enhanced in glioma tissue,296,322 with evidence in humans showing that enhanced xCT expression in tumors positively correlates with degree of tumor invasion and with shortened survival.323 Moreover, pharmacological inhibition of Sxc reduces seizure frequency in glioma-bearing mice and peritumoral glutamate levels in human patients.324,325 Increased xCT expression was found in post-mortem samples of dorsolateral prefrontal cortex of patients diagnosed with schizophrenia,326 although the significance to disease pathogenesis or symptomology remains to be determined. Finally, reduced activity of Sxc in the nucleus accumbens of rats followed repeated cocaine exposure was demonstrated convincingly to be associated with pathological changes in extracellular glutamate levels as well as their compulsive drug seeking behavior.278,305,327329 Likewise, xCT levels were reduced in nucleus accumbens and the ventral tegmental area of rats self-administering nicotine.330 In this same study, human smokers treated with N-acetyl-cysteine, a cysteine prodrug that activates Sxc, reported they smoked fewer cigarettes.330

With respect to more conventional behaviors, mice lacking Sxc (both transgenic xCT−/− as well as Slc7a11sut/sut mice) show reduced alternations in the three arm spontaneous alternation task, indicating a deficit in spatial working memory.276,331 Additionally, impaired functioning in both amygdala and hippocampal-dependent fear conditioning tasks, as well as in a hippocampal-dependent passive avoidance tasks, representing aberrant learning and/or memory, has been reported in male Slc7a11sut/sut mice.332 Notably, CA1-Schaeffer collateral cellular long-term potentiation—a synaptic mechanism thought to underlie learning and memory–is reduced in these same mice.332 Other studies demonstrate that physiological Sxc signaling influences behavioral anxiety and despair. For instance, as compared to wild-type, male transgenic mice null for xCT occupy the illuminated portion of the light/dark box and spend an increased amount of time in the open during the open field test.333 As compared to wild-type control mice, they also show reduced immobility and enhanced climbing behaviors in the tail suspension and forced swim tests.333

Overall, evidence across multiple species using different paradigms indicate that physiological Sxc activity contributes importantly to the maintenance of E/I balance in brain.

Conclusions

Stable global neuronal activity is achieved in forebrain by a coordinated and dynamically regulated balance between excitatory (chiefly glutamatergic) and inhibitory (chiefly GABAergic) inputs. This coordination is essential for the normal functioning of most complex brain processes, with imbalances contributing to the pathobiology of neurodevelopmental disorders, neurodegenerative/neurological disease, as well as, acute neurological disorders. Most studies researching E/I imbalance focus on neurotransmitter levels and/or concentrate on receptor signaling. However, the capacity of glutamate and GABA transporters to both modulate release and uptake of neurotransmitter, as well as neural network activity, in a cell-type specific manner underscores their important contribution to maintenance of physiological balance. Hence, a comprehensive understanding of how these transporters work normally and how their physiological function may be altered under pathophysiological conditions are the first steps to identifying novel therapeutic avenues and targets to prevent or mitigate imbalance.

Footnotes

AUTHORS’ CONTRIBUTIONS: SMSS and SJH both contributed equally to the writing of this review.

DECLARATION OF CONFLICTING INTERESTS: The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

FUNDING: The author(s) disclosed receipt of the following financial support for the research, authorship, and/or publication of this article: This work was supported by 1R01NS105767 and 2R01NS051445 from the NINDS of the National Institutes of Health.

ORCID iD: Sandra J Hewett https://orcid.org/0000-0002-2987-3791

References

  • 1.Cannon WB. The wisdom of the body. Norton and Co. New York, NY, 1932 [Google Scholar]
  • 2.Zhou S, Yu Y. Synaptic EI balance underlies efficient neural coding. Front Neurosci 2018; 12:46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Megıas M, Emri Z, Freund T, Gulyas A. Total number and distribution of inhibitory and excitatory synapses on hippocampal CA1 pyramidal cells. Neuroscience 2001; 102:527–40 [DOI] [PubMed] [Google Scholar]
  • 4.Pratt KG, Aizenman CD. Homeostatic regulation of intrinsic excitability and synaptic transmission in a developing visual circuit. J Neurosci 2007; 27:8268–77 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Beck H, Yaari Y. Plasticity of intrinsic neuronal properties in CNS disorders. Nat Rev Neurosci 2008; 9:357. [DOI] [PubMed] [Google Scholar]
  • 6.Turrigiano GG, Leslie KR, Desai NS, Rutherford LC, Nelson SB. Activity-dependent scaling of quantal amplitude in neocortical neurons. Nature 1998; 391:892–6 [DOI] [PubMed] [Google Scholar]
  • 7.Gao R, Penzes P. Common mechanisms of excitatory and inhibitory imbalance in schizophrenia and autism spectrum disorders. Curr Mol Med 2015; 15:146–67 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Nelson SB, Valakh V. Excitatory/inhibitory balance and circuit homeostasis in autism spectrum disorders. Neuron 2015; 87:684–98 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Luscher B, Fuchs T. GABAergic control of depression-related brain states. Advances in pharmacology. Amsterdam: Elsevier, 2015, pp. 97–144 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Coghlan S, Horder J, Inkster B, Mendez MA, Murphy DG, Nutt DJ. GABA system dysfunction in autism and related disorders: from synapse to symptoms. Neurosci Biobehav Rev 2012; 36:2044–55 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Olsen RW, Avoli M. GABA and epileptogenesis. Epilepsia 1997; 38:399–407 [DOI] [PubMed] [Google Scholar]
  • 12.Moghaddam B, Javitt D. From revolution to evolution: the glutamate hypothesis of schizophrenia and its implication for treatment. Neuropsychopharmacology 2012; 37:4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Meldrum BS. The role of glutamate in epilepsy and other CNS disorders. Neurology 1994; 44(11 Suppl 8):S14–23 [PubMed] [Google Scholar]
  • 14.Krnjević K, Phillis J. Iontophoretic studies of neurones in the mammalian cerebral cortex. J Physiol 1963; 165:274–304 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Krnjević K, Whittaker V. Excitation and depression of cortical neurones by brain fractions released from micropipettes. J Physiol 1965; 179:298–322 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Awapara J, Landua AJ, Fuerst R, Seale B. Free γ-aminobutyric acid in brain. J Biol Chem 1950; 187:35–9 [PubMed] [Google Scholar]
  • 17.Elliott K. γ-Aminobutyric acid and other inhibitory substances. Br Med Bull 1965; 21:70–5 [DOI] [PubMed] [Google Scholar]
  • 18.Platel J-C, Stamboulian S, Nguyen I, Bordey A. Neurotransmitter signaling in postnatal neurogenesis: the first leg. Brain Res Rev 2010; 63:60–71 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Ge S, Pradhan DA, Ming G-L, Song H. GABA sets the tempo for activity-dependent adult neurogenesis. Trends Neurosci 2007; 30:1–8 [DOI] [PubMed] [Google Scholar]
  • 20.Martin DL, Rimvall K. Regulation of γ‐aminobutyric acid synthesis in the brain. J Neurochem 1993; 60:395–407 [DOI] [PubMed] [Google Scholar]
  • 21.Waagepetersen HS, Sonnewald U, Schousboe A. The GABA paradox: multiple roles as metabolite, neurotransmitter, and neurodifferentiative agent. J Neurochem 1999; 73:1335–42 [DOI] [PubMed] [Google Scholar]
  • 22.Lamigeon C, Bellier J, Sacchettoni S, Rujano M, Jacquemont B. Enhanced neuronal protection from oxidative stress by coculture with glutamic acid decarboxylase‐expressing astrocytes. J Neurochem 2001; 77:598–606 [DOI] [PubMed] [Google Scholar]
  • 23.Westergaard N, Sonnewald U, Petersen SB, Schousboe A. Glutamate and glutamine metabolism in cultured GABAergic neurons studied by 13C NMR spectroscopy may indicate compartmentation and mitochondrial heterogeneity. Neurosci Lett 1995; 185:24–8 [DOI] [PubMed] [Google Scholar]
  • 24.Waagepetersen H, Bakken I, Larsson O, Sonnewald U, Schousboe A. Comparison of lactate and glucose metabolism in cultured neocortical neurons and astrocytes using 13C-NMR spectroscopy. Dev Neurosci 1998; 20:310–20 [DOI] [PubMed] [Google Scholar]
  • 25.Chaudhry FA, Reimer RJ, Bellocchio EE, Danbolt NC, Osen KK, Edwards RH, Storm-Mathisen J. The vesicular GABA transporter, VGAT, localizes to synaptic vesicles in sets of glycinergic as well as GABAergic neurons. J Neurosci 1998; 18:9733–50 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.McIntire SL, Reimer RJ, Schuske K, Edwards RH, Jorgensen EM. Identification and characterization of the vesicular GABA transporter. Nature 1997; 389:870. [DOI] [PubMed] [Google Scholar]
  • 27.Bormann J. The ‘ABC’of GABA receptors. Trends Pharmacol Sci 2000; 21:16–9 [DOI] [PubMed] [Google Scholar]
  • 28.Rivera C, Voipio J, Payne JA, Ruusuvuori E, Lahtinen H, Lamsa K, Pirvola U, Saarma M, Kaila K. The K+/Cl− co-transporter KCC2 renders GABA hyperpolarizing during neuronal maturation. Nature 1999; 397:251. [DOI] [PubMed] [Google Scholar]
  • 29.Hedblom E, Kirkness EF. A novel class of GABAA receptor subunit in tissues of the reproductive system. J Biol Chem 1997; 272:15346–50 [DOI] [PubMed] [Google Scholar]
  • 30.Bonnert TP, McKernan RM, Farrar S, Le Bourdellès B, Heavens RP, Smith DW, Hewson L, Rigby MR, Sirinathsinghji DJ, Brown N. θ, a novel γ-aminobutyric acid type a receptor subunit. Proc Natl Acad Sci U S A 1999; 96:9891–6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Zhang D, Pan Z-H, Awobuluyi M, Lipton SA. Structure and function of GABAC receptors: a comparison of native versus recombinant receptors. Trends Pharmacol Sci 2001; 22:121–32 [DOI] [PubMed] [Google Scholar]
  • 32.Macdonald RL, Olsen RW. GABAA receptor channels. Annu Rev Neurosci 1994; 17:569–602 [DOI] [PubMed] [Google Scholar]
  • 33.Bormann J, Feigenspan A. GABAC receptors. Trends Neurosci 1995; 18:515–9 [DOI] [PubMed] [Google Scholar]
  • 34.Pritchett DB, Luddens H, Seeburg PH. Type I and type II GABAA-benzodiazepine receptors produced in transfected cells. Science 1989; 245:1389–92 [DOI] [PubMed] [Google Scholar]
  • 35.Chiara DC, Jayakar SS, Zhou X, Zhang X, Savechenkov PY, Bruzik KS, Miller KW, Cohen JB. Specificity of intersubunit general anesthetic binding sites in the transmembrane domain of the human α1β3γ2 GABAA receptor. J Biol Chem 2013; 288:19343–57 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Andrews P, Johnston G. GABA agonists and antagonists. Biochem Pharmacol 1979; 28:2697–702 [DOI] [PubMed] [Google Scholar]
  • 37.Heaney CF, Kinney JW. Role of GABAB receptors in learning and memory and neurological disorders. Neurosci Biobehav Rev 2016; 63:1–28 [DOI] [PubMed] [Google Scholar]
  • 38.Bowery N, Doble A, Hill D, Hudson A, Shaw J, Turnbull M. Baclofen: a selective agonist for a novel type of GABA receptor proceedings. Br J Pharmacol 1979; 67:444P. [PMC free article] [PubMed] [Google Scholar]
  • 39.Pérez-Garci E, Gassmann M, Bettler B, Larkum ME. The GABA B1b isoform mediates long-lasting inhibition of dendritic Ca 2+ spikes in layer 5 somatosensory pyramidal neurons. Neuron 2006; 50:603–16 [DOI] [PubMed] [Google Scholar]
  • 40.Sakaba T, Neher E. Direct modulation of synaptic vesicle priming by GABA B receptor activation at a glutamatergic synapse. Nature 2003; 424:775. [DOI] [PubMed] [Google Scholar]
  • 41.Kabashima N, Shibuya I, Ibrahim N, Ueta Y, Yamashita H. Inhibition of spontaneous EPSCs and IPSCs by presynaptic GABAB receptors on rat supraoptic magnocellular neurons. J Physiol 1997; 504:113–26 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Radian R, Bendahan A, Kanner B. Purification and identification of the functional sodium-and chloride-coupled gamma-aminobutyric acid transport glycoprotein from rat brain. J Biol Chem 1986; 261:15437–41 [PubMed] [Google Scholar]
  • 43.Guastella J, Nelson N, Nelson H, Czyzyk L, Keynan S, Miedel MC, Davidson N, Lester HA, Kanner BI. Cloning and expression of a rat brain GABA transporter. Science 1990; 249:1303–6 [DOI] [PubMed] [Google Scholar]
  • 44.Liu Q-R, Lopez-Corcuera B, Mandiyan S, Nelson H, Nelson N. Molecular characterization of four pharmacologically distinct gamma-aminobutyric acid transporters in mouse brain corrected. J Biol Chem 1993; 268:2106–12 [PubMed] [Google Scholar]
  • 45.Lopez-Corcuera B, Liu Q-R, Mandiyan S, Nelson H, Nelson N. Expression of a mouse brain cDNA encoding novel gamma-aminobutyric acid transporter. J Biol Chem 1992; 267:17491–3 [PubMed] [Google Scholar]
  • 46.Yamauchi A, Uchida S, Kwon HM, Preston A, Robey RB, Garcia-Perez A, Burg M, Handler J. Cloning of a Na (+)-and Cl (−)-dependent betaine transporter that is regulated by hypertonicity. J Biol Chem 1992; 267:649–52 [PubMed] [Google Scholar]
  • 47.Borden LA, Smith KE, Hartig PR, Branchek TA, Weinshank RL. Molecular heterogeneity of the gamma-aminobutyric acid (GABA) transport system. Cloning of two novel high affinity GABA transporters from rat brain. J Biol Chem 1992; 267:21098–104 [PubMed] [Google Scholar]
  • 48.Liu Q-R, Mandiyan S, Nelson H, Nelson N. A family of genes encoding neurotransmitter transporters. Proc Natl Acad Sci U S A 1992; 89:6639–43 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Minelli A, Brecha N, Karschin C, DeBiasi S, Conti F. GAT-1, a high-affinity GABA plasma membrane transporter, is localized to neurons and astroglia in the cerebral cortex. J Neurosci 1995; 15:7734–46 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Minelli A, DeBiasi S, Brecha NC, Zuccarello LV, Conti F. GAT-3, a high-affinity GABA plasma membrane transporter, is localized to astrocytic processes, and it is not confined to the vicinity of GABAergic synapses in the cerebral cortex. J Neurosci 1996; 16:6255–64 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Melone M, Ciappelloni S, Conti F. A quantitative analysis of cellular and synaptic localization of GAT-1 and GAT-3 in rat neocortex. Brain Struct Funct 2015; 220:885–97 [DOI] [PubMed] [Google Scholar]
  • 52.Ribak CE, Tong WM, Brecha NC. GABA plasma membrane transporters, GAT‐1 and GAT‐3, display different distributions in the rat hippocampus. J Comparat Neurol 1996; 367:595–606 [DOI] [PubMed] [Google Scholar]
  • 53.Yan X-X, Cariaga WA, Ribak CE. Immunoreactivity for GABA plasma membrane transporter, GAT-1, in the developing rat cerebral cortex: transient presence in the somata of neocortical and hippocampal neurons. Brain Res Dev Brain Res 1997; 99:1–19 [DOI] [PubMed] [Google Scholar]
  • 54.Frahm C, Engel D, Piechotta A, Heinemann U, Draguhn A. Presence of γ-aminobutyric acid transporter mRNA in interneurons and principal cells of rat hippocampus. Neurosci Lett 2000; 288:175–8 [DOI] [PubMed] [Google Scholar]
  • 55.Zhou Y, Holmseth S, Guo C, Hassel B, Höfner G, Huitfeldt HS, Wanner KT, Danbolt NC. Deletion of the γ-aminobutyric acid transporter 2 (GAT2 and SLC6A13) gene in mice leads to changes in liver and brain taurine contents. J Biol Chem 2012; 287:35733–46 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Zhou Y, Holmseth S, Hua R, Lehre AC, Olofsson AM, Poblete-Naredo I, Kempson SA, Danbolt NC. The betaine-GABA transporter (BGT1, slc6a12) is predominantly expressed in the liver and at lower levels in the kidneys and at the brain surface. Am J Physiol Renal Physiol 2011; 302:F316–F28 [DOI] [PubMed] [Google Scholar]
  • 57.Borden LA, Smith KE, Gustafson EL, Branchek TA, Weinshank RL. Cloning and expression of a betaine/GABA transporter from human brain. J Neurochem 1995; 64:977–84 [DOI] [PubMed] [Google Scholar]
  • 58.Eulenburg V, Gomeza J. Neurotransmitter transporters expressed in glial cells as regulators of synapse function. Brain Res Rev 2010; 63:103–12 [DOI] [PubMed] [Google Scholar]
  • 59.Coulter DA, Eid T. Astrocytic regulation of glutamate homeostasis in epilepsy. Glia 2012; 60:1215–26 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Fonnum F, Fyske E. Uptake and storage of GABA in synaptic vesicles. GABA in the nervous system: the view at fifty years. Philadelphia: Lippincott Williams and Wilkins, 2000, pp.51–64 [Google Scholar]
  • 61.Bak LK, Schousboe A, Waagepetersen HS. The glutamate/GABA‐glutamine cycle: aspects of transport, neurotransmitter homeostasis and ammonia transfer. J Neurochem 2006; 98:641–53 [DOI] [PubMed] [Google Scholar]
  • 62.Jensen K, Chiu C-S, Sokolova I, Lester HA, Mody I. GABA transporter-1 (GAT1)-deficient mice: differential tonic activation of GABAA versus GABAB receptors in the hippocampus. J Neurophysiol 2003; 90:2690–701 [DOI] [PubMed] [Google Scholar]
  • 63.Nelson H, Mandiyan S, Nelson N. Cloning of the human brain GABA transporter. FEBS Lett 1990; 269:181–4 [DOI] [PubMed] [Google Scholar]
  • 64.Frahm C, Engel D, Draguhn A. Efficacy of background GABA uptake in rat hippocampal slices. Neuroreport 2001; 12:1593–6 [DOI] [PubMed] [Google Scholar]
  • 65.Nusser Z, Mody I. Selective modulation of tonic and phasic inhibitions in dentate gyrus granule cells. J Neurophysiol 2002; 87:2624–8 [DOI] [PubMed] [Google Scholar]
  • 66.Wetherington JP, Lambert NA. Differential desensitization of responses mediated by presynaptic and postsynaptic A1 adenosine receptors. J Neurosci 2002; 22:1248–55 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Gong X, Shao Y, Li B, Chen L, Wang C, CY. γ-aminobutyric acid transporter-1 is involved in anxiety-like behaviors and cognitive function in knockout mice. Exp Ther Med 2015; 10:653–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Liu G-X, Cai G-Q, Cai Y-Q, Sheng Z-J, Jiang J, Mei Z, Wang Z-G, Guo L, Fei J. Reduced anxiety and depression-like behaviors in mice lacking GABA transporter subtype 1. Neuropsychopharmacology 2007; 32:1531. [DOI] [PubMed] [Google Scholar]
  • 69.Liu GX, Liu S, Cai GQ, Sheng ZJ, Cai YQ, Jiang J, Sun X, Ma SK, Wang L, Wang ZG. Reduced aggression in mice lacking GABA transporter subtype 1. J Neurosci Res 2007; 85:649–55 [DOI] [PubMed] [Google Scholar]
  • 70.Xu YF, Cai YQ, Cai GQ, Jiang J, Sheng ZJ, Wang ZG, Fei J. Hypoalgesia in mice lacking GABA transporter subtype 1. J Neurosci Res 2008; 86:465–70 [DOI] [PubMed] [Google Scholar]
  • 71.Gong N, Li Y, Cai G-Q, Niu R-F, Fang Q, Wu K, Chen Z, Lin L-N, Xu L, Fei J. GABA transporter-1 activity modulates hippocampal theta oscillation and theta burst stimulation-induced long-term potentiation. J Neurosci 2009; 29:15836–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Borden LA, Dhar TM, Smith KE, Weinshank RL, Branchek TA, Gluchowski C. Tiagabine, SK&F 89976-A, CI-966, and NNC-711 are selective for the cloned GABA transporter GAT-1. Eur J Pharmacol 1994; 269:219–24 [DOI] [PubMed] [Google Scholar]
  • 73.Sałat K, Podkowa A, Malikowska N, Kern F, Pabel J, Wojcieszak E, Kulig K, Wanner KT, Strach B, Wyska E. Novel, highly potent and in vivo active inhibitor of GABA transporter subtype 1 with anticonvulsant, anxiolytic, antidepressant and antinociceptive properties. Neuropharmacology 2017; 113:331–42 [DOI] [PubMed] [Google Scholar]
  • 74.Sałat K, Podkowa A, Kowalczyk P, Kulig K, Dziubina A, Filipek B, Librowski T. Anticonvulsant active inhibitor of GABA transporter subtype 1, tiagabine, with activity in mouse models of anxiety, pain and depression. Pharmacol Rep 2015; 67:465–72 [DOI] [PubMed] [Google Scholar]
  • 75.Sałat K, Podkowa A, Mogilski S, Zaręba P, Kulig K, Sałat R, Malikowska N, Filipek B. The effect of GABA transporter 1 (GAT1) inhibitor, tiagabine, on scopolamine-induced memory impairments in mice. Pharmacol Rep 2015; 67:1155–62 [DOI] [PubMed] [Google Scholar]
  • 76.Cope DW, Di Giovanni G, Fyson SJ, Orbán G, Errington AC, Lőrincz ML, Gould TM, Carter DA, Crunelli V. Enhanced tonic GABAA inhibition in typical absence epilepsy. Nat Med 2009; 15:1392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Kaufman KR, Kugler SL, Sachdeo RC. Tiagabine in the management of postencephalitic epilepsy and impulse control disorder. Epilepsy Behav 2002; 3:190–4 [DOI] [PubMed] [Google Scholar]
  • 78.Hoffman DA. Tiagabine for rage, aggression, and anxiety. J Neuropsychiatry Clin Neurosci 2005; 17:252. [DOI] [PubMed] [Google Scholar]
  • 79.Rosenthal M. Tiagabine for the treatment of generalized anxiety disorder: a randomized, open-label, clinical trial with paroxetine as a positive control. J Clin Psychiatry 2003; 64:1245–49 [DOI] [PubMed] [Google Scholar]
  • 80.Pollack MH, Roy-Byrne PP, Van Ameringen M, Snyder H, Brown C, Ondrasik J, Rickels K. The selective GABA reuptake inhibitor tiagabine for the treatment of generalized anxiety disorder: results of a placebo-controlled study. J Clin Psychiatry 2005; 66:1401–8 [DOI] [PubMed] [Google Scholar]
  • 81.Gonzalez G, Sevarino K, Sofuoglu M, Poling J, Oliveto A, Gonsai K, George TP, Kosten TR. Tiagabine increases cocaine‐free urines in cocaine‐dependent methadone‐treated patients: results of a randomized pilot study. Addiction 2003; 98:1625–32 [DOI] [PubMed] [Google Scholar]
  • 82.Novak V, Kanard R, Kissel JT, Mendell JR. Treatment of painful sensory neuropathy with tiagabine: a pilot study. Clin Autonomic Res 2001; 11:357–61 [DOI] [PubMed] [Google Scholar]
  • 83.Bagley EE, Gerke MB, Vaughan CW, Hack SP, Christie MJ. GABA transporter currents activated by protein kinase a excite midbrain neurons during opioid withdrawal. Neuron 2005; 45:433–45 [DOI] [PubMed] [Google Scholar]
  • 84.Bagley EE, Hacker J, Chefer VI, Mallet C, McNally GP, Chieng BC, Perroud J, Shippenberg TS, Christie MJ. Drug-induced GABA transporter currents enhance GABA release to induce opioid withdrawal behaviors. Nat Neurosci 2011; 14:1548–54 [DOI] [PubMed] [Google Scholar]
  • 85.Wu Y, Wang W, Díez-Sampedro A, Richerson GB. Nonvesicular inhibitory neurotransmission via reversal of the GABA transporter GAT-1. Neuron 2007; 56:851–65 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Wu Y, Wang W, Richerson GB. GABA transaminase inhibition induces spontaneous and enhances depolarization-evoked GABA efflux via reversal of the GABA transporter. J Neurosci 2001; 21:2630–39 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Medina-Ceja L, Sandoval-García F, Morales-Villagrán A, López-Pérez SJ. Rapid compensatory changes in the expression of EAAT-3 and GAT-1 transporters during seizures in cells of the CA1 and dentate gyrus. J Biomed Sci 2012; 19:78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Sperk G, Schwarzer C, Heilman J, Furtinger S, Reimer RJ, Edwards RH, Nelson N. Expression of plasma membrane GABA transporters but not of the vesicular GABA transporter in dentate granule cells after kainic acid seizures. Hippocampus 2003; 13:806–15 [DOI] [PubMed] [Google Scholar]
  • 89.Honmou O, Kocsis JD, Richerson GB. Gabapentin potentiates the conductance increase induced by nipecotic acid in CA1 pyramidal neurons in vitro. Epilepsy Res 1995; 20:193–202 [DOI] [PubMed] [Google Scholar]
  • 90.Wu Y, Wang W, Richerson GB. Vigabatrin induces tonic inhibition via GABA transporter reversal without increasing vesicular GABA release. J Neurophysiol 2003; 89:2021–34 [DOI] [PubMed] [Google Scholar]
  • 91.Schmidt D, Gram L, Brodie M, Krämer G, Perucca E, Kälviäinen R, Elger C. Tiagabine in the treatment of epilepsy – a clinical review with a guide for the prescribing physician. Epilepsy Res 2000; 41:245–51 [DOI] [PubMed] [Google Scholar]
  • 92.Dalby NO, Nielsen EB. Comparison of the preclinical anticonvulsant profiles of tiagabine, lamotrigine, gabapentin and vigabatrin. Epilepsy Res 1997; 28:63–72 [DOI] [PubMed] [Google Scholar]
  • 93.Dalby NO. GABA-level increasing and anticonvulsant effects of three different GABA uptake inhibitors. Neuropharmacology 2000; 39:2399–407 [DOI] [PubMed] [Google Scholar]
  • 94.Ikegaki N, Saito N, Hashima M, Tanaka C. Production of specific antibodies against GABA transporter subtypes (GAT1, GAT2, GAT3) and their application to immunocytochemistry. Brain Res Mol Brain Res 1994; 26:47–54 [DOI] [PubMed] [Google Scholar]
  • 95.Liu QR, Lopez-Corcuera B, Nelson H, Mandiyan S, Nelson N. Cloning and expression of a cDNA encoding the transporter of taurine and beta-alanine in mouse brain. Proc Natl Acad Sci U S A 1992; 89:12145–9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Takanaga H, Ohtsuki S, Hosoya K-I, Terasaki T. GAT2/BGT-1 as a system responsible for the transport of γ-aminobutyric acid at the mouse blood–brain barrier. J Cereb Blood Flow Metab 2001; 21:1232–39 [DOI] [PubMed] [Google Scholar]
  • 97.Lehre A, Rowley N, Zhou Y, Holmseth S, Guo C, Holen T, Hua R, Laake P, Olofsson A, Poblete-Naredo I. Deletion of the betaine–GABA transporter (BGT1; slc6a12) gene does not affect seizure thresholds of adult mice. Epilepsy Res 2011; 95:70–81 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Smith MD, Saunders GW, Clausen RP, Frolund B, Krogsgaard-Larsen P, Larsson OM, Schousboe A, Wilcox KS, White HS. Inhibition of the betaine-GABA transporter (mGAT2/BGT-1) modulates spontaneous electrographic bursting in the medial entorhinal cortex (mEC). Epilepsy Res 2008; 79:6–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Kwakowsky A, C-F, Guzman B, Govindpani K, Waldvogel HJ, Faull RL. Gamma-aminobutyric acid a receptors in Alzheimer's disease: highly localized remodeling of a complex and diverse signaling pathway. Neural Regen Res 2018; 13:1362–63 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Govindpani K, Guzman C-, Vinnakota FB, Waldvogel C, Faull HJ, Kwakowsky RL. A. A. Towards a better understanding of GABAergic remodeling in Alzheimer's disease. Int J Mol Sci 2017; 18:1813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Fuhrer TE, Palpagama TH, Waldvogel HJ, Synek BJL, Turner C, Faull RL, Kwakowsky A. Impaired expression of GABA transporters in the human Alzheimer's disease hippocampus, subiculum, entorhinal cortex and superior temporal gyrus. Neuroscience 2017; 351:108–18 [DOI] [PubMed] [Google Scholar]
  • 102.Kersanté F, Rowley SC, Pavlov I, Gutièrrez‐Mecinas M, Semyanov A, Reul JM, Walker MC, Linthorst AC. A functional role for both γ‐aminobutyric acid (GABA) transporter‐1 and GABA transporter‐3 in the modulation of extracellular GABA and GABAergic tonic conductances in the rat hippocampus. J Physiol 2013; 591:2429–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Albrecht A, Ivens S, Papageorgiou IE, Çalışkan G, Saiepour N, Brück W, Richter ‐Levin G, Heinemann U, Stork O. Shifts in excitatory/inhibitory balance by juvenile stress: a role for neuron–astrocyte interaction in the dentate gyrus. Glia 2016; 64:911–22 [DOI] [PubMed] [Google Scholar]
  • 104.Cohen-Kfir E, Lee W, Eskandari S, Nelson N. Zinc inhibition of γ-aminobutyric acid transporter 4 (GAT4) reveals a link between excitatory and inhibitory neurotransmission. Proc Natl Acad Sci U S A 2005; 102:6154–59 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Assaf S, Chung S-H. Release of endogenous Zn2+ from brain tissue during activity. Nature 1984; 308:734. [DOI] [PubMed] [Google Scholar]
  • 106.Howell GA, Welch MG, Frederickson CJ. Stimulation-induced uptake and release of zinc in hippocampal slices. Nature 1984; 308:736. [DOI] [PubMed] [Google Scholar]
  • 107.Kinney GA. GAT-3 transporters regulate inhibition in the neocortex. J Neurophysiol 2005; 94:4533–37 [DOI] [PubMed] [Google Scholar]
  • 108.Hirao T, Morimoto K, Yamamoto Y, Watanabe T, Sato H, Sato K, Sato S, Yamada N, Tanaka K, Suwaki H. Time-dependent and regional expression of GABA transporter mRNAs following amygdala-kindled seizures in rats. Brain Res Mol Brain Res 1998; 54:49–55 [DOI] [PubMed] [Google Scholar]
  • 109.Augier E, Barbier E, Dulman RS, Licheri V, Augier G, Domi E, Barchiesi R, Farris S, Nätt D, Mayfield RD. A molecular mechanism for choosing alcohol over an alternative reward. Science 2018; 360:1321–26 [DOI] [PubMed] [Google Scholar]
  • 110.Fonnum F. Glutamate: a neurotransmitter in mammalian brain. J Neurochem 1984; 42:1–11 [DOI] [PubMed] [Google Scholar]
  • 111.Ottersen OP, Storm-Mathisen J. Glutamate‐and GABA‐containing neurons in the mouse and rat brain, as demonstrated with a new immunocytochemical technique. J Comp Neurol 1984; 229:374–92 [DOI] [PubMed] [Google Scholar]
  • 112.McGeer PL, Eccles JC, McGeer EG. Putative excitatory neurons: glutamate and aspartate. Molecular neurobiology of the mammalian brain. Berlin: Springer, 1978, pp. 183–98 [Google Scholar]
  • 113.Weil-Malherbe H. Significance of glutamic acid for the metabolism of nervous tissue. Physiol Rev 1950; 30:549–68 [DOI] [PubMed] [Google Scholar]
  • 114.Roberts E, Frankel S. γ Aminobutyric acid in brain: its formation from glutamic acid. J Biol Chem 1950; 187:55–63 [PubMed] [Google Scholar]
  • 115.Meister A. Biochemistry of glutamate: glutamine and glutathione. Glutamic Acid 1979;69–84 [Google Scholar]
  • 116.Hayashi T. A physiological study of epileptic seizures following cortical stimulation in animals and its application to human clinics. Jpn J Physiol 1952; 3:46–64 [DOI] [PubMed] [Google Scholar]
  • 117.Okamoto S. Epileptogenic action of glutamate directly applied into the brains of animals and inhibitory effects of protein and tissue emulsions on its action. J Physiol Soc Jpn 1951; 13:555–62 [Google Scholar]
  • 118.Hayashi T. Effects of sodium glutamate on the nervous system. Keio j Med 1954; 3:183–92 [Google Scholar]
  • 119.Watkins J, Evans R. Excitatory amino acid transmitters. Annu Rev Pharmacol Toxicol 1981; 21:165–204 [DOI] [PubMed] [Google Scholar]
  • 120.Braitenberg V, Schüz A. Cortex: statistics and geometry of neuronal connectivity. Berlin: Springer Science & Business Media, 2013 [Google Scholar]
  • 121.Hassel B, Dingledine R. Glutamate and glutamate receptors. Basic neurochemistry. 8th ed. Amsterdam: Elsevier, 2012, pp. 342–66 [Google Scholar]
  • 122.Choi DW. Excitotoxic cell death. J Neurobiol 1992; 23:1261–76 [DOI] [PubMed] [Google Scholar]
  • 123.Olney JW. Brain lesions, obesity, and other disturbances in mice treated with monosodium glutamate. Science 1969; 164:719–21 [DOI] [PubMed] [Google Scholar]
  • 124.Nadler JV, Perry BW, Cotman CW. Intraventricular kainic acid preferentially destroys hippocampal pyramidal cells. Nature 1978; 271:676–7 [DOI] [PubMed] [Google Scholar]
  • 125.McKenna MC. The glutamate-glutamine cycle is not stoichiometric: fates of glutamate in brain. J Neurosci Res 2007; 85:3347–58 [DOI] [PubMed] [Google Scholar]
  • 126.Peng L, Hertz L, Huang R, Sonnewald U, Petersen SB, Westergaard N, Larsson O, Schousboe A. Utilization of glutamine and of TCA cycle constituents as precursors for transmitter glutamate and GABA. Dev Neurosci 1993; 15:367–77 [DOI] [PubMed] [Google Scholar]
  • 127.Omote H, Miyaji T, Juge N, Moriyama Y. Vesicular neurotransmitter transporter: bioenergetics and regulation of glutamate transport. Biochemistry 2011; 50:5558–65 [DOI] [PubMed] [Google Scholar]
  • 128.Burger PM, Mehl E, Cameron PL, Maycox PR, Baumert M, Lottspeich F, De Camilli P, Jahn R. Synaptic vesicles immunoisolated from rat cerebral cortex contain high levels of glutamate. Neuron 1989; 3:715–20 [DOI] [PubMed] [Google Scholar]
  • 129.Riveros N, Fiedler J, Lagos N, Mun C, Orrego F. Glutamate in rat brain cortex synaptic vesicles: influence of the vesicle isolation procedure. Brain Res 1986; 386:405–8 [DOI] [PubMed] [Google Scholar]
  • 130.Naito S, Ueda T. Characterization of glutamate uptake into synaptic vesicles. J Neurochem 1985; 44:99–109 [DOI] [PubMed] [Google Scholar]
  • 131.Nicholls D, Attwell D. The release and uptake of excitatory amino acids. Trends Pharmacol Sci 1990; 11:462–8 [DOI] [PubMed] [Google Scholar]
  • 132.Szatkowski M, Barbour B, Attwell D. Non-vesicular release of glutamate from glial cells by reversed electrogenic glutamate uptake. Nature 1990; 348:443–6 [DOI] [PubMed] [Google Scholar]
  • 133.Longuemare M, Swanson RA. Excitatory amino acid release from astrocytes during energy failure by reversal of sodium‐dependent uptake. J Neurosci Res 1995; 40:379–86 [DOI] [PubMed] [Google Scholar]
  • 134.Kimelberg H, Goderie S, Higman S, Pang S, Waniewski R. Swelling-induced release of glutamate, aspartate, and taurine from astrocyte cultures. J Neurosci 1990; 10:1583–91 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Mongin AA, Orlov SN. Mechanisms of cell volume regulation and possible nature of the cell volume sensor. Pathophysiology 2001; 8:77–88 [DOI] [PubMed] [Google Scholar]
  • 136.Ye Z-C, Wyeth MS, Baltan-Tekkok S, Ransom BR. Functional hemichannels in astrocytes: a novel mechanism of glutamate release. J Neurosci 2003; 23:3588–96 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Quist AP, Rhee SK, Lin H, Lal R. Physiological role of gap-junctional hemichannels: extracellular calcium-dependent isosmotic volume regulation. J Cell Biol 2000; 148:1063–74 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Bruzzone S, Guida L, Zocchi E, Franco L, De FA. Connexin 43 hemi channels mediate Ca2+-regulated transmembrane NAD+ fluxes in intact cells. FASEB J 2001; 15:10–2 [DOI] [PubMed] [Google Scholar]
  • 139.Kamermans M, Fahrenfort I, Schultz K, Janssen-Bienhold U, Sjoerdsma T, Weiler R. Hemichannel-mediated inhibition in the outer retina. Science 2001; 292:1178–80 [DOI] [PubMed] [Google Scholar]
  • 140.Plotkin L, Bellido T. Bisphosphonate-induced, hemichannel-mediated, anti-apoptosis through the SRC/ERK pathway: a gap junction-independent action of connexin43. Cell Commun Adhes 2001; 8:377–82 [DOI] [PubMed] [Google Scholar]
  • 141.John SA, Kondo R, Wang S-Y, Goldhaber JI, Weiss JN. Connexin-43 hemichannels opened by metabolic inhibition. J Biol Chem 1999; 274:236–40 [DOI] [PubMed] [Google Scholar]
  • 142.Contreras JE, Sánchez HA, Eugenín EA, Speidel D, Theis M, Willecke K, Bukauskas FF, Bennett MV, Sáez JC. Metabolic inhibition induces opening of unapposed connexin 43 gap junction hemichannels and reduces gap junctional communication in cortical astrocytes in culture. Proc Natl Acad Sci U S A 2002; 99:495–500 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Duan S, Anderson CM, Keung EC, Chen Y, Chen Y, Swanson RA. P2X7 receptor-mediated release of excitatory amino acids from astrocytes. J Neurosci 2003; 23:1320–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Zhang Y, Chen K, Sloan SA, Bennett ML, Scholze AR, O'Keeffe S, Phatnani HP, Guarnieri P, Caneda C, Ruderisch N. An RNA-sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex. J Neurosci 2014; 34:11929–47 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Cahoy JD, Emery B, Kaushal A, Foo LC, Zamanian JL, Christopherson KS, Xing Y, Lubischer JL, Krieg PA, Krupenko SA. A transcriptome database for astrocytes, neurons, and oligodendrocytes: a new resource for understanding brain development and function. J Neurosci 2008; 28:264–78 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Barres BA. The mystery and magic of glia: a perspective on their roles in health and disease. Neuron 2008; 60:430–40 [DOI] [PubMed] [Google Scholar]
  • 147.Li D, Hérault K, Silm K, Evrard A, Wojcik S, Oheim M, Herzog E, Ropert N. Lack of evidence for vesicular glutamate transporter expression in mouse astrocytes. J Neurosci 2013; 33:4434–55 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Schwarz Y, Zhao N, Kirchhoff F, Bruns D. Astrocytes control synaptic strength by two distinct v-SNARE-dependent release pathways. Nat Neurosci 2017; 20:1529. [DOI] [PubMed] [Google Scholar]
  • 149.Bezzi P, Gundersen V, Galbete JL, Seifert G, Steinhäuser C, Pilati E, Volterra A. Astrocytes contain a vesicular compartment that is competent for regulated exocytosis of glutamate. Nat Neurosci 2004; 7:613. [DOI] [PubMed] [Google Scholar]
  • 150.Parpura V, Fang Y, Basarsky T, Jahn R, Haydon PG. Expression of synaptobrevin II, cellubrevin and syntaxin but not SNAP-25 in cultured astrocytes. FEBS Lett 1995; 377:489–92 [DOI] [PubMed] [Google Scholar]
  • 151.Parpura V, Basarsky TA, Liu F, Jeftinija K, Jeftinija S, Haydon PG. Glutamate-mediated astrocyte–neuron signalling. Nature 1994; 369:744. [DOI] [PubMed] [Google Scholar]
  • 152.Bezzi P, Carmignoto G, Pasti L, Vesce S, Rossi D, Rizzini BL, Pozzan T, Volterra A. Prostaglandins stimulate calcium-dependent glutamate release in astrocytes. Nature 1998; 391:281. [DOI] [PubMed] [Google Scholar]
  • 153.Sun W, McConnell E, Pare J-F, Xu Q, Chen M, Peng W, Lovatt D, Han X, Smith Y, Nedergaard M. Glutamate-dependent neuroglial calcium signaling differs between young and adult brain. Science 2013; 339:197–200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Savtchouk I, Volterra A. Gliotransmission: beyond black-and-white. J Neurosci 2018; 38:14–25 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Fiacco TA, McCarthy KD. Multiple lines of evidence indicate that gliotransmission does not occur under physiological conditions. J Neurosci 2018; 38:3–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Bannai S, Kitamura E. Transport interaction of L-cystine and L-glutamate in human diploid fibroblasts in culture. J Biol Chem 1980; 255:2372–6 [PubMed] [Google Scholar]
  • 157.Traynelis SF, Wollmuth LP, McBain CJ, Menniti FS, Vance KM, Ogden KK, Hansen KB, Yuan H, Myers SJ, Dingledine R. Glutamate receptor ion channels: structure, regulation, and function. Pharmacol Rev 2010; 62:405–96 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Niswender CM, Conn PJ. Metabotropic glutamate receptors: physiology, pharmacology, and disease. Ann Rev Pharmacol Toxicol 2010; 50:295–322 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Lodge D. The history of the pharmacology and cloning of ionotropic glutamate receptors and the development of idiosyncratic nomenclature. Neuropharmacology 2009; 56:6–21 [DOI] [PubMed] [Google Scholar]
  • 160.Lomeli H, Sprengel R, Laurie DJ, Köhr G, Herb A, Seeburg PH, Wisden W. The rat Delta‐1 and Delta‐2 subunits extend the excitatory amino acid receptor family. FEBS Lett 1993; 315:318–22 [DOI] [PubMed] [Google Scholar]
  • 161.Yamazaki M, Araki K, Shibata A, Mishina M. Molecular cloning of a cDNA encoding a novel member of the mouse glutamate receptor channel family. Biochem Biophys Res Commun 1992; 183:886–92 [DOI] [PubMed] [Google Scholar]
  • 162.Naur P, Hansen KB, Kristensen AS, Dravid SM, Pickering DS, Olsen L, Vestergaard B, Egebjerg J, Gajhede M, Traynelis SF. Ionotropic glutamate-like receptor δ2 binds D-serine and glycine. Proc Natl Acad Sci U S A 2007; 104:14116–21 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Craig AM, Blackstone CD, Huganir RL, Banker G. The distribution of glutamate receptors in cultured rat hippocampal neurons: postsynaptic clustering of AMPA selective subunits. Neuron 1993; 10:1055–68 [DOI] [PubMed] [Google Scholar]
  • 164.Jones MV, Westbrook GL. The impact of receptor desensitization on fast synaptic transmission. Trends in Neurosci 1996; 19:96–101 [DOI] [PubMed] [Google Scholar]
  • 165.Greger IH, Khatri L, Kong X, Ziff EB. AMPA receptor tetramerization is mediated by Q/R editing. Neuron 2003; 40:763–74 [DOI] [PubMed] [Google Scholar]
  • 166.Yang Y, Wang X-B, Zhou Q. Perisynaptic GluR2-lacking AMPA receptors control the reversibility of synaptic and spines modifications. Proc Natl Acad Sci U S A 2010; 107:11999–2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Plant K, Pelkey KA, Bortolotto ZA, Morita D, Terashima A, McBain CJ, Collingridge GL, Isaac JT. Transient incorporation of native GluR2-lacking AMPA receptors during hippocampal long-term potentiation. Nat Neurosci 2006; 9:602. [DOI] [PubMed] [Google Scholar]
  • 168.Kasai H, Matsuzaki M, Noguchi J, Yasumatsu N, Nakahara H. Structure–stability–function relationships of dendritic spines. Trends Neurosci 2003; 26:360–8 [DOI] [PubMed] [Google Scholar]
  • 169.Matsuzaki M, Ellis-Davies GC, Nemoto T, Miyashita Y, Iino M, Kasai H. Dendritic spine geometry is critical for AMPA receptor expression in hippocampal CA1 pyramidal neurons. Nat Neurosci 2001; 4:1086–92 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Lüscher C, Xia H, Beattie EC, Carroll RC, von Zastrow M, Malenka RC, Nicoll RA. Role of AMPA receptor cycling in synaptic transmission and plasticity. Neuron 1999; 24:649–58 [DOI] [PubMed] [Google Scholar]
  • 171.Porter RH, Eastwood SL, Harrison PJ. Distribution of kainate receptor subunit mRNAs in human hippocampus, neocortex and cerebellum, and bilateral reduction of hippocampal GluR6 and KA2 transcripts in schizophrenia. Brain Res 1997; 751:217–31 [DOI] [PubMed] [Google Scholar]
  • 172.Bowie D. External anions and cations distinguish between AMPA and kainate receptor gating mechanisms. J Physiol 2002; 539:725–33 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Paternain AV, Cohen A, Stern-Bach Y, Lerma J. A role for extracellular Na+ in the channel gating of native and recombinant kainate receptors. J Neurosci 2003; 23:8641–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Sommer B, Köhler M, Sprengel R, Seeburg PH. RNA editing in brain controls a determinant of ion flow in glutamate-gated channels. Cell 1991; 67:11–9 [DOI] [PubMed] [Google Scholar]
  • 175.Köhler M, Burnashev N, Sakmann B, Seeburg PH. Determinants of Ca2+ permeability in both T. M1 and TM2 of high affinity kainate receptor channels: diversity by RNA editing. Neuron 1993; 10:491–500 [DOI] [PubMed] [Google Scholar]
  • 176.Castillo PE, Malenka RC, Nicoll RA. Kainate receptors mediate a slow postsynaptic current in hippocampal CA3 neurons. Nature 1997; 388:182. [DOI] [PubMed] [Google Scholar]
  • 177.Frerking M, Malenka R, Nicoll R. Synaptic activation of kainate receptors on hippocampal interneurons. Nat Neurosci 1998; 1:479. [DOI] [PubMed] [Google Scholar]
  • 178.Frerking M, Schmitz D, Zhou Q, Johansen J, Nicoll RA. Kainate receptors depress excitatory synaptic transmission at CA3→ CA1 synapses in the hippocampus via a direct presynaptic action. J Neurosci 2001; 21:2958–66 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Schmitz D, Frerking M, Nicoll RA. Synaptic activation of presynaptic kainate receptors on hippocampal mossy fiber synapses. Neuron 2000; 27:327–38 [DOI] [PubMed] [Google Scholar]
  • 180.Chittajallu R, Vignes M, Dev KK, Barnes JM, Collingridge GL, Henley JM. Regulation of glutamate release by presynaptic kainate receptors in the hippocampus. Nature 1996; 379:78. [DOI] [PubMed] [Google Scholar]
  • 181.Rodríguez-Moreno A, Lerma J. Kainate receptor modulation of GABA release involves a metabotropic function. Neuron 1998; 20:1211–8 [DOI] [PubMed] [Google Scholar]
  • 182.Sanz-Clemente A, Nicoll RA, Roche KW. Diversity in NMDA receptor composition: many regulators, many consequences. Neuroscientist 2013; 19:62–75 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Henson MA, Roberts AC, Perez-Otano I, Philpot BD. Influence of the NR3A subunit on NMDA receptor functions. Prog Neurobiol 2010; 91:23–37 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Glasgow NG, Siegler Retchless B, Johnson JW. Molecular bases of NMDA receptor subtype-dependent properties. J Physiol 2015; 593:83–95 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Mayer ML, Westbrook GL, Guthrie PB. Voltage-dependent block by Mg2+ of NMDA responses in spinal cord neurones. Nature 1984; 309:261. [DOI] [PubMed] [Google Scholar]
  • 186.Snyder SH, Kim PD, Amino A, As PN, F, On D. Neurochem Res 2000; 25:553–60 [DOI] [PubMed] [Google Scholar]
  • 187.Hunt DL, Castillo PE. Synaptic plasticity of NMDA receptors: mechanisms and functional implications. Curr Opin Neurobiol 2012; 22:496–508 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Pérez-Otaño I, Ehlers MD. Homeostatic plasticity and NMDA receptor trafficking. Trends Neurosci 2005; 28:229–38 [DOI] [PubMed] [Google Scholar]
  • 189.Pin J-P, Acher F. The metabotropic glutamate receptors: structure, activation mechanism and pharmacology. Curr Drug Targets CNS Neurol Disord 2002; 1:297–317 [DOI] [PubMed] [Google Scholar]
  • 190.Anwyl R. Metabotropic glutamate receptors: electrophysiological properties and role in plasticity. Brain Res Brain Res Rev 1999; 29:83–120 [DOI] [PubMed] [Google Scholar]
  • 191.Danbolt N, Furness D, Zhou Y. Neuronal vs glial glutamate uptake: resolving the conundrum. Neurochem Int 2016; 98:29–45 [DOI] [PubMed] [Google Scholar]
  • 192.Hediger MA, Clémençon B, Burrier RE, Bruford EA. The ABCs of membrane transporters in health and disease (SLC series): introduction. Mol Aspects Med 2013; 34:95–107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Kanner BI, Schuldiner S. Mechanism of transport and storage of neurotransmitter. CRC Crit Rev Biochem 1987; 22:1–38 [DOI] [PubMed] [Google Scholar]
  • 194.Barbour B, Brew H, Attwell D. Electrogenic uptake of glutamate and aspartate into glial cells isolated from the salamander (ambystoma) retina. J Physiol 1991; 436:169–93 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Zerangue N, Kavanaugh MP. Flux coupling in a neuronal glutamate transporter. Nature 1996; 383:634. [DOI] [PubMed] [Google Scholar]
  • 196.Levy LM, Warr O, Attwell D. Stoichiometry of the glial glutamate transporter GLT-1 expressed inducibly in a Chinese hamster ovary cell line selected for low endogenous Na+-dependent glutamate uptake. J Neurosci 1998; 18:9620–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Barbour B, Brew H, Attwell D. Electrogenic glutamate uptake in glial cells is activated by intracellular potassium. Nature 1988; 335:433–5 [DOI] [PubMed] [Google Scholar]
  • 198.Bergles DE, Tzingounis AV, Jahr CE. Comparison of coupled and uncoupled currents during glutamate uptake by GLT-1 transporters. J Neurosci 2002; 22:10153–62 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Wadiche JI, Arriza JL, Amara SG, Kavanaugh MP. Kinetics of a human glutamate transporter. Neuron 1995; 14:1019–27 [DOI] [PubMed] [Google Scholar]
  • 200.Jen JC, Wan J, Palos TP, Howard BD, Baloh RW. Mutation in the glutamate transporter EAAT1 causes episodic ataxia, hemiplegia, and seizures. Neurology 2005; 65:529–34 [DOI] [PubMed] [Google Scholar]
  • 201.Storck T, Schulte S, Hofmann K, Stoffel W. Structure, expression, and functional analysis of a Na (+)-dependent glutamate/aspartate transporter from rat brain. Proc Natl Acad Sci U S A 1992; 89:10955–9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Tanaka K. Cloning and expression of a glutamate transporter from mouse brain. Neurosci Lett 1993; 159:183–6 [DOI] [PubMed] [Google Scholar]
  • 203.Lehre KP, Levy LM, Ottersen OP, Storm-Mathisen J, Danbolt NC. Differential expression of two glial glutamate transporters in the rat brain: quantitative and immunocytochemical observations. J Neurosci 1995; 15:1835–53 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Danbolt NC, Pines G, Kanner BI. Purification and reconstitution of the sodium-and potassium-coupled glutamate transport glycoprotein from rat brain. Biochemistry 1990; 29:6734–40 [DOI] [PubMed] [Google Scholar]
  • 205.Zhou Y, Hassel B, Eid T, Danbolt NC. Axon-terminals expressing EAAT2 (GLT-1; Slc1a2) are common in the forebrain and not limited to the hippocampus. Neurochem Int 2019; 123:101--113. [DOI] [PubMed] [Google Scholar]
  • 206.Chen W, Mahadomrongkul V, Berger UV, Bassan M, DeSilva T, Tanaka K, Irwin N, Aoki C, Rosenberg PA. The glutamate transporter GLT1a is expressed in excitatory axon terminals of mature hippocampal neurons. J Neurosci 2004; 24:1136–48 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Furness D, Dehnes Y, Akhtar A, Rossi D, Hamann M, Grutle N, Gundersen V, Holmseth S, Lehre K, Ullensvang K. A quantitative assessment of glutamate uptake into hippocampal synaptic terminals and astrocytes: new insights into a neuronal role for excitatory amino acid transporter 2 (EAAT2). Neuroscience 2008; 157:80–94 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Melone M, Bellesi M, Conti F. Synaptic localization of GLT-1a in the rat somatic sensory cortex. Glia 2009; 57:108–17 [DOI] [PubMed] [Google Scholar]
  • 209.Kanai Y, Hediger MA. Primary structure and functional characterization of a high-affinity glutamate transporter. Nature 1992; 360:467. [DOI] [PubMed] [Google Scholar]
  • 210.Bjørjås M, Gjesdal O, Erickson J, Torp R, Levy L, Ottersen O, Degree M, Storm-Mathisen J, Seeberg E, Danbolt N. Cloning and expression of a neuronal rat brain glutamate transporter. Brain Res Mol Brain Res 1996; 36:163–8 [DOI] [PubMed] [Google Scholar]
  • 211.Rothstein JD, Martin L, Levey AI, Dykes-Hoberg M, Jin L, Wu D, Nash N, Kuncl RW. Localization of neuronal and glial glutamate transporters. Neuron 1994; 13:713–25 [DOI] [PubMed] [Google Scholar]
  • 212.Holmseth S, Dehnes Y, Huang YH, Follin-Arbelet VV, Grutle NJ, Mylonakou MN, Plachez C, Zhou Y, Furness DN, Bergles DE. The density of EAAC1 (EAAT3) glutamate transporters expressed by neurons in the mammalian CNS. J Neurosci 2012; 32:6000–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Fairman W, Vandenberg R, Arriza J, Kavanaught M, Amara S. An excitatory amino-acid transporter with properties of a ligand-gated chloride channel. Nature 1995; 375:599. [DOI] [PubMed] [Google Scholar]
  • 214.Itoh M, Watanabe Y, Watanabe M, Tanaka K, Wada K, Takashima S. Expression of a glutamate transporter subtype, EAAT4, in the developing human cerebellum. Brain Res 1997; 767:265–71 [DOI] [PubMed] [Google Scholar]
  • 215.Hu WH, Walters WM, Xia XM, Karmally SA, Bethea JR. Neuronal glutamate transporter EAAT4 is expressed in astrocytes. Glia 2003; 44:13–25 [DOI] [PubMed] [Google Scholar]
  • 216.Arriza JL, Eliasof S, Kavanaugh MP, Amara SG. Excitatory amino acid transporter 5, a retinal glutamate transporter coupled to a chloride conductance. Proc Natl Acad Sci U S A 1997; 94:4155–60 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Danbolt N, Storm-Mathisen J, Kanner B. An [Na++ K+] coupledl-glutamate transporter purified from rat brain is located in glial cell processes. Neuroscience 1992; 51:295–310 [DOI] [PubMed] [Google Scholar]
  • 218.Haugeto Ø, Ullensvang K, Levy LM, Chaudhry FA, Honoré T, Nielsen M, Lehre KP, Danbolt NC. Brain glutamate transporter proteins form homomultimers. J Biol Chem 1996; 271:27715–22 [DOI] [PubMed] [Google Scholar]
  • 219.Otis TS, Kavanaugh MP. Isolation of current components and partial reaction cycles in the glial glutamate transporter EAAT2. J Neurosci 2000; 20:2749–57 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.de Vries B, Mamsa H, Stam AH, Wan J, Bakker SL, Vanmolkot KR, Haan J, Terwindt GM, Boon EM, Howard BD, Frants RR, Baloh RW, Ferrari MD, Jen JC, van den Maagdenberg AM. Episodic ataxia associated with EAAT1 mutation C186S affecting glutamate reuptake. Arch Neurol 2009; 66:97–101 [DOI] [PubMed] [Google Scholar]
  • 221.Chivukula AS, Suslova M, Kortzak D, Kovermann P, Fahlke C. Functional consequences of SLC1A3 mutations associated with episodic ataxia 6. Hum Mutat 2020; doi: 10.1002/humu.24089. [DOI] [PubMed] [Google Scholar]
  • 222.Watanabe T, Morimoto K, Hirao T, Suwaki H, Watase K, Tanaka K. Amygdala-kindled and pentylenetetrazole-induced seizures in glutamate transporter GLAST-deficient mice. Brain Res 1999; 845:92–6 [DOI] [PubMed] [Google Scholar]
  • 223.Karlsson RM, Tanaka K, Heilig M, Holmes A. Loss of glial glutamate and aspartate transporter (excitatory amino acid transporter 1) causes locomotor hyperactivity and exaggerated responses to psychotomimetics: rescue by haloperidol and metabotropic glutamate 2/3 agonist. Biol Psychiatry 2008; 64:810–4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Karlsson RM, Tanaka K, Saksida LM, Bussey TJ, Heilig M, Holmes A. Assessment of glutamate transporter GLAST (EAAT1)-deficient mice for phenotypes relevant to the negative and executive/cognitive symptoms of schizophrenia. Neuropsychopharmacology 2009; 34:1578–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Tanaka K, Watase K, Manabe T, Yamada K, Watanabe M, Takahashi K, Iwama H, Nishikawa T, Ichihara N, Kikuchi T. Epilepsy and exacerbation of brain injury in mice lacking the glutamate transporter GLT-1. Science 1997; 276:1699–702 [DOI] [PubMed] [Google Scholar]
  • 226.Petr GT, Sun Y, Frederick NM, Zhou Y, Dhamne SC, Hameed MQ, Miranda C, Bedoya EA, Fischer KD, Armsen W. Conditional deletion of the glutamate transporter GLT-1 reveals that astrocytic GLT-1 protects against fatal epilepsy while neuronal GLT-1 contributes significantly to glutamate uptake into synaptosomes. J Neurosci 2015; 35:5187–201 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Kirschner MA, Copeland NG, Gilbert DJ, Jenkins NA, Amara SG. Mouse excitatory amino acid transporter EAAT2: isolation, characterization, and proximity to neuroexcitability loci on mouse chromosome 2. Genomics 1994; 24:218–24 [DOI] [PubMed] [Google Scholar]
  • 228.Cavus I, Kasoff WS, Cassaday MP, Jacob R, Gueorguieva R, Sherwin RS, Krystal JH, Spencer DD, Abi-Saab WM. Extracellular metabolites in the cortex and hippocampus of epileptic patients. Ann Neurol 2005; 57:226–35 [DOI] [PubMed] [Google Scholar]
  • 229.Sarac S, Afzal S, Broholm H, Madsen FF, Ploug T, Laursen H. Excitatory amino acid transporters EAAT-1 and EAAT-2 in temporal lobe and hippocampus in intractable temporal lobe epilepsy. APMIS 2009; 117:291–301 [DOI] [PubMed] [Google Scholar]
  • 230.Rakhade SN, Loeb JA. Focal reduction of neuronal glutamate transporters in human neocortical epilepsy. Epilepsia 2008; 49:226–36 [DOI] [PubMed] [Google Scholar]
  • 231.Hoogland G, van Oort RJ, Proper EA, Jansen GH, van Rijen PC, van Veelen CW, van Nieuwenhuizen O, Troost D, de Graan PN. Alternative splicing of glutamate transporter EAAT2 RNA in neocortex and hippocampus of temporal lobe epilepsy patients. Epilepsy Res 2004; 59:75–82 [DOI] [PubMed] [Google Scholar]
  • 232.Sha L, Chen T, Deng Y, Du T, Ma K, Zhu W, Shen Y, Xu Q. Hsp90 inhibitor HSP990 in very low dose upregulates EAAT2 and exerts potent antiepileptic activity. Theranostics 2020; 10:8415–29 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Kong Q, Takahashi K, Schulte D, Stouffer N, Lin Y, Lin CL. Increased glial glutamate transporter EAAT2 expression reduces epileptogenic processes following pilocarpine-induced status epilepticus. Neurobiol Dis 2012; 47:145–54 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Rao PS, Goodwani S, Bell RL, Wei Y, Boddu SH, Sari Y. Effects of ampicillin, cefazolin and cefoperazone treatments on GLT-1 expressions in the mesocorticolimbic system and ethanol intake in alcohol-preferring rats. Neuroscience 2015; 295:164–74 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Rao PS, Bell RL, Engleman EA, Sari Y. Targeting glutamate uptake to treat alcohol use disorders. Front Neurosci 2015; 9:144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Takizawa T, Ayata C, Chen SP. Therapeutic implications of cortical spreading depression models in migraine. Prog Brain Res 2020; 255:29–67 [DOI] [PubMed] [Google Scholar]
  • 237.Eikermann-Haerter K, Ayata C. Cortical spreading depression and migraine. Curr Neurol Neurosci Rep 2010; 10:167–73 [DOI] [PubMed] [Google Scholar]
  • 238.Lauritzen M. Cortical spreading depression in migraine. Cephalalgia 2001; 21:757–60 [DOI] [PubMed] [Google Scholar]
  • 239.Iijima T, Mies G, Hossmann KA. Repeated negative DC deflections in rat cortex following middle cerebral artery occlusion are abolished by MK-801: effect on volume of ischemic injury. J Cereb Blood Flow Metab 1992; 12:727–33 [DOI] [PubMed] [Google Scholar]
  • 240.Back T, Ginsberg MD, Dietrich WD, Watson BD. Induction of spreading depression in the ischemic hemisphere following experimental Middle cerebral artery occlusion: effect on infarct morphology. J Cereb Blood Flow Metab 1996; 16:202–13 [DOI] [PubMed] [Google Scholar]
  • 241.Aizawa H, Sun W, Sugiyama K, Itou Y, Aida T, Cui W, Toyoda S, Terai H, Yanagisawa M, Tanaka K. Glial glutamate transporter GLT-1 determines susceptibility to spreading depression in the mouse cerebral cortex. Glia 2020; 68:2631–42 [DOI] [PubMed] [Google Scholar]
  • 242.Scimemi A, Tian H, Diamond JS. Neuronal transporters regulate glutamate clearance, NMDA receptor activation, and synaptic plasticity in the hippocampus. J Neurosci 2009; 29:14581–95 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Diamond JS. Neuronal glutamate transporters limit activation of NMDA receptors by neurotransmitter spillover on CA1 pyramidal cells. J Neurosci 2001; 21:8328–38 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Petralia RS, Wang YX, Hua F, Yi Z, Zhou A, Ge L, Stephenson FA, Wenthold RJ. Organization of NMDA receptors at extrasynaptic locations. Neuroscience 2010; 167:68–87 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Delgado-Acevedo C, Estay SF, Radke AK, Sengupta A, Escobar AP, Henriquez-Belmar F, Reyes CA, Haro-Acuna V, Utreras E, Sotomayor-Zarate R, Cho A, Wendland JR, Kulkarni AB, Holmes A, Murphy DL, Chavez AE, Moya PR. Behavioral and synaptic alterations relevant to obsessive-compulsive disorder in mice with increased EAAT3 expression. Neuropsychopharmacology 2019; 44:1163–73 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Papouin T, Oliet SH. Organization, control and function of extrasynaptic NMDA receptors. Philos Trans R Soc Lond B Biol Sci 2014; 369:20130601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Hardingham GE, Fukunaga Y, Bading H. Extrasynaptic NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell death pathways. Nat Neurosci 2002; 5:405–14 [DOI] [PubMed] [Google Scholar]
  • 248.Okamoto S, Pouladi MA, Talantova M, Yao D, Xia P, Ehrnhoefer DE, Zaidi R, Clemente A, Kaul M, Graham RK, Zhang D, Vincent Chen HS, Tong G, Hayden MR, Lipton SA. Balance between synaptic versus extrasynaptic NMDA receptor activity influences inclusions and neurotoxicity of mutant huntingtin. Nat Med 2009; 15:1407–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Rothstein JD, Dykes-Hoberg M, Pardo CA, Bristol LA, Jin L, Kuncl RW, Kanai Y, Hediger MA, Wang Y, Schielke JP, Welty DF. Knockout of glutamate transporters reveals a major role for astroglial transport in excitotoxicity and clearance of glutamate. Neuron 1996; 16:675–86 [DOI] [PubMed] [Google Scholar]
  • 250.Sepkuty JP, Cohen AS, Eccles C, Rafiq A, Behar K, Ganel R, Coulter DA, Rothstein JA. Neuronal glutamate transporter contributes to neurotransmitter GABA synthesis and epilepsy. J Neurosci 2002; 22:6372–9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Stoffel W, Korner R, Wachtmann D, Keller BU. Functional analysis of glutamate transporters in excitatory synaptic transmission of GLAST1 and GLAST1/EAAC1 deficient mice. Brain Res Mol Brain Res 2004; 128:170–81 [DOI] [PubMed] [Google Scholar]
  • 252.Crino PB, Jin H, Shumate MD, Robinson MB, Coulter DA, Brooks-Kayal AR. Increased expression of the neuronal glutamate transporter (EAAT3/EAAC1) in hippocampal and neocortical epilepsy. Epilepsia 2002; 43:211–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Conti F, DeBiasi S, Minelli A, Rothstein JD, Melone M. EAAC1, a high-affinity glutamate tranporter, is localized to astrocytes and gabaergic neurons besides pyramidal cells in the rat cerebral cortex. Cereb Cortex 1998; 8:108–16 [DOI] [PubMed] [Google Scholar]
  • 254.Myles-Worsley M, Tiobech J, Browning SR, Korn J, Goodman S, Gentile K, Melhem N, Byerley W, Faraone SV, Middleton FA. Deletion at the SLC1A1 glutamate transporter gene co-segregates with schizophrenia and bipolar schizoaffective disorder in a 5-generation family. 2013; 162B:87–95 [DOI] [PubMed] [Google Scholar]
  • 255.Afshari P, Yao WD, Middleton FA. Reduced Slc1a1 expression is associated with neuroinflammation and impaired sensorimotor gating and cognitive performance in mice: Implications for schizophrenia. PLoS One 2017; 12:e0183854. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Greenberg BD, Ziemann U, Cora-Locatelli G, Harmon A, Murphy DL, Keel JC, Wassermann EM. Altered cortical excitability in obsessive-compulsive disorder. Neurology 2000; 54:142–7 [DOI] [PubMed] [Google Scholar]
  • 257.Escobar AP, Wendland JR, Chavez AE, Moya PR. The neuronal glutamate transporter EAAT3 in obsessive-compulsive disorder. Front Pharmacol 2019; 10:1362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Lin CL, Tzingounis AV, Jin L, Furuta A, Kavanaugh MP, Rothstein JD. Molecular cloning and expression of the rat EAAT4 glutamate transporter subtype. Brain Res Mol Brain Res 1998; 63:174–9 [DOI] [PubMed] [Google Scholar]
  • 259.Massie A, Cnops L, Smolders I, McCullumsmith R, Kooijman R, Kwak S, Arckens L, Michotte Y. High-affinity Na+/K+-dependent glutamate transporter EAAT4 is expressed throughout the rat fore- and midbrain. J Comp Neurol 2008; 511:155–72 [DOI] [PubMed] [Google Scholar]
  • 260.Dalet A, Bonsacquet J, Gaboyard-Niay S, Calin-Jageman I, Chidavaenzi RL, Venteo S, Desmadryl G, Goldberg JM, Lysakowski A, Chabbert C. Glutamate transporters EAAT4 and EAAT5 are expressed in vestibular hair cells and calyx endings. PLoS One 2012; 7:e46261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Mim C, Balani P, Rauen T, Grewer C. The glutamate transporter subtypes EAAT4 and EAATs 1-3 transport glutamate with dramatically different kinetics and voltage dependence but share a common uptake mechanism. J Gen Physiol 2005; 126:571–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Huang YH, Dykes-Hoberg M, Tanaka K, Rothstein JD, Bergles DE. Climbing fiber activation of EAAT4 transporters and kainate receptors in cerebellar Purkinje cells. J Neurosci 2004; 24:103–11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Bergles DE, Dzubay JA, Jahr CE. Glutamate transporter currents in Bergmann glial cells follow the time course of extrasynaptic glutamate. Proc Natl Acad Sci U S A 1997; 94:14821–5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Mashimo M, Okubo Y, Yamazawa T, Yamasaki M, Watanabe M, Murayama T, Iino M. Inositol 1,4,5-trisphosphate signaling maintains the activity of glutamate uptake in Bergmann glia. Eur J Neurosci 2010; 32:1668–77 [DOI] [PubMed] [Google Scholar]
  • 265.Yamashita A, Makita K, Kuroiwa T, Tanaka K. Glutamate transporters GLAST and EAAT4 regulate postischemic Purkinje cell death: an in vivo study using a cardiac arrest model in mice lacking GLAST or EAAT4. Neurosci Res 2006; 55:264–70 [DOI] [PubMed] [Google Scholar]
  • 266.Takayasu Y, Iino M, Kakegawa W, Maeno H, Watase K, Wada K, Yanagihara D, Miyazaki T, Komine O, Watanabe M, Tanaka K, Ozawa S. Differential roles of glial and neuronal glutamate transporters in Purkinje cell synapses. J Neurosci 2005; 25:8788–93 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Wersinger E, Schwab Y, Sahel JA, Rendon A, Pow DV, Picaud S, Roux MJ. The glutamate transporter EAAT5 works as a presynaptic receptor in mouse rod bipolar cells. J Physiol 2006; 577:221–34 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Pow DV, Barnett NL. Developmental expression of excitatory amino acid transporter 5: a photoreceptor and bipolar cell glutamate transporter in rat retina. Neurosci Lett 2000; 280:21–4 [DOI] [PubMed] [Google Scholar]
  • 269.Gameiro A, Braams S, Rauen T, Grewer C. The discovery of slowness: low-capacity transport and slow anion channel gating by the glutamate transporter EAAT5. Biophys J 2011; 100:2623–32 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Schneider N, Cordeiro S, Machtens JP, Braams S, Rauen T, Fahlke C. Functional properties of the retinal glutamate transporters GLT-1c and EAAT5. J Biol Chem 2014; 289:1815–24 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Herman MA, Jahr CE. Extracellular glutamate concentration in hippocampal slice. J Neurosci 2007; 27:9736–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Benveniste H, Drejer J, Schousboe A, Diemer NH. Elevation of the extracellular concentrations of glutamate and aspartate in rat hippocampus during transient cerebral ischemia monitored by intracerebral microdialysis. J Neurochem 1984; 43:1369–74 [DOI] [PubMed] [Google Scholar]
  • 273.Meur KL, Galante M, Angulo MC, Audinat E. Tonic activation of NMDA receptors by ambient glutamate of non‐synaptic origin in the rat hippocampus. J Physiol 2007; 580:373–83 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Cavelier P, Attwell D. Tonic release of glutamate by a DIDS‐sensitive mechanism in rat hippocampal slices. J Physiol 2005; 564:397–410 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Chiu DN, Jahr CE. Extracellular glutamate in the nucleus accumbens is nanomolar in both synaptic and non-synaptic compartments. Cell Rep 2017; 18:2576–83 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.De Bundel D, Schallier A, Loyens E, Fernando R, Miyashita H, Van Liefferinge J, Vermoesen K, Bannai S, Sato H, Michotte Y, Smolders I, Massie A. Loss of system x(c)(-) does not induce oxidative stress but decreases extracellular glutamate in hippocampus and influences spatial working memory and limbic seizure susceptibility. J Neurosci 2011; 31:5792–803 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Dash MB, Douglas CL, Vyazovskiy VV, Cirelli C, Tononi G. Long-term homeostasis of extracellular glutamate in the rat cerebral cortex across sleep and waking states. J Neurosci 2009; 29:620–9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Baker DA, McFarland K, Lake RW, Shen H, Tang XC, Toda S, Kalivas PW. Neuroadaptations in cystine-glutamate exchange underlie cocaine relapse. Nat Neurosci 2003; 6:743–9 [DOI] [PubMed] [Google Scholar]
  • 279.Massie A, Schallier A, Kim SW, Fernando R, Kobayashi S, Beck H, De Bundel D, Vermoesen K, Bannai S, Smolders I. Dopaminergic neurons of system xc-deficient mice are highly protected against 6-hydroxydopamine-induced toxicity. FASEB J 2011; 25:1359–69 [DOI] [PubMed] [Google Scholar]
  • 280.Sun W, Shchepakin D, Kalachev LV, Kavanaugh MP. Glutamate transporter control of ambient glutamate levels. Neurochem Int 2014; 73:146–51 [DOI] [PubMed] [Google Scholar]
  • 281.McKenna MC, Sonnewald U, Huang X, Stevenson J, Zielke HR. Exogenous glutamate concentration regulates the metabolic fate of glutamate in astrocytes. J Neurochem 1996; 66:386–93 [DOI] [PubMed] [Google Scholar]
  • 282.Divito CB, Underhill SM. Excitatory amino acid transporters: roles in glutamatergic neurotransmission. Neurochem Int 2014; 73:172–80 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Fairman WA, Amara SG. Functional diversity of excitatory amino acid transporters: ion channel and transport modes. Am J Physiol 1999; 277:F481–6 [DOI] [PubMed] [Google Scholar]
  • 284.Grewer C, Gameiro A, Rauen T. SLC1 glutamate transporters. Pflugers Arch 2014; 466:3–24 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 285.Wadiche JI, Amara SG, Kavanaugh MP. Ion fluxes associated with excitatory amino acid transport. Neuron 1995; 15:721–8 [DOI] [PubMed] [Google Scholar]
  • 286.Sonders MS, Amara SG. Channels in transporters. Curr Opin Neurobiol 1996; 6:294–302 [DOI] [PubMed] [Google Scholar]
  • 287.Grewer C, Rauen T. Electrogenic glutamate transporters in the CNS: molecular mechanism, pre-steady-state kinetics, and their impact on synaptic signaling. J Membr Biol 2005; 203:1–20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Bannai S. Exchange of cystine and glutamate across plasma membrane of human fibroblasts. J Biol Chem 1986; 261:2256–63 [PubMed] [Google Scholar]
  • 289.Piani D, Fontana A. Involvement of the cystine transport system xc-in the macrophage-induced glutamate-dependent cytotoxicity to neurons. J Immunol 1994; 152:3578–85 [PubMed] [Google Scholar]
  • 290.Jackman NA, Uliasz TF, Hewett JA, Hewett SJ. Regulation of system x(c)(-)activity and expression in astrocytes by interleukin-1beta: implications for hypoxic neuronal injury. Glia 2010; 58:1806–15 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Murphy T, Schnaar R, Coyle J. Immature cortical neurons are uniquely sensitive to glutamate toxicity by inhibition of cystine uptake. FASEB J 1990; 4:1624–33 [PubMed] [Google Scholar]
  • 292.Burdo J, Dargusch R, Schubert D. Distribution of the cystine/glutamate antiporter system x(c)(-) in the brain, kidney, and duodenum. J Histochem Cytochem 2006; 54:549–57 [DOI] [PubMed] [Google Scholar]
  • 293.Dun Y, Mysona B, Van Ells T, Amarnath L, Ola MS, Ganapathy V, Smith SB. Expression of the cystine-glutamate exchanger (x c−) in retinal ganglion cells and regulation by nitric oxide and oxidative stress. Cell Tissue Res 2006; 324:189–202 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Lewerenz J, Letz J, Methner A. Activation of stimulatory heterotrimeric G proteins increases glutathione and protects neuronal cells against oxidative stress. J Neurochem 2003; 87:522–31 [DOI] [PubMed] [Google Scholar]
  • 295.Gochenauer GE, Robinson MB. Dibutyryl‐cAMP (dbcAMP) up‐regulates astrocytic chloride‐dependent l‐[3H] glutamate transport and expression of both system xc− subunits. J Neurochem 2001; 78:276–86 [DOI] [PubMed] [Google Scholar]
  • 296.Ye Z-C, Sontheimer H. Glioma cells release excitotoxic concentrations of glutamate. Cancer Res 1999; 59:4383–91 [PubMed] [Google Scholar]
  • 297.Cho Y, Bannai S. Uptake of glutamate and cystine in C‐6 glioma cells and in cultured astrocytes. J Neurochem 1990; 55:2091–7 [DOI] [PubMed] [Google Scholar]
  • 298.Bender A, Reichelt W, Norenberg M. Characterization of cystine uptake in cultured astrocytes. Neurochem Int 2000; 37:269–76 [DOI] [PubMed] [Google Scholar]
  • 299.Pow DV. Visualising the activity of the cystine-glutamate antiporter in glial cells using antibodies to aminoadipic acid, a selectively transported substrate. Glia 2001; 34:27–38 [DOI] [PubMed] [Google Scholar]
  • 300.Zhang Y, Sloan SA, Clarke LE, Caneda C, Plaza CA, Blumenthal PD, Vogel H, Steinberg GK, Edwards MS, Li G. Purification and characterization of progenitor and mature human astrocytes reveals transcriptional and functional differences with mouse. Neuron 2016; 89:37–53 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Ottestad-Hansen S, Hu QX, Follin-Arbelet VV, Bentea E, Sato H, Massie A, Zhou Y, Danbolt NC. The cystine-glutamate exchanger (xCT, Slc7a11) is expressed in significant concentrations in a subpopulation of astrocytes in the mouse brain. Glia 2018; 66:951–70 [DOI] [PubMed] [Google Scholar]
  • 302.Cavelier P, Hamann M, Rossi D, Mobbs P, Attwell D. Tonic excitation and inhibition of neurons: ambient transmitter sources and computational consequences. Prog Biophys Mol Biol 2005; 87:3–16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Jabaudon D, Shimamoto K, Yasuda-Kamatani Y, Scanziani M, Gahwiler BH, Gerber U. Inhibition of uptake unmasks rapid extracellular turnover of glutamate of nonvesicular origin. Proc Natl Acad Sci U S A 1999; 96:8733–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Warr O, Takahashi M, Attwell D. Modulation of extracellular glutamate concentration in rat brain slices by cystine-glutamate exchange. J Physiol 1999; 514:783–93 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305.Baker DA, Shen H, Kalivas PW. Cystine/glutamate exchange serves as the source for extracellular glutamate: modifications by repeated cocaine administration. Amino Acids 2002; 23:161–2 [DOI] [PubMed] [Google Scholar]
  • 306.Baker DA, Xi ZX, Shen H, Swanson CJ, Kalivas PW. The origin and neuronal function of in vivo nonsynaptic glutamate. J Neurosci 2002; 22:9134–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Melendez RI, Vuthiganon J, Kalivas PW. Regulation of extracellular glutamate in the prefrontal cortex: focus on the cystine glutamate exchanger and group I metabotropic glutamate receptors. J Pharmacol Exp Ther 2005; 314:139–47 [DOI] [PubMed] [Google Scholar]
  • 308.Augustin H, Grosjean Y, Chen K, Featherstone DE. Nonvesicular release of glutamate by glial xCT transporters suppresses glutamate receptor clustering in vivo. J Neurosci 2007; 27:111–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 309.Featherstone DE, Shippy SA. Regulation of synaptic transmission by ambient extracellular glutamate. Neuroscientist 2008; 14:171–81 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Zorumski C, Mennerick S, Que J. Modulation of excitatory synaptic transmission by low concentrations of glutamate in cultured rat hippocampal neurons. J Physiol 1996; 494:465–77 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311.Williams LE, Featherstone DE. Regulation of hippocampal synaptic strength by glial xCT. J Neurosci 2014; 34:16093–102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Sah P, Hestrin S, Nicoll R. Tonic activation of NMDA receptors by ambient glutamate enhances excitability of neurons. Science 1989; 246:815–8 [DOI] [PubMed] [Google Scholar]
  • 313.Dalby NO, Mody I. Activation of NMDA receptors in rat dentate gyrus granule cells by spontaneous and evoked transmitter release. J Neurophysiol 2003; 90:786–97 [DOI] [PubMed] [Google Scholar]
  • 314.Angulo MC, Kozlov AS, Charpak S, Audinat E. Glutamate released from glial cells synchronizes neuronal activity in the hippocampus. J Neurosci 2004; 24:6920–7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.LoTurco JJ, Blanton M, Kriegstein AR. Initial expression and endogenous activation of NMDA channels in early neocortical development. J Neurosci 1991; 11:792–9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.LoTurco JJ, Owens DF, Heath MJ, Davis MB, Kriegstein AR. GABA and glutamate depolarize cortical progenitor cells and inhibit DNA synthesis. Neuron 1995; 15:1287–98 [DOI] [PubMed] [Google Scholar]
  • 317.Manent J-B, Demarque M, Jorquera I, Pellegrino C, Ben-Ari Y, Aniksztejn L, Represa A. A noncanonical release of GABA and glutamate modulates neuronal migration. J Neurosci 2005; 25:4755–65 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 318.Nguyen L, Rigo J-M, Rocher V, Belachew S, Malgrange B, Rogister B, Leprince P, Moonen G. Neurotransmitters as early signals for central nervous system development. Cell Tissue Res 2001; 305:187–202 [DOI] [PubMed] [Google Scholar]
  • 319.Soria FN, Perez-Samartin A, Martin A, Gona KB, Llop J, Szczupak B, Chara JC, Matute C, Domercq M. Extrasynaptic glutamate release through cystine/glutamate antiporter contributes to ischemic damage. J Clin Invest 2014; 124:3645–55 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Fogal B, Li J, Lobner D, McCullough LD, Hewett SJ. System x(c)- activity and astrocytes are necessary for interleukin-1beta-mediated hypoxic neuronal injury. J Neurosci 2007; 27:10094–105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 321.Jackman NA, Melchior SE, Hewett JA, Hewett SJ. Non-cell autonomous influence of the astrocyte system xc- on hypoglycaemic neuronal cell death. ASN Neuro 2012; 4:e00074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Ye Z-C, Rothstein JD, Sontheimer H. Compromised glutamate transport in human glioma cells: reduction–mislocalization of sodium-dependent glutamate transporters and enhanced activity of cystine–glutamate exchange. J Neurosci 1999; 19:10767–77 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Takeuchi S, Wada K, Toyooka T, Shinomiya N, Shimazaki H, Nakanishi K, Nagatani K, Otani N, Osada H, Uozumi Y. Increased xCT expression correlates with tumor invasion and outcome in patients with glioblastomas. Neurosurgery 2012; 72:33–41 [DOI] [PubMed] [Google Scholar]
  • 324.Robert SM, Buckingham SC, Campbell SL, Robel S, Holt KT, Ogunrinu-Babarinde T, Warren PP, White DM, Reid MA, Eschbacher JM. SLC7A11 expression is associated with seizures and predicts poor survival in patients with malignant glioma. Sci Transl Med 2015; 7:289ra86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Buckingham SC, Campbell SL, Haas BR, Montana V, Robel S, Ogunrinu T, Sontheimer H. Glutamate release by primary brain tumors induces epileptic activity. Nat Med 2011; 17:1269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Baker DA, Madayag A, Kristiansen LV, Meador-Woodruff JH, Haroutunian V, Raju I. Contribution of cystine-glutamate antiporters to the psychotomimetic effects of phencyclidine. Neuropsychopharmacology 2008; 33:1760–72 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Kau KS, Madayag A, Mantsch JR, Grier MD, Abdulhameed O, Baker DA. Blunted cystine-glutamate antiporter function in the nucleus accumbens promotes cocaine-induced drug seeking. Neuroscience 2008; 155:530–7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Hakami AY, Sari Y. beta-Lactamase inhibitor, clavulanic acid, attenuates ethanol intake and increases glial glutamate transporters expression in alcohol preferring rats. Neurosci Lett 2017; 657:140–5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 329.Reissner KJ, Gipson CD, Tran PK, Knackstedt LA, Scofield MD, Kalivas PW. Glutamate transporter GLT-1 mediates N-acetylcysteine inhibition of cocaine reinstatement. Addict Biol 2015; 20:316–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Knackstedt LA, LaRowe S, Mardikian P, Malcolm R, Upadhyaya H, Hedden S, Markou A, Kalivas PW. The role of cystine-glutamate exchange in nicotine dependence in rats and humans. Biol Psychiatry 2009; 65:841–5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.McCullagh EA, Featherstone DE. Behavioral characterization of system xc-mutant mice. Behav Brain Res 2014; 265:1–11 [DOI] [PubMed] [Google Scholar]
  • 332.Li Y, Tan Z, Li Z, Sun Z, Duan S, Li W. Impaired long-term potentiation and long-term memory deficits in xCT-deficient sut mice. Biosci Rep 2012; 32:315–21 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Bentea E, Demuyser T, Van Liefferinge J, Albertini G, Deneyer L, Nys J, Merckx E, Michotte Y, Sato H, Arckens L. Absence of system xc-in mice decreases anxiety and depressive-like behavior without affecting sensorimotor function or spatial vision. Prog Neuropsychopharmacol Biol Psychiatry 2015; 59:49–58 [DOI] [PubMed] [Google Scholar]

Articles from Experimental Biology and Medicine are provided here courtesy of Frontiers Media SA

RESOURCES