Open access peer-reviewed chapter - ONLINE FIRST

Glutamate Transporters in Health and Disease

Written By

Katelyn L. Reeb, Simran K. Gill, Rhea Temmermand and Andréia C.K. Fontana

Submitted: 28 February 2024 Reviewed: 01 April 2024 Published: 04 June 2024

DOI: 10.5772/intechopen.1005544

Two Sides of the Same Coin - Glutamate in Health and Disease IntechOpen
Two Sides of the Same Coin - Glutamate in Health and Disease Edited by Kaneez Fatima-Shad

From the Edited Volume

Two Sides of the Same Coin - Glutamate in Health and Disease [Working Title]

Prof. Kaneez Fatima Shad

Chapter metrics overview

29 Chapter Downloads

View Full Metrics

Abstract

Glutamate transporters, or excitatory amino acid transporters (EAATs), are key proteins that regulate the excitatory tone in the central nervous system (CNS) by clearing synaptic glutamate, maintaining extracellular glutamate concentrations low enough to prevent receptor desensitization and/or glutamate-mediated excitotoxicity. Dysregulation of the function and/or expression of the EAATs is implicated in several diseases, including epilepsy, stroke, traumatic brain injury, drug abuse disorders, neurodegenerative disorders, and neuropathic pain, among others. In this chapter, we will discuss the regulatory mechanisms of EAATs in health and disease states. We will discuss post-translational modifications, trafficking deficits, reverse transport, and other regulatory processes. We will also discuss current approaches on potential therapeutic strategies targeting these transporters for many neuropsychiatric diseases.

Keywords

  • glutamate
  • glutamate transporters
  • EAAT
  • EAAT2
  • expression enhancers
  • GLT-1
  • ischemia
  • neuropathic pain
  • stroke
  • drug use disorder
  • allosteric modulation

1. Introduction

Glutamate transporters, or excitatory amino acid transporters (EAATs), play a crucial role in regulating excitatory activity in the central nervous system (CNS). Thus, studying their regulation is essential for understanding health and diseased states. In physiological states, glutamate is involved in memory, learning, and other processes, and EAATs exert a tight control of the synaptic concentration of glutamate, through clearance into glial and neuronal cells. In diseased states, glutamate transporter function and/or expression can be dysregulated, leading to devastating effects in the CNS. Under conditions of prolonged glutamate activation, cells become overexcited by glutamate and go through degeneration and, ultimately, death, in a process called excitotoxicity [1]. Glutamate-mediated excitotoxicity has been shown to be involved in numerous conditions, such as ischemic stroke, epilepsy, and traumatic brain injury. Further, dysregulated levels of glutamate transporters that result in abnormal synaptic glutamate concentration are observed in several pathologies, such as drug abuse disorders, neurodegenerative disorders, and neuropathic pain, among others. In this chapter, we will discuss the regulatory mechanisms of EAATs in physiological states and the implication of glutamate transporter dysregulation in neurological and neuropsychiatric disorders.

Advertisement

2. Glutamate transporters in health

In the mammalian CNS, glutamate serves as the main excitatory neurotransmitter and is a critical signal for neural communication and plasticity. Once released into the synaptic cleft, glutamate promotes specific signaling pathways in post-synaptic neurons by interacting with ionotropic and metabotropic glutamate receptors, which initiate downstream signaling. As extracellular glutamate cannot be enzymatically degraded, glutamate must be removed from the synaptic cleft through glutamate transporters [2]. Therefore, EAATs are imperative for proper neuronal functioning, as they clear synaptic glutamate and maintain excitatory balance.

2.1 Subtypes and localization in the nervous system

There are two main classes of glutamate transporters: EAATs, which are dependent on an electrochemical gradient of sodium (Na+) ions, and the vesicular glutamate transporters (VGLUT-1-3) and cystine-glutamate antiporters (xCT), which are Na+-independent (Table 1). EAATs and xCT are both found in cell membranes; however, xCT has much lower expression than the EAATs, whereas vGLUTs are found in the membrane of glutamate-containing synaptic vesicles. In this chapter, we will focus on EAATs. For a review of the xCT system, see [3]; for a review of vGLUTs, see [4].

Glutamate transporter subtype-Human homologGlutamate transporter subtype-Rodent homologGeneCell typeAnatomic localization
EAAT1GLASTSLC1A3Astrocytes, oligodendrocytesCerebellum, cortex, spinal cord
Also, in testis and bone
EAAT2GLT-1SLC1A2Mainly astrogliaThroughout brain and spinal cord
Also, in liver
EAAT2bGLT-1bSLC1A2Mainly astroglia, also expressed in neuronsThroughout brain and spinal cord
Also in intestine, kidney, liver, and heart
EAAT3EAAC1SLC1A1Neurons (dendrites and axon terminals)Hippocampus, cerebellum, striatum
EAAT4EAAT4SLC1A6Neurons (Purkinje cells)Cerebellum, hippocampus, and basal ganglia
Also, in placenta
EAAT5EAAT5SLC1A7Neurons (photoreceptors and bipolar cells)Retina
Also, in liver
VGLUT1VGLUT1SLC17A7NeuronsCerebral cortex, hippocampus, and cerebellum
VGLUT2VGLUT2SLC17A6NeuronsThalamus and brainstem
VGLUT3VGLUT3SLC17A8NeuronsCerebral cortex, hippocampus, striatum, and raphe nuclei
xCTxCTSLC7A11Neurons and gliaHippocampus, cortex, hypothalamus, and dentate gyrus

Table 1.

Glutamate transporters: nomenclature (human/rodent/genes), cell type and anatomic localization.

EAATs are classified into five subtypes (rat/human homolog/gene): GLAST/EAAT1/SLC1A3, GLT-1/EAAT2/ SLC1A2, EAAC1/EAAT3/ SLC1A1, EAAT4/SLC1A6, and EAAT5/SLC1A7 [2]. EAAT1 and EAAT2 are mainly localized in astrocytes, whereas EAAT3-5 are neuronal [5]. The main transporters responsible for the uptake of synaptic glutamate are astrocytic EAAT1 and EAAT2; thus, they are key for preventing the accumulation of synaptic glutamate an avoid excitotoxicity [6]. EAAT1 is predominantly expressed in astrocytes within the cerebellar Purkinje cell layer [7]. EAAT2 is the predominant subtype of glutamate transporter in the CNS [8] and is expressed in astrocytes, select presynaptic neurons, and oligodendrocytes within the brain and spinal cord [9]. EAAT2 expression contributes to around 95% of total glutamate transport activity and represents approximately 1% of the total brain protein in the CNS [10], therefore, playing a key role in the maintenance of extracellular glutamate homeostasis. While both EAAT1 and EAAT2 are expressed within the same astrocytic plasma membrane, they are regulated differently; exogenous glutamate levels affect the cell-surface expression of EAAT1, but not EAAT2; EAAT2 is regulated by neuronal soluble factors, unlike EAAT1 [11]. Several splice variants of EAAT2 have been identified, with EAAT2b (or GLT-1b) being the most extensively studied. However, apparent functional differences between each of the variants are not readily apparent [12]. EAAT3 exhibits ubiquitous expression in the brain, playing a crucial role in regulating local glutamate concentrations, as it is predominantly located within post-synapsis, which allows for the buffering of nearby glutamate receptors, influencing excitatory neurotransmission and synaptic plasticity [13]. EAAT4, which is expressed in cerebellar neurons, is primarily responsible for maintaining low extracellular glutamate levels and preventing neurotoxicity in the cerebellum, together with glial EAAT1 [14]. EAAT5, a neuronal transporter mainly expressed in retina, plays a vital role in controlling glutamate release and mediating light responses in depolarizing bipolar cells in the retina [15].

See Figure 1 for an overview of the localization of EAATs in the CNS.

Figure 1.

Localization of neuronal and astrocytic excitatory amino acid transporters (EAATs) in the CNS. A. Overview of a glutamate tripartite synapse (pre- and post-synaptic neurons and astrocytes) showing the localization of astrocytic transporters EAAT1 and EAAT2 and neuronal transporter EAAT3, in the cerebrum and hippocampus. After glutamate (Glu, red dots) is released from the presynaptic terminal and stimulates postsynaptic glutamate receptors AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic) and NMDA (N-methyl-D-aspartate), and metabotropic glutamate receptors (mGluRs), it undergoes reuptake through EAATs present in astrocytes and neurons. In astrocytes, glutamate is converted into glutamine (Gln, green dots), which is shuttled back to neurons. In addition to glutamate, EAAT3 also transports the glutathione precursor cysteine (not shown) into neurons, which is required to produce glutathione. B. Overview of a glutamate tripartite synapse (pre- and post-synaptic neurons and astrocytes) showing the localization of astrocytic transporters EAAT1 and EAAT2 and neuronal transporters EAAT3 and EAAT4 in the cerebellum. Note that EAAT4 is only expressed in synapses in the cerebellum. C. Overview of a glutamate tripartite synapse (pre- and post-synaptic neurons and astrocytes) showing the localization of astrocytic transporter EAAT1 and EAAT2. Note that EAAT2, in the retina, is mainly expressed in neurons such as photoreceptors and bipolar cells, and neuronal transporter EAAT5 is only expressed in bipolar cells neurons in the retina. EAAT5 is restricted to. Created with BioRender.

2.2 Stoichiometry and structure of the EAATs

EAATs function through an electrogenic process that is dependent on the co-transport of three Na+ (sodium) ions and one proton with glutamate and the counter-transport of one K+ (potassium) ion out of the cell, in addition to an uncoupled Cl (chloride) current (Figure 2A) [16, 17]. This process is also known as an “induced fit mechanism,” as Na+ binding is required for glutamate binding, which also contributes to the high selectivity of EAATs for glutamate [18]. As EAATs are secondary active transporters, the ionic gradients of these substrates facilitate the glutamate transport cycle. EAATs rely indirectly on Na+/K+-ATPase (NKA) to generate these ion gradients [19].

Figure 2.

Molecular properties of EAATs. A. Schematic displaying substrate stoichiometry associated with glutamate transport, displayed as trimer in the membrane. Glutamate (Glu) uptake is driven by co-transport of three Na+ ions and one proton (H+) and by counter-transport of one K+ ion. EAATs also show an uncoupled Cl conductance (shown as dotted arrow). B. Transmembrane topology of glutamate transporters consisting of eight transmembrane domains and two hairpin loops (HP1 and HP2). Transmembrane domains are shown in dark blue, and scaffold domains in light blue. C. Glutamate transport cycle of a single protomer. Transport domain (magenta) of protomer moves along the scaffold domain (green) to go from an outward-facing conformation (far left) to an inward conformation (far right), passing through an intermediate conformation in which chloride conductance can occur. Created with BioRender.

EAATs are transmembrane integral proteins that traverse the cell membrane eight times (Figure 2B). EAATs consist of three protomers that can independently transport glutamate [16]. They have four key domains: a scaffold domain, or trimerization domain (comprised of transmembrane domains 1, 2, 4, 5), which remains stationary; a transporter domain (transmembrane domains 3, 6, 7, 8) that moves as a large rigid body along the scaffold domain in a twisting elevator-like motion to transport glutamate and its cosubstrates into the cell, and two hairpin domains (hairpin domains 1 and 2) that act as intracellular and extracellular gates [20]. These transporters follow an “alternating-access model” through this motion, which brings the transporter from an outward-facing conformation, where glutamate and cosubstrates bind, to an inward-facing conformation, where they are released (Figure 2C) [21]. While transporters and ion channels have historically been viewed as distinct proteins, EAATs also function as anion channels, which open during an intermediate state of the transport cycle [22]. This anion channel in EAATs is selective for chloride and is stoichiometrically uncoupled from glutamate transport [23]. Additionally, this chloride conductance may mediate neuronal excitability and ion homeostasis [24]. The malfunction of this chloride channel has also been linked to neurological diseases such as episodic ataxia [25]. Recent cryo-EM studies have reported the structures of EAAT1 [26], EAAT3 [27], and EAAT2 [28, 29] in several states in the presence of substrate glutamate or inhibitors. These studies are highly relevant as they revealed the structural basis of coupled substrate and ion binding.

2.3 Modulators of the activity of EAATs

In the last decades, compounds that modulate EAAT functions, directly or indirectly, have been identified and developed (some examples are outlined in Table 2).

CompoundMechanismsReferences
Indirect activators
Riluzole (2-amino 6-(trifluoromethoxy)benzothiazole)
  • Increases activity and expression of EAAT2

  • Blocks sodium channels

[30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43]
Positive allosteric modulators (PAMs)
Parawixin1 (unknown chemical structure)
  • Activates EAAT2 activity

[44]
GT949 (3-((4-Cyclohexylpiperazin-1-yl) (1-phenethyl-1H-tetrazol-5-yl) methyl)-6-methoxyquinolin-2(1H)-one and GT951 6-methoxy-3-((1-phenethyl-1H-tetrazol-5-yl) (4-(3-(trifluoromethyl) phenyl) piperazin-1-yl) methyl) quinolin-2(1H)-one)
  • EAAT2 PAMs, activity, no effect on NMDA-mediated currents

  • In vitro neuroprotection in glutamate-mediated excitotoxicity models

[45, 46]
Parawixin10 (S)-N1-(3-(2-amino-5-guanidinopentanamido)propyl)-N4-(3-(2-(4-hydroxy-1H-indol-3-yl)acetamido)propyl)-N1,N1,N4,N4-tetramethylbutane-1,4-diaminium iodide)
  • Increases glutamate uptake in rat brain synaptosomes.

  • PAM of EAAT1 and EAAT2

  • Offers neuroprotection in in vitro stroke models and in vivo epilepsy models

[47, 48, 49, 50]
[(R)-AS-1] [(R)- N-Benzyl-2-(2,5-dioxopyrrolidin-1-yl)propanamide)]
  • Selective EAAT2 PAM

  • Offers in vivo protection from seizures

[51]
Competitive Inhibitors
DL-TBOA and analogs (DL-threo-beta-benzyloxy aspartate)
  • Non-transportable blocker of all subtypes of EAATs

[52, 53, 54, 55, 56, 57, 58, 59, 60]
TFB-TBOA [(3S)-3-[[3-[[4-(Trifluoromethyl)benzoyl] amino] phenyl] methoxy]-L-aspartic acid]
  • Selective inhibitor of EAAT1 and EAAT2

[61, 62, 63, 64]
Non-competitive inhibitors
Several putative negative allosteric modulators
  • Selective and non-selective inhibition of EAAT1, EAAT2, and/or EAAT3

[65]

Table 2.

Examples of modulators of excitatory amino acid (EAAT) activity.

The table includes indirect activators (compounds that interact indirectly with EAATs augmenting their catalytic activity, usually acting through multiple mechanisms), positive allosteric modulators (PAMs, compounds that interact directly with EAATs at an allosteric site and putatively increase their activity), competitive inhibitors (compounds that bind to the same binding sites as glutamate, competing with glutamate for binding to the transporter protein; some of these inhibitors are substrates themselves, some are exchanged with internal glutamate and induce glutamate release), and non-competitive inhibitors (or negative allosteric modulators, NAMs, that act by binding EAATs at an allosteric site, often leading to a conformational change that impacts transporter dynamics). In addition, we provide some of their known mechanisms and include further references.

Several compounds indirectly interact with EAATs, augmenting their catalytic activity, usually acting through multiple mechanisms, such as riluzole [30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 66]. Inhibitors of EAAT activity are grouped into competitive (bind to the same binding sites as glutamate) and non-competitive (bind somewhere else, i.e., to an allosteric site). Some of these inhibitors are substrates themselves, and some are exchanged with internal glutamate, thereby inducing glutamate release [52].

An emerging approach to drug design focuses on allosteric modulators, which bind to allosteric sites and alter the conformation of the orthosteric binding site, affecting transport by either enhancing (positive allosteric modulation, PAM) or inhibiting (negative allosteric modulation, NAM) binding affinity and transport. Allosteric modulation may offer advantages such as targeted drug therapy and increased specificity, thus offering promising prospects for drug discovery and the development of novel therapeutics [67]. Recently, allosteric sites have been described in EAATs [45, 46]. The development of selective EAAT PAMs and NAMs may be useful tools to decipher how drugs can affect the transport cycle and to investigate the intrinsic properties and functions of the EAATs [68].

Further research on allosteric modulators of EAATs is necessary to address several outstanding questions. These include determining the precise location and physiological significance of the allosteric binding sites, investigating whether these compounds can stabilize specific conformations of the transporter, understanding how different conformations of EAATs transmit signals into the cell, and exploring the possibility of allosterically modulating transporter-mediated efflux. Ultimately, this knowledge could be invaluable for advancing our understanding of EAAT modulation and for informing the drug development process [69].

2.4 Modulators of the expression of EAATs

The expression of EAATs is altered in certain disease states, leading to changes in neuronal excitation. This regulation can occur through transcriptional and translational processes [70]. However, the exact mechanisms underlying EAATs expression regulation are not well understood; hence, synthetic modulators of expression could serve as valuable tools for investigating the functions and regulation of EAATs in both physiological and disease contexts. There are exogenous drugs, endogenous molecules, and proteins that increase EAATs expression (transcriptional and level, i.e., increase gene transcription, and translational, i.e., increase the protein expression) and inhibitors of EAAT expression. Additional research is needed to evaluate whether a compound that enhances GLT-1 expression and has good brain penetrance, favorable pharmacokinetic properties, and low risk of side effects and toxicity can be developed into a therapeutic drug [71]. Some commonly studied EAAT expression modulators are outlined in Table 3.

CompoundMechanismReferences
Expression enhancers
Ceftriaxone ((6R,7R)-7-{[(2Z)-2-(2-amino-1,3-thiazol-4-yl)- > 2-(methoxyimino) acetyl]amino}MI-3-{[(2-methyl-5,6-dioxo-1,2,5,6-tetrahydro-1,2,4-triazin-3-yl)thio]methyl}-8-oxo-5-thia-1-azabicyclo[4.2.0]oct-2-ene-2-carboxylic acid)
  • β-Lactam antibiotic

  • Selective enhancer of EAAT2 expression, through transcriptional activation via mechanisms involving PI3K/Akt/NF-κB

[72, 73, 74, 75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100]
Clavulanic acid ((2R,5R, Z)-3-(2-hydroxyethylidene)-7-oxo-4-oxa-1-aza-bicyclo [3.2.0] heptane-2-carboxylic acid)
  • Structural analog β-lactam but lacks antibiotic effects

  • Increases EAAT2 expression (has better oral availability and blood-brain barrier penetration than ceftriaxone)

[77, 79, 101, 102, 103, 104, 105, 106]
LDN/OSU-0212320 (Thiopyridazine and pyridazine derivatives)
  • Increases EAAT2 expression through translational activation

[107, 108, 109, 110, 111, 112]
Amitriptyline (3-(10,11-dihydro-5H-dibenzo [a, d] cycloheptene-5-ylidene)-N, N-dimethylpropan-1-amine)
  • Tricyclic antidepressant

  • Induces EAAT2 expression and ameliorates neuropathic pain

[113, 114]
N-acetylcysteine ((2R)-2-acetamido-3-sulfanylpropanoic acid))
  • Increases EAAT2 expression-

  • Reduces drug-seeking/taking behavior for cocaine, nicotine, and ethanol

[115, 116, 117, 118, 119, 120, 121]
Minocycline ((2E,4S,4aR,5aS,12aR)-2-(Amino-hydroxy
-methylidene)-4,7-bis(dimethylamino)-10,11,12a-trihydroxy-4a,5,5a,6-tetrahydro-4H-tetracene-1,3,12-trione)
  • Broad-spectrum tetracycline antibiotic

  • Ameliorates downregulation of GLT-1 expression in neuropathy model

[122]
Expression inhibitors
PMA (phorbol 12-myristate 13-acetate)
  • Activator of PKC

  • Decreases the activity and expression of GLT-1 through clathrin-mediated endocytosis

  • (Also has actions on EAAT3)

[123, 124, 125, 126, 127, 128]
Synthetic cathinone MDPV (3,4-methylenedioxypyrovalerone methylenedioxy pyrovalerone)
  • Downregulates GLT-1 expression, and norepinephrine-dopamine reuptake

[129]
Amphetamine (1-phenylpropan-2-amine)
  • Potent CNS stimulant

  • Downregulates EAAT3 expression

  • Enters dopamine neurons via DAT and triggers endocytosis of EAAT3, in a process involving dynamin- and Rho-mediated mechanisms

[130, 131]

Table 3.

Examples of modulators of Excitatory Amino Acid Transporters (EAAT) expression.

The table includes examples of drugs that enhance EAATs expression (at the transcriptional and translational level), and examples of drugs that inhibit EAATs expression, along with some of their known mechanisms and further references.

2.5 Trafficking and post-translational modifications of EAATs

The expression and function of EAATs are regulated at the genetic, epigenetic, transcriptional, post-transcriptional, and translational levels [132]. Dysregulation of these processes can lead to severe outcomes such as high levels of extracellular glutamate and excitotoxicity [6]. The activity of transporters can be regulated by many means, such as ubiquitination, phosphorylation glycosylation, and sulfhydryl oxidation, among others [133, 134, 135].

Post-translational modification of GLT-1 by ubiquitin conjugation to lysine residues has been found to mediate the transporters constitutive internalization and degradation. This regulated clathrin-mediated endocytosis in basal conditions determines the availability of transporter on the cell surface and therefore relates directly to activity levels and rate of glutamate transport [123, 136]. Additionally, the palmitoylation of GLT-1 has been shown to drive glutamate uptake kinetics. Palmitoylation involves the covalent attachment of palmitic acid to one of more cysteine residues of the target protein. Studies suggest that palmitoylation of GLT-1 is important for glutamate uptake capacity, as a reduction in palmitoylated GLT-1 leads to impairments in glutamate clearance [137]. The sumoylation of GLT-1 in physiological signaling has also been identified, which involves the attachment of a SUMO protein to a lysine residue on the target protein and was found to regulate GLT-1 subcellular localization [138]. Notably, several groups have reported aberrations in the constitutive trafficking, expression, and activity levels of glutamate transporters due to alterations in post-translational modifications in many disease states, which will be further discussed in Section 3.

Advertisement

3. Glutamate transporters in disease

Dysregulation of EAATs plays a significant role in many neuropsychiatric diseases/disorders. Below, we briefly discuss two mechanisms of EAAT dysregulation involved in several disorders: glutamate efflux via transporter reversal and downregulation of the expression of glutamate transporters. We also briefly describe studies of knockout of EAAT that were pivotal for our understanding of the role of glutamate transporters in health and disease.

3.1 Glutamate transporter reversal

Under physiological conditions, the direction of glutamate transport is inward; however, in pathological conditions, the function of Na+/K+-ATPase (NKA) becomes dysfunctional, with subsequent disruption of Na+/K+ electrochemical gradient needed for glutamate translocation, resulting in increases in extracellular K+ and decreases in Na+ decreases, which leads to glutamate transport in the outward direction [139]. In support of this, massive increases in extracellular K+ and indiscriminate release of glutamate have been reported following concussive and lateral fluid percussion brain injuries [140], ischemia [141], and oxidative stress in a SOD1 mutation amyotrophic lateral sclerosis (ALS) model [142, 143].

It remains to be clarified whether EAAT functional activators or expression enhancers will facilitate glutamate clearance under excitotoxic conditions or will intensify reverse transport. However, previous preclinical studies suggest neuroprotective properties of these compound classes against ischemia, which encourages future drug development using this strategy. Nevertheless, as glutamatergic signaling plays crucial roles in brain development, cell survival, and synaptogenesis, these pharmacological strategies encounter similar challenges to other drugs targeting glutamate-mediated mechanisms, such as the potential for significant adverse effects [71, 144].

3.2 Glutamate transporter downregulation

Reduced expression and function of EAATs has been reported in numerous neurological disorders. Although the exact mechanism of downregulation has yet to be fully established, it results in impairment of the overall function of glutamate transporters, which plays an important role in the etiology of neurological diseases (see below).

For an overview of the role of EAATs at glutamatergic synapses at physiological (A) and pathological (B) states, see Figure 3.

Figure 3.

Tripartite glutamatergic synapse in physiological and pathological conditions. Schematic representing glutamatergic transmission in the context of physiological (A) and pathological (B) states that are associated with glutamate-induced excitotoxicity and EAAT2 protein dysregulation. A. In B. Hyper-glutamatergic signaling results in excess glutamate (red dots) release, overactivation of ionotropic glutamate receptors AMPA and NMDA, and aberrant calcium (blue dots) influx leading to the activation of various cell death pathways. In many disease states, glutamatergic dysfunction can lead to transporter reversal (shown by the arrow pointing to glutamate efflux from the astrocytes to the extracellular environment) and downregulation (shown as lower EAAT2 expression than in A), further exacerbating excitotoxic outcomes, such as in ischemic stroke. Na+/K+ ATPase (NKA) dysfunction also contributes to the dysregulation of EAAT activity. Downregulation of EAAT2 expression is also observed in pathologies such as in drugs of abuse disorders (depicted in the figure as drug-seeking behavior) and neuropathic pain. Created with BioRender.

3.3 Studies with knockout of EAATs

Studies with knockout of glutamate transporters reveal a major role for EAATs in clearance of glutamate, excitotoxicity, and associated neurotransmission. A key study by Tanaka’s group in the 1990s demonstrated that EAAT2 gene knockout resulted in lethal spontaneous seizures and increased susceptibility to acute cortical injury in mice [145]. Later studies confirmed severe disturbances in mice lacking GLT-1 and GLAST [146], including elevated extracellular glutamate levels, exacerbated hippocampal neuronal damage after brain injury [147], impairment of several essential aspects of neuronal development [148], neurodegeneration and progressive paralysis [149]. Mice lacking EAAT1 have decreased cerebellar function, reduced motor coordination, hearing loss, and disturbed retinal function. Antisense knockdown of GLAST compromises retinal function [150], and constitutive deletion of GLAST results in markedly reduced alcohol consumption and preference [151]. Mice lacking EAAT3 develop dicarboxylic aminoaciduria (Kegg’s disease), exhibit reduced spontaneous locomotor activity, and may age prematurely [152]. Additionally, they experience depletion of glutathione and neuronal cell loss [153], suggesting that EAAT3 plays a critical role in providing cysteine for glutathione synthesis [154]. Humans lacking EAAT3 develop dicarboxylic aminoaciduria, a rare metabolic disorder that causes the body to excrete too much aspartate and glutamate in urine, and human EAAT3 polymorphisms have been reported to be associated with obsessive-compulsive disorders [155]. In EAAT4 knockout mice, Purkinje cells are more likely to die [156], and in a double knockout of EAAT1 and EAAT4, both highly expressed in the cerebellum, a differential effect was observed in the spontaneous firing pattern and survival of Purkinje cells [14], demonstrating the essential role of these transporters in the cerebellum. Finally, EAAT5 was shown to shape the retinal light responses, in an EAAT5 knockout model [157].

3.4 Stroke

Stroke is the third leading cause of death in the United States, accounting for 700,000 fatalities each year [158]. Following an ischemic event, inadequate blood flow to the brain prevents the delivery of oxygen and glucose. The deprivation of these substrates to neurons causes cell death and lasting brain damage [159]. Following a focal ischemic event, the resulting damage manifests into two distinct regions that are classified as the infarction core and the penumbra. The core is the region of the brain in which the primary occlusion occurs, which undergoes rapid and irreversible cellular death within minutes of ischemic onset, usually due to necrosis [160]. The penumbra is classified as the region that surrounds the core, which receives reduced blood supply but remains partially metabolically active and therefore contains salvageable tissue. However, this region is at risk of cell death if blood flow is not quickly restored [159], making the penumbra the primary area of interest for new targeted therapies.

Glutamatergic dysfunction is a key factor contributing to neuronal cell death following ischemic stroke [161, 162]. The lack of oxygen due to cessation of blood flow results in depleted ATP stores, thereby disrupting the ionic gradients responsible for regulating neuronal firing, resulting in increased action potentials and aberrant glutamatergic signaling. Excessive release of glutamate during ischemia leads to overactivation of postsynaptic ionotropic glutamate receptors. Released glutamate may also diffuse out of the synaptic cleft, causing activation of distant (extrasynaptic) receptors and subsequent elevated calcium influx into the cytosol. This affects calcium-sensitive organelles such as the mitochondria and endoplasmic reticulum [163] and causes the release of calcineurin and calpains, mediators of cell death [164]. Mitochondrial dysfunction and oxidative stress also play key parts in the excitotoxic phenotype [165]. Thus, minimizing glutamate-induced excitotoxicity by regulating the aberrant signaling cascade following stroke can be therapeutically beneficial in improving post-stroke outcomes.

As EAATs are responsible for removing glutamate from the synaptic space to help end neurotransmission, these proteins play a key role in mitigating excitotoxic outcomes. Recent research has focused on better understanding the regulation of EAATs during or after an ischemic event as well as on strategies to modulate their expression and/or activity to bolster glutamate clearance and promote cell recovery. Many groups have identified temporal and spatial alterations in glutamate transporter expression over the course of ischemic injury [166]. Additionally, the activity and function of glutamate transporters can be significantly affected by the severity of ischemic insult, with severe insults even causing a reversal in glutamate transport due to disruptions in the ionic gradients that drive these transporters [167]. Membrane translocation of glutamate transporters after ischemia has also been suggested, with elevations in glutamate release causing rapid changes in the diffusion and clustering of EAAT2 along the plasma membrane [168]. Taken together, the regulatory movement and response of glutamate transporters to ischemic insult is complex and multifaceted.

An extensively researched pathway in addressing ischemic injury involves the modulation of glutamate transporter expression. Pharmacological preconditioning with several β-lactam antibiotics, most notably ceftriaxone, an EAAT2 expression enhancer, has shown promising results as a treatment option in several neurological disorders including ischemic stroke. Previous work has found that ceftriaxone, when administered 48 hours before OGD, reduced neuronal death by 20–50%, thereby providing neuroprotection [72]. In this model, daily administration of ceftriaxone 5 days prior to middle cerebral artery occlusion (MCAO) provided neuroprotection through a reduction in infarct volume as well as neuroinflammatory and apoptotic factors through EAAT2 upregulation [73]. Additionally, a study found that daily administration of ceftriaxone as well as N-acetylcysteine 5 days prior to focal cerebral ischemia also significantly increased EAAT2 expression levels in addition to a reduction of infarct volume [115]. Other compounds that protect neurons from glutamate-induced excitotoxic death, such as LDN-OSU0212320, decrease infarct volume in mice when administered 24 hours prior to photothrombotic ischemia; however, only in male mice, suggesting that this treatment may not be effective in the female population and further emphasizing the importance of sex differences in targeted treatments [107, 108]. A significant drawback associated with β-lactam antibiotics in enhancing EAAT2 expression is the extended time required for drug onset (≥24 hours) [169]. Therefore, in the context of clinical ischemic injury, more fast-acting compounds need to be developed. EAAT PAMs offer a different approach to restoring glutamate clearance through the enhancement of glutamate uptake; however, this hypothesis remains to be tested in animal models of stroke.

3.5 Traumatic brain injury

Studies on traumatic brain injury (TBI) in both humans and animals have shown an acute increase in tissue glutamate concentrations that persist at elevated levels for up to 5 days in humans. This suggests a delay or insufficient glutamate clearance by glutamate transporters following TBI [71, 170]. Many studies have established that the subsequent glutamate-mediated excitotoxicity plays a significant role in acute post-injury neurodegenerative events [171]. In addition, decreased GLT-1 expression was shown in several TBI preclinical studies [147], which is consistent with the decreased EAAT2 activity observed in TBI patients [171]. Furthermore, antisense knockdown of GLT-1 in rat aggravates neuronal damage following TBI [172].

TBI is a complex pathology of many etiologies, which varies depending on the severity of the injury and the location of the affected brain tissue. In cases of concussion, where brain injury is typically reversible and symptoms resolve over time [173], the opportunity to reduce acute glutamate excitotoxicity by targeting GLT-1 offers potential for neuroprotection. However, for other types of TBI, additional pharmacological approaches may be necessary to address the diverse pathophysiological mechanisms involved [174].

3.6 Epilepsy

Epilepsy is a group of disorders characterized by recurrent spontaneous seizures that appear to stem from intricate processes involving various neurotransmitter systems, including glutamate [175, 176]. This leads to an imbalance between neuronal excitatory and inhibitory activities, ultimately culminating in epileptogenesis [177]. Numerous drugs are available as anti-seizure medications; however, none of the current treatments are considered disease-modifying, as they only suppress seizures without addressing the development and progression of epilepsy.

As previously mentioned, a key preclinical study revealed that mice lacking GLT-1 are prone to exhibit seizures [145]. Additionally, conditional deletion of GLT-1 revealed that GLT-1 protects against fatal epilepsy [178]. Loss of GLAST or GLT-1 led to elevated extracellular glutamate levels, neurodegeneration, and progressive paralysis, and loss of EAAC1 caused mild neurotoxicity and resulted in epilepsy [149]. In patients with medial temporal lobe epilepsy (mTLE), increased extracellular glutamate levels were observed [179] as well as decreased levels of EAAT1, EAAT2, and EAAT3 [180], reviewed in Ref. [176]. Decreased levels of GLT-1 were also observed in animal models of epilepsy such as pilocarpine-induced [181], albumin-induced [182], tuberous sclerosis-induced [183], FeCl3-induced models [184], and chest compression-induced audiogenic model [185].

However, some studies did not observe changes in EAAT2 levels in patients [186] and in animal models including kindling [187], anticonvulsant ketogenic diet [188], and spontaneously epileptic rats [189], suggesting that EAAT2 may be implicated in the etiology of only some types of epilepsy. Another concept suggests that a deficiency in glutamine synthetase in astrocytes may be the molecular mechanism underlying extracellular glutamate accumulation and seizure generation [190].

Nonetheless, the upregulation of EAAT2 via transcriptional and translational regulation has demonstrated success in vivo by reducing spontaneous recurrent seizures and providing neuroprotection [191].

Parawixin10, a compound isolated from Parawixia bistriata spider venom, is a non-selective PAM of EAAT1 and EAAT2, with in vitro neuroprotective properties [47] and in vivo neuroprotection in epilepsy models of intrahippocampal injection of NMDA [48], kainic acid and pentylenetetrazol (PTZ) [49], and pilocarpine [50]. These studies served as proof of concept that EAAT1-2 PAMs may offer neuroprotection and anticonvulsant properties. More recently, the selective EAAT2 PAM [(R)-AS-1] revealed favorable anticonvulsant and safety profiles, and significant protection in mice against seizures in acute and chronic animal models of seizures [51]. These novel compounds may have disease-modifying potential in acquired epilepsy, which will require future experimental testing.

Several studies demonstrated mutations in EAAT1 are implicated in episodic ataxia 6 (EA6), a chronic condition characterized by epilepsy, nystagmus, and tinnitus. These mutations result in impaired glutamate uptake and alterations in anion conductance; however, how exactly these mutations affect EAAT1 expression, subcellular localization, function, and the complex neurological phenotype of EA6 remains to be understood [25].

Collectively, these studies suggest that astrocytic glutamate uptake plays a critical role in protecting neurons from hyperexcitability. However, discrepancies in some findings indicate that the exact mechanisms remain elusive. Nevertheless, the idea that modulating astrocytic EAATs represents a potential therapeutic approach to provide neuroprotection, to prevent spontaneous recurrent seizures, and to halt epileptogenesis [192].

3.7 Pain

Pain, an aversive sensory experience arising from actual or perceived tissue damage, constitutes a physiologic response to noxious stimuli or disease, serving as a protective mechanism to prompt seeking care or preventing further harm [193]. Pain perception involves complex interactions among cellular and molecular components, including neurons, glia, glutamate receptors, and transporters that utilize glutamate, the primary transmitter released by sensory afferents in the nervous system [194, 195, 196]. However, when pain extends beyond the acute injury phase, persisting for more than 3 months, it transitions into a chronic disease. Chronic pain, a prevalent motive for medical intervention, correlates with heightened risks of poor mental health, opioid dependency, and diminished quality of life.

During chronic pain development, excess glutamate released in the peripheral and central nervous system contributes to elevated extracellular glutamate levels, which overactivates glutamate receptors and exacerbates pain symptoms. Prolonged overactivation leads to neuroplasticity in pain pathways, amplifying pain signals, a phenomenon known as central sensitization [197]. Glutamate transporters, particularly EAAT2, counteract this process by removing extracellular glutamate and reducing pain signaling [196, 198]. In the pain pathway, EAAT2 is expressed in the anterior cingulate cortex, somatosensory cortex, hippocampus, and dorsal horn of the spinal cord [199].

After injury occurs, the expression of EAATs changes, potentially contributing to chronic pain conditions. Numerous pain models have been used to study EAATs during neuropathic pain development but with varying results. For example, in nerve-injury models of pain, studies found an initial increase in EAAT2 3–7 days after surgery, followed by a steep downregulation below baseline [200]. Other studies did not observe this initial upregulation in EAAT1 or EAAT2 early after injury and instead observed an upregulation of EAAT3 and a downregulation of EAAT1 and EAAT2 [101]. Because of these inconsistencies, the relationship between neuropathic pain development and EAAT expression remains unclear.

Various mechanisms are being explored for chronic pain therapy, and targeting the glutamatergic system to reduce pain transmission shows promising potential [196, 201]. Although glutamate receptor antagonism provides anti-nociception, it may be associated with severe side effects, including sedation, hallucinations, and cardiac instability, limiting its use in outpatient settings [202]. A potential novel therapeutic approach involves the modulation of astrocytic EAATs, located on cells that surround glutamatergic neurons, aiming to restore homeostasis by lowering the concentration of extracellular glutamate. In this sense, downregulation or inhibition of EAATs has been reported to increase pain [203], whereas increased EAAT2 expression can mitigate pain [204, 205, 206]. For example, overexpression of GLT-1 in the spinal cord attenuated the induction of inflammatory and neuropathic pain, suggesting this could be a strategy for pain [207]. Furthermore, positive allosteric modulation of EAATs can enhance their efficiency in various neurological diseases, offering neuroprotection, and is a potential target for chronic pain.

In preclinical pain models, certain drug classes have been shown to have anti-nociceptive properties by modulating EAATs through the enhancement of EAAT expression and removal of excessive glutamate [196, 203, 208]. Some of these drugs include β-lactam antibiotics (such as ceftriaxone), β-lactamase inhibitors (such as clavulanic acid), tetracycline antibiotics (such as minocycline), anticonvulsants (including VPA and riluzole), and tricyclic antidepressants (such as amitriptyline) [196]. Ceftriaxone has undergone extensive research as a potential therapy for modulating EAAT2 expression. Studies in the chronic constriction injury (CCI) pain model demonstrated that ceftriaxone upregulated GLT-1 expression and glutamate uptake in the spinal dorsal horn and resulted in anti-nociceptive effects [74, 75]. Furthermore, its impact extended to a model of multiple sclerosis, where ceftriaxone not only reversed tactile allodynia but also halted the progression of motor weakness and paralysis. Notably, in both models, ceftriaxone reversed the reduction in EAAT2 expression and astrocyte activation in the lumbar spinal cord, suggesting its potential to suppress glial activation and alleviate pain [75].

Concerns about the antibiotic activity of ceftriaxone led to the investigation of clavulanic acid, which is devoid of antibiotic properties. In the CCI model, both clavulanic acid and ceftriaxone demonstrated anti-nociception, increased EAAT2 expression in the rat spinal cord, and reversed EAAT2 downregulation at day 14 post-surgery [209]. Clavulanic acid may be a candidate for relieving pain in diabetic peripheral neuropathy, and its benefits could be attributed to increased EAAT2 expression [101].

Another approach to modulate EAAT2 that could be beneficial for various pain conditions is through positive allosteric modulation (PAM), which involves the enhancement of extracellular glutamate uptake. Several preclinical pain studies demonstrated a downregulation of EAAT2, thus enhancing the activity of the remaining transporters via PAM represents a promising new avenue for future drug development [196, 210]. Selective EAAT2 PAMs will restore glutamate homeostasis by enhancing the uptake of excessive glutamate, rather than blocking glutamatergic transmission; thus, this approach is expected to be safer and devoid of the dissociative effects observed with NMDA receptor antagonists [196]. Additionally, EAAT2 PAM compounds differ significantly from transcriptional or translational upregulation of EAAT2 expression, such as ceftriaxone, which display varying efficacies across individuals, raising concerns about safety, efficacy, and adverse effects. EAAT2 PAMs work directly on the transporter, rapidly increasing glutamate uptake efficiency, resulting in a quicker onset compared to drugs that modulate expression. They are highly selective, providing a more robust and safe clinical profile compared to epigenetic-modifying expression enhancers that are often associated with adverse side effects [196, 211]; however, further studies are needed to confirm these expectations.

3.8 Substance use disorders

The maintenance of low extracellular glutamate levels in the CNS is important for proper cognitive functions such as learning and memory [212]. Both learning and memory are involved in the cycle of addiction, as is evident by relapses [213]. Substance use disorder is characterized by compulsion to use a substance, inability to limit intake of that substance, and the appearance of negative affect following the use of the drug. Regions such as the prefrontal cortex, amygdala, and hippocampus, which are involved in learning and memory, have projections connecting them to key regions involved in addiction, such as the ventral tegmental area and nucleus accumbens [214]. Drug use activates glutamatergic neurotransmission in the mesocorticolimbic system [215]. Following repeated use, drug-specific changes occur in these brain regions that drive the susceptibility to relapse and chronic drug use [216]. Furthermore, in abuse models, dysregulation of glutamate levels is observed in key brain regions associated with addiction, such as the prefrontal cortex, hippocampus, and nucleus accumbens [217]. All drugs of abuse disrupt glutamate homeostasis during use, increasing synaptic glutamate release. Numerous studies have demonstrated the importance of glutamatergic signaling in the nucleus accumbens on drug-seeking [218]. The importance of changes in glutamatergic tone that occur in addiction is further evident as pharmacologically restoring glutamate levels reduces drug-seeking behavior [219].

Targeting EAAT2 specifically has displayed promising results in the context of addiction. The use of drugs such as cocaine, cannabinoids, amphetamine, ethanol, nicotine, and opiates results in significant changes in EAAT2 levels [216]. Preclinical studies have shown that ceftriaxone inhibits cue- and drug-induced reinstatement seeking for cocaine [220], heroin [221], methamphetamine [222], and nicotine [223]. Additionally, administration of ceftriaxone attenuates the development of cocaine-induced conditioned place and prevents the decrease in EAAT2 expression in the nucleus accumbens [224]. Ceftriaxone also reduces alcohol intake [76]. Clavulanic acid reduces the reinforcing efficacy of cocaine and reduces cocaine-conditioned place preference [77]. Additionally, clavulanic acid reduced morphine-conditioned place preference, morphine-induced hypothermia, and locomotor sensitization [102]. Riluzole prevents cocaine reinstatement while restoring the expression of EAAT2 in the nucleus accumbens [225]. However, clinical trials found that riluzole was not effective for cocaine use disorder, when taken during the period of cocaine dependency [226]. Interestingly, riluzole has shown some promising effects in clinical trials for methamphetamine dependence, as it decreased symptoms, such as craving, withdrawal, and depression, experienced by men in an outpatient setting [227]. This study was supported by previous findings that riluzole reduces methamphetamine-induced locomotor sensitization [228]. Additionally, riluzole affects morphine- and amphetamine-conditioned place preference [229]. Riluzole also reduced ethanol self-administration in mice [230]. N-acetylcysteine (NAC), an antioxidant cystine pro-drug, is an over-the-counter supplement that suppresses NF-κB [116], a transcriptional regulator of EAAT2 [117]. While NAC has low oral availability, NAC has been shown to restore EAAT2 expression and to reduce drug-seeking/taking behavior for cocaine [118], nicotine [119], and ethanol [120].

Another promising approach for targeting EAAT2 in the context of addiction is by increasing the efficiency of the transport through PAMs. EAAT2 PAMs offer the unique advantage of having a quick onset, which may be of critical importance during craving.

3.9 Neurodegenerative disorders

Alzheimer’s disease (AD) is a progressive age-related neurodegenerative disorder characterized by abnormal deposition of fibrillar amyloid β (Aβ) protein, intracellular neurofibrillary tangles, oxidative damage, and tau protein hyperphosphorylation that contributes to neuronal dysfunction [231, 232, 233]. The aberrant glutamate stimulation that results in synaptic dysfunction has been proposed as one of several mechanisms of synaptic damage in AD. This is supported by studies reporting reduced GLT-1 expression and function in AD [234, 235] and post-mortem analysis of human brain tissue from patients with AD (as well as other neurodegenerative disorders), suggesting that EAAT2 loss or dysfunction could be an early trigger evolving into chronic reactive astrogliosis, oxidative stress, and neuronal death [236]. Also, a study demonstrated that Aβ1–42 prompts rapid GLT-1 mislocalization and internalization in astrocytes, reducing the rate of glutamate clearance [237]. Another study found that GLT-1 loss led to a compensatory increase in insulin-degrading enzyme activity in the liver, implicating partial GLT-1 loss in insulin/Akt signaling abnormalities observed in AD [238]. Further, a study suggested that decreased expression of glutamine synthetase but no changes in GLT-1 expression in a model of AD [239]. Thus, it is not yet entirely clear whether GLT-1/EAAT2 dysfunction plays a pathogenic role in AD and whether targeting this transporter may be developed as a neuroprotective strategy for the pathogenesis of AD.

Currently, available treatments for AD include acetylcholinesterase inhibitors [240], NMDA receptor antagonist memantine [241], and monoclonal antibody therapies that target the removal of β-amyloid from the brain such as lecanemab [242]. However, these treatments offer only symptomatic relief and primarily target late-stage aspects of the disease. Additionally, the efficacy of monoclonal antibodies has not been proven. Therefore, a definite treatment for this disease is yet to be identified.

Amyotrophic lateral sclerosis (ALS): ALS is a debilitating disease characterized by progressive loss of voluntary motor neurons, leading to muscle atrophy, weight loss, and respiratory failure [243]. The pathogenesis of ALS involves inflammation, oxidative stress, apoptosis, dysfunction of mitochondria, aggregation of SOD1 protein, and dysfunction of astroglia, which include a severe loss of EAAT2 in both the motor cortex and spinal cord [244, 245]. Additionally, several mouse models of ALS exhibit marked loss or inactivation of glutamate transporters [246]. Moreover, selective loss of EAAT2 has also been demonstrated in both sporadic and familial cases of ALS [247]. However, a large-scale clinical trial testing the efficacy of ceftriaxone (an EAAT2 expression enhancer) in ALS patients reported no significant difference in survival between placebo- and ceftriaxone-treated patients [248].

While glutamate transport dysfunction and excitotoxicity may contribute to the late stage of ALS disease progression, there is no consistent evidence supporting EAAT2 as a primary factor in motor neuron degeneration in ALS. A study found that overexpression of GLT-1 in the cervical spinal cord of SOD1G93A mice with ALS did not protect motor neurons, preserve diaphragm function, or prolong animal survival, thus challenging the notion that EAAT2 is a primary factor in motor neuron degeneration in ALS [249]. It is thought that simultaneous targeting of calcium overload, endoplasmic reticulum stress, and mitochondrial dysfunction pathways may be necessary to halt ALS progression [250].

Huntington’s disease (HD): HD is a devastating neurodegenerative disorder characterized by degeneration of multiple brain areas, involving dopamine, glutamate, and GABA neurotransmitter systems. It is caused by a mutated form of the huntingtin gene, resulting in the accumulation of mutant protein aggregates [251], resulting in oxidative stress, mitochondrial dysfunction, and excitotoxicity [252]. Reduced EAAT2 mRNA levels have been observed in HD patients [253], and transgenic mouse models, while increased GLT-1 expression can improve behavioral symptoms in HD mouse models [254]. Additionally, EAAT3 expression has been shown to be reduced in HD; however, this appears to be associated with issues with cysteine transport and oxidative stress rather than increases in extracellular glutamate concentration.

These studies suggest that alterations in EAAT2 and EAAT3 expression or function can influence the progression of Huntington’s disease (HD). However, effective therapies for HD have yet to be discovered and developed [71, 255, 256].

Human idiopathic Parkinson’s disease (PD): is a progressive neurodegenerative movement disorder characterized by the degeneration of dopaminergic neurons, leading to increased firing rates of glutamatergic excitatory projections to the substantia nigra [257]. This disturbance may contribute to glutamate-mediated excitotoxicity, exacerbating nigrostriatal degeneration in PD. Studies using unilateral 6-hydroxydopamine (6-OHDA) animal models of PD have demonstrated a link between disturbed glutamatergic neurotransmission and glutamate transporter functioning in the striatum. A study revealed posttranslational modifications on GLT-1, resulting in transporter trafficking by Nedd4-2 in a PD model. Additionally, rapamycin protects mice against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced loss of dopaminergic neurons in a PD model, through preservation of EAAT2, an effect mediated by NF-κB. Moreover, it was reported that increased expression of GLT-1 with ceftriaxone ameliorated locomotor impairments in a PD model [258].

Dyskinesias are a motor complication that develops as common side effect of current PD treatments such as L-DOPA (L-3,4-dihydroxyphenylalanine), which is used to replenish dopamine levels. Dyskinesias are linked to elevated extracellular glutamate levels in the basal ganglia, thus targeting EAAT2 modulation may represent a potential therapeutic target to treat them.

Collectively, it seems that dysregulation or loss of EAATs function, especially, EAAT2, can lead to glutamatergic excitotoxicity and neuronal death, contributing to neurodegenerative diseases such as AD, ALS, HD, and PD. However, our current understanding of the contribution of EAATs in these diseases is primarily based on experimental models and post-mortem brain tissue analysis. Detecting EAAT2 in living human brains could greatly improve diagnosis and therapy for these neurodegenerative disorders, and this could allow the possibility of utilizing EAAT2 activation for therapeutic interventions.

3.10 Other disorders

HIV-associated neurocognitive disorder (HAND): A common neuropathology observed in the brains of HIV-infected individuals is the excess release of glutamate upon HIV infection of macrophage/microglial cells [71, 259]. This has been linked to neurotoxicity mediated by various HIV proteins, including gp120 and transactivator of transcription (TAT). NMDA receptor antagonists were not effective at mitigating glutamate excitotoxicity in HAND due to side effects. Therefore, alternative approaches are being pursued, such as modulating the activity or expression of glutamate transporters. A study found that methamphetamine and HIV treatment activate trace amine-associated receptor 1 (TAAR1) in human astrocytes, leading to reduced EAAT2 mRNA levels and impaired glutamate clearance. CCL2 impairs spatial memory and cognition, potentially through upregulating mRNA expression linked to inflammation, excitotoxicity, and neuronal apoptosis in HAND. However, our understanding of the host factors contributing to the neurotoxic effects of HIV-1 on the CNS is evolving, and identification of strategies to mitigate the neurotoxic effects of viral and host proteins is crucial to develop neuroprotection strategies to alleviate the detrimental impact of HIV-1 on the brain.

Autism: The etiology of autism spectrum disorder is complex and involves genetic predisposition, environmental influences, and other yet unknown factors. It is thought that glutamate excitotoxicity, mitochondrial dysfunction, and degeneration are key components of autism. A report by the Autism Genome Project Consortium identified a linkage peak for autism in the region of chromosome 11, where the gene for EAAT2 is situated [260]. There is evidence suggesting astroglial dysfunction in the autistic brain and activators of GLT-1 expression have been shown to ameliorate certain symptoms of autism and reduce epilepsy seizures, suggesting that GLT-1 may be a novel therapeutic strategy for autism.

Additionally, EAAT2 dysfunction has been implicated in the pathogenesis of major depressive disorders [261], mood disorders [262, 263], glioma [264, 265], multiple sclerosis [266], and schizophrenia [267], among others. Dysfunction of EAAT3 has also been observed in schizophrenia [13] and OCD [268].

Collectively, these studies suggest that therapies aimed at the modulation of the function and/or expression of EAATs, particularly EAAT2, could have broad therapeutic potential for various CNS disorders.

Advertisement

4. Conclusions

EAATs are key proteins that regulate the excitatory tone in CNS and are important for many physiological functions. Dysfunction in their activity or expression has profound effects that have been implicated in the etiology of many acute and chronic pathologies. In the past decade, it has become evident that drugs targeting NMDA receptors and secondary damages from glutamate-mediated excitotoxicity are limited and ineffective, often resulting in unwanted side effects; thus, there is a need for better therapeutics. In this regard, enhancing glutamate transporter expression or function pharmacologically holds great promise for therapeutic interventions, particularly targeting EAAT2 [71, 269].

Drug development targeting EAATs started with the discovery of pharmacological agents that were developed to study the intrinsic properties and function of the EAATs, specifically the potent EAAT2 inhibitors TBOA and analogs [68, 71]. TBOA was the first non-transportable blocker for all subtypes of EAATs identified, which helped elucidate several functions of the EAATs, and encouraged the search for other modulators of the function of EAATs. However, there is also a need to identify subtype-selective enhancers and inhibitors for all subtypes of glutamate transporters, to fully understand how to fine-tune the extracellular concentration of glutamate in the CNS. Several EAAT2 PAMs have already been identified, which can be administered acutely and thus are advantageous over the expression enhancers that generally require prophylactic administration. It is yet to be established whether small molecule allosteric activators of EAAT2, like many biologics, can undergo traditional pharmacokinetic analysis and demonstrate efficacy in animal models of CNS disease. The future may uncover whether this class of compounds is effective in chronic conditions and when used in combination therapies. Moreover, the recent publication of cryo-EM structures of EAAT2 presents an opportunity to launch a structural-based drug design initiative aimed at screening and developing high-affinity EAAT2 PAMs. These compounds can have therapeutic potential and serve as imaging tools to detect changes in EAAT2 density in neurodegenerative diseases. Furthermore, the optimal timing for therapeutic intervention with EAAT2-targeting drugs remains uncertain. Hence, discovering noninvasive methods to enhance our understanding of EAAT2 function and expression in the living brain is imperative. Finally, there is still more to understand about the molecular behavior of glutamate transporters in dysfunctions. Current understanding suggests that a complex process is involved in the downregulation, reversal, and post-translational modifications of EAATs. The intracellular signaling pathways that accompany the changes in EAATs on disease states also need to be further understood.

In conclusion, EAAT dysfunction plays a significant role in many neuropsychiatric disorders. Recent advancements have propelled the pursuit of targeting these transporters as a strategy to prevent and treat glutamatergic dysfunctions. However, the full potential of modulating these transporters as a therapeutic target is yet to be fully understood, appreciated, and explored. We expect to see, in the upcoming years, the identification of an arsenal of selective pharmacological compounds that target these fascinating proteins and the investigation of their translational possibilities [71, 210].

Advertisement

Acknowledgments

ACKF received funding from NIH (RO1 NS111767), and KLR received financial support from the PhRMA foundation.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Olney JW. Brain lesions, obesity, and other disturbances in mice treated with monosodium glutamate. Science. 1969;164(3880):719-721
  2. 2. Danbolt NC. Glutamate uptake. Progress in Neurobiology. 2001;65(1):1-105
  3. 3. Lewerenz J, Hewett SJ, Huang Y, Lambros M, Gout PW, Kalivas PW, et al. The cystine/glutamate antiporter system x(c)(−) in health and disease: From molecular mechanisms to novel therapeutic opportunities. Antioxidants & Redox Signaling. 2013;18(5):522-555
  4. 4. El Mestikawy S, Wallén-Mackenzie A, Fortin GM, Descarries L, Trudeau LE. From glutamate co-release to vesicular synergy: Vesicular glutamate transporters. Nature Reviews Neuroscience. 2011;12(4):204-216
  5. 5. Roberts RC, Roche JK, McCullumsmith RE. Localization of excitatory amino acid transporters EAAT1 and EAAT2 in human postmortem cortex: A light and electron microscopic study. Neuroscience. 2014;277:522-540
  6. 6. Pajarillo E, Rizor A, Lee J, Aschner M, Lee E. The role of astrocytic glutamate transporters GLT-1 and GLAST in neurological disorders: Potential targets for neurotherapeutics. Neuropharmacology. 2019;161:107559-107573
  7. 7. Storck T, Schulte S, Hofmann K, Stoffel W. Structure, expression, and functional analysis of a Na(+)-dependent glutamate/aspartate transporter from rat brain. Proceedings of the National Academy of Sciences of the United States of America. 1992;89(22):10955-10959
  8. 8. Suchak SK, Baloyianni NV, Perkinton MS, Williams RJ, Meldrum BS, Rattray M. The 'glial' glutamate transporter, EAAT2 (Glt-1) accounts for high affinity glutamate uptake into adult rodent nerve endings. Journal of Neurochemistry. 2003;84(3):522-532
  9. 9. Pines G, Danbolt NC, Bjoras M, Zhang Y, Bendahan A, Eide L, et al. Cloning and expression of a rat brain L-glutamate transporter. Nature. 1992;360(6403):464-467
  10. 10. Haugeto O, Ullensvang K, Levy LM, Chaudhry FA, Honore T, Nielsen M, et al. Brain glutamate transporter proteins form homomultimers. The Journal of Biological Chemistry. 1996;271(44):27715-27722
  11. 11. Gegelashvili G, Danbolt NC, Schousboe A. Neuronal soluble factors differentially regulate the expression of the GLT1 and GLAST glutamate transporters in cultured astroglia. Journal of Neurochemistry. 1997;69(6):2612-2615
  12. 12. Peacey E, Miller CC, Dunlop J, Rattray M. The four major N- and C-terminal splice variants of the excitatory amino acid transporter GLT-1 form cell surface homomeric and heteromeric assemblies. Molecular Pharmacology. 2009;75(5):1062-1073
  13. 13. Bjorn-Yoshimoto WE, Underhill SM. The importance of the excitatory amino acid transporter 3 (EAAT3). Neurochemistry International. 2016;98:4-18
  14. 14. Perkins EM, Clarkson YL, Suminaite D, Lyndon AR, Tanaka K, Rothstein JD, et al. Loss of cerebellar glutamate transporters EAAT4 and GLAST differentially affects the spontaneous firing pattern and survival of Purkinje cells. Human Molecular Genetics. 2018;27(15):2614-2627
  15. 15. Arriza JL, Eliasof S, Kavanaugh MP, Amara SG. Excitatory amino acid transporter 5, a retinal glutamate transporter coupled to a chloride conductance. Proceedings of the National Academy of Sciences of the United States of America. 1997;94(8):4155-4160
  16. 16. Alleva C, Machtens JP, Kortzak D, Weyand I, Fahlke C. Molecular basis of coupled transport and anion conduction in excitatory amino acid transporters. Neurochemical Research. 2022;47(1):9-22
  17. 17. Kosugi T, Kawahara K. Reversed actrocytic GLT-1 during ischemia is crucial to excitotoxic death of neurons, but contributes to the survival of astrocytes themselves. Neurochemical Research. 2006;31(7):933-943
  18. 18. Ewers D, Becher T, Machtens JP, Weyand I, Fahlke C. Induced fit substrate binding to an archeal glutamate transporter homologue. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(30):12486-12491
  19. 19. Rose EM, Koo JC, Antflick JE, Ahmed SM, Angers S, Hampson DR. Glutamate transporter coupling to Na, K-ATPase. The Journal of Neuroscience. 2009;29(25):8143-8155
  20. 20. Reyes N, Ginter C, Boudker O. Transport mechanism of a bacterial homologue of glutamate transporters. Nature. 2009;462(7275):880-885
  21. 21. Akyuz N, Altman RB, Blanchard SC, Boudker O. Transport dynamics in a glutamate transporter homologue. Nature. 2013;502(7469):114-118
  22. 22. Machtens JP, Kortzak D, Lansche C, Leinenweber A, Kilian P, Begemann B, et al. Mechanisms of anion conduction by coupled glutamate transporters. Cell. 2015;160(3):542-553
  23. 23. Bergles DE, Tzingounis AV, Jahr CE. Comparison of coupled and uncoupled currents during glutamate uptake by GLT-1 transporters. The Journal of Neuroscience. 2002;22(23):10153
  24. 24. Billups B, Rossi D, Attwell D. Anion conductance behavior of the glutamate uptake carrier in salamander retinal glial cells. The Journal of Neuroscience. 1996;16(21):6722-6731
  25. 25. Winter N, Kovermann P, Fahlke C. A point mutation associated with episodic ataxia 6 increases glutamate transporter anion currents. Brain. 2012;135(Pt 11):3416-3425
  26. 26. Canul-Tec JC, Assal R, Cirri E, Legrand P, Brier S, Chamot-Rooke J, et al. Structure and allosteric inhibition of excitatory amino acid transporter 1. Nature. 2017;544(7651):446-451
  27. 27. Qiu B, Matthies D, Fortea E, Yu Z, Boudker O. Cryo-EM structures of excitatory amino acid transporter 3 visualize coupled substrate, sodium, and proton binding and transport. Science Advances. 2021;7(10):eabf5814
  28. 28. Kato T, Kusakizako T, Jin C, Zhou X, Ohgaki R, Quan L, et al. Structural insights into inhibitory mechanism of human excitatory amino acid transporter EAAT2. Nature Communications. 2022;13(1):4714
  29. 29. Zhang Z, Chen H, Geng Z, Yu Z, Li H, Dong Y, et al. Structural basis of ligand binding modes of human EAAT2. Nature Communications. 2022;13(1):3329
  30. 30. Zhang C, Raghupathi R, Saatman KE, Smith DH, Stutzmann JM, Wahl F, et al. Riluzole attenuates cortical lesion size, but not hippocampal neuronal loss, following traumatic brain injury in the rat. Journal of Neuroscience Research. 1998;52(3):342-349
  31. 31. Azbill RD, Mu X, Springer JE. Riluzole increases high-affinity glutamate uptake in rat spinal cord synaptosomes. Brain Research. 2000;871(2):175-180
  32. 32. Mu X, Azbill RD, Springer JE. Riluzole and methylprednisolone combined treatment improves functional recovery in traumatic spinal cord injury. Journal of Neurotrauma. 2000;17(9):773-780
  33. 33. Brothers HM, Bardou I, Hopp SC, Kaercher RM, Corona AW, Fenn AM, et al. Riluzole partially rescues age-associated, but not LPS-induced, loss of glutamate transporters and spatial memory. Journal of Neuroimmune Pharmacology. 2013;8(5):1098-1105
  34. 34. Miller RG, Mitchell JD, Moore DH. Riluzole for amyotrophic lateral sclerosis (ALS)/motor neuron disease (MND). The Cochrane Database of Systematic Reviews. 2000;2(2):CD001447
  35. 35. Frizzo ME, Dall'Onder LP, Dalcin KB, Souza DO. Riluzole enhances glutamate uptake in rat astrocyte cultures. Cellular and Molecular Neurobiology. 2004;24(1):123-128
  36. 36. Fumagalli E, Funicello M, Rauen T, Gobbi M, Mennini T. Riluzole enhances the activity of glutamate transporters GLAST, GLT1 and EAAC1. European Journal of Pharmacology. 2008;578(2-3):171-176
  37. 37. Banasr M, Chowdhury GM, Terwilliger R, Newton SS, Duman RS, Behar KL, et al. Glial pathology in an animal model of depression: Reversal of stress-induced cellular, metabolic and behavioral deficits by the glutamate-modulating drug riluzole. Molecular Psychiatry. 2010;15(5):501-511
  38. 38. Carbone M, Duty S, Rattray M. Riluzole elevates GLT-1 activity and levels in striatal astrocytes. Neurochemistry International. 2012;60(1):31-38
  39. 39. Wilson JR, Fehlings MG. Riluzole for acute traumatic spinal cord injury: A promising neuroprotective treatment strategy. World Neurosurgery. 2014;81(5-6):825-829
  40. 40. Waibel S, Reuter A, Malessa S, Blaugrund E, Ludolph AC. Rasagiline alone and in combination with riluzole prolongs survival in an ALS mouse model. Journal of Neurology. 2004;251(9):1080-1084
  41. 41. Del Signore SJ, Amante DJ, Kim J, Stack EC, Goodrich S, Cormier K, et al. Combined riluzole and sodium phenylbutyrate therapy in transgenic amyotrophic lateral sclerosis mice. Amyotrophic Lateral Sclerosis. 2009;10(2):85-94
  42. 42. Whitcomb DJ, Molnar E. Is riluzole a new drug for Alzheimer's disease? Journal of Neurochemistry. 2015;135(2):207-209
  43. 43. Al-Horani RA. Riluzole and its prodrugs for the treatment of Alzheimer's disease. Pharmaceutical Patent Analyst. 2023;12(2):79-85
  44. 44. Fontana AC, Guizzo R, de Oliveira BR, Meirelles ESAR, Coimbra NC, Amara SG, et al. Purification of a neuroprotective component of Parawixia bistriata spider venom that enhances glutamate uptake. British Journal of Pharmacology. 2003;139(7):1297-1309
  45. 45. Duffield M, Patel A, Mortensen OV, Schnur D, Gonzalez-Suarez AD, Torres-Salazar D, et al. Transport rate of EAAT2 is regulated by amino acid located at the interface between the scaffolding and substrate transport domains. Neurochemistry International. 2020;139:104792
  46. 46. Kortagere S, Mortensen OV, Xia J, Lester W, Fang Y, Srikanth YVV, et al. Identification of novel allosteric modulators of glutamate transporter EAAT2. ACS Chemical Neuroscience. 2017;9(3):522-534
  47. 47. Forster YM, Green JL, Khatiwada A, Liberato JL, Reddy PAN, Salvino JM, et al. Elucidation of the structure and synthesis of neuroprotective low molecular mass components of the Parawixia bistriata spider venom. ACS Chemical Neuroscience. 2020;11(11):1573-1596
  48. 48. Fachim HA, Mortari MR, Gobbo-Netto L, Dos Santos WF. Neuroprotective activity of parawixin 10, a compound isolated from Parawixia bistriata spider venom (Araneidae: Araneae) in rats undergoing intrahippocampal NMDA microinjection. Pharmacognosy Magazine. 2015;11(43):579-585
  49. 49. Fachim HA, Cunha AO, Pereira AC, Beleboni RO, Gobbo-Neto L, Lopes NP, et al. Neurobiological activity of Parawixin 10, a novel anticonvulsant compound isolated from Parawixia bistriata spider venom (Araneidae: Araneae). Epilepsy & Behavior. 2011;22(2):158-164
  50. 50. Liberato JL, Godoy LD, Mortari MR, Gobbo-Neto L, Lopes NP, Santos WF. Parawixin10: A new natural compound from Parawixia bistriata spider venom that presents neuroprotective, memory-saving, and disease-modifying effects in the pilocarpine model of TLE in Wistar rats. Epilepsy & Behavior. 2014;38:196
  51. 51. Abram M, Jakubiec M, Reeb K, Cheng MH, Gedschold R, Rapacz A, et al. Discovery of (R)-N-benzyl-2-(2,5-dioxopyrrolidin-1-yl)propanamide [(R)-AS-1], a novel orally bioavailable EAAT2 modulator with drug-like properties and potent antiseizure activity in vivo. Journal of Medicinal Chemistry. 2022;65(17):11703-11725
  52. 52. O'Shea RD, Fodera MV, Aprico K, Dehnes Y, Danbolt NC, Crawford D, et al. Evaluation of drugs acting at glutamate transporters in organotypic hippocampal cultures: New evidence on substrates and blockers in excitotoxicity. Neurochemical Research. 2002;27(1-2):5-13
  53. 53. Lebrun B, Sakaitani M, Shimamoto K, Yasuda-Kamatani Y, Nakajima T. New beta-hydroxyaspartate derivatives are competitive blockers for the bovine glutamate/aspartate transporter. The Journal of Biological Chemistry. 1997;272(33):20336-20339
  54. 54. Shimamoto K. Glutamate transporter blockers for elucidation of the function of excitatory neurotransmission systems. Chemical Record. 2008;8(3):182-199
  55. 55. Shigeri Y, Seal RP, Shimamoto K. Molecular pharmacology of glutamate transporters, EAATs and VGLUTs. Brain Research. Brain Research Reviews. 2004;45(3):250-265
  56. 56. Shimamoto K, Lebrun B, Yasuda-Kamatani Y, Sakaitani M, Shigeri Y, Yumoto N, et al. DL-threo-beta-benzyloxyaspartate, a potent blocker of excitatory amino acid transporters. Molecular Pharmacology. 1998;53(2):195-201
  57. 57. Izumi Y, Shimamoto K, Benz AM, Hammerman SB, Olney JW, Zorumski CF. Glutamate transporters and retinal excitotoxicity. GLIA. 2002;39(1):58-68
  58. 58. Bonde C, Noraberg J, Noer H, Zimmer J. Ionotropic glutamate receptors and glutamate transporters are involved in necrotic neuronal cell death induced by oxygen-glucose deprivation of hippocampal slice cultures. Neuroscience. 2005;136(3):779-794
  59. 59. 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. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(15):8733-8738
  60. 60. Vandenberg RJ, Ryan RM. Mechanisms of glutamate transport. Physiological Reviews. 2013;93(4):1621-1657
  61. 61. Shimamoto K, Sakai R, Takaoka K, Yumoto N, Nakajima T, Amara SG, et al. Characterization of novel L-threo-beta-benzyloxyaspartate derivatives, potent blockers of the glutamate transporters. Molecular Pharmacology. 2004;65(4):1008-1015
  62. 62. Bozzo L, Chatton JY. Inhibitory effects of (2S, 3S)-3-[3-[4-(trifluoromethyl)benzoylamino]benzyloxy]aspartate (TFB-TBOA) on the astrocytic sodium responses to glutamate. Brain Research. 2010;1316:27-34
  63. 63. Tsukada S, Iino M, Takayasu Y, Shimamoto K, Ozawa S. Effects of a novel glutamate transporter blocker, (2S, 3S)-3-[3-[4-(trifluoromethyl)benzoylamino]benzyloxy]aspartate (TFB-TBOA), on activities of hippocampal neurons. Neuropharmacology. 2005;48(4):479-491
  64. 64. Martinez D, Rogers RC, Hermann GE, Hasser EM, Kline DD. Astrocytic glutamate transporters reduce the neuronal and physiological influence of metabotropic glutamate receptors in nucleus tractus solitarii. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology. 2020;318(3):R545-Rr64
  65. 65. Fontana ACK, Poli ANR, Gour J, Srikanth YVV, Anastasi N, Ashok D, et al. Synthesis and structure-activity relationships for glutamate transporter allosteric modulators. Journal of Medicinal Chemistry. 2024;67(8):6119-6143
  66. 66. McIntosh TK, Smith DH, Voddi M, Perri BR, Stutzmann JM. Riluzole, a novel neuroprotective agent, attenuates both neurologic motor and cognitive dysfunction following experimental brain injury in the rat. Journal of Neurotrauma. 1996;13(12):767-780
  67. 67. Abdel-Magid AF. Allosteric modulators: An emerging concept in drug discovery. ACS Medicinal Chemistry Letters. 2015;6(2):104-107
  68. 68. Bridges RJ, Kavanaugh MP, Chamberlin AR. A pharmacological review of competitive inhibitors and substrates of high-affinity, sodium-dependent glutamate transport in the central nervous system. Current Pharmaceutical Design. 1999;5(5):363-379
  69. 69. Niello M, Gradisch R, Loland CJ, Stockner T, Sitte HH. Allosteric modulation of neurotransmitter transporters as a therapeutic strategy. Trends in Pharmacological Sciences. 2020;41(7):446-463
  70. 70. Anderson CM, Swanson RA. Astrocyte glutamate transport: Review of properties, regulation, and physiological functions. GLIA. 2000;32(1):1-14
  71. 71. Fontana AC. Current approaches to enhance glutamate transporter function and expression. Journal of Neurochemistry. 2015;134(6):982-1007
  72. 72. Rothstein JD, Patel S, Regan MR, Haenggeli C, Huang YH, Bergles DE, et al. Beta-lactam antibiotics offer neuroprotection by increasing glutamate transporter expression. Nature. 2005;433(7021):73-77
  73. 73. Chu K, Lee ST, Sinn DI, Ko SY, Kim EH, Kim JM, et al. Pharmacological induction of ischemic tolerance by glutamate transporter-1 (EAAT2) upregulation. Stroke; a Journal of Cerebral Circulation. 2007;38(1):177-182
  74. 74. Hu Y, Li W, Lu L, Cai J, Xian X, Zhang M, et al. An anti-nociceptive role for ceftriaxone in chronic neuropathic pain in rats. Pain. 2010;148(2):284-301
  75. 75. Ramos KM, Lewis MT, Morgan KN, Crysdale NY, Kroll JL, Taylor FR, et al. Spinal upregulation of glutamate transporter GLT-1 by ceftriaxone: Therapeutic efficacy in a range of experimental nervous system disorders. Neuroscience. 2010;169(4):1888-1900
  76. 76. 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;295:164-174
  77. 77. Kim J, John J, Langford D, Walker E, Ward S, Rawls SM. Clavulanic acid enhances glutamate transporter subtype I (GLT-1) expression and decreases reinforcing efficacy of cocaine in mice. Amino Acids. 2016;48(3):689-696
  78. 78. Melzer N, Meuth SG, Torres-Salazar D, Bittner S, Zozulya AL, Weidenfeller C, et al. A beta-lactam antibiotic dampens excitotoxic inflammatory CNS damage in a mouse model of multiple sclerosis. PLoS One. 2008;3(9):e3149
  79. 79. Rawls SM, Karaca F, Madhani I, Bhojani V, Martinez RL, Abou-Gharbia M, et al. β-lactamase inhibitors display anti-seizure properties in an invertebrate assay. Neuroscience. 2010;169(4):1800-1804
  80. 80. Rao PS, Saternos H, Goodwani S, Sari Y. Effects of ceftriaxone on GLT1 isoforms, xCT and associated signaling pathways in P rats exposed to ethanol. Psychopharmacology. 2015;232(13):2333-2342
  81. 81. Lipski J, Wan CK, Bai JZ, Pi R, Li D, Donnelly D. Neuroprotective potential of ceftriaxone in in vitro models of stroke. Neuroscience. 2007;146(2):617-629
  82. 82. Thone-Reineke C, Neumann C, Namsolleck P, Schmerbach K, Krikov M, Schefe JH, et al. The beta-lactam antibiotic, ceftriaxone, dramatically improves survival, increases glutamate uptake and induces neurotrophins in stroke. Journal of Hypertension. 2008;26(12):2426-2435
  83. 83. Pan XD, Wei J, Xiao GM. Effects of beta-lactam antibiotics ceftriaxone on expression of glutamate in hippocampus after traumatic brain injury in rats. Zhejiang da xue xue bao Yi xue ban = Journal of Zhejiang University Medical sciences. 2011;40(5):522-526
  84. 84. Wei J, Pan X, Pei Z, Wang W, Qiu W, Shi Z, et al. The beta-lactam antibiotic, ceftriaxone, provides neuroprotective potential via anti-excitotoxicity and anti-inflammation response in a rat model of traumatic brain injury. Journal of Trauma and Acute Care Surgery. 2012;73(3):654-660
  85. 85. Cui C, Cui Y, Gao J, Sun L, Wang Y, Wang K, et al. Neuroprotective effect of ceftriaxone in a rat model of traumatic brain injury. Neurological Sciences. 2014;35(5):695-700
  86. 86. Goodrich GS, Kabakov AY, Hameed MQ , Dhamne SC, Rosenberg PA, Rotenberg A. Ceftriaxone treatment after traumatic brain injury restores expression of the glutamate transporter, GLT-1, reduces regional gliosis, and reduces post-traumatic seizures in the rat. Journal of Neurotrauma. 2013;30(16):1434-1441
  87. 87. Leung TC, Lui CN, Chen LW, Yung WH, Chan YS, Yung KK. Ceftriaxone ameliorates motor deficits and protects dopaminergic neurons in 6-hydroxydopamine-lesioned rats. ACS Chemical Neuroscience. 2012;3(1):22-30
  88. 88. Kelsey JE, Neville C. The effects of the beta-lactam antibiotic, ceftriaxone, on forepaw stepping and L-DOPA-induced dyskinesia in a rodent model of Parkinson's disease. Psychopharmacology. 2014;231(12):2405-2415
  89. 89. Miller BR, Dorner JL, Shou M, Sari Y, Barton SJ, Sengelaub DR, et al. Up-regulation of GLT1 expression increases glutamate uptake and attenuates the Huntington's disease phenotype in the R6/2 mouse. Neuroscience. 2008;153(1):329-337
  90. 90. Rebec GV. Dysregulation of corticostriatal ascorbate release and glutamate uptake in transgenic models of Huntington's disease. Antioxidants & Redox Signaling. 2013;19(17):2115-2128
  91. 91. Feng D, Wang W, Dong Y, Wu L, Huang J, Ma Y, et al. Ceftriaxone alleviates early brain injury after subarachnoid hemorrhage by increasing excitatory amino acid transporter 2 expression via the PI3K/Akt/NF-kappaB signaling pathway. Neuroscience. 2014;268:21-32
  92. 92. Hu YY, Xu J, Zhang M, Wang D, Li L, Li WB. Ceftriaxone modulates uptake activity of glial glutamate transporter-1 against global brain ischemia in rats. Journal of Neurochemistry. 2015;132(2):194-205
  93. 93. Jagadapillai R, Mellen NM, Sachleben LR Jr, Gozal E. Ceftriaxone preserves glutamate transporters and prevents intermittent hypoxia-induced vulnerability to brain excitotoxic injury. PLoS One. 2014;9(7):e100230
  94. 94. Beghi E, Bendotti C, Mennini T. New ideas for therapy in ALS: Critical considerations. Amyotrophic Lateral Sclerosis. 2006;7(2):126-127; discussion 7
  95. 95. Nederkoorn PJ, Westendorp WF, Hooijenga IJ, de Haan RJ, Dippel DW, Vermeij FH, et al. Preventive antibiotics in stroke study: Rationale and protocol for a randomised trial. International Journal of Stroke. 2011;6(2):159-163
  96. 96. Berry JD, Shefner JM, Conwit R, Schoenfeld D, Keroack M, Felsenstein D, et al. Design and initial results of a multi-phase randomized trial of ceftriaxone in amyotrophic lateral sclerosis. PLoS One. 2013;8(4):e61177
  97. 97. Cudkowicz M, Shefner J, Consortium N. STAGE 3 clinical trial of ceftriaxone in subjects with ALS (S36.001). Neurology. 2013;80:S36.001
  98. 98. Hota SK, Barhwal K, Ray K, Singh SB, Ilavazhagan G. Ceftriaxone rescues hippocampal neurons from excitotoxicity and enhances memory retrieval in chronic hypobaric hypoxia. Neurobiology of Learning and Memory. 2008;89(4):522-532
  99. 99. Lee SG, Su ZZ, Emdad L, Gupta P, Sarkar D, Borjabad A, et al. Mechanism of ceftriaxone induction of excitatory amino acid transporter-2 expression and glutamate uptake in primary human astrocytes. The Journal of Biological Chemistry. 2008;283(19):13116-13123
  100. 100. Verma R, Mishra V, Sasmal D, Raghubir R. Pharmacological evaluation of glutamate transporter 1 (GLT-1) mediated neuroprotection following cerebral ischemia/reperfusion injury. European Journal of Pharmacology. 2010;638(1-3):65-71
  101. 101. Kolahdouz M, Jafari F, Falanji F, Nazemi S, Mohammadzadeh M, Molavi M, et al. Clavulanic acid attenuating effect on the diabetic neuropathic pain in rats. Neurochemical Research. 2021;46(7):1759-1770
  102. 102. Schroeder JA, Tolman NG, McKenna FF, Watkins KL, Passeri SM, Hsu AH, et al. Clavulanic acid reduces rewarding, hyperthermic and locomotor-sensitizing effects of morphine in rats: A new indication for an old drug? Drug and Alcohol Dependence. 2014;142:41-45
  103. 103. Philogene-Khalid HL, Morrison MF, Darbinian N, Selzer ME, Schroeder J, Rawls SM. The GLT-1 enhancer clavulanic acid suppresses cocaine place preference behavior and reduces GCPII activity and protein levels in the rat nucleus accumbens. Drug and Alcohol Dependence. 2022;232:109306
  104. 104. Ochoa-Aguilar A, Ventura- Martinez R, Sotomayor-Sobrino MA, Jaimez R, Coffeen U, Jiménez- González A, et al. Ceftriaxone and clavulanic acid induce antiallodynia and anti-inflammatory effects in rats using the carrageenan model. Journal of Pain Research. 2018;11:977-985
  105. 105. Kristensen PJ, Gegelashvili G, Munro G, Heegaard AM, Bjerrum OJ. The β-lactam clavulanic acid mediates glutamate transport-sensitive pain relief in a rat model of neuropathic pain. European Journal of Pain. 2018;22(2):282-294
  106. 106. Hajhashemi V, Dehdashti K. Antinociceptive effect of clavulanic acid and its preventive activity against development of morphine tolerance and dependence in animal models. Results in Pharma Sciences. 2014;9(5):315-321
  107. 107. Kong Q , Chang L-C, Takahashi K, Liu Q , Schulte DA, Lai L, et al. Small-molecule activator of glutamate transporter EAAT2 translation provides neuroprotection. The Journal of Clinical Investigation. 2014;124(3):1255-1267
  108. 108. Tejeda-Bayron FA, Rivera-Aponte DE, Malpica-Nieves CJ, Maldonado-Martinez G, Maldonado HM, Skatchkov SN, et al. Activation of glutamate transporter-1 (GLT-1) confers sex-dependent neuroprotection in brain ischemia. Brain Sciences. 2021;11(1):76
  109. 109. Colton CK, Kong Q , Lai L, Zhu MX, Seyb KI, Cuny GD, et al. Identification of translational activators of glial glutamate transporter EAAT2 through cell-based high-throughput screening: An approach to prevent excitotoxicity. Journal of Biomolecular Screening. 2010;15(6):653-662
  110. 110. Xing X, Chang LC, Kong Q , Colton CK, Lai L, Glicksman MA, et al. Structure-activity relationship study of pyridazine derivatives as glutamate transporter EAAT2 activators. Bioorganic & Medicinal Chemistry Letters. 2011;21(19):5774-5777
  111. 111. Takahashi K, Kong Q , Lin Y, Stouffer N, Schulte DA, Lai L, et al. Restored glial glutamate transporter EAAT2 function as a potential therapeutic approach for Alzheimer's disease. The Journal of Experimental Medicine. 2015;212(3):319-332
  112. 112. Alotaibi G, Rahman S. Effects of glial glutamate transporter activator in formalin-induced pain behaviour in mice. European Journal of Pain (London, England). 2019;23(4):765-783
  113. 113. Mao QX, Yang TD. Amitriptyline upregulates EAAT1 and EAAT2 in neuropathic pain rats. Brain Research Bulletin. 2010;81(4-5):424-427
  114. 114. Moore RA, Derry S, Aldington D, Cole P, Wiffen PJ. Amitriptyline for neuropathic pain in adults. The Cochrane Database of Systematic Reviews. 2015;2015(7):Cd008242
  115. 115. Krzyzanowska W, Pomierny B, Bystrowska B, Pomierny-Chamiolo L, Filip M, Budziszewska B, et al. Ceftriaxone- and N-acetylcysteine-induced brain tolerance to ischemia: Influence on glutamate levels in focal cerebral ischemia. PLoS One. 2017;12(10):e0186243
  116. 116. Oka S, Kamata H, Kamata K, Yagisawa H, Hirata H. N-acetylcysteine suppresses TNF-induced NF-kappaB activation through inhibition of IkappaB kinases. FEBS Letters. 2000;472(2-3):196-202
  117. 117. Sitcheran R, Gupta P, Fisher PB, Baldwin AS. Positive and negative regulation of EAAT2 by NF-kappaB: A role for N-myc in TNFalpha-controlled repression. The EMBO Journal. 2005;24(3):510-520
  118. 118. Jastrzębska J, Frankowska M, Filip M, Atlas D. N-acetylcysteine amide (AD4) reduces cocaine-induced reinstatement. Psychopharmacology. 2016;233(18):3437-3448
  119. 119. Ramirez-Niño AM, D'Souza MS, Markou A. N-acetylcysteine decreased nicotine self-administration and cue-induced reinstatement of nicotine seeking in rats: Comparison with the effects of N-acetylcysteine on food responding and food seeking. Psychopharmacology. 2013;225(2):473-482
  120. 120. Quintanilla ME, Rivera-Meza M, Berríos-Cárcamo P, Salinas-Luypaert C, Herrera-Marschitz M, Israel Y. Beyond the "First Hit": Marked inhibition by N-acetyl cysteine of chronic ethanol intake but not of early ethanol intake. Parallel effects on ethanol-induced saccharin motivation. Alcoholism, Clinical and Experimental Research. 2016;40(5):1044-1051
  121. 121. Reissner KJ, Gipson CD, Tran PK, Knackstedt LA, Scofield MD, Kalivas PW. Glutamate transporter GLT-1 mediates N-acetylcysteine inhibition of cocaine reinstatement. Addiction Biology. 2015;20(2):316-323
  122. 122. Nie H, Zhang H, Weng HR. Minocycline prevents impaired glial glutamate uptake in the spinal sensory synapses of neuropathic rats. Neuroscience. 2010;170(3):901-912
  123. 123. Sheldon AL, Gonzalez MI, Krizman-Genda EN, Susarla BT, Robinson MB. Ubiquitination-mediated internalization and degradation of the astroglial glutamate transporter, GLT-1. Neurochemistry International. 2008;53(6-8):296-308
  124. 124. Park HJ, Baik HJ, Kim DY, Lee GY, Woo JH, Zuo Z, et al. Doxepin and imipramine but not fluoxetine reduce the activity of the rat glutamate transporter EAAT3 expressed in Xenopus oocytes. BMC Anesthesiology. 2015;15:116
  125. 125. Susarla BT, Robinson MB. Internalization and degradation of the glutamate transporter GLT-1 in response to phorbol ester. Neurochemistry International. 2008;52(4-5):709-722
  126. 126. Dowd LA, Robinson MB. Rapid stimulation of EAAC1-mediated Na+−dependent L-glutamate transport activity in C6 glioma cells by phorbol ester. Journal of Neurochemistry. 1996;67(2):508-516
  127. 127. Davis KE, Straff DJ, Weinstein EA, Bannerman PG, Correale DM, Rothstein JD, et al. Multiple signaling pathways regulate cell surface expression and activity of the excitatory amino acid carrier 1 subtype of Glu transporter in C6 glioma. The Journal of Neuroscience. 1998;18(7):2475-2485
  128. 128. Ryan RM, Ingram SL, Scimemi A. Regulation of glutamate, GABA and dopamine transporter uptake, surface mobility and expression. Frontiers in Cellular Neuroscience. 2021;15:670346
  129. 129. Gregg RA, Hicks C, Nayak SU, Tallarida CS, Nucero P, Smith GR, et al. Synthetic cathinone MDPV downregulates glutamate transporter subtype I (GLT-1) and produces rewarding and locomotor-activating effects that are reduced by a GLT-1 activator. Neuropharmacology. 2016;108:111-119
  130. 130. Underhill SM, Wheeler DS, Li M, Watts SD, Ingram SL, Amara SG. Amphetamine modulates glutamatergic neurotransmission through endocytosis of the excitatory amino acid transporter EAAT3 in dopamine neurons. Neuron. 2014;83(2):404-416
  131. 131. Li M-H, Underhill SM, Reed C, Phillips TJ, Amara SG, Ingram SL. Amphetamine and methamphetamine increase NMDAR-GluN2B synaptic currents in midbrain dopamine neurons. Neuropsychopharmacology. 2017;42(7):1539-1547
  132. 132. Grewer C, Gameiro A, Rauen T. SLC1 glutamate transporters. Pflügers Archiv. 2014;466(1):3-24
  133. 133. Conradt M, Stoffel W. Inhibition of the high-affinity brain glutamate transporter GLAST-1 via direct phosphorylation. Journal of Neurochemistry. 1997;68(3):1244-1251
  134. 134. Gegelashvili G, Dehnes Y, Danbolt NC, Schousboe A. The high-affinity glutamate transporters GLT1, GLAST, and EAAT4 are regulated via different signalling mechanisms. Neurochemistry International. 2000;37(2-3):163-170
  135. 135. Karki P, Smith K, Johnson J Jr, Aschner M, Lee EY. Genetic dys-regulation of astrocytic glutamate transporter EAAT2 and its implications in neurological disorders and manganese toxicity. Neurochemical Research. 2015;40(2):380-388
  136. 136. Peterson AR, Binder DK. Post-translational regulation of GLT-1 in neurological diseases and its potential as an effective therapeutic target. Frontiers in Molecular Neuroscience. 2019;12:164
  137. 137. Huang K, Kang MH, Askew C, Kang R, Sanders SS, Wan J, et al. Palmitoylation and function of glial glutamate transporter-1 is reduced in the YAC128 mouse model of Huntington disease. Neurobiology of Disease. 2010;40(1):207-215
  138. 138. Foran E, Rosenblum L, Bogush A, Pasinelli P, Trotti D. Sumoylation of the astroglial glutamate transporter EAAT2 governs its intracellular compartmentalization. GLIA. 2014;62(8):1241-1253
  139. 139. Szatkowski M, Barbour B, Attwell D. Non-vesicular release of glutamate from glial cells by reversed electrogenic glutamate uptake. Nature. 1990;348(6300):443-446
  140. 140. Katayama Y, Becker DP, Tamura T, Hovda DA. Massive increases in extracellular potassium and the indiscriminate release of glutamate following concussive brain injury. Journal of Neurosurgery. 1990;73(6):889-900
  141. 141. Phillis JW, O'Regan MH. Mechanisms of glutamate and aspartate release in the ischemic rat cerebral cortex. Brain Research. 1996;730(1-2):150-164
  142. 142. Volterra A, Trotti D, Tromba C, Floridi S, Racagni G. Glutamate uptake inhibition by oxygen free radicals in rat cortical astrocytes. The Journal of Neuroscience. 1994;14(5 Pt 1):2924-2932
  143. 143. Rothstein JD. Excitotoxicity and neurodegeneration in amyotrophic lateral sclerosis. Clinical Neuroscience. 1995;3(6):348-359
  144. 144. Lipton SA. Failures and successes of NMDA receptor antagonists: Molecular basis for the use of open-channel blockers like memantine in the treatment of acute and chronic neurologic insults. NeuroRx. 2004;1(1):101-110
  145. 145. Tanaka K, Watase K, Manabe T, Yamada K, Watanabe M, Takahashi K, et al. Epilepsy and exacerbation of brain injury in mice lacking the glutamate transporter GLT-1. Science. 1997;276(5319):1699-1702
  146. 146. Hakuba N, Koga K, Gyo K, Usami SI, Tanaka K. Exacerbation of noise-induced hearing loss in mice lacking the glutamate transporter GLAST. The Journal of Neuroscience. 2000;20(23):8750-8753
  147. 147. Rao VL, Baskaya MK, Dogan A, Rothstein JD, Dempsey RJ. Traumatic brain injury down-regulates glial glutamate transporter (GLT-1 and GLAST) proteins in rat brain. Journal of Neurochemistry. 1998;70(5):2020-2027
  148. 148. Matsugami TR, Tanemura K, Mieda M, Nakatomi R, Yamada K, Kondo T, et al. From the Cover: Indispensability of the glutamate transporters GLAST and GLT1 to brain development. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(32):12161-12166
  149. 149. Rothstein JD, Dykes-Hoberg M, Pardo CA, Bristol LA, Jin L, Kuncl RW, et al. Knockout of glutamate transporters reveals a major role for astroglial transport in excitotoxicity and clearance of glutamate. Neuron. 1996;16(3):675-686
  150. 150. Barnett NL, Pow DV. Antisense knockdown of GLAST, a glial glutamate transporter, compromises retinal function. Investigative Ophthalmology & Visual Science. 2000;41(2):585-591
  151. 151. Karlsson RM, Adermark L, Molander A, Perreau-Lenz S, Singley E, Solomon M, et al. Reduced alcohol intake and reward associated with impaired endocannabinoid signaling in mice with a deletion of the glutamate transporter GLAST. Neuropharmacology. 2012;63(2):181-189
  152. 152. Peghini P, Janzen J, Stoffel W. Glutamate transporter EAAC-1-deficient mice develop dicarboxylic aminoaciduria and behavioral abnormalities but no neurodegeneration. The EMBO Journal. 1997;16(13):3822-3832
  153. 153. Aoyama K, Suh SW, Hamby AM, Liu J, Chan WY, Chen Y, et al. Neuronal glutathione deficiency and age-dependent neurodegeneration in the EAAC1 deficient mouse. Nature Neuroscience. 2006;9(1):119-126
  154. 154. Watts SD, Torres-Salazar D, Divito CB, Amara SG. Cysteine transport through excitatory amino acid transporter 3 (EAAT3). PLoS One. 2014;9(10):e109245
  155. 155. Bailey CG, Ryan RM, Thoeng AD, Ng C, King K, Vanslambrouck JM, et al. Loss-of-function mutations in the glutamate transporter SLC1A1 cause human dicarboxylic aminoaciduria. The Journal of Clinical Investigation. 2011;121(1):446-453
  156. 156. Nikkuni O, Takayasu Y, Iino M, Tanaka K, Ozawa S. Facilitated activation of metabotropic glutamate receptors in cerebellar Purkinje cells in glutamate transporter EAAT4-deficient mice. Neuroscience Research. 2007;59(3):296-303
  157. 157. Lukasiewcz PD, Bligard GW, DeBrecht JD. EAAT5 glutamate transporter-mediated inhibition in the vertebrate retina. Frontiers in Cellular Neuroscience. 2021;15:662859
  158. 158. Beresford IJ, Parsons AA, Hunter AJ. Treatments for stroke. Expert Opinion on Emerging Drugs. 2003;8(1):103-122
  159. 159. Puig B, Brenna S, Magnus T. Molecular communication of a dying neuron in stroke. International Journal of Molecular Sciences. 2018;19(9):2834
  160. 160. Wetterling F, Chatzikonstantinou E, Tritschler L, Meairs S, Fatar M, Schad LR. Investigating potentially salvageable penumbra tissue in an in vivo model of transient ischemic stroke using sodium, diffusion, and perfusion magnetic resonance imaging. BMC Neuroscience. 2016;17(1):82
  161. 161. Krzyzanowska W, Pomierny B, Filip M, Pera J. Glutamate transporters in brain ischemia: To modulate or not? Acta Pharmacologica Sinica. 2014;35(4):444-462
  162. 162. Levy LM, Lehre KP, Walaas SI, Storm-Mathisen J, Danbolt NC. Down-regulation of glial glutamate transporters after glutamatergic denervation in the rat brain. The European Journal of Neuroscience. 1995;7(10):2036-2041
  163. 163. Dong X, Wang Y, Qin Z. Molecular mechanisms of excitotoxicity and their relevance to pathogenesis of neurodegenerative diseases. Acta Pharmacologica Sinica. 2009;30:379-387
  164. 164. Wu HY, Tomizawa K, Oda Y, Wei FY, Lu YF, Matsushita M, et al. Critical role of calpain-mediated cleavage of calcineurin in excitotoxic neurodegeneration. The Journal of Biological Chemistry. 2004;279(6):4929-4940
  165. 165. Beckman JS, Chen J, Crow JP, Ye YZ. Reactions of nitric oxide, superoxide and peroxynitrite with superoxide dismutase in neurodegeneration. Progress in Brain Research. 1994;103:371-380
  166. 166. Shen Y, Lu H, Xu R, Tian H, Xia X, Zhou FH, et al. The expression of GLAST and GLT1 in a transient cerebral ischemia Mongolian gerbil model. Neuropsychiatric Disease and Treatment. 2020;16:789-800
  167. 167. Wang D, Zhao Y, Zhang Y, Zhang T, Shang X, Wang J, et al. Hypothermia protects against oxygen-glucose deprivation-induced neuronal injury by down-regulating the reverse transport of glutamate by astrocytes as mediated by neurons. Neuroscience. 2013;237:130-138
  168. 168. Al Awabdh S, Gupta-Agarwal S, Sheehan DF, Muir J, Norkett R, Twelvetrees AE, et al. Neuronal activity mediated regulation of glutamate transporter GLT-1 surface diffusion in rat astrocytes in dissociated and slice cultures. GLIA. 2016;64(7):1252-1264
  169. 169. Abulseoud OA, Alasmari F, Hussein AM, Sari Y. Ceftriaxone as a novel therapeutic agent for hyperglutamatergic states: Bridging the gap between preclinical results and clinical translation. Frontiers in Neuroscience. 2022;16:841036
  170. 170. Palmer AM, Marion DW, Botscheller ML, Swedlow PE, Styren SD, DeKosky ST. Traumatic brain injury-induced excitotoxicity assessed in a controlled cortical impact model. Journal of Neurochemistry. 1993;61(6):2015-2024
  171. 171. Ikematsu K, Tsuda R, Kondo T, Nakasono I. The expression of excitatory amino acid transporter 2 in traumatic brain injury. Forensic Science International. 2002;130(2-3):83-89
  172. 172. Rao VL, Dogan A, Bowen KK, Todd KG, Dempsey RJ. Antisense knockdown of the glial glutamate transporter GLT-1 exacerbates hippocampal neuronal damage following traumatic injury to rat brain. The European Journal of Neuroscience. 2001;13(1):119-128
  173. 173. Collins M, Lovell MR, Iverson GL, Ide T, Maroon J. Examining concussion rates and return to play in high school football players wearing newer helmet technology: A three-year prospective cohort study. Neurosurgery. 2006;58(2):275-286; discussion-86
  174. 174. Matute C, Domercq M, Sanchez-Gomez MV. Glutamate-mediated glial injury: Mechanisms and clinical importance. GLIA. 2006;53(2):212-224
  175. 175. Jabs R, Seifert G, Steinhauser C. Astrocytic function and its alteration in the epileptic brain. Epilepsia. 2008;49(Suppl. 2):3-12
  176. 176. Green JL, Dos Santos WF, Fontana ACK. Role of glutamate excitotoxicity and glutamate transporter EAAT2 in epilepsy: Opportunities for novel therapeutics development. Biochemical Pharmacology. 2021;193:114786
  177. 177. Coutinho-Netto J, Abdul-Ghani AS, Collins JF, Bradford HF. Is glutamate a trigger factor in epileptic hyperactivity? Epilepsia. 1981;22(3):289-296
  178. 178. Petr GT, Sun Y, Frederick NM, Zhou Y, Dhamne SC, Hameed MQ , et al. 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. The Journal of Neuroscience. 2015;35(13):5187-5201
  179. 179. During MJ, Spencer DD. Extracellular hippocampal glutamate and spontaneous seizure in the conscious human brain. Lancet. 1993;341(8861):1607-1610
  180. 180. Mathern GW, Mendoza D, Lozada A, Pretorius JK, Dehnes Y, Danbolt NC, et al. Hippocampal GABA and glutamate transporter immunoreactivity in patients with temporal lobe epilepsy. Neurology. 1999;52(3):453-472
  181. 181. Lopes MW, Soares FM, de Mello N, Nunes JC, Cajado AG, de Brito D, et al. Time-dependent modulation of AMPA receptor phosphorylation and mRNA expression of NMDA receptors and glial glutamate transporters in the rat hippocampus and cerebral cortex in a pilocarpine model of epilepsy. Experimental Brain Research. 2013;226(2):153-163
  182. 182. David Y, Cacheaux LP, Ivens S, Lapilover E, Heinemann U, Kaufer D, et al. Astrocytic dysfunction in epileptogenesis: Consequence of altered potassium and glutamate homeostasis? The Journal of Neuroscience. 2009;29(34):10588-10599
  183. 183. Wong M, Ess KC, Uhlmann EJ, Jansen LA, Li W, Crino PB, et al. Impaired glial glutamate transport in a mouse tuberous sclerosis epilepsy model. Annals of Neurology. 2003;54(2):251-256
  184. 184. Ueda Y, Doi T, Nagatomo K, Willmore LJ, Nakajima A. Functional role for redox in the epileptogenesis: Molecular regulation of glutamate in the hippocampus of FeCl3-induced limbic epilepsy model. Experimental Brain Research. 2007;181(4):571-577
  185. 185. Lu Z, Zhang W, Zhang N, Jiang J, Luo Q , Qiu Y. The expression of glutamate transporters in chest compression-induced audiogenic epilepsy: A comparative study. Neurological Research. 2008;30(9):915-919
  186. 186. Tessler S, Danbolt NC, Faull RL, Storm-Mathisen J, Emson PC. Expression of the glutamate transporters in human temporal lobe epilepsy. Neuroscience. 1999;88(4):1083-1091
  187. 187. Akbar MT, Torp R, Danbolt NC, Levy LM, Meldrum BS, Ottersen OP. Expression of glial glutamate transporters GLT-1 and GLAST is unchanged in the hippocampus in fully kindled rats. Neuroscience. 1997;78(2):351-359
  188. 188. Bough KJ, Paquet M, Pare JF, Hassel B, Smith Y, Hall RA, et al. Evidence against enhanced glutamate transport in the anticonvulsant mechanism of the ketogenic diet. Epilepsy Research. 2007;74(2-3):232-236
  189. 189. Guo F, Sun F, Yu JL, Wang QH, Tu DY, Mao XY, et al. Abnormal expressions of glutamate transporters and metabotropic glutamate receptor 1 in the spontaneously epileptic rat hippocampus. Brain Research Bulletin. 2010;81(4-5):510-516
  190. 190. Eid T, Lee TW, Patrylo P, Zaveri HP. Astrocytes and glutamine synthetase in epileptogenesis. Journal of Neuroscience Research. 2019;97(11):1345-1362
  191. 191. 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. Neurobiology of Disease. 2012;47(2):145-154
  192. 192. Wetherington J, Serrano G, Dingledine R. Astrocytes in the epileptic brain. Neuron. 2008;58(2):168-178
  193. 193. Swieboda P, Filip R, Prystupa A, Drozd M. Assessment of pain: Types, mechanism and treatment. Annals of Agricultural and Environmental Medicine. 2013;Spec no. 1:2-7
  194. 194. Tao YX, Gu J, Stephens RL Jr. Role of spinal cord glutamate transporter during normal sensory transmission and pathological pain states. Molecular Pain. 2005;1:30
  195. 195. Yang S, Chang MC. Chronic pain: Structural and functional changes in brain structures and associated negative affective states. International Journal of Molecular Sciences. 2019;20(13):3130
  196. 196. Temmermand R, Barrett JE, Fontana ACK. Glutamatergic systems in neuropathic pain and emerging non-opioid therapies. Pharmacological Research. 2022;185:106492
  197. 197. Ji RR, Nackley A, Huh Y, Terrando N, Maixner W. Neuroinflammation and central sensitization in chronic and widespread pain. Anesthesiology. 2018;129(2):343-366
  198. 198. Gegelashvili G, Bjerrum OJ. High-affinity glutamate transporters in chronic pain: An emerging therapeutic target. Journal of Neurochemistry. 2014;131(6):712-730
  199. 199. Liaw WJ, Stephens RL Jr, Binns BC, Chu Y, Sepkuty JP, Johns RA, et al. Spinal glutamate uptake is critical for maintaining normal sensory transmission in rat spinal cord. Pain. 2005;115(1-2):60-70
  200. 200. Wang W, Wang W, Wang Y, Huang J, Wu S, Li YQ. Temporal changes of astrocyte activation and glutamate transporter-1 expression in the spinal cord after spinal nerve ligation-induced neuropathic pain. The Anatomical Record (Hoboken). 2008;291(5):513-518
  201. 201. Pereira V, Goudet C. Emerging trends in pain modulation by metabotropic glutamate receptors. Frontiers in Molecular Neuroscience. 2019;11:1-23
  202. 202. Hewitt DJ. The use of NMDA-receptor antagonists in the treatment of chronic pain. The Clinical Journal of Pain. 2000;16(Suppl. 2):S73-S79
  203. 203. Gegelashvili G, Bjerrum OJ. Glutamate transport system as a novel therapeutic target in chronic pain: Molecular mechanisms and pharmacology. Advances in Neurobiology. 2017;16:225-253
  204. 204. Amin B, Avaznia M, Noorani R, Mehri S, Hosseinzadeh H. Upregulation of glutamate transporter 1 by clavulanic acid administration and attenuation of allodynia and hyperalgesia in neuropathic rats. Basic and Clinical Neuroscience. 2019;10(4):345-354
  205. 205. Kristensen PJ, Gegelashvili G, Munro G, Heegaard AM, Bjerrum OJ. The beta-lactam clavulanic acid mediates glutamate transport-sensitive pain relief in a rat model of neuropathic pain. European Journal of Pain (London, England). 2018;22(2):282-294
  206. 206. Eljaja L, Bjerrum OJ, Honoré PH, Abrahamsen B. Effects of the excitatory amino acid transporter subtype 2 (EAAT-2) inducer ceftriaxone on different pain modalities in rat. Scandinavian Journal of Pain. 2018;2(3):132-136
  207. 207. Maeda S, Kawamoto A, Yatani Y, Shirakawa H, Nakagawa T, Kaneko S. Gene transfer of GLT-1, a glial glutamate transporter, into the spinal cord by recombinant adenovirus attenuates inflammatory and neuropathic pain in rats. Molecular Pain. 2008;4:65
  208. 208. Yousuf MS, Kerr BJ. The role of regulatory transporters in neuropathic pain. Advances in Pharmacology (San Diego, Calif). 2016;75:245-271
  209. 209. Hajhashemi V, Hosseinzadeh H, Amin B. Antiallodynia and antihyperalgesia effects of ceftriaxone in treatment of chronic neuropathic pain in rats. Acta Neuropsychiatrica. 2013;25(1):27-32
  210. 210. Fontana IC, Souza DG, Souza DO, Gee A, Zimmer ER, Bongarzone S. A Medicinal chemistry perspective on excitatory amino acid transporter 2 dysfunction in neurodegenerative diseases. Journal of Medicinal Chemistry. 2023;66(4):2330-2346
  211. 211. Limpert AS, Cosford ND. Translational enhancers of EAAT2: Therapeutic implications for neurodegenerative disease. The Journal of Clinical Investigation. 2014;124(3):964-967
  212. 212. Zhou Y, Danbolt NC. Glutamate as a neurotransmitter in the healthy brain. Journal of Neural Transmission (Vienna). 2014;121(8):799-817
  213. 213. Kalivas PW. The glutamate homeostasis hypothesis of addiction. Nature Reviews. Neuroscience. 2009;10(8):561-572
  214. 214. Christie MJ, Summers RJ, Stephenson JA, Cook CJ, Beart PM. Excitatory amino acid projections to the nucleus accumbens septi in the rat: A retrograde transport study utilizingd [3H] aspartate and [3H]GABA. Neuroscience. 1987;22(2):425-439
  215. 215. Wise RA. Roles for nigrostriatal--not just mesocorticolimbic--dopamine in reward and addiction. Trends in Neurosciences. 2009;32(10):517-524
  216. 216. Spencer S, Kalivas PW. Glutamate transport: A new bench to bedside mechanism for treating drug abuse. The International Journal of Neuropsychopharmacology. 2017;20(10):797-812
  217. 217. Niedzielska-Andres E, Pomierny-Chamioło L, Andres M, Walczak M, Knackstedt LA, Filip M, et al. Cocaine use disorder: A look at metabotropic glutamate receptors and glutamate transporters. Pharmacology & Therapeutics. 2021;221:107797
  218. 218. Griffin WC, Ramachandra VS, Knackstedt LA, Becker HC. Repeated cycles of chronic intermittent ethanol exposure increases basal glutamate in the nucleus accumbens of mice without affecting glutamate transport. Frontiers in Pharmacology. 2015;6:27
  219. 219. Reissner KJ, Kalivas PW. Using glutamate homeostasis as a target for treating addictive disorders. Behavioural Pharmacology. 2010;21(5-6):514-522
  220. 220. Sari Y, Smith KD, Ali PK, Rebec GV. Upregulation of GLT1 attenuates cue-induced reinstatement of cocaine-seeking behavior in rats. The Journal of Neuroscience. 2009;29(29):9239-9243
  221. 221. Shen HW, Scofield MD, Boger H, Hensley M, Kalivas PW. Synaptic glutamate spillover due to impaired glutamate uptake mediates heroin relapse. The Journal of Neuroscience. 2014;34(16):5649-5657
  222. 222. Abulseoud OA, Miller JD, Wu J, Choi DS, Holschneider DP. Ceftriaxone upregulates the glutamate transporter in medial prefrontal cortex and blocks reinstatement of methamphetamine seeking in a condition place preference paradigm. Brain Research. 2012;1456:14-21
  223. 223. Sari Y, Toalston JE, Rao PS, Bell RL. Effects of ceftriaxone on ethanol, nicotine or sucrose intake by alcohol-preferring (P) rats and its association with GLT-1 expression. Neuroscience. 2016;326:117-125
  224. 224. Niedzielska-Andres E, Mizera J, Sadakierska-Chudy A, Pomierny-Chamioło L, Filip M. Changes in the glutamate biomarker expression in rats vulnerable or resistant to the rewarding effects of cocaine and their reversal by ceftriaxone. Behavioural Brain Research. 2019;370:111945
  225. 225. Sepulveda-Orengo MT, Healey KL, Kim R, Auriemma AC, Rojas J, Woronoff N, et al. Riluzole impairs cocaine reinstatement and restores adaptations in intrinsic excitability and GLT-1 expression. Neuropsychopharmacology. 2018;43(6):1212-1223
  226. 226. Ciraulo DA, Sarid-Segal O, Knapp CM, Ciraulo AM, LoCastro J, Bloch DA, et al. Efficacy screening trials of paroxetine, pentoxifylline, riluzole, pramipexole and venlafaxine in cocaine dependence. Addiction. 2005; 100(Suppl. 1):12-22
  227. 227. Farahzadi MH, Moazen-Zadeh E, Razaghi E, Zarrindast MR, Bidaki R, Akhondzadeh S. Riluzole for treatment of men with methamphetamine dependence: A randomized, double-blind, placebo-controlled clinical trial. Journal of Psychopharmacology. 2019;33(3):305-315
  228. 228. Itzhak Y, Martin JL. Effect of riluzole and gabapentin on cocaine- and methamphetamine-induced behavioral sensitization in mice. Psychopharmacology. 2000;151(2-3):226-233
  229. 229. Tzschentke TM, Schmidt WJ. Blockade of morphine- and amphetamine-induced conditioned place preference in the rat by riluzole. Neuroscience Letters. 1998;242(2):114-116
  230. 230. Besheer J, Lepoutre V, Hodge CW. Preclinical evaluation of riluzole: Assessments of ethanol self-administration and ethanol withdrawal symptoms. Alcoholism, Clinical and Experimental Research. 2009;33(8):1460-1468
  231. 231. Kumar A, Singh A, Ekavali. A review on Alzheimer's disease pathophysiology and its management: An update. Pharmacological Reports: PR. 2015;67(2):195-203
  232. 232. Sehar U, Rawat P, Reddy AP, Kopel J, Reddy PH. Amyloid beta in aging and Alzheimer’s disease. International Journal of Molecular Sciences. 2022;23(21):12924
  233. 233. de Paula VJR, Guimarães FM, Diniz BS, Forlenza OV. Neurobiological pathways to Alzheimer's disease: Amyloid-beta, TAU protein or both? Dementia & Neuropsychologia. 2009;3(3):188-194
  234. 234. Hardy J, Cowburn R, Barton A, Reynolds G, Lofdahl E, O'Carroll AM, et al. Region-specific loss of glutamate innervation in Alzheimer's disease. Neuroscience Letters. 1987;73(1):77-80
  235. 235. Mookherjee P, Green PS, Watson GS, Marques MA, Tanaka K, Meeker KD, et al. GLT-1 loss accelerates cognitive deficit onset in an Alzheimer's disease animal model. Journal of Alzheimer's Disease: JAD. 2011;26(3):447-455
  236. 236. Woltjer RL, Duerson K, Fullmer JM, Mookherjee P, Ryan AM, Montine TJ, et al. Aberrant detergent-insoluble excitatory amino acid transporter 2 accumulates in Alzheimer disease. Journal of Neuropathology and Experimental Neurology. 2010;69(7):667-676
  237. 237. Scimemi A, Meabon JS, Woltjer RL, Sullivan JM, Diamond JS, Cook DG. Amyloid-beta1-42 slows clearance of synaptically released glutamate by mislocalizing astrocytic GLT-1. The Journal of Neuroscience. 2013;33(12):5312-5318
  238. 238. Meeker KD, Meabon JS, Cook DG. Partial loss of the glutamate transporter GLT-1 alters brain akt and insulin signaling in a mouse model of Alzheimer's disease. Journal of Alzheimer's Disease: JAD. 2015;45(2):509-520
  239. 239. Kulijewicz-Nawrot M, Sykova E, Chvatal A, Verkhratsky A, Rodriguez JJ. Astrocytes and glutamate homoeostasis in Alzheimer's disease: A decrease in glutamine synthetase, but not in glutamate transporter-1, in the prefrontal cortex. ASN Neuro. 2013;5(4):273-282
  240. 240. Hogan DB. Long-term efficacy and toxicity of cholinesterase inhibitors in the treatment of Alzheimer disease. Canadian Journal of Psychiatry. 2014;59(12):618-623
  241. 241. Zimmer ER, Kalinine E, Haas CB, Torrez VR, Souza DO, Muller AP, et al. Pretreatment with memantine prevents Alzheimer-like alterations induced by intrahippocampal okadaic acid administration in rats. Current Alzheimer Research. 2012;9(10):1182-1190
  242. 242. Gettman L. Lecanemab-irmb (Leqembi™) for treatment of Alzheimer's disease. The Senior Care Pharmacist. 2024;39(2):75-77
  243. 243. Zhu Y, Fotinos A, Mao LL, Atassi N, Zhou EW, Ahmad S, et al. Neuroprotective agents target molecular mechanisms of disease in ALS. Drug Discovery Today. 2015;20(1):65-75
  244. 244. Trotti D, Aoki M, Pasinelli P, Berger UV, Danbolt NC, Brown RH Jr, et al. Amyotrophic lateral sclerosis-linked glutamate transporter mutant has impaired glutamate clearance capacity. The Journal of Biological Chemistry. 2001;276(1):576-582
  245. 245. Barber SC, Mead RJ, Shaw PJ. Oxidative stress in ALS: A mechanism of neurodegeneration and a therapeutic target. Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease. 2006;1762(11):1051-1067
  246. 246. Bruijn LI, Becher MW, Lee MK, Anderson KL, Jenkins NA, Copeland NG, et al. ALS-linked SOD1 mutant G85R mediates damage to astrocytes and promotes rapidly progressive disease with SOD1-containing inclusions. Neuron. 1997;18(2):327-338
  247. 247. Van Den Bosch L, Van Damme P, Bogaert E, Robberecht W. The role of excitotoxicity in the pathogenesis of amyotrophic lateral sclerosis. Biochimica et Biophysica Acta. 2006;1762(11-12):1068-1082
  248. 248. Cudkowicz ME, Titus S, Kearney M, Yu H, Sherman A, Schoenfeld D, et al. Safety and efficacy of ceftriaxone for amyotrophic lateral sclerosis: A multi-stage, randomised, double-blind, placebo-controlled trial. Lancet Neurology. 2014;13(11):1083-1091
  249. 249. Li K, Hala TJ, Seetharam S, Poulsen DJ, Wright MC, Lepore AC. GLT1 overexpression in SOD1(G93A) mouse cervical spinal cord does not preserve diaphragm function or extend disease. Neurobiology of Disease. 2015;78:12-23
  250. 250. Verche VL, Ikiz B, Jacquier A, Przedborski S, Re DB. Glutamate pathway implication in amyotrophic lateral sclerosis: What is the signal in the noise? Journal of Receptor, Ligand and Channel Research. 2011;4:1-22
  251. 251. Walker FO. Huntington's disease. The Lancet. 2007;369(9557):218-228
  252. 252. Massieu L, García O. The role of excitotoxicity and metabolic failure in the pathogenesis of neurological disorders. Neurobiology (Budapest, Hungary). 1998;6(1):99-108
  253. 253. Arzberger T, Krampfl K, Leimgruber S, Weindl A. Changes of NMDA receptor subunit (NR1, NR2B) and glutamate transporter (GLT1) mRNA expression in Huntington's disease--an in situ hybridization study. Journal of Neuropathology and Experimental Neurology. 1997;56(4):440-454
  254. 254. Estrada-Sanchez AM, Montiel T, Segovia J, Massieu L. Glutamate toxicity in the striatum of the R6/2 Huntington's disease transgenic mice is age-dependent and correlates with decreased levels of glutamate transporters. Neurobiology of Disease. 2009;34(1):78-86
  255. 255. Todd AC, Hardingham GE. The regulation of astrocytic glutamate transporters in health and neurodegenerative diseases. International Journal of Molecular Sciences. 2020;21(24):9607
  256. 256. Frank S. Treatment of Huntington's disease. Neurotherapeutics. 2014;11(1):153-160
  257. 257. Assous M, Had-Aissouni L, Gubellini P, Melon C, Nafia I, Salin P, et al. Progressive parkinsonism by acute dysfunction of excitatory amino acid transporters in the rat substantia nigra. Neurobiology of Disease. 2014;65:69-81
  258. 258. Chotibut T, Davis RW, Arnold JC, Frenchek Z, Gurwara S, Bondada V, et al. Ceftriaxone increases glutamate uptake and reduces striatal tyrosine hydroxylase loss in 6-OHDA Parkinson's model. Molecular Neurobiology. 2014;49(3):1282-1292
  259. 259. Yadav A, Collman RG. CNS inflammation and macrophage/microglial biology associated with HIV-1 infection. Journal of Neuroimmune Pharmacology. 2009;4(4):430-447
  260. 260. Afaf E-A, Hussain AD. Biomarkers-Directed Strategies to Treat Autism. In: Mu W, Frank AW, editors. Role of Biomarkers in Medicine. Rijeka: IntechOpen; 2016. p. Ch. 10
  261. 261. Takahashi KK. Glutamate transporter EAAT2: Regulation, function, and potential as a therapeutic target for neurological and psychiatric disease. Cellular and Molecular Life Sciences: CMLS. 2015;72(18):3489-3506
  262. 262. Chen JX, Yao LH, Xu BB, Qian K, Wang HL, Liu ZC, et al. Glutamate transporter 1-mediated antidepressant-like effect in a rat model of chronic unpredictable stress. Journal of Huazhong University of Science and Technology Medical sciences = Hua zhong ke ji da xue xue bao Yi xue Ying De wen ban = Huazhong keji daxue xuebao Yixue Yingdewen ban. 2014;34(6):838-844
  263. 263. Nakagawa T, Kaneko S. SLC1 glutamate transporters and diseases: Psychiatric diseases and pathological pain. Current Molecular Pharmacology. 2013;6(2):66-73
  264. 264. Sorensen MF, Heimisdottir SB, Sorensen MD, Mellegaard CS, Wohlleben H, Kristensen BW, et al. High expression of cystine-glutamate antiporter xCT (SLC7A11) is an independent biomarker for epileptic seizures at diagnosis in glioma. Journal of Neuro-Oncology. 2018;138(1):49-53
  265. 265. Huberfeld G, Vecht CJ. Seizures and gliomas - towards a single therapeutic approach. Nature Reviews. Neurology. 2016;12(4):204-216
  266. 266. Werner P, Pitt D, Raine CS. Multiple sclerosis: Altered glutamate homeostasis in lesions correlates with oligodendrocyte and axonal damage. Annals of Neurology. 2001;50(2):169-180
  267. 267. Parkin GM, Udawela M, Gibbons A, Dean B. Glutamate transporters, EAAT1 and EAAT2, are potentially important in the pathophysiology and treatment of schizophrenia and affective disorders. World Journal of Psychiatry. 2018;8(2):51-63
  268. 268. Escobar AP, Wendland JR, Chávez AE, Moya PR. The neuronal glutamate transporter EAAT3 in obsessive-compulsive disorder. Frontiers in Pharmacology. 2019;10:1362
  269. 269. Takahashi K, Foster JB, Lin CL. Glutamate transporter EAAT2: Regulation, function, and potential as a therapeutic target for neurological and psychiatric disease. Cellular and Molecular Life Sciences. 2015;72(18):3489-3506

Written By

Katelyn L. Reeb, Simran K. Gill, Rhea Temmermand and Andréia C.K. Fontana

Submitted: 28 February 2024 Reviewed: 01 April 2024 Published: 04 June 2024