Skip to main content

REVIEW article

Front. Cell. Neurosci., 13 April 2021
Sec. Cellular Neurophysiology
This article is part of the Research Topic Neurotransmitter transporters: from the biophysics of the transport process to its implications for synapse and circuit function View all 14 articles

Regulation of Glutamate, GABA and Dopamine Transporter Uptake, Surface Mobility and Expression

  • 1School of Medical Sciences, Faculty of Medicine and Health, University of Sydney, Sydney, NSW, Australia
  • 2Department of Neurological Surgery, Oregon Health & Science University, Portland, OR, United States
  • 3Department of Biology, SUNY Albany, Albany, NY, United States

Neurotransmitter transporters limit spillover between synapses and maintain the extracellular neurotransmitter concentration at low yet physiologically meaningful levels. They also exert a key role in providing precursors for neurotransmitter biosynthesis. In many cases, neurons and astrocytes contain a large intracellular pool of transporters that can be redistributed and stabilized in the plasma membrane following activation of different signaling pathways. This means that the uptake capacity of the brain neuropil for different neurotransmitters can be dynamically regulated over the course of minutes, as an indirect consequence of changes in neuronal activity, blood flow, cell-to-cell interactions, etc. Here we discuss recent advances in the mechanisms that control the cell membrane trafficking and biophysical properties of transporters for the excitatory, inhibitory and modulatory neurotransmitters glutamate, GABA, and dopamine.

Glutamate Transporters

In addition to being one of the most abundant amino acids and the main excitatory neurotransmitter in the brain, glutamate controls synapse formation, and maturation, acts as an energy substrate for oxidative metabolism, contributes to the antioxidant properties of glutathione, and can be used as a building block for non-ribosomal peptide synthesis and as the precursor for the biosynthesis of the inhibitory neurotransmitter GABA (Sieber and Marahiel, 2005; Brosnan and Brosnan, 2013; Martinez-Lozada et al., 2016; Walker and van der Donk, 2016). In addition, in the developing brain, glutamate guides cell proliferation, migration, differentiation, and survival of neural progenitor cells (Jansson et al., 2013). Because of the plethora of effects that glutamate can exert, its lifetime in the extracellular space needs to be finely controlled through the activity of a family of Na+- and K+-dependent secondary active transporters. Geneticists, structural biologists, physiologists and clinicians all have their preferred nomenclature to refer to the five known glutamate transporters subtypes, summarized as follows: (i) Slc1a1/SLC1A1/EAAC1/EAAT3; (ii) Slc1a2/SLC1A2/GLT1/EAAT2; (iii) Slc1a3/SLC1A3/GLAST/EAAT1; (iv) Slc1a6/SLC1A6/EAAT4; (v) Slc1a7/SLC1A7/EAAT5. The use of each terminology is physiologically meaningful, as it refers to the genes encoding each transporter in rodents (Slc1a1, Slc1a2, Slc1a3, Slc1a6, Slc1a7) and in humans (SLC1A1, SLC1A2, SLC1A3, SLC1A6, SLC1A7), or the protein product in rodents (EAAC1, GLT1, GLAST, EAAT4-5) and humans (EAAT1-5). For simplicity, in this review, we refer to them as EAAT1-5. Different subtypes of glutamate transporters are differentially expressed in neuronal and glial cells (Danbolt, 2001). Accordingly, the glutamate transporters EAAT1-2 are mostly expressed in astrocytes, whereas EAAT3-5 are mostly expressed in neurons. Two of the neuronal transporters are predominantly expressed in the cerebellum (EAAT4) or retina (EAAT5), with a low yet experimentally measurable concentration in the forebrain (Dehnes et al., 1998; Danbolt, 2001). The others (i.e., EAAT1, EAAT2, and EAAT3) are expressed broadly throughout the brain, at varying levels.

Structural information of the glutamate transporters comes from prokaryote homologs such as GltPh and GltTk which share ∼35 amino acid identity with human EAAT2 (Yernool et al., 2004; Boudker et al., 2007; Reyes et al., 2009; Verdon and Boudker, 2012; Guskov et al., 2016; Scopelliti et al., 2018) and more recent crystal and cryo-EM structures of human transporters including EAAT1, EAAT3, and ASCT2 (Canul-Tec et al., 2017; Garaeva et al., 2018; Garaeva et al., 2019; Yu et al., 2019; Wang and Boudker, 2020) (Figure 1). All members of this family appear to assemble as trimers, with each monomer capable of transporting substrate and coupled ions, generating stoichiometric and non-stoichiometric currents, independently of the two other monomers (Grewer et al., 2005; Koch et al., 2007; Leary et al., 2007). The transporters are composed of a “transport domain” which binds and transports substrate and coupled ions, and a “scaffold domain” that forms inter-protomer contacts and interacts with the lipid membrane (Boudker et al., 2007; Reyes et al., 2009). GltPh transports aspartate together with three Na+ ions into the cytoplasm using a twisting elevator mechanism (Reyes et al., 2009; Ryan and Vandenberg, 2016) and generates a stoichiometrically uncoupled Cl conductance (Boudker et al., 2007; Ryan and Mindell, 2007; Reyes et al., 2009).

FIGURE 1
www.frontiersin.org

Figure 1. Crystal structure of GltPh, a bacterial homolog of glutamate transporters (PDB: 2NWX). (A) Side view of the glutamate transporter trimer parallel to the membrane, using a ribbon (top) and a surface representation sliced through the center of the transporter basin (bottom). (B) As in A, viewed from the extracellular side of the membrane. Aspartate (magenta) and sodium ions (blue) are represented as spheres. (C) Summary of the main feature of the stoichiometric and non-stoichiometric current mediated by glutamate transporters in prokaryotes and eukaryotes, respectively. Image generated using The PyMOL Molecular Graphics System (Version 2.3.2; Schrödinger, 2010).

In contrast to prokaryotes, for all eukaryotic transporters, glutamate uptake is driven by the electrochemical gradient for Na+ and H+ ions, and the rate limiting step is the counter-transport of one K+ ion across the membrane (Zerangue and Kavanaugh, 1996). The stoichiometry of the transport process is the inward movement of 1 glutamate (which carries a negative charge): 3 Na+: 1 H+, followed by the counter-transport of 1 K+, leading to the net influx of two positive charges per transport cycle (Zerangue and Kavanaugh, 1996; Levy et al., 1998). In addition, these transporters mediate a glutamate-dependent anion flux, which under physiological conditions is carried by Cl ions (Picaud et al., 1995; Wadiche et al., 1995b; Eliasof and Jahr, 1996). The chloride channel forms at the interface of the transport and scaffold domains during the transport cycle (Ryan et al., 2004; Cater et al., 2014, 2016; Cheng et al., 2017; Kolen et al., 2020; Chen et al., 2021) and the direction of this flux is determined by the driving force for chloride (i.e., the difference between membrane potential and reversal potential for chloride). The function of the chloride current remains incompletely understood but it has been suggested that it might serve to counterbalance the influx of positive charges due to glutamate transport and prevent cell depolarization (Grewer and Rauen, 2005). This hypothesis would only hold true in the presence of a positive driving force for chloride, causing an inward movement of chloride into the cell. This occurs when the reversal potential for chloride is more hyperpolarized than the membrane potential. This condition, however, is not fulfilled in developing neurons. Here, the reversal potential for chloride is more depolarized than in mature neurons, due to a delayed expression of the potassium chloride co-transporter KCC2, and a consequently higher intracellular chloride concentration in developing compared to mature neurons (Rivera et al., 1999). Other physiological roles for the chloride conductance have been demonstrated with EAAT5 in the retina, where it regulates cell activity and neurotransmitter release (Picaud et al., 1995; Veruki et al., 2006) and in cerebellar astrocytes, where the chloride currents of EAAT1/2 affects the resting intracellular chloride concentration (Sonders and Amara, 1996; Untiet et al., 2017).

Similarities and Distinctive Biophysical Features of Different Glutamate Transporter Subtypes

The general mechanisms of substrate transport, identified through studies of uptake of radiolabeled substrates or voltage-clamp recordings of transporter currents under steady-state conditions, are roughly similar among different glutamate transporter types. For all of them, the transport efficiency is in the order of ∼50%, which means that − perhaps surprisingly − these molecules translocate into the cell cytoplasm only 50% of all glutamate molecules they bind (Tzingounis and Wadiche, 2007). The remaining 50% of glutamate molecules that unbind from the transporters are released back in the extracellular space, and ultimately bind to nearby glutamate receptors, neuronal or glial transporters. This low transport efficiency can lead to an apparently paradoxical prolongation of glutamate lifetime in the extracellular space (Scimemi et al., 2009). Since the rates of glutamate binding to receptors and transporters are relatively similar, the likelihood that a glutamate molecule unbound from a transporter will eventually bind to a receptor or another transporter depends on the relative abundance of receptors compared to that of transporters. Typically, the surface density of expression of transporters in glial membranes is four orders of magnitude higher than that of glutamate receptors in extra-synaptic neuronal membrane (Lehre and Danbolt, 1998). For this reason, the most likely ultimate fate of glutamate molecules unbound from transporters is to be bound again by other transporter molecules (Leary et al., 2011).

Non-rate-limiting partial reaction steps, obtained by perturbing the steady-state and by subsequently following the kinetics of the relaxation to a new steady-state (i.e., the pre-steady-state kinetics), are similar for EAAT1-3 (Grewer and Rauen, 2005). However, the pre-steady-state kinetics and steady-state turnover rats are significantly slower for EAAT4 (Mim et al., 2005). There are also notable differences in the steady state affinity of different transporter types (km EAAT1: 48 ± 10 μM; EAAT2: 97 ± 4 μM; EAAT3: 62 ± 8 μM; EAAT4: 0.6 μM; Wadiche and Kavanaugh, 1998; Grewer et al., 2000; Bergles et al., 2002), and in the ratio of substrate transport versus anion permeation (Arriza et al., 1994; Seal and Amara, 1999; Mim et al., 2005; Torres-Salazar and Fahlke, 2007). Interestingly, and in contrast to EAAT1-3, the apparent affinity for glutamate is voltage dependent for EAAT4 and increases with negative voltages, suggesting higher glutamate buffering capacity for EAAT4 than other glutamate transporters (Mim et al., 2005). The fact that EAAT4 has a 10-fold higher affinity for glutamate but a 10-fold slower translocation rate than other transporters has led to hypothesize that the main functional role of EAAT4 is accounted for by its ability to generate a stoichiometrically uncoupled anion current (Fairman et al., 1995; Lin et al., 1998). Others have suggested that these biophysical properties would allow EAAT4 to clear glutamate away from synapses, where its concentration is lower than at the outer boundary of the synaptic cleft (Mim et al., 2005). Consistent with this hypothesis, one of the most prominent roles of EAAT4 is to limit metabotropic glutamate receptor activation in cerebellar Purkinje cells, in sub-cellular domains where the density of expression of these receptors and EAAT4 are both high (Wadiche and Jahr, 2005).

Glutamate transport via EAAT4 has a unique voltage-dependence. Its maximum transport activity is detected at –20 mV < Vm < 0 mV and the transporter inactivates at more negative membrane potentials (Mim et al., 2005). Membrane hyperpolarization promotes glutamate transport via other glutamate transporters, which have reversal potentials of 9.3 ± 0.7 mV (EAAT1), >80 mV (EAAT2) and 38.0 ± 2.7 mV (EAAT3) (Arriza et al., 1994). At hyperpolarized potentials, not only transport, but also the anion conductance of EAAT4 is inhibited (Mim et al., 2005). This means that at membrane potentials close to the resting potential of neurons, glutamate is bound strongly to all transporters, but its transport via EAAT4 is inhibited (Mim et al., 2005).

There are differences in the sodium requirement for activation of the anion conductance between neuronal and glial glutamate transporters (Wadiche et al., 1995a; Grewer et al., 2000, 2001; Otis and Kavanaugh, 2000). For EAAT3, the anion conductance can be activated by glutamate and Na+ ions from both sides of the membrane (Watzke and Grewer, 2001). The activation of the anion conductance by sodium alone has only been demonstrated for EAAT3-4, while EAAT1-2 mediate glutamate- and sodium-independent anion conducting states (Divito et al., 2017). For EAAT4-5, the anion conductance is particularly large compared to their glutamate transport capacity (Sonders and Amara, 1996; Seal et al., 2001). Consequently, transport currents generated by EAAT1-3 are easily measured experimentally using heterologous expression systems, whereas those mediated by EAAT4-5 are relatively small (Wadiche et al., 1995a; Grewer et al., 2001; Mitrovic et al., 2001; Watzke et al., 2001).

The existence of functional differences in the properties of glutamate transporter subtypes indicates that the function of these molecules is much more complex than previously thought, and that an evaluation of the physiological implications of glutamate transporters cannot bypass an understanding of the biophysical properties of these molecules in their native environments.

The Surface Mobility and Cellular Distribution of EAAT2

Out of all glutamate transporter types, EAAT2 has the highest density of expression in the adult brain, and is responsible for the largest proportion of glutamate transport (Minelli et al., 2001). In astrocytes, 70–75% of all EAAT2 is expressed on the plasma membrane (Michaluk et al., 2020). In the hippocampus, cerebellum and neocortex, EAAT2 is expressed mostly in astrocytic processes, in the vicinity and at a distance from the synaptic cleft (Danbolt et al., 1992; Levy et al., 1993; Rothstein et al., 1994; Torp et al., 1994; Chaudhry et al., 1995; Lehre et al., 1995). EAAT2 exists in at least three splice variants which differ only in their C-terminal domain. EAAT2a is the predominant variant, and represents ∼90%, whereas EAAT2b and EAAT2c represent ∼6% and ∼1% of total EAAT2 protein, respectively (Chen et al., 2002; Rauen et al., 2004; Holmseth et al., 2009). These variants have similar regional distribution and although they are all primarily expressed in astrocytes, they have also been detected in neurons (Schmitt et al., 2002; Chen et al., 2004; Maragakis et al., 2004; Bassan et al., 2008; Gonzalez-Gonzalez et al., 2008a; Holmseth et al., 2009). EAAT2a is more abundantly expressed than EAAT2b, and is a presynaptic glutamate transporter (Berger et al., 2005; Rimmele and Rosenberg, 2016). Within the hippocampus, 14–29% of axon terminals express EAAT2a (Chen et al., 2004). Though EAAT2a and EAAT2b share similar functional properties, they differ for the sequence of their extreme intracellular C-terminus (Chen et al., 2002; Sullivan et al., 2004). Unlike EAAT2a, EAAT2b has a PDZ binding domain that makes it capable of binding to proteins like PICK1, PSD95, and DLG1 (Bassan et al., 2008; Gonzalez-Gonzalez et al., 2008a; Underhill et al., 2015). These, in turn, can alter the currents mediated by the EAAT2b transporter (Sogaard et al., 2013). EAAT2a forms heteromers with EAAT2b, which have been proposed to stabilize EAAT2a around synapses (Haugeto et al., 1996; Peacey et al., 2009).

EAAT2 is highly mobile on the plasma membrane of astrocytes (D = 0.15–0.23 μm2/s) (though a lower value of D = 0.039 μm2/s has been measured using quantum dots; Al Awabdh et al., 2016), and the rate with which it diffuses along the plasma membrane can be increased by glutamate binding to EAAT2 or to AMPA, NMDA and metabotropic glutamate receptors, and can be reduced by blocking glutamate binding to EAAT2 (Benediktsson et al., 2012; Murphy-Royal et al., 2015; Michaluk et al., 2020). The surface diffusion of the EAAT2 variants EAAT2a and EAAT2b is more confined at astrocytic processes proximal to synapses, especially for EAAT2b (Al Awabdh et al., 2016). Consistent with previous work, glutamate increases the surface mobility of both EAAT2 variants, whereas blocking synaptic activity reduces it (Al Awabdh et al., 2016). The more confined diffusion of EAAT2b may be attributed to the presence of the PDZ domain, which allows EAAT2b to interact with scaffolding proteins and anchor it to macromolecular complexes that astrocytes form in subcellular domains opposite to neuronal presynaptic terminals (Al Awabdh et al., 2016). There is also evidence that membrane raft association regulates the targeting and function of glutamate transporters, especially for EAAT2 (Butchbach et al., 2004).

Glutamate transporters have been suggested to become incorporated into the plasma membrane through exocytosis (Cheng et al., 2002; Chowdhury et al., 2002; Robinson, 2002; Fournier et al., 2004; Karylowski et al., 2004; Zeigerer et al., 2004), and several components of this molecular machinery have been identified in astrocytes (Parpura et al., 1995; Hepp et al., 1999; Zhang et al., 2004). Consistent with these findings, calcium-dependent exocytosis of EAAT2 has been detected using FM dyes in cultured astrocytes (Stenovec et al., 2008). The same experiments showed that EAAT2 has a punctate distribution along the astrocytic plasma membrane, suggesting that the glutamate uptake capacity of these cells varies depending on the local density of expression of glutamate transporters (Stenovec et al., 2008). Interestingly, recent work indicate that the lifetime of EAAT2 on the astrocyte plasma membrane is ∼22 s (Michaluk et al., 2020). Whereas lateral diffusion could serve as a mechanism for transporter turnover away from synapses, where the diffusivity of these molecules is higher, endo/exocytosis from/to the plasma membrane to intracellular organelles could contribute to transporter turnover in astrocytic processes nearby synapses, where the diffusivity of these molecules is lower (Michaluk et al., 2020).

Although 80–90% of EAAT2 is localized in astrocytes, there is a small proportion (5–10%) in neuronal axon terminals (Rauen and Kanner, 1994; Torp et al., 1994; Schmitt et al., 1996; Torp et al., 1997; Chen et al., 2004; Furness et al., 2008; Melone et al., 2009, 2011, 2019). In neurons, EAAT2 co-localizes with the α1 and α3 isoforms of the Na+/K+-ATPase, but this co-localization is looser than that with the α2 isoform in astrocytes, suggesting a less efficient interaction between EAAT2 and the Na+/K+-ATPase in neurons (Melone et al., 2019). Neuronal EAAT2 appears to be required to provide glutamate to synaptic mitochondria, and is therefore linked to energy metabolism (Petr et al., 2015; Fischer et al., 2018; McNair et al., 2019, 2020; Sharma et al., 2019). By contrast, astrocytic EAAT2 is crucial to ensure survival, resistance to epilepsy, and prevent cognitive decline. Loss of neuronal and astrocytic EAAT2 have both implications on long-term memory and spatial reference learning, but the time scale over which they exert these effects is different (Sharma et al., 2019). Loss of astrocytic EAAT2 leads to early deficits, whereas loss of neuronal EAAT2 leads to late-onset deficits in long-term memory and spatial reference learning (Sharma et al., 2019). These findings are important because they identify neuronal and astrocytic EAAT2 as contributing to different aspects of cognitive function and, potentially, as different therapeutic targets in cognitive decline (Petr et al., 2015; Fischer et al., 2018; McNair et al., 2019, 2020; Sharma et al., 2019).

Multiple studies have shown that there are interesting relationships between astrocytic coverage, glutamate transporter expression and synaptic size (Ventura and Harris, 1999; Genoud et al., 2006; Witcher et al., 2007; Lushnikova et al., 2009; Witcher et al., 2010; Patrushev et al., 2013; Medvedev et al., 2014; Gavrilov et al., 2018; Herde et al., 2020). For example, in the rodent hippocampus, only 40–60% of synapses have astrocytic processes, and these cover only 53% of their perimeter (Ventura and Harris, 1999; Witcher et al., 2007, 2010). The size of a spine correlates with its release probability (Schikorski and Stevens, 1997). Astrocytes are closer to smaller spines (Medvedev et al., 2014), but the overall astrocytic coverage does not change with spine size (Gavrilov et al., 2018). Although there is less EAAT2 at smaller spines, its density is higher (Herde et al., 2020). These findings have important implications for understanding how glutamate transporters control spillover at synapses with different size (Scimemi et al., 2004), and the implications of this phenomenon on synaptic and astrocyte plasticity (Wenzel et al., 1991; Matsusaki et al., 2001; Bernardinelli et al., 2014; Perez-Alvarez et al., 2014).

Transcriptional, Translational, and Post-translational Regulation of Glutamate Transporters

Despite current agreements on the presence of EAAT2 in neurons, in the 1990s, some groups failed to find EAAT2 proteins in subsets of cortical neurons (Rothstein et al., 1994; Chaudhry et al., 1995; Lehre et al., 1995), in contrast to others (Torp et al., 1994; Schmitt et al., 1996; Torp et al., 1997). The discrepancy between these findings was attributed to the existence of a post-transcriptional and post-translational control of EAAT2 expression (Gegelashvili and Schousboe, 1997; Anderson and Swanson, 2000; Plachez et al., 2000). In fact, like most membrane proteins, glutamate transporters can be regulated at the gene expression, protein targeting and trafficking, and post-translational level.

The gene structure and organization of most glutamate transporters identified in the late 1990s showed that the promoter regions are highly conserved between mouse and human (Hagiwara et al., 1996; Stoffel et al., 1996). These regions do not contain a TATA box but a GC box and, in humans, an E box (Martinez-Lozada et al., 2016). The genes encoding different glutamate transporters contain different excision/splicing sites for different exons, which can generate variant mRNA species and proteins. Accordingly, multiple studies have reported the existence of several mRNA size classes encoding EAAT1 and EAAT3 (Pines et al., 1992; Tanaka, 1993; Mukainaka et al., 1995; Nakayama et al., 1996; Palos et al., 1996; Gegelashvili and Schousboe, 1997). This might be due to differences in polyadenylation and the existence of mRNAs with different coding capacity (Gegelashvili and Schousboe, 1997).

When glutamate transporters and their human homologs were first cloned, knowledge on the elements regulating their transcription largely relied on information collected from ASCT1, a neutral amino acid transporter of the same family of high affinity transporters as glutamate transporters (Hofmann et al., 1994). ASCT1 has 39–44% amino acid sequence identity, similar hydropathy profiles, trans-membrane organization and conservation of crucial function-related motifs compared to EAAT2 (Gegelashvili and Schousboe, 1997). According to these initial reports, the promoter region of EAAT2 was thought to contain at least five consensus sequences for the Sp1 transcription factor (krox24, krox20, Egr3, NGFI-C), involved in cell differentiation (Hofmann et al., 1994; Gegelashvili and Schousboe, 1997). By the early 2000’s, cloning and bioinformatics works established that the promoter region for EAAT1-2 is highly conserved: not only does it not have a TATA box, but lacks well-defined cis-elements and contains five consensus sequences for Sp1 and GC-rich repeats, also found in humans (Su et al., 2003). The search for regulatory transcription factors has led to the identification of the Nuclear Factor of Activated T cells (NFAT), the N-myc proto-oncogene protein (N-myc), and the Nuclear Factor κB (NF-κB), and a consensus NF-κB binding sequence in the 5′-UTR region of the Slc1a2 gene in humans (Meyer et al., 1996; Su et al., 2003). The Tumor Necrosis Factor α (TNFα), a cytokine involved in the acute phase of inflammatory reactions, decreases EAAT2 mRNA expression by increasing NF-κB activation (Su et al., 2003), perhaps by promoting NF-kB interactions with other transcription factors (Sitcheran et al., 2005). The ability of transcription factors to regulate gene expression can change the sensitivity of glutamate transporter expression to other regulatory proteins. For example, TNFα regulates the activity of the Yin Yang 1 (YY1) transcription factor which, when bound to the EAAT2 promoter, changes the effect of NF-κB from activation to suppression (Karki et al., 2014). In turn, NF-κB regulates YY1 expression. These findings suggest the existence of complex interplays between NF-κB and YY1 for the transcriptional regulation of EAAT2 expression.

Over the last two decades, other groups have identified factors, cis-regulatory elements and epigenetic mechanisms regulating EAAT1-2 transcription, but many unknowns remain about the molecular mechanisms regulating EAAT3-5 expression. In Table 1, we provide a summary of current studies on modulation of glutamate transporters listed in PubMed. One can easily note that in many cases, the results are conflicting. One may argue that inconsistencies are inevitable when studying a given transporter in different cell types and animal species. However, in some cases these considerations do not allow to resolve conflicting results from different laboratories. For this reason, in this review, we limit our discussion to forms of modulation of glutamate transporters for which some consensus exists.

TABLE 1
www.frontiersin.org

Table 1. Regulating factors of glutamate transporter uptake and expression.

There is a general agreement on the fact that in cultured neurons, PACAP, cAMP, and PKA-dependent pathways increase EAAT1 expression (Martinez-Lozada et al., 2016) and uptake (Hertz et al., 1978; Gegelashvili et al., 1996). Presumably this effect is mediated by activation of the transcription factor cAMP-response element binding protein (CREB), but the transcription factors or the cis elements of the promoter responsible for this effect have not been identified (Martinez-Lozada et al., 2016). This form of transcriptional regulation may provide a pathway through which activation of G-protein membrane receptors coupled to cAMP and PKA-dependent signaling pathways indirectly affect glutamate uptake, as it has been shown in other contexts for D2 dopamine receptors or α1 and β-adrenergic receptors (Kerkerian et al., 1987).

Stable epigenetic alterations of glutamate transporter gene expression are heritable in the short term, but do not involve DNA mutations. These include methylation and histone modifications, and there are multiple CpG regions in the promoter for EAAT2 where methylation can occur (Zschocke et al., 2005). One of the consequences of methylation is that it alters the ability of glucocorticoids to change EAAT2 expression. Accordingly, in the cerebellum, where the EAAT2 promoter is hyper-methylated, glucocorticoids are unable to change EAAT2 expression (Zschocke et al., 2005). By contrast, in the forebrain, where the EAAT2 promoter is hypo-methylated, glucocorticoid up-regulate EAAT2 expression (Zschocke et al., 2005).

Phosphorylation and glycosylation are two documented forms of post-translational modifications for glutamate transporters (Gegelashvili and Schousboe, 1997). Accordingly, increasing protein kinase C (PKC) activation by phorbol esters leads to increased phosphorylation of EAAT2 at the residue Ser113 and increased glutamate uptake (Roginski et al., 1993). N-glycosylation promotes EAAT3 expression, but has no effect on EAAT1 (Conradt et al., 1995; Ferrer-Martinez et al., 1995).

Arachidonic acid, produced via activation of NMDA receptors in neurons and metabotropic glutamate receptors in astrocytes, can directly interact with different types of glutamate transporters, with different effects. For example, it reduces glutamate uptake via EAAT1, enhances glutamate uptake via EAAT2 and leads to a moderate increase of glutamate uptake via EAAT3 (Chan et al., 1983; Volterra et al., 1992; Trotti et al., 1995). Arachidonic acid can lead to the activation of PKC, promoting glutamate uptake via EAAT2. Metabolites of arachidonic acid can be a source of reactive oxygen species (ROS), which inhibit glutamate uptake but increase the transporters’ steady state affinity for glutamate (Volterra et al., 1994b).

There is a growing awareness that these forms of regulation are complex not only because they likely differ among species, cell types and brain regions, but also because they interact with one another in ways that can be difficult to reproduce in reduced preparations but that affect the function of these transporters in vivo. These currently unknowns are likely going to be addressed by using experimental approaches that allow manipulation of different regulation factor in situ, and in a cell-specific manner.

Activity-Dependent Modulation of Glutamate Transporter Trafficking

The first observation that glutamate itself can modulate its uptake came from data showing that glutamate uptake is increased in astrocyte cultures supplemented with conditioned media from neuronal cultures (Drejer et al., 1983; Voisin et al., 1993). Pure astrocytic cultures only express EAAT1, but co-culture of neurons and astrocytes increases EAAT1 expression and induces EAAT2 expression (Gegelashvili and Schousboe, 1997; Swanson et al., 1997). This effect depends on p42/44 MAP kinases activation via the tyrphostin-sensitive Receptor Tyrosine Kinase (RTK) signaling pathway, and is abolished by inhibitors of PI3K, tyrosine kinase and NF-kB (Swanson et al., 1997; Zelenaia et al., 2000). Similarly, decreased glutamate uptake via EAAT1-2 (not EAAT3) occurs in response to axotomy of glutamatergic neurons of cortical lesions (McGeer et al., 1977; Shifman, 1991; Ginsberg et al., 1995; Levy et al., 1995). Together, these forms of activity-dependent regulation of EAAT2 expression allow this transporter to be more abundant and less mobile in astrocytic processes close to active glutamatergic synapses, an effect that can provide an effective strategy to limit glutamate spillover away from the synaptic cleft.

Similarly, to glutamate, glutamate transporter substrates like D-Aspartate and ligands like L-trans-PDC and TBOA can also produce a redistribution of EAAT1 on the cell membrane (Shin et al., 2009). Consistent with these findings, treating astrocyte cultures with kainate, dbcAMP or AMPA receptor agonists increases D-aspartate uptake and EAAT1 protein expression (Gegelashvili et al., 1996). It is unclear whether these effects are due to binding to AMPA receptors or are due to release of diffusible molecules (e.g., arachidonic acid, diacylglycerol, nitric oxide) through more complex intracellular signaling cascades (Gegelashvili and Schousboe, 1997).

Physiological Roles of Glutamate Transporters

The role of glutamate transporters in regulation of phasic and tonic extracellular glutamate levels is critical for neuronal signaling and controlling excitotoxicity (Rothstein et al., 1996). Spillover of glutamate and inter-synaptic cross-talk have been associated with multiple neuropsychiatric disorders, including schizophrenia, epilepsy, addiction, depression and obsessive compulsive disorder (O’Donovan et al., 2017; Bellini et al., 2018; Malik and Willnow, 2019). Knock-out mouse models of different subtypes of glutamate transporters have revealed major motor deficits with decreased levels of glial transporters EAAT1-2 (Rothstein et al., 1996), with more subtle behavioral deficits with loss of the neuronal transporter EAAT3 (Peghini et al., 1997; Bellini et al., 2018).

In humans, decreased expression and reduced function of EAAT2 are associated with amyotrophic lateral sclerosis (ALS) (Rothstein et al., 1995; Bruijn et al., 1997; Rosenblum and Trotti, 2017). In transgenic mice expressing an N-terminal fragment of mutant huntingtin (R6/2), there is an age-dependent downregulation of EAAT2, which leads to a progressive increase in the extracellular glutamate (Behrens et al., 2002). Other glutamate transporters, however, remain unchanged, suggesting that EAAT2-mediated excitotoxicity might contribute to Huntington’s disease (Behrens et al., 2002). In the few identified polymorphisms of the gene encoding EAAT3, dicarboxylic aminoaciduria, a deficit in kidney function was prominent, as well as family linkage to schizophrenia and obsessive–compulsive disorders (OCD) (Bailey et al., 2011; Porton et al., 2013). In mice, loss or increased EAAT3 are associated with increased anxiety and OCD-like behaviors, suggesting that preserving an optimal expression of this transporter is key for the function of neuronal circuits affected by this disease (Zike et al., 2017; Bellini et al., 2018; Delgado-Acevedo et al., 2019). Significant lower levels of EAAT2 have also been reported in vitro (Scimemi et al., 2013) and in vivo, in animal models of Alzheimer’s disease (Wilson et al., 2003; Dabir et al., 2006; Mookherjee et al., 2011; Schallier et al., 2011; Hefendehl et al., 2016), as well as in humans affected by this disease (Li et al., 1997; Jacob et al., 2007; Scott et al., 2011; Leng et al., 2021), whereas promoting EAAT2 expression improves cognitive function (Fan et al., 2018). Several mutations in the gene that encodes EAAT1 have also been linked to the neurological disease Episodic Ataxia Type 6 (Choi et al., 2017; Chivukula et al., 2020). The most well-studied mutation (P290R) results in reduced glutamate transport activity and a large increase in the uncoupled anion conductance, the latter property being suggested to be responsible for the phenotype in studies using a Drosophila melanogaster model (Winter et al., 2012; Parinejad et al., 2016). Further understanding of how these transporters are regulated in specific circuits and synapses may allow for design of therapeutics that correct for specific deficits in transporter function.

GABA Transporters

In mammals, GABA transporters are classified into four subtypes, based on amino acid sequence homology and pharmacological properties. These include GAT1-3 and the betaine GABA transporter BGT1. Out of these, GAT1 and GAT3 account for the largest proportion of GABA uptake in the CNS, and for this reason, they will be the focus of our attention in this review. Structural information of the GABA transporters comes from prokaryote homologs such as LeuTAa (Figure 2; Yamashita et al., 2005). In the neocortex, GAT1 is expressed robustly in GABAergic axon terminals, astrocytic processes, oligodendrocytes and microglial cells (Fattorini et al., 2020). The expression of GAT1 is pronounced in the axon terminal of chandelier GABAergic neurons (known as “cartridges”), which provide inhibitory inputs to the axon initial segment of pyramidal cells (Woo et al., 1998). In young mice (P9), GAT1 is also transiently expressed somatically, but this somatic expression is lost in juvenile mice (P29) (Yan et al., 1997; Yan and Ribak, 1998). Presynaptic boutons in the cerebellum and hippocampus express 800–1,300 μm–2 GAT1 molecules, with a preferential perisynaptic localization (Chiu et al., 2002; Melone et al., 2015). These density values drop to 640 μm–2 GAT1 molecules along the length of the axon (Chiu et al., 2002), whereas the surface density of GAT1 in astrocytic membranes is 3.5 times higher than in axon terminals (Melone et al., 2015). GAT3 is mainly localized in peri-synaptic astrocytic processes, but has also been detected in brainstem and cortical neurons (Clark et al., 1992; Melone et al., 2003, 2005, 2015). These findings are important because they provide anatomical evidence to the fact that no GABA transporter can be described as being purely neuronal or glial, although they may have a preferential distribution in a given cell type depending on the age and brain region of different animal models.

FIGURE 2
www.frontiersin.org

Figure 2. Crystal structure of LeuTAA, a bacterial homolog of GABA transporters (PDB: 2A65). (A) Side view of the LeuTAa transporter parallel to the membrane, using a ribbon (top) and a surface representation sliced through the center of the transporter (bottom). (B) As in (A), viewed from the extracellular side of the membrane. Leu (magenta), sodium (blue), and chloride ions (green) are represented as spheres. (C) Summary of the main feature of the stoichiometric and non-stoichiometric current mediated by Leu and GABA transporters in prokaryotes and eukaryotes, respectively. Image generated using The PyMOL Molecular Graphics System (Version 2.3.2; Schrödinger, 2010).

Both GAT1 and GAT3 transporters translocate the zwitterion GABA across the membrane by coupling its movement to the co-transport of two Na+ and one Cl ion, leading to the net influx of one positive charge per transport cycle (1 GABA: 2 Na+: 1 Cl) (Radian and Kanner, 1983; Keynan and Kanner, 1988). In addition to this stoichiometric current, GABA transporters mediate an agonist-independent leak current carried by alkali ions, which can be detected in mammalian expression systems but not in Xenopus laevis oocytes (Cammack and Schwartz, 1996; Eckstein-Ludwig et al., 1999; Lu and Hilgemann, 1999; MacAulay et al., 2002; Matthews et al., 2009). This current can generate a local change in membrane voltage and/or membrane resistance, which can be more or less pronounced depending on the magnitude and direction of the driving force for the permeant ions and the local density of the transporters. Changes in membrane resistance are important because they act as a local shunt capable of hampering action potential propagation and cell excitability.

As in the case of glutamate transporters, we provide a summary of current studies on modulation of GABA transporters listed in PubMed (Table 2).

TABLE 2
www.frontiersin.org

Table 2. Regulating factors of GABA transporter Uptake and expression.

The Surface Mobility and Cellular Distribution of GAT1

The GABA transporter GAT1 has a membrane pool of 61–63% (Chiu et al., 2002). Fluorescence Recovery After Photobleaching (FRAP) experiments in neuroblastoma 2a cells show that 50% of membrane GAT1 is immobile, likely due to the existence of tight interactions between GAT1 and the actin cytoskeleton, mediated by the adaptor protein ezrin (Imoukhuede et al., 2009). Accordingly, depolymerizing actin or interrupting the GAT1 PDZ-interacting domain increases the transporter mobility, whereas depolymerizing microtubules does not (Imoukhuede et al., 2009). Numerous mechanisms contribute to cycle this pool of transporter to and from the plasma membrane. Specifically, there are two regions in the C-terminal domain of GAT1 that are responsible for supporting GAT1 export from the endoplasmic reticulum and for putting GAT1 under the control of the exocyst (Farhan et al., 2004; Moss et al., 2009). In addition to these, the MAGUK protein Pals1 is co-expressed with GAT1 in COS7 cells and contributes to stabilize GAT1 on the cell membrane (McHugh et al., 2004).

Regulation of GABA Uptake

The control of GABA uptake can be expressed through changes in the rate with which GABA transporters are trafficked and redistributed in the plasma membrane, with consequences on the number of transporter molecules available for binding extracellular GABA, or by modulating the biophysical properties of GABA transporters (e.g., Vmax, km, etc.). Neurons regulate the surface expression of GAT1 in parallel with that of extracellular neurotransmitter levels. GABA transporters interact with the SNARE proteins syntaxin 1A and Munc-18 through molecular interactions modulated by PKC (Beckman et al., 1998; Deken et al., 2000; Geerlings et al., 2001; Horton and Quick, 2001). These transporters are associated with presynaptic vesicles similar to synaptic vesicles, capable of undergoing clathrin-mediated internalization and endosomal sorting (Barbaresi et al., 2001). Despite being morphologically similar to synaptic vesicles, and despite the fact that they can be released at similar rates in a calcium dependent manner, they lack synaptophysin and vesicular GABA transporters (Deken et al., 2003). It has been suggested that these GABA transporter-containing vesicles might represent a population of endocytic or exocytic vesicles acting as cargos for the assembly of synaptic domains distinct from the active zone (Deken et al., 2003).

PKC modulation is important to change the number of functional GABA transporters expressed on the plasma membrane (Corey et al., 1994; Quick et al., 1997). In Xenopus laevis oocytes, PKC activation with PMA promotes GAT1 translocation to the cell membrane at low basal GAT1 expression levels, but this effect is not detected at high levels of expression of GAT1. The rate of GABA uptake can also be altered by PKC (Corey et al., 1994; Quick et al., 1997). Accordingly, inhibiting PKC reduces GABA uptake through a reduction of Vmax (not km), whereas inhibiting protein phosphatase 2B increases it. These forms of PKC-dependent modulation differ in cultured neurons and isolated nerve terminals, where surface expression of GAT1 is decreased by PKC dependent phosphorylation (Beckman et al., 1999; Wang and Quick, 2005; Cristovao-Ferreira et al., 2009).

Another protein kinase, PKA, stimulates GABA transport via GAT1 and activates a GAT1-mediated cationic current during opioid withdrawal (Bagley et al., 2005). This enhanced PKA signaling contributes to increase neuronal action potential firing rates in opioid-sensitive PAG neurons during opioid withdrawal (Bagley et al., 2005).

In the area CA1 of the rat hippocampus, chronic stimulation of cannabinoid receptors CB1/2 reduces GAT1 gene expression (Higuera-Matas et al., 2012). Conversely, an endogenous agonist of endocannabinoid receptors, 2-arachidonoylglycerol (2-AG), increases GABA uptake (Romero et al., 1998; Venderova et al., 2005). CB1 receptors have been suggested to interact or co-localize with β2 adrenergic receptors (Hudson et al., 2010), the activation of which also increases GAT1 expression and GABA uptake (Martins et al., 2018). The ability of β2 adrenergic receptors to promote GABA uptake (and that of β1 adrenergic receptors to inhibit it) are mediated by a PKA pathway controlled by cannabinoid receptors, because the modulation of GABA update by adrenergic receptors is inhibited by the cannabinoid receptor agonist WIN55,212-2 (Martins et al., 2018).

In astrocytes, GABA uptake via GAT1, not GAT3, is modulated by neurotrophic factors like BDNF (Vaz et al., 2011). This is due to the ability of BDNF to inhibit dynamin/clathrin-dependent constitutive internalization of GAT1, effectively increasing the lifetime of GAT1 on the cell membrane. The effect of BDNF is mediated by activation of a truncated form of the TrkB receptor, is coupled to a PLC-γ/PKC-δ and ERK/MAPK pathway, and requires activation of adenosine A2A receptors (Vaz et al., 2011). The cross talk between A2A receptors and BDNF is likely due to the fact that activation of A2A receptors activates TrkB and induces its translocation to lipid rafts (Tebano et al., 2008; Assaife-Lopes et al., 2010). This type of modulation differs in neurons, where BDNF can still inhibit GABA uptake via GAT1 in conjunction with A2A receptors, but here the BDNF modulation persists in the presence of A2A receptor antagonists or upon removal of extracellular adenosine (Vaz et al., 2008).

Physiological Roles of GABA Transporters

Genetic variants in the solute carrier family 6 member 1 (SLC6A1) gene, encoding GAT1, are associated with various neurodevelopmental disorders, including epilepsy with myoclonic atonic seizures, autism spectrum disorder and intellectual disability (Bhat et al., 2020; Goodspeed et al., 2020). Knockout mouse models of GAT1, as well as GAT1 inhibitors, have shown a range of physiological effects that indicate that GAT1, through its exquisite regulation of GABA in the brain, may be an interesting target for therapies for neuropsychiatric diseases (Salat and Kulig, 2011; Egawa and Fukuda, 2013; Bhat et al., 2020). GAT1 inhibitors, such as tiagabine, NO-711 and DDPM-2571, have anti-seizure, antinociceptive, antiallodynic and anxiolytic properties (Laughlin et al., 2002; Todorov et al., 2005; Pakulska, 2007; Xu et al., 2008; Salat et al., 2017). Loss of GAT1 in the nucleus accumbens has also been observed in mice treated with chronic social defeat stress paralleling observations in patients with major depressive disorder (MDD) (Heshmati et al., 2020).

During epileptic seizures, the ionic gradient that typically supports GABA uptake from the extracellular space can be dissipated or even inverted, promoting GABA release through the reversed activity of GABA transporters like GAT3 (Raiteri et al., 2002; Wu et al., 2003; Richerson and Wu, 2004; Kinney, 2005). Under these conditions, the reversal of GABA uptake could provide a useful mechanism to curtail seizure propagation. With some exceptions (Xie et al., 2017), single nucleotide polymorphisms in SLC6A11, the gene encoding human GAT3, have been detected in patients with antiepileptic drug resistance (Kim et al., 2011). It is possible that a dysfunction of GAT3 could alter both phasic and tonic GABAergic transmission in the epileptic brain.

GATs regulate neurotransmission in other complex ways. As mentioned above, substrate transport through GAT elicits an inward current that can directly excite neurons (Bagley et al., 2005, 2011). GAT1 also generates a sodium-dependent capacitive current that can also contribute to shunting (Mager et al., 1993). Efflux of GABA into the extracellular space contributes to tonic currents mediated by GABAA receptors in hippocampal neurons (Wu et al., 2007) and glia (Barakat and Bordey, 2002). Tonic currents are critical in neurodevelopment (Egawa and Fukuda, 2013) and are increased in pathological conditions, such as inflammation (Tonsfeldt et al., 2016). Regulation of GAT activity and trafficking has a dramatic effect on tonic currents, especially those mediated by extrasynaptic GABAA receptors (Scimemi et al., 2005; Scimemi, 2014a,b). In addition to controlling integration of inputs onto GABAergic neurons, these tonic currents control the coincidence detection window of excitatory inputs onto pyramidal neurons (Sylantyev et al., 2020) and contribute to regulation of dopamine release in the dorsal striatum (Roberts et al., 2020). GAT3 activity in astrocytes regulates release of ATP and adenosine that contributes to heterosynaptic depression in the hippocampus (Boddum et al., 2016), highlighting an additional mechanism for regulating synaptic activity.

Dopamine Uptake: One Transporter, Three Currents, Multiple Regulation Sites

Dopamine uptake via the membrane transporter DAT is stoichiometrically coupled to the co-transport of two Na+ and one Cl ion (Figure 3) (Krueger, 1990; McElvain and Schenk, 1992; Gu et al., 1994; Penmatsa et al., 2013, 2015). Voltage-clamp recordings of DAT-mediated currents in Xenopus oocytes, however, show that the mean net charge to dopamine ratio is significantly larger than the one predicted by the stoichiometric coupling ratio (Sonders et al., 1997; Ingram et al., 2002). This discrepancy can be accounted for by the fact that DAT, like glutamate and GABA transporters (as well as serotonin and norepinephrine transporters; Bruns et al., 1993; Cammack et al., 1994; Mager et al., 1994; Galli et al., 1995; Jayanthi et al., 2002), also mediates two stoichiometrically uncoupled conductances: one that requires dopamine binding to the transporter, and one that does not and is constitutively active. For simplicity, we refer to them as the uncoupled and the leak conductance, respectively.

FIGURE 3
www.frontiersin.org

Figure 3. Crystal structure of Drosophila melanogaster dDAT dopamine transporter (PDB: 4XP1). (A) Side view of the Drosophila dopamine transporter parallel to the membrane, using a ribbon (top) and a surface representation sliced through the center of the transporter (bottom). (B) As in A, viewed from the extracellular side of the membrane. L-Dopa (magenta), sodium (blue) and chloride ions (green) are represented as spheres. (C) Summary of the main feature of the stoichiometric and non-stoichiometric current mediated by LeuT (prokaryotic dopamine transporter homolog) and DAT transporters in eukaryotes, respectively. Image generated using The PyMOL Molecular Graphics System (Version 2.3.2; Schrödinger, 2010).

It is interesting to note that the concentration of dopamine required for half-maximal transport is ∼580 nM, whereas the one required for activation of the uncoupled current is only ∼35 nM (Ingram et al., 2002). The extracellular concentration of dopamine in the brain has been known to vary on a sub-second time scale due to phasic firing of dopaminergic neurons, and to be modulated by administration of substances of abuse. Until recently, fast-scan cyclic voltammetry (FSCV) has been considered the technique that provides the best combination of temporal resolution, sensitivity and chemical selectivity (Roberts et al., 2013), but even in this case it has been traditionally used to obtain relative, as opposed to absolute values of dopamine concentrations in the brain. In 2018, the development of cellular-scale probes have enabled stable recording of sub-second extracellular dopamine oscillations (Schwerdt et al., 2018). Although the recorded levels of dopamine were specific for each mouse, they ranged between 40–450 nM (Schwerdt et al., 2018). This means that the uncoupled DAT current is likely to be a main contributor to cell excitability and neurotransmitter release in DAT-expressing neurons. The uncoupled DAT-mediated current is carried by chloride and, surprisingly, its activation increases cell excitability in cultured dopaminergic neurons (Ingram et al., 2002; Carvelli et al., 2004).

In physiological conditions, the leak current is carried by K+ and perhaps H+ ions. This leak current is voltage-dependent, outward-rectifying and reverses at ∼–10 mV (Sonders et al., 1997). Although its activation does not require ion or agonist binding to the transporter, it can be blocked by dopamine and other DAT ligands. The ability of substrates to block the leak current is not consistently observed among other members of the Na+/Cl-dependent cotransporter family, with the exception of GAT1 (Mager et al., 1994; Cammack and Schwartz, 1996; MacAulay et al., 2002; Kanner, 2003) and the serotonin (Mager et al., 1994) and norepinephrine transporters (Galli et al., 1995). These findings identify multiple mechanisms through which dopamine transporters can change the membrane potential and the signaling properties of neurons through mechanisms that are distinct from its ability to take up dopamine from the extracellular space.

A summary of current studies on modulation of dopamine transporters listed in PubMed is provided in Table 3.

TABLE 3
www.frontiersin.org

Table 3. Regulating factors of dopamine transporter Uptake and expression.

The Pharmacology of Dopamine Transporters

DAT is a target of psychostimulants like cocaine and amphetamine. Cocaine binds to DAT and increases extracellular dopamine levels by blocking its transport activity (Chen and Reith, 2000; Norregaard and Gether, 2001; Goldberg et al., 2003; Torres et al., 2003). In contrast, amphetamine is a substrate for this transporter (Kahlig et al., 2005). The inward transport of amphetamine increases the number of inward-facing transporter binding sites, which results in an increased rate of dopamine efflux through what at first sight may look like an exchange process (Sulzer et al., 1995; Kahlig et al., 2005). In reality, this phenomenon may be more complex, as evidenced by the fact that N-terminus phosphorylation of DAT alters only the amphetamine-induced dopamine efflux without altering dopamine uptake (Khoshbouei et al., 2004). Dopamine and amphetamine are both substrates for DAT, and compete with each other for binding to the transporter (Sulzer et al., 1993; Hoffman et al., 1998). They can both trigger dopamine efflux through reversed uptake, by inverting the electrochemical gradient for Na+ and Cl (Sitte et al., 1998). Whereas dopamine inhibits the DAT channel-like behavior, amphetamine activates it. This effect is not due to differences in the ability of dopamine and amphetamine to change the intracellular Na+ concentration, because it can still be detected in outside-out patches, where the intracellular Na+ concentration is controlled (Kahlig et al., 2005).

The Surface Mobility and Cellular Distribution of Dopamine Transporters

Dopamine transporters are monoamine neurotransmitter transporters located in the pre-synaptic terminal of dopaminergic neurons, away from the synaptic area (Nirenberg et al., 1997). These cells are located in the ventral tegmental area and substantia nigra, project to the cingulate and medial pre-frontal cortex, nucleus accumbens, olfactory tubercle, lateral habenula, and striatum. Through their axonal projections, dopaminergic neurons control motivation, reward reinforcement and movement, in addition to attention, planning and memory. Therefore, the regulation of DAT function has extensive functional implications for all these behaviors. It is now evident that the role of DAT is not limited to regulating the extracellular concentration of dopamine but is also involved in the homeostatic maintenance of presynaptic function (Torres et al., 2003). Like other membrane proteins, DAT is associated with intracellular proteins that ensure the appropriate location of the transporter in specific domains of the cell membrane at a given time. Accordingly, DAT has been shown to bind to the SNARE protein complex syntaxin 1A (Lee et al., 2004), the PDZ domain protein PICK1 (Torres et al., 2001; Bjerggaard et al., 2004), CaMKII (Fog et al., 2006), and the multiple-LIM-domain-containing adaptor protein HIC-5 (Carneiro et al., 2002).

DAT is equally distributed in and out of rafts (Foster et al., 2008), nanometer-wide and temporally dynamic lipid microdomains enriched in cholesterol, sphingolipids and glycosylphosphatidylinositol (GPI)-anchored proteins (Hancock, 2006). Structural studies of the Drosophila melanogaster DAT (dDAT) reveal that cholesterol can bind in a crevice formed by transmembrane domains 5, 7, and 1a, which is thought to prevent the conformational changes required for the transporter to transition from outward-facing to inward-facing (Penmatsa et al., 2013, 2015). This is in agreement with functional studies that show DAT can be regulated by cholesterol, due to its ability to promote an outward-facing conformation of the transporter (Hong and Amara, 2010; Jones et al., 2012).

By using Fluorescence Correlation Spectroscopy (FCS) and FRAP, Adkins et al. (2007) measured the surface diffusion coefficient of DAT in two types of cells: HEK293 and N2a cells. Due to differences in laser beam waist and sampling areas between FRAP and FCS, FCS is better suited to detect fast protein movements within a confined domain, whereas FRAP allows to detect long-range diffusion between domains in the membrane (Adkins et al., 2007). In HEK293 cells, DAT diffuses with a diffusion coefficient of 3.6 × 10–9 cm2/s, consistent with a relatively freely diffusible protein. This diffusion coefficient is lower in N2a cells, where DAT is partially immobilized (D < 10–10 cm2/s). This difference is likely due to the existence of cell-specific direct or indirect interactions between DAT and cytoskeletal proteins or with membrane rafts. Accordingly, single particle tracking studies confirm that the median diffusion coefficient for DAT is 1.6 × 10–10 cm2/s in Flp-In 293 cells (Kovtun et al., 2015).

Overall, these findings indicate that DAT transporters diffuse substantially slower and in a more confined manner than the glutamate transporter GLT-1, probably because of the presence of different types of protein interactions in different membrane transporters or because of the presence/lack of specific interacting substrates (Murphy-Royal et al., 2015).

Transcriptional, Translational, and Post-translational Regulation of Dopamine Transporters

The human DAT gene was first cloned in 1991, and it is localized to chromosome 5p15.3 and has a single transcriptional start site (Kilty et al., 1991; Vandenbergh et al., 1992a,b). DAT has a half-life of about 2 days on the cell membrane, suggesting the existence of dynamic processes of transcriptional and translational regulation (Kimmel et al., 2000; Kahlig and Galli, 2003). Transcription factors like Nurr1 and Pitx3 have an expression pattern that match that of dopaminergic neurons, and are known to be crucial for the development, survival and maintenance of midbrain dopaminergic neurons (Lee et al., 2010; Rodriguez-Traver et al., 2016; Salemi et al., 2016). Accordingly, disrupting the Nurr1 gene alters the development of dopaminergic neurons (Zetterstrom et al., 1997; Castillo et al., 1998; Saucedo-Cardenas et al., 1998). Nurr1 binds with high-affinity to an NGFI-B responsive element within the promoter regions of DAT and other dopamine-related genes (Sacchetti et al., 1999). The ability of Nurr1 to increase DAT expression, however, relies on mechanisms that are independent of the NGFI-B responsive element. There are a variety of sequence motifs identified through in silico studies, which point to the fact that DAT can be regulated epigenetically via DNA methylation and histone acetylation in vitro and in vivo (Wang et al., 2007). Like other housekeeping genes, the DAT gene lacks conserved TATA and CAAT boxes (confirming that DAT is susceptible to regulation by histone acetylation), and its core promoter is GC-rich (confirming that DAT expression can be regulated via DNA methylation) (Choi and Kim, 2008).

These discoveries prompted a number of biochemical and mutagenesis studies on heterologous expression systems, which led to the identification of key structural features and functional domains of the transporter. A major breakthrough occurred when the first X-ray crystal structure of the bacterial leucine transporter LeuTAa, which shares 20–25% homology with all monoamine transporters was solved at 1.65 Å resolution (Yamashita et al., 2005). This was followed by the structure of Drosophila melanogaster DAT, which shares 50–55% homology with monoamine transporters (Penmatsa et al., 2013, 2015; Wang et al., 2015). Together, these studies suggest that all monoamine transporters have 12 α-helix spanning domains and an alternating access substrate translocation mechanism (Kristensen et al., 2011). Despite the structural similarity of the core transmembrane regions of all monoamine transporters, their extracellular loops, N- and C-termini differ significantly in length and sequence (Kristensen et al., 2011). This is important because these regions are the site of post-translational modifications and can be the site of protein–protein interactions that control transporter localization, stability and activity (Aggarwal and Mortensen, 2017). Post-translational modifications, binding partner interactions, modulation by cholesterol and membrane raft associations are all capable of modulating DAT activity and ultimately dopamine clearance from the extracellular space.

The N-terminus of the DAT protein is subject to phosphorylation and ubiquitination (Karam and Javitch, 2018). Although five serine residues at positions 2, 4, 7, 12, 13 have been identified as targets for phosphorylation by PKC, the only verified phosphorylation site is Ser7 (Moritz et al., 2013). The localization of these sites at the distal end of a long and flexible domain suggests that these residues may be also regulated by partner interactions, but these effects have not been demonstrated. A second verified phosphorylation site is Thr53, which is followed by a Pro residue, making it specific for proline-directed kinases such as ERK. Thr53 is also flanked by an SH3 domain, which is a ligand for protein scaffolding (Saksela and Permi, 2012). In between these two phosphorylation sites, Lys27 is a residue that undergoes ubiquitination catalyzed by the ubiquitin E3 ligases Nedd4-2 and Parkin. This modification is increased by PKC activation and contributes to promote DAT endocytosis (Hong and Amara, 2010; Vina-Vilaseca and Sorkin, 2010). The N-terminal domain also contains binding sites for the regulatory partners syntaxin 1 (res 1–33, which reduces dopamine uptake) and D2 dopamine receptors (res. 1–15) (Lee et al., 2007; Binda et al., 2008; Carvelli et al., 2008). On the C-terminal domain of the DAT protein, Cys580 provides an S-palmitoylation site, and a FREK motif at residues 587–590 binds the small Ras-like GTPase Rin1 and contributes to PKC-mediated endocytosis (Boudanova et al., 2008b; Navaroli et al., 2011). Other regulatory domains in the C-terminal region include binding sites for CaMK (res. 612–617) and α-synuclein (res. 606–620) (Lee et al., 2001; Moszczynska et al., 2007). The C-terminal domain also contains a PDZ domain-binding sequence, where interactions with scaffolding proteins like PICK1 occur. Notably, these interactions are neither necessary nor sufficient for surface targeting of DAT, and have yet unidentified functional consequences (Bjerggaard et al., 2004). Other members of the “DAT interactome” include PP2Ac (Baumann et al., 2000), Hic-5 (Carneiro et al., 2002), RACK (Lee et al., 2004), D2 dopamine receptors (Bolan et al., 2007), flotillin-1 (Cremona et al., 2011), Rin (Navaroli et al., 2011), and the k-opioid receptor (Kivell et al., 2014; Bermingham and Blakely, 2016).

The regulatory mechanisms described so far make it easy to spot that protein kinases like PKC, one of the best characterized regulatory proteins for DAT, can regulate dopamine uptake through a variety of processes, including an endocytotic mechanism driven by the phosphorylation of DAT accessory proteins, a kinetic down-regulation mediated by Ser7 phosphorylation, and an increased dopamine efflux mediated by altered surface transporter activity. Together, these modifications allow PKC to reduce dopamine uptake. The actions of PKC are opposite to those induced by palmitoylation, which lead to reduced DAT degradation (Chen et al., 2013). It is important to keep in mind that the primary sites of PKC-stimulated phosphorylation are the membrane rafts, the microdomains rich in cholesterol and sphingolipids. Therefore, changing PKC-mediated phosphorylation of DAT can affect its raft distribution and protein interactions. In contrast to PKC, ERK provides a tonic mechanism to increase dopamine uptake perhaps via phosphorylation of Thr53 (Foster et al., 2012). In addition to PKC and ERK, there is a host of other kinases that are capable of regulating DAT activity, including PKA, PKG, CaMKII, MAPK, PI3K/Akt and tyrosine kinases. PKCβ may also be involved in the mechanism of D2 receptor regulation of DAT by ERK (Chen et al., 2013).

Physiological Roles of Dopamine Transporters

DAT regulates extracellular concentrations of the neuromodulator dopamine and thus, subsequent activation of dopamine receptors that can enhance or inhibit neurons. DAT is the target of psychoactive and psychotherapeutic drugs such as methylphenidate and amphetamines that are used in treatment of attention deficit and hyperactivity disorder (ADHD) symptoms, as well as drugs of abuse (Schmitt et al., 2013; German et al., 2015). Genetic variants of SLC6A3, the gene encoding DAT in humans, have been identified in patients with neuropsychiatric, neurodevelopmental and neurodegenerative disorders (Mazei-Robison et al., 2005; Serretti and Mandelli, 2008; Mick and Faraone, 2009; Sakrikar et al., 2012; Bowton et al., 2014; Hansen et al., 2014; Mergy et al., 2014; Herborg et al., 2018; Campbell et al., 2019; DiCarlo et al., 2019). Studies of these variants have observed deficits in uptake, transporter-associated currents, trafficking and regulation (Gowrishankar et al., 2014). Further studies of these mutants should give more insight into the complex regulation of dopamine-modulated circuits. An intriguing new area is the role that DAT plays in transport of ligands that act on the trace amine-associated receptor subtype 1 (Taar1), an intracellularly localized G protein-coupled receptor that can, in turn, regulate internalization of both the DAT and EAAT3 in dopaminergic neurons (Underhill et al., 2014, 2019), enhancing extrasynaptic glutamate signaling (Li et al., 2017). Taar1 is known to regulate monoaminergic signaling and dysregulation of TAAR1 signaling may play important roles in neuropsychiatric disorders (Schwartz et al., 2018; Dodd et al., 2020).

Conclusion

Transporters for excitatory, inhibitory, and modulatory neurotransmitters like glutamate, GABA and dopamine are complex molecular machines that do much more than act as vacuums that clear neurotransmitters out of the extracellular space. This complexity arises in part from the core biophysical properties of these transporters, which are capable of generating different types of ionic currents. The additional levels of complexity comes from the fact that their residence time on the membrane, their trafficking from intracellular compartments and their kinetics can be modulated at different levels, including transcriptional, epigenetic, translational and post-translational levels. All these forms of modulation can change across species, cell types and brain regions. Since these regulatory mechanisms also change over time, their efficacy likely follows the metabolic state of a given neuron or astrocyte. The knowledge accumulated over the last few years and the growth of novel experimental approaches will undoubtedly provide new insights into cell-specific changes in the way the activity of neurotransmitter transporters can control cell excitability and metabolism across the brain.

Author Contributions

RMR and SLI wrote the manuscript. AS wrote the manuscript and coordinated the investigative team. All authors contributed to the article and approved the submitted version.

Funding

This work was supported by the Australian National Health and Medical Research Council Project grant APP1164494 to RMR, NIH grant R01 DA042565 to SLI, and NSF grants IOS1655365 and IOS2011998 to AS.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We performed detailed PubMed searches to include information listed in Tables 13. We apologize to authors of any work we might have missed.

References

Adkins, E. M., Samuvel, D. J., Fog, J. U., Eriksen, J., Jayanthi, L. D., Vaegter, C. B., et al. (2007). Membrane mobility and microdomain association of the dopamine transporter studied with fluorescence correlation spectroscopy and fluorescence recovery after photobleaching. Biochemistry 46, 10484–10497. doi: 10.1021/bi700429z

PubMed Abstract | CrossRef Full Text | Google Scholar

Adolph, O., Koster, S., Rath, M., Georgieff, M., Weigt, H. U., Engele, J., et al. (2007). Rapid increase of glial glutamate uptake via blockade of the protein kinase A pathway. Glia 55, 1699–1707. doi: 10.1002/glia.20583

PubMed Abstract | CrossRef Full Text | Google Scholar

Aggarwal, S., and Mortensen, O. V. (2017). Overview of monoamine transporters. Curr. Protoc. Pharmacol. 79, 12.16.1–12.16.17. doi: 10.1002/cpph.32

PubMed Abstract | CrossRef Full Text | Google Scholar

Al Awabdh, S., Gupta-Agarwal, S., Sheehan, D. F., Muir, J., Norkett, R., Twelvetrees, A. E., et al. (2016). Neuronal activity mediated regulation of glutamate transporter GLT-1 surface diffusion in rat astrocytes in dissociated and slice cultures. Glia 64, 1252–1264. doi: 10.1002/glia.22997

PubMed Abstract | CrossRef Full Text | Google Scholar

Anderson, B. B., Chen, G., Gutman, D. A., and Ewing, A. G. (1998). Dopamine levels of two classes of vesicles are differentially depleted by amphetamine. Brain Res. 788, 294–301. doi: 10.1016/s0006-8993(98)00040-7

CrossRef Full Text | Google Scholar

Anderson, C. M., and Swanson, R. A. (2000). Astrocyte glutamate transport: review of properties, regulation, and physiological functions. Glia 32, 1–14. doi: 10.1002/1098-1136(200010)32:1<1::aid-glia10>3.0.co;2-w

CrossRef Full Text | Google Scholar

Anderson, J. G., Cooney, P. T., and Erikson, K. M. (2007). Brain manganese accumulation is inversely related to gamma-amino butyric acid uptake in male and female rats. Toxicol. Sci. 95, 188–195. doi: 10.1093/toxsci/kfl130

PubMed Abstract | CrossRef Full Text | Google Scholar

Anderson, J. G., Fordahl, S. C., Cooney, P. T., Weaver, T. L., Colyer, C. L., and Erikson, K. M. (2008). Manganese exposure alters extracellular GABA, GABA receptor and transporter protein and mRNA levels in the developing rat brain. Neurotoxicology 29, 1044–1053. doi: 10.1016/j.neuro.2008.08.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Arriza, J. L., Fairman, W. A., Wadiche, J. I., Murdoch, G. H., Kavanaugh, M. P., and Amara, S. G. (1994). Functional comparisons of three glutamate transporter subtypes cloned from human motor cortex. J. Neurosci. 14, 5559–5569. doi: 10.1523/jneurosci.14-09-05559.1994

PubMed Abstract | CrossRef Full Text | Google Scholar

Assaife-Lopes, N., Sousa, V. C., Pereira, D. B., Ribeiro, J. A., Chao, M. V., and Sebastiao, A. M. (2010). Activation of adenosine A2A receptors induces TrkB translocation and increases BDNF-mediated phospho-TrkB localization in lipid rafts: implications for neuromodulation. J. Neurosci. 30, 8468–8480. doi: 10.1523/jneurosci.5695-09.2010

PubMed Abstract | CrossRef Full Text | Google Scholar

Bagley, E. E., Gerke, M. B., Vaughan, C. W., Hack, S. P., and Christie, M. J. (2005). GABA transporter currents activated by protein kinase A excite midbrain neurons during opioid withdrawal. Neuron 45, 433–445. doi: 10.1016/j.neuron.2004.12.049

PubMed Abstract | CrossRef Full Text | Google Scholar

Bagley, E. E., Hacker, J., Chefer, V. I., Mallet, C., McNally, G. P., Chieng, B. C., et al. (2011). Drug-induced GABA transporter currents enhance GABA release to induce opioid withdrawal behaviors. Nat. Neurosci. 14, 1548–1554. doi: 10.1038/nn.2940

PubMed Abstract | CrossRef Full Text | Google Scholar

Bahena-Trujillo, R., and Arias-Montano, J. A. (1999). [3H] gamma-aminobutyric acid transport in rat substantia nigra pars reticulata synaptosomes: pharmacological characterization and phorbol ester-induced inhibition. Neurosci. Lett. 274, 119–122. doi: 10.1016/s0304-3940(99)00692-8

CrossRef Full Text | Google Scholar

Bailey, C. G., Ryan, R. M., Thoeng, A. D., Ng, C., King, K., Vanslambrouck, J. M., et al. (2011). Loss-of-function mutations in the glutamate transporter SLC1A1 cause human dicarboxylic aminoaciduria. J. Clin. Invest. 121, 446–453. doi: 10.1172/jci44474

PubMed Abstract | CrossRef Full Text | Google Scholar

Barakat, L., and Bordey, A. (2002). GAT-1 and reversible GABA transport in Bergmann glia in slices. J. Neurophysiol. 88, 1407–1419. doi: 10.1152/jn.2002.88.3.1407

PubMed Abstract | CrossRef Full Text | Google Scholar

Barbaresi, P., Gazzanelli, G., and Malatesta, M. (2001). gamma-Aminobutyric acid transporters in the cat periaqueductal gray: a light and electron microscopic immunocytochemical study. J. Comp. Neurol. 429, 337–354. doi: 10.1002/1096-9861(20000108)429:2<337::aid-cne12>3.0.co;2-z

CrossRef Full Text | Google Scholar

Bassan, M., Liu, H., Madsen, K. L., Armsen, W., Zhou, J., Desilva, T., et al. (2008). Interaction between the glutamate transporter GLT1b and the synaptic PDZ domain protein PICK1. Eur. J. Neurosci. 27, 66–82. doi: 10.1111/j.1460-9568.2007.05986.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Baumann, M. H., Ayestas, M. A., Dersch, C. M., Brockington, A., Rice, K. C., and Rothman, R. B. (2000). Effects of phentermine and fenfluramine on extracellular dopamine and serotonin in rat nucleus accumbens: therapeutic implications. Synapse 36, 102–113. doi: 10.1002/(sici)1098-2396(200005)36:2<102::aid-syn3>3.0.co;2-#

CrossRef Full Text | Google Scholar

Beckman, M. L., Bernstein, E. M., and Quick, M. W. (1998). Protein kinase C regulates the interaction between a GABA transporter and syntaxin 1A. J. Neurosci. 18, 6103–6112. doi: 10.1523/jneurosci.18-16-06103.1998

PubMed Abstract | CrossRef Full Text | Google Scholar

Beckman, M. L., Bernstein, E. M., and Quick, M. W. (1999). Multiple G protein-coupled receptors initiate protein kinase C redistribution of GABA transporters in hippocampal neurons. J. Neurosci. 19:RC9.

Google Scholar

Behrens, P. F., Franz, P., Woodman, B., Lindenberg, K. S., and Landwehrmeyer, G. B. (2002). Impaired glutamate transport and glutamate-glutamine cycling: downstream effects of the Huntington mutation. Brain 125, 1908–1922. doi: 10.1093/brain/awf180

PubMed Abstract | CrossRef Full Text | Google Scholar

Bellini, S., Fleming, K. E., De, M., McCauley, J. P., Petroccione, M. A., D’Brant, L. Y., et al. (2018). Neuronal glutamate transporters control dopaminergic signaling and compulsive behaviors. J. Neurosci. 38, 937–961. doi: 10.1523/jneurosci.1906-17.2017

PubMed Abstract | CrossRef Full Text | Google Scholar

Benediktsson, A. M., Marrs, G. S., Tu, J. C., Worley, P. F., Rothstein, J. D., Bergles, D. E., et al. (2012). Neuronal activity regulates glutamate transporter dynamics in developing astrocytes. Glia 60, 175–188. doi: 10.1002/glia.21249

PubMed Abstract | CrossRef Full Text | Google Scholar

Berger, U. V., DeSilva, T. M., Chen, W., and Rosenberg, P. A. (2005). Cellular and subcellular mRNA localization of glutamate transporter isoforms GLT1a and GLT1b in rat brain by in situ hybridization. J. Comp. Neurol. 492, 78–89. doi: 10.1002/cne.20737

PubMed Abstract | CrossRef Full Text | Google Scholar

Bergles, D. E., Tzingounis, A. V., and Jahr, C. E. (2002). Comparison of coupled and uncoupled currents during glutamate uptake by GLT-1 transporters. J. Neurosci. 22, 10153–10162. doi: 10.1523/jneurosci.22-23-10153.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Bermingham, D. P., and Blakely, R. D. (2016). Kinase-dependent regulation of monoamine neurotransmitter transporters. Pharmacol. Rev. 68, 888–953. doi: 10.1124/pr.115.012260

PubMed Abstract | CrossRef Full Text | Google Scholar

Bernabe, A., Mendez, J. A., Hernandez-Kelly, L. C., and Ortega, A. (2003). Regulation of the Na+-dependent glutamate/aspartate transporter in rodent cerebellar astrocytes. Neurochem. Res. 28, 1843–1849.

Google Scholar

Bernardinelli, Y., Randall, J., Janett, E., Nikonenko, I., Konig, S., Jones, E. V., et al. (2014). Activity-dependent structural plasticity of perisynaptic astrocytic domains promotes excitatory synapse stability. Curr. Biol. 24, 1679–1688. doi: 10.1016/j.cub.2014.06.025

PubMed Abstract | CrossRef Full Text | Google Scholar

Bernstein, E. M., and Quick, M. W. (1999). Regulation of gamma-aminobutyric acid (GABA) transporters by extracellular GABA. J. Biol. Chem. 274, 889–895. doi: 10.1074/jbc.274.2.889

PubMed Abstract | CrossRef Full Text | Google Scholar

Berry, C. B., Hayes, D., Murphy, A., Wiessner, M., Rauen, T., and McBean, G. J. (2005). Differential modulation of the glutamate transporters GLT1, GLAST and EAAC1 by docosahexaenoic acid. Brain Res. 1037, 123–133. doi: 10.1016/j.brainres.2005.01.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Bhat, S., El-Kasaby, A., Freissmuth, M., and Sucic, S. (2020). Functional and biochemical consequences of disease variants in neurotransmitter transporters: a special emphasis on folding and trafficking deficits. Pharmacol. Ther. 2020:107785. doi: 10.1016/j.pharmthera.2020.107785

PubMed Abstract | CrossRef Full Text | Google Scholar

Binda, F., Dipace, C., Bowton, E., Robertson, S. D., Lute, B. J., Fog, J. U., et al. (2008). Syntaxin 1A interaction with the dopamine transporter promotes amphetamine-induced dopamine efflux. Mol. Pharmacol. 74, 1101–1108. doi: 10.1124/mol.108.048447

PubMed Abstract | CrossRef Full Text | Google Scholar

Bjerggaard, C., Fog, J. U., Hastrup, H., Madsen, K., Loland, C. J., Javitch, J. A., et al. (2004). Surface targeting of the dopamine transporter involves discrete epitopes in the distal C terminus but does not require canonical PDZ domain interactions. J. Neurosci. 24, 7024–7036. doi: 10.1523/jneurosci.1863-04.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Boddum, K., Jensen, T. P., Magloire, V., Kristiansen, U., Rusakov, D. A., Pavlov, I., et al. (2016). Astrocytic GABA transporter activity modulates excitatory neurotransmission. Nat. Commun. 7:13572.

Google Scholar

Bohmer, C., Philippin, M., Rajamanickam, J., Mack, A., Broer, S., Palmada, M., et al. (2004). Stimulation of the EAAT4 glutamate transporter by SGK protein kinase isoforms and PKB. Biochem. Biophys. Res. Commun. 324, 1242–1248. doi: 10.1016/j.bbrc.2004.09.193

PubMed Abstract | CrossRef Full Text | Google Scholar

Bolan, E. A., Kivell, B., Jaligam, V., Oz, M., Jayanthi, L. D., Han, Y., et al. (2007). D2 receptors regulate dopamine transporter function via an extracellular signal-regulated kinases 1 and 2-dependent and phosphoinositide 3 kinase-independent mechanism. Mol. Pharmacol. 71, 1222–1232. doi: 10.1124/mol.106.027763

PubMed Abstract | CrossRef Full Text | Google Scholar

Boudanova, E., Navaroli, D. M., and Melikian, H. E. (2008a). Amphetamine-induced decreases in dopamine transporter surface expression are protein kinase C-independent. Neuropharmacology 54, 605–612. doi: 10.1016/j.neuropharm.2007.11.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Boudanova, E., Navaroli, D. M., Stevens, Z., and Melikian, H. E. (2008b). Dopamine transporter endocytic determinants: carboxy terminal residues critical for basal and PKC-stimulated internalization. Mol. Cell Neurosci. 39, 211–217. doi: 10.1016/j.mcn.2008.06.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Boudker, O., Ryan, R. M., Yernool, D., Shimamoto, K., and Gouaux, E. (2007). Coupling substrate and ion binding to extracellular gate of a sodium-dependent aspartate transporter. Nature 445, 387–393. doi: 10.1038/nature05455

PubMed Abstract | CrossRef Full Text | Google Scholar

Bowton, E., Saunders, C., Reddy, I. A., Campbell, N. G., Hamilton, P. J., Henry, L. K., et al. (2014). SLC6A3 coding variant Ala559Val found in two autism probands alters dopamine transporter function and trafficking. Transl. Psychiatry 4:e464. doi: 10.1038/tp.2014.90

PubMed Abstract | CrossRef Full Text | Google Scholar

Brosnan, J. T., and Brosnan, M. E. (2013). Glutamate: a truly functional amino acid. Amino Acids 45, 413–418. doi: 10.1007/s00726-012-1280-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Bruijn, L. I., Becher, M. W., Lee, M. K., Anderson, K. L., Jenkins, N. A., Copeland, N. G., et al. (1997). ALS-linked SOD1 mutant G85R mediates damage to astrocytes and promotes rapidly progressive disease with SOD1-containing inclusions. Neuron 18, 327–338. doi: 10.1016/s0896-6273(00)80272-x

CrossRef Full Text | Google Scholar

Bruns, D., Engert, F., and Lux, H. D. (1993). A fast activating presynaptic reuptake current during serotonergic transmission in identified neurons of Hirudo. Neuron 10, 559–572. doi: 10.1016/0896-6273(93)90159-o

CrossRef Full Text | Google Scholar

Bull, N. D., and Barnett, N. L. (2002). Antagonists of protein kinase C inhibit rat retinal glutamate transport activity in situ. J. Neurochem. 81, 472–480. doi: 10.1046/j.1471-4159.2002.00819.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Butchbach, M. E., Tian, G., Guo, H., and Lin, C. L. (2004). Association of excitatory amino acid transporters, especially EAAT2, with cholesterol-rich lipid raft microdomains: importance for excitatory amino acid transporter localization and function. J. Biol. Chem. 279, 34388–34396. doi: 10.1074/jbc.m403938200

PubMed Abstract | CrossRef Full Text | Google Scholar

Cammack, J. N., Rakhilin, S. V., and Schwartz, E. A. (1994). A GABA transporter operates asymmetrically and with variable stoichiometry. Neuron 13, 949–960. doi: 10.1016/0896-6273(94)90260-7

CrossRef Full Text | Google Scholar

Cammack, J. N., and Schwartz, E. A. (1996). Channel behavior in a gamma-aminobutyrate transporter. Proc. Natl. Acad. Sci. U.S.A. 93, 723–727. doi: 10.1073/pnas.93.2.723

PubMed Abstract | CrossRef Full Text | Google Scholar

Campbell, N. G., Shekar, A., Aguilar, J. I., Peng, D., Navratna, V., Yang, D., et al. (2019). Structural, functional, and behavioral insights of dopamine dysfunction revealed by a deletion in SLC6A3. Proc. Natl. Acad. Sci. U.S.A. 116, 3853–3862. doi: 10.1073/pnas.1816247116

PubMed Abstract | CrossRef Full Text | Google Scholar

Canul-Tec, J. C., Assal, R., Cirri, E., Legrand, P., Brier, S., Chamot-Rooke, J., et al. (2017). Structure and allosteric inhibition of excitatory amino acid transporter 1. Nature 544, 446–451. doi: 10.1038/nature22064

PubMed Abstract | CrossRef Full Text | Google Scholar

Carneiro, A. M., Ingram, S. L., Beaulieu, J. M., Sweeney, A., Amara, S. G., Thomas, S. M., et al. (2002). The multiple LIM domain-containing adaptor protein Hic-5 synaptically colocalizes and interacts with the dopamine transporter. J. Neurosci. 22, 7045–7054. doi: 10.1523/jneurosci.22-16-07045.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Carvelli, L., Blakely, R. D., and DeFelice, L. J. (2008). Dopamine transporter/syntaxin 1A interactions regulate transporter channel activity and dopaminergic synaptic transmission. Proc. Natl. Acad. Sci. U.S.A. 105, 14192–14197. doi: 10.1073/pnas.0802214105

PubMed Abstract | CrossRef Full Text | Google Scholar

Carvelli, L., McDonald, P. W., Blakely, R. D., and DeFelice, L. J. (2004). Dopamine transporters depolarize neurons by a channel mechanism. Proc. Natl. Acad. Sci. U.S.A. 101, 16046–16051. doi: 10.1073/pnas.0403299101

PubMed Abstract | CrossRef Full Text | Google Scholar

Carvelli, L., Moron, J. A., Kahlig, K. M., Ferrer, J. V., Sen, N., Lechleiter, J. D., et al. (2002). PI 3-kinase regulation of dopamine uptake. J. Neurochem. 81, 859–869. doi: 10.1046/j.1471-4159.2002.00892.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Casado, M., Bendahan, A., Zafra, F., Danbolt, N. C., Aragon, C., Gimenez, C., et al. (1993). Phosphorylation and modulation of brain glutamate transporters by protein kinase C. J. Biol. Chem. 268, 27313–27317. doi: 10.1016/s0021-9258(19)74251-3

CrossRef Full Text | Google Scholar

Castillo, S. O., Baffi, J. S., Palkovits, M., Goldstein, D. S., Kopin, I. J., Witta, J., et al. (1998). Dopamine biosynthesis is selectively abolished in substantia nigra/ventral tegmental area but not in hypothalamic neurons in mice with targeted disruption of the Nurr1 gene. Mol. Cell Neurosci. 11, 36–46. doi: 10.1006/mcne.1998.0673

PubMed Abstract | CrossRef Full Text | Google Scholar

Cater, R. J., Vandenberg, R. J., and Ryan, R. M. (2014). The domain interface of the human glutamate transporter EAAT1 mediates chloride permeation. Biophys. J. 107, 621–629. doi: 10.1016/j.bpj.2014.05.046

PubMed Abstract | CrossRef Full Text | Google Scholar

Cater, R. J., Vandenberg, R. J., and Ryan, R. M. (2016). Tuning the ion selectivity of glutamate transporter-associated uncoupled conductances. J. Gen. Physiol. 148, 13–24. doi: 10.1085/jgp.201511556

PubMed Abstract | CrossRef Full Text | Google Scholar

Cervinski, M. A., Foster, J. D., and Vaughan, R. A. (2005). Psychoactive substrates stimulate dopamine transporter phosphorylation and down-regulation by cocaine-sensitive and protein kinase C-dependent mechanisms. J. Biol. Chem. 280, 40442–40449. doi: 10.1074/jbc.m501969200

PubMed Abstract | CrossRef Full Text | Google Scholar

Chan, P. H., Kerlan, R., and Fishman, R. A. (1983). Reductions of gamma-aminobutyric acid and glutamate uptake and (Na+ + K+)-ATPase activity in brain slices and synaptosomes by arachidonic acid. J. Neurochem. 40, 309–316. doi: 10.1111/j.1471-4159.1983.tb11284.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Chang, M. Y., Lee, S. H., Kim, J. H., Lee, K. H., Kim, Y. S., Son, H., et al. (2001). Protein kinase C-mediated functional regulation of dopamine transporter is not achieved by direct phosphorylation of the dopamine transporter protein. J. Neurochem. 77, 754–761. doi: 10.1046/j.1471-4159.2001.00284.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Chaudhry, F. A., Lehre, K. P., van Lookeren Campagne, M., Ottersen, O. P., Danbolt, N. C., and Storm-Mathisen, J. (1995). Glutamate transporters in glial plasma membranes: highly differentiated localizations revealed by quantitative ultrastructural immunocytochemistry. Neuron 15, 711–720. doi: 10.1016/0896-6273(95)90158-2

CrossRef Full Text | Google Scholar

Chen, I., Pant, S., Wu, Q., Cater, R., Sobti, M., Vandenberg, R. J., et al. (2021). Glutamate transporters contain a conserved chloride channel with two hydrophobic gates. bioRxiv [Preprint], doi: 10.1101/2020.05.25.115360

CrossRef Full Text | Google Scholar

Chen, N., and Reith, M. E. (2000). Structure and function of the dopamine transporter. Eur. J. Pharmacol. 405, 329–339.

Google Scholar

Chen, R., Daining, C. P., Sun, H., Fraser, R., Stokes, S. L., Leitges, M., et al. (2013). Protein kinase Cbeta is a modulator of the dopamine D2 autoreceptor-activated trafficking of the dopamine transporter. J. Neurochem. 125, 663–672. doi: 10.1111/jnc.12229

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, R., Furman, C. A., Zhang, M., Kim, M. N., Gereau, R. W. T., Leitges, M., et al. (2009). Protein kinase Cbeta is a critical regulator of dopamine transporter trafficking and regulates the behavioral response to amphetamine in mice. J. Pharmacol. Exp. Ther. 328, 912–920. doi: 10.1124/jpet.108.147959

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, T., Tanaka, M., Wang, Y., Sha, S., Furuya, K., Chen, L., et al. (2017). Neurosteroid dehydroepiandrosterone enhances activity and trafficking of astrocytic GLT-1 via sigma1 receptor-mediated PKC activation in the hippocampal dentate gyrus of rats. Glia 65, 1491–1503. doi: 10.1002/glia.23175

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, W., Aoki, C., Mahadomrongkul, V., Gruber, C. E., Wang, G. J., Blitzblau, R., et al. (2002). Expression of a variant form of the glutamate transporter GLT1 in neuronal cultures and in neurons and astrocytes in the rat brain. J. Neurosci. 22, 2142–2152. doi: 10.1523/jneurosci.22-06-02142.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, W., Mahadomrongkul, V., Berger, U. V., Bassan, M., DeSilva, T., Tanaka, K., et al. (2004). The glutamate transporter GLT1a is expressed in excitatory axon terminals of mature hippocampal neurons. J. Neurosci. 24, 1136–1148. doi: 10.1523/jneurosci.1586-03.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheng, C., Glover, G., Banker, G., and Amara, S. G. (2002). A novel sorting motif in the glutamate transporter excitatory amino acid transporter 3 directs its targeting in Madin-Darby canine kidney cells and hippocampal neurons. J. Neurosci. 22, 10643–10652. doi: 10.1523/jneurosci.22-24-10643.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheng, M. H., Torres-Salazar, D., Gonzalez-Suarez, A. D., Amara, S. G., and Bahar, I. (2017). Substrate transport and anion permeation proceed through distinct pathways in glutamate transporters. eLife 6:e25850.

Google Scholar

Chi, L., and Reith, M. E. (2003). Substrate-induced trafficking of the dopamine transporter in heterologously expressing cells and in rat striatal synaptosomal preparations. J. Pharmacol. Exp. Ther. 307, 729–736. doi: 10.1124/jpet.103.055095

PubMed Abstract | CrossRef Full Text | Google Scholar

Chiu, C. S., Jensen, K., Sokolova, I., Wang, D., Li, M., Deshpande, P., et al. (2002). Number, density, and surface/cytoplasmic distribution of GABA transporters at presynaptic structures of knock-in mice carrying GABA transporter subtype 1-green fluorescent protein fusions. J. Neurosci. 22, 10251–10266. doi: 10.1523/jneurosci.22-23-10251.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Chivukula, A. S., Suslova, M., Kortzak, D., Kovermann, P., and Fahlke, C. (2020). Functional consequences of SLC1A3 mutations associated with episodic ataxia 6. Hum. Mutat. 41, 1892–1905. doi: 10.1002/humu.24089

PubMed Abstract | CrossRef Full Text | Google Scholar

Choi, J. K., and Kim, Y. J. (2008). Epigenetic regulation and the variability of gene expression. Nat. Genet. 40, 141–147. doi: 10.1038/ng.2007.58

PubMed Abstract | CrossRef Full Text | Google Scholar

Choi, K. D., Jen, J. C., Choi, S. Y., Shin, J. H., Kim, H. S., Kim, H. J., et al. (2017). Late-onset episodic ataxia associated with SLC1A3 mutation. J. Hum. Genet. 62, 443–446. doi: 10.1038/jhg.2016.137

PubMed Abstract | CrossRef Full Text | Google Scholar

Chowdhury, H. H., Kreft, M., and Zorec, R. (2002). Rapid insulin-induced exocytosis in white rat adipocytes. Pflugers Arch. 445, 352–356. doi: 10.1007/s00424-002-0938-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Clark, J. A., Deutch, A. Y., Gallipoli, P. Z., and Amara, S. G. (1992). Functional expression and CNS distribution of a beta-alanine-sensitive neuronal GABA transporter. Neuron 9, 337–348. doi: 10.1016/0896-6273(92)90172-a

CrossRef Full Text | Google Scholar

Conradt, M., and Stoffel, W. (1997). Inhibition of the high-affinity brain glutamate transporter GLAST-1 via direct phosphorylation. J. Neurochem. 68, 1244–1251. doi: 10.1046/j.1471-4159.1997.68031244.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Conradt, M., Storck, T., and Stoffel, W. (1995). Localization of N-glycosylation sites and functional role of the carbohydrate units of GLAST-1, a cloned rat brain L-glutamate/L-aspartate transporter. Eur. J. Biochem. 229, 682–687. doi: 10.1111/j.1432-1033.1995.0682j.x

CrossRef Full Text | Google Scholar

Copeland, B. J., Vogelsberg, V., Neff, N. H., and Hadjiconstantinou, M. (1996). Protein kinase C activators decrease dopamine uptake into striatal synaptosomes. J. Pharmacol. Exp. Ther. 277, 1527–1532.

Google Scholar

Cordeiro, J. M., Meireles, S. M., Vale, M. G., Oliveira, C. R., and Goncalves, P. P. (2000). Ca(2+) regulation of the carrier-mediated gamma-aminobutyric acid release from isolated synaptic plasma membrane vesicles. Neurosci. Res. 38, 385–395. doi: 10.1016/s0168-0102(00)00193-0

CrossRef Full Text | Google Scholar

Corey, J. L., Davidson, N., Lester, H. A., Brecha, N., and Quick, M. W. (1994). Protein kinase C modulates the activity of a cloned gamma-aminobutyric acid transporter expressed in Xenopus oocytes via regulated subcellular redistribution of the transporter. J. Biol. Chem. 269, 14759–14767. doi: 10.1016/s0021-9258(17)36690-5

CrossRef Full Text | Google Scholar

Cremona, M. L., Matthies, H. J., Pau, K., Bowton, E., Speed, N., Lute, B. J., et al. (2011). Flotillin-1 is essential for PKC-triggered endocytosis and membrane microdomain localization of DAT. Nat. Neurosci. 14, 469–477. doi: 10.1038/nn.2781

PubMed Abstract | CrossRef Full Text | Google Scholar

Cristovao-Ferreira, S., Navarro, G., Brugarolas, M., Perez-Capote, K., Vaz, S. H., Fattorini, G., et al. (2013). A1R-A2AR heteromers coupled to Gs and G i/0 proteins modulate GABA transport into astrocytes. Purinerg. Signal. 9, 433–449. doi: 10.1007/s11302-013-9364-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Cristovao-Ferreira, S., Vaz, S. H., Ribeiro, J. A., and Sebastiao, A. M. (2009). Adenosine A2A receptors enhance GABA transport into nerve terminals by restraining PKC inhibition of GAT-1. J. Neurochem. 109, 336–347. doi: 10.1111/j.1471-4159.2009.05963.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Dabir, D. V., Robinson, M. B., Swanson, E., Zhang, B., Trojanowski, J. Q., Lee, V. M., et al. (2006). Impaired glutamate transport in a mouse model of tau pathology in astrocytes. J. Neurosci. 26, 644–654. doi: 10.1523/jneurosci.3861-05.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Dall’Igna, O. P., Bobermin, L. D., Souza, D. O., and Quincozes-Santos, A. (2013). Riluzole increases glutamate uptake by cultured C6 astroglial cells. Int. J. Dev. Neurosci. 31, 482–486. doi: 10.1016/j.ijdevneu.2013.06.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Danbolt, N. C. (2001). Glutamate uptake. Prog. Neurobiol. 65, 1–105.

Google Scholar

Danbolt, N. C., Storm-Mathisen, J., and Kanner, B. I. (1992). An [Na+ + K+]coupled L-glutamate transporter purified from rat brain is located in glial cell processes. Neuroscience 51, 295–310. doi: 10.1016/0306-4522(92)90316-t

CrossRef Full Text | Google Scholar

Daniels, G. M., and Amara, S. G. (1999). Regulated trafficking of the human dopamine transporter. Clathrin-mediated internalization and lysosomal degradation in response to phorbol esters. J. Biol. Chem. 274, 35794–35801. doi: 10.1074/jbc.274.50.35794

PubMed Abstract | CrossRef Full Text | Google Scholar

Davis, K. E., Straff, D. J., Weinstein, E. A., Bannerman, P. G., Correale, D. M., Rothstein, J. D., et al. (1998). Multiple signaling pathways regulate cell surface expression and activity of the excitatory amino acid carrier 1 subtype of Glu transporter in C6 glioma. J. Neurosci. 18, 2475–2485. doi: 10.1523/jneurosci.18-07-02475.1998

PubMed Abstract | CrossRef Full Text | Google Scholar

Dehnes, Y., Chaudhry, F. A., Ullensvang, K., Lehre, K. P., Storm-Mathisen, J., and Danbolt, N. C. (1998). The glutamate transporter EAAT4 in rat cerebellar Purkinje cells: a glutamate-gated chloride channel concentrated near the synapse in parts of the dendritic membrane facing astroglia. J. Neurosci. 18, 3606–3619. doi: 10.1523/jneurosci.18-10-03606.1998

PubMed Abstract | CrossRef Full Text | Google Scholar

Deken, S. L., Beckman, M. L., Boos, L., and Quick, M. W. (2000). Transport rates of GABA transporters: regulation by the N-terminal domain and syntaxin 1A. Nat. Neurosci. 3, 998–1003. doi: 10.1038/79939

PubMed Abstract | CrossRef Full Text | Google Scholar

Deken, S. L., Wang, D., and Quick, M. W. (2003). Plasma membrane GABA transporters reside on distinct vesicles and undergo rapid regulated recycling. J. Neurosci. 23, 1563–1568. doi: 10.1523/jneurosci.23-05-01563.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Delgado-Acevedo, C., Estay, S. F., Radke, A. K., Sengupta, A., Escobar, A. P., Henriquez-Belmar, F., et al. (2019). Behavioral and synaptic alterations relevant to obsessive-compulsive disorder in mice with increased EAAT3 expression. Neuropsychopharmacology 44, 1163–1173. doi: 10.1038/s41386-018-0302-7

PubMed Abstract | CrossRef Full Text | Google Scholar

DiCarlo, G. E., Aguilar, J. I., Matthies, H. J., Harrison, F. E., Bundschuh, K. E., West, A., et al. (2019). Autism-linked dopamine transporter mutation alters striatal dopamine neurotransmission and dopamine-dependent behaviors. J. Clin. Invest. 129, 3407–3419. doi: 10.1172/jci127411

PubMed Abstract | CrossRef Full Text | Google Scholar

Divito, C. B., Borowski, J. E., Glasgow, N. G., Gonzalez-Suarez, A. D., Torres-Salazar, D., Johnson, J. W., et al. (2017). Glial and neuronal glutamate transporters differ in the Na(+) requirements for activation of the substrate-independent anion conductance. Front. Mol. Neurosci. 10:150. doi: 10.3389/fnmol.2017.00150

PubMed Abstract | CrossRef Full Text | Google Scholar

Dodd, S., Carvalho, A. F., Puri, B. K., Maes, M., Bortolasci, C. C., Morris, G., et al. (2020). Trace amine-associated receptor 1 (TAAR1): a new drug target for psychiatry? Neurosci. Biobehav. Rev. 120, 537–541. doi: 10.1016/j.neubiorev.2020.09.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Doolen, S., and Zahniser, N. R. (2001). Protein tyrosine kinase inhibitors alter human dopamine transporter activity in Xenopus oocytes. J. Pharmacol. Exp. Ther. 296, 931–938.

Google Scholar

Dowd, L. A., and Robinson, M. B. (1996). Rapid stimulation of EAAC1-mediated Na+-dependent L-glutamate transport activity in C6 glioma cells by phorbol ester. J. Neurochem. 67, 508–516. doi: 10.1046/j.1471-4159.1996.67020508.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Drejer, J., Meier, E., and Schousboe, A. (1983). Novel neuron-related regulatory mechanisms for astrocytic glutamate and GABA high affinity uptake. Neurosci. Lett. 37, 301–306. doi: 10.1016/0304-3940(83)90448-2

CrossRef Full Text | Google Scholar

Eckstein-Ludwig, U., Fei, J., and Schwarz, W. (1999). Inhibition of uptake, steady-state currents, and transient charge movements generated by the neuronal GABA transporter by various anticonvulsant drugs. Br. J. Pharmacol. 128, 92–102. doi: 10.1038/sj.bjp.0702794

PubMed Abstract | CrossRef Full Text | Google Scholar

Egawa, K., and Fukuda, A. (2013). Pathophysiological power of improper tonic GABA(A) conductances in mature and immature models. Front. Neural Circ. 7:170. doi: 10.3389/fncir.2013.00170

PubMed Abstract | CrossRef Full Text | Google Scholar

Eliasof, S., and Jahr, C. E. (1996). Retinal glial cell glutamate transporter is coupled to an anionic conductance. Proc. Natl. Acad. Sci. U.S.A. 93, 4153–4158. doi: 10.1073/pnas.93.9.4153

PubMed Abstract | CrossRef Full Text | Google Scholar

Eriksen, J., Rasmussen, S. G., Rasmussen, T. N., Vaegter, C. B., Cha, J. H., Zou, M. F., et al. (2009). Visualization of dopamine transporter trafficking in live neurons by use of fluorescent cocaine analogs. J. Neurosci. 29, 6794–6808. doi: 10.1523/jneurosci.4177-08.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Espinoza-Rojo, M., Lopez-Bayghen, E., and Ortega, A. (2000). GLAST: gene expression regulation by phorbol esters. Neuroreport 11, 2827–2832. doi: 10.1097/00001756-200008210-00043

PubMed Abstract | CrossRef Full Text | Google Scholar

Fairman, W. A., Vandenberg, R. J., Arriza, J. L., Kavanaugh, M. P., and Amara, S. G. (1995). An excitatory amino-acid transporter with properties of a ligand-gated chloride channel. Nature 375, 599–603. doi: 10.1038/375599a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Fan, H. P., Fan, F. J., Bao, L., and Pei, G. (2006). SNAP-25/syntaxin 1A complex functionally modulates neurotransmitter gamma-aminobutyric acid reuptake. J. Biol. Chem. 281, 28174–28184. doi: 10.1074/jbc.m601382200

PubMed Abstract | CrossRef Full Text | Google Scholar

Fan, S., Xian, X., Li, L., Yao, X., Hu, Y., Zhang, M., et al. (2018). Ceftriaxone improves cognitive function and upregulates GLT-1-related glutamate-glutamine cycle in APP/PS1 mice. J. Alzheimers Dis. 66, 1731–1743. doi: 10.3233/jad-180708

PubMed Abstract | CrossRef Full Text | Google Scholar

Fang, H., Huang, Y., and Zuo, Z. (2006). Enhancement of substrate-gated Cl- currents via rat glutamate transporter EAAT4 by PMA. Am. J. Physiol. Cell Physiol. 290, C1334–C1340.

Google Scholar

Fang, Q., Hu, W. W., Wang, X. F., Yang, Y., Lou, G. D., Jin, M. M., et al. (2014). Histamine up-regulates astrocytic glutamate transporter 1 and protects neurons against ischemic injury. Neuropharmacology 77, 156–166. doi: 10.1016/j.neuropharm.2013.06.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Farhan, H., Korkhov, V. M., Paulitschke, V., Dorostkar, M. M., Scholze, P., Kudlacek, O., et al. (2004). Two discontinuous segments in the carboxyl terminus are required for membrane targeting of the rat gamma-aminobutyric acid transporter-1 (GAT1). J. Biol. Chem. 279, 28553–28563. doi: 10.1074/jbc.m307325200

PubMed Abstract | CrossRef Full Text | Google Scholar

Fattorini, G., Melone, M., and Conti, F. (2020). A reappraisal of GAT-1 localization in Neocortex. Front. Cell Neurosci. 14:9. doi: 10.3389/fncel.2020.00009

PubMed Abstract | CrossRef Full Text | Google Scholar

Ferrer-Martinez, A., Felipe, A., Nicholson, B., Casado, J., Pastor-Anglada, M., and McGivan, J. (1995). Induction of the high-affinity Na(+)-dependent glutamate transport system XAG- by hypertonic stress in the renal epithelial cell line NBL-1. Biochem. J. 310(Pt 2), 689–692. doi: 10.1042/bj3100689

PubMed Abstract | CrossRef Full Text | Google Scholar

Figiel, M., Maucher, T., Rozyczka, J., Bayatti, N., and Engele, J. (2003). Regulation of glial glutamate transporter expression by growth factors. Exp. Neurol. 183, 124–135. doi: 10.1016/s0014-4886(03)00134-1

CrossRef Full Text | Google Scholar

Fischer, J. F., and Cho, A. K. (1979). Chemical release of dopamine from striatal homogenates: evidence for an exchange diffusion model. J. Pharmacol. Exp. Ther. 208, 203–209.

Google Scholar

Fischer, K. D., Houston, A. C. W., Desai, R. I., Doyle, M. R., Bergman, J., Mian, M., et al. (2018). Behavioral phenotyping and dopamine dynamics in mice with conditional deletion of the glutamate transporter GLT-1 in neurons: resistance to the acute locomotor effects of amphetamine. Psychopharmacology 235, 1371–1387. doi: 10.1007/s00213-018-4848-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Fog, J. U., Khoshbouei, H., Holy, M., Owens, W. A., Vaegter, C. B., Sen, N., et al. (2006). Calmodulin kinase II interacts with the dopamine transporter C terminus to regulate amphetamine-induced reverse transport. Neuron 51, 417–429. doi: 10.1016/j.neuron.2006.06.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Foster, J. D., Adkins, S. D., Lever, J. R., and Vaughan, R. A. (2008). Phorbol ester induced trafficking-independent regulation and enhanced phosphorylation of the dopamine transporter associated with membrane rafts and cholesterol. J. Neurochem. 105, 1683–1699. doi: 10.1111/j.1471-4159.2008.05262.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Foster, J. D., Yang, J. W., Moritz, A. E., Challasivakanaka, S., Smith, M. A., Holy, M., et al. (2012). Dopamine transporter phosphorylation site threonine 53 regulates substrate reuptake and amphetamine-stimulated efflux. J. Biol. Chem. 287, 29702–29712. doi: 10.1074/jbc.m112.367706

PubMed Abstract | CrossRef Full Text | Google Scholar

Fournier, K. M., Gonzalez, M. I., and Robinson, M. B. (2004). Rapid trafficking of the neuronal glutamate transporter, EAAC1: evidence for distinct trafficking pathways differentially regulated by protein kinase C and platelet-derived growth factor. J. Biol. Chem. 279, 34505–34513. doi: 10.1074/jbc.m404032200

PubMed Abstract | CrossRef Full Text | Google Scholar

Furman, C. A., Chen, R., Guptaroy, B., Zhang, M., Holz, R. W., and Gnegy, M. (2009). Dopamine and amphetamine rapidly increase dopamine transporter trafficking to the surface: live-cell imaging using total internal reflection fluorescence microscopy. J. Neurosci. 29, 3328–3336. doi: 10.1523/jneurosci.5386-08.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Furness, D. N., Dehnes, Y., Akhtar, A. Q., Rossi, D. J., Hamann, M., Grutle, N. J., et al. (2008). A quantitative assessment of glutamate uptake into hippocampal synaptic terminals and astrocytes: new insights into a neuronal role for excitatory amino acid transporter 2 (EAAT2). Neuroscience 157, 80–94. doi: 10.1016/j.neuroscience.2008.08.043

PubMed Abstract | CrossRef Full Text | Google Scholar

Galli, A., DeFelice, L. J., Duke, B. J., Moore, K. R., and Blakely, R. D. (1995). Sodium-dependent norepinephrine-induced currents in norepinephrine-transporter-transfected HEK-293 cells blocked by cocaine and antidepressants. J. Exp. Biol. 198, 2197–2212.

Google Scholar

Gamboa, C., and Ortega, A. (2002). Insulin-like growth factor-1 increases activity and surface levels of the GLAST subtype of glutamate transporter. Neurochem. Int. 40, 397–403. doi: 10.1016/s0197-0186(01)00106-1

CrossRef Full Text | Google Scholar

Ganel, R., and Crosson, C. E. (1998). Modulation of human glutamate transporter activity by phorbol ester. J. Neurochem. 70, 993–1000. doi: 10.1046/j.1471-4159.1998.70030993.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Garaeva, A. A., Guskov, A., Slotboom, D. J., and Paulino, C. (2019). A one-gate elevator mechanism for the human neutral amino acid transporter ASCT2. Nat. Commun. 10:3427.

Google Scholar

Garaeva, A. A., Oostergetel, G. T., Gati, C., Guskov, A., Paulino, C., and Slotboom, D. J. (2018). Cryo-EM structure of the human neutral amino acid transporter ASCT2. Nat. Struct. Mol. Biol. 25, 515–521. doi: 10.1038/s41594-018-0076-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Garcia, B. G., Wei, Y., Moron, J. A., Lin, R. Z., Javitch, J. A., and Galli, A. (2005). Akt is essential for insulin modulation of amphetamine-induced human dopamine transporter cell-surface redistribution. Mol. Pharmacol. 68, 102–109. doi: 10.1124/mol.104.009092

PubMed Abstract | CrossRef Full Text | Google Scholar

Garcia-Tardon, N., Gonzalez-Gonzalez, I. M., Martinez-Villarreal, J., Fernandez-Sanchez, E., Gimenez, C., and Zafra, F. (2012). Protein kinase C (PKC)-promoted endocytosis of glutamate transporter GLT-1 requires ubiquitin ligase Nedd4-2-dependent ubiquitination but not phosphorylation. J. Biol. Chem. 287, 19177–19187. doi: 10.1074/jbc.m112.355909

PubMed Abstract | CrossRef Full Text | Google Scholar

Gavrilov, N., Golyagina, I., Brazhe, A., Scimemi, A., Turlapov, V., and Semyanov, A. (2018). Astrocytic coverage of dendritic spines, dendritic shafts, and axonal boutons in hippocampal neuropil. Front. Cell Neurosci. 12:248. doi: 10.3389/fncel.2018.00248

PubMed Abstract | CrossRef Full Text | Google Scholar

Geerlings, A., Nunez, E., Lopez-Corcuera, B., and Aragon, C. (2001). Calcium- and syntaxin 1-mediated trafficking of the neuronal glycine transporter GLYT2. J. Biol. Chem. 276, 17584–17590. doi: 10.1074/jbc.m010602200

PubMed Abstract | CrossRef Full Text | Google Scholar

Gegelashvili, G., Civenni, G., Racagni, G., Danbolt, N. C., Schousboe, I., and Schousboe, A. (1996). Glutamate receptor agonists up-regulate glutamate transporter GLAST in astrocytes. Neuroreport 8, 261–265. doi: 10.1097/00001756-199612200-00052

PubMed Abstract | CrossRef Full Text | Google Scholar

Gegelashvili, G., and Schousboe, A. (1997). High affinity glutamate transporters: regulation of expression and activity. Mol. Pharmacol. 52, 6–15. doi: 10.1124/mol.52.1.6

PubMed Abstract | CrossRef Full Text | Google Scholar

Genoud, C., Quairiaux, C., Steiner, P., Hirling, H., Welker, E., and Knott, G. W. (2006). Plasticity of astrocytic coverage and glutamate transporter expression in adult mouse cortex. PLoS Biol. 4:e343. doi: 10.1371/journal.pbio.0040343

PubMed Abstract | CrossRef Full Text | Google Scholar

German, C. L., Baladi, M. G., McFadden, L. M., Hanson, G. R., and Fleckenstein, A. E. (2015). Regulation of the dopamine and vesicular monoamine transporters: pharmacological targets and implications for disease. Pharmacol. Rev. 67, 1005–1024. doi: 10.1124/pr.114.010397

PubMed Abstract | CrossRef Full Text | Google Scholar

Ginsberg, S. D., Martin, L. J., and Rothstein, J. D. (1995). Regional deafferentation down-regulates subtypes of glutamate transporter proteins. J. Neurochem. 65, 2800–2803. doi: 10.1046/j.1471-4159.1995.65062800.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Goldberg, N. R., Beuming, T., Soyer, O. S., Goldstein, R. A., Weinstein, H., and Javitch, J. A. (2003). Probing conformational changes in neurotransmitter transporters: a structural context. Eur. J. Pharmacol. 479, 3–12. doi: 10.1016/j.ejphar.2003.08.052

PubMed Abstract | CrossRef Full Text | Google Scholar

Goncalves, P. P., Carvalho, A. P., and Vale, M. G. (1997). Regulation of [gamma-3H]aminobutyric acid transport by Ca2 + in isolated synaptic plasma membrane vesicles. Brain Res. Mol. Brain Res. 51, 106–114. doi: 10.1016/s0169-328x(97)00223-4

CrossRef Full Text | Google Scholar

Goncalves, P. P., Meireles, S. M., and Vale, M. G. (1999). Regulation of the gamma-aminobutyric acid transporter activity by protein phosphatases in synaptic plasma membranes. Neurosci. Res. 33, 41–47. doi: 10.1016/s0168-0102(98)00107-2

CrossRef Full Text | Google Scholar

Gonzalez, M. I., Kazanietz, M. G., and Robinson, M. B. (2002). Regulation of the neuronal glutamate transporter excitatory amino acid carrier-1 (EAAC1) by different protein kinase C subtypes. Mol. Pharmacol. 62, 901–910. doi: 10.1124/mol.62.4.901

PubMed Abstract | CrossRef Full Text | Google Scholar

Gonzalez, M. I., Lopez-Colom, A. M., and Ortega, A. (1999). Sodium-dependent glutamate transport in Muller glial cells: regulation by phorbol esters. Brain Res. 831, 140–145. doi: 10.1016/s0006-8993(99)01438-9

CrossRef Full Text | Google Scholar

Gonzalez-Gonzalez, I. M., Garcia-Tardon, N., Cubelos, B., Gimenez, C., and Zafra, F. (2008a). The glutamate transporter GLT1b interacts with the scaffold protein PSD-95. J. Neurochem. 105, 1834–1848. doi: 10.1111/j.1471-4159.2008.05281.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Gonzalez-Gonzalez, I. M., Garcia-Tardon, N., Gimenez, C., and Zafra, F. (2008b). PKC-dependent endocytosis of the GLT1 glutamate transporter depends on ubiquitylation of lysines located in a C-terminal cluster. Glia 56, 963–974. doi: 10.1002/glia.20670

PubMed Abstract | CrossRef Full Text | Google Scholar

Goodspeed, K., Pérez-Palma, E., Iqbal, S., Cooper, D., Scimemi, A., Johannesen, K. M., et al. (2020). Current knowledge of SLC6A1-related neurodevelopmental disorders. Brain Commun. 2:fcaa170.

Google Scholar

Gorentla, B. K., and Vaughan, R. A. (2005). Differential effects of dopamine and psychoactive drugs on dopamine transporter phosphorylation and regulation. Neuropharmacology 49, 759–768. doi: 10.1016/j.neuropharm.2005.08.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Gowrishankar, R., Hahn, M. K., and Blakely, R. D. (2014). Good riddance to dopamine: roles for the dopamine transporter in synaptic function and dopamine-associated brain disorders. Neurochem. Int. 73, 42–48. doi: 10.1016/j.neuint.2013.10.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Granas, C., Ferrer, J., Loland, C. J., Javitch, J. A., and Gether, U. (2003). N-terminal truncation of the dopamine transporter abolishes phorbol ester- and substance P receptor-stimulated phosphorylation without impairing transporter internalization. J. Biol. Chem. 278, 4990–5000. doi: 10.1074/jbc.m205058200

PubMed Abstract | CrossRef Full Text | Google Scholar

Grewer, C., Balani, P., Weidenfeller, C., Bartusel, T., Tao, Z., and Rauen, T. (2005). Individual subunits of the glutamate transporter EAAC1 homotrimer function independently of each other. Biochemistry 44, 11913–11923. doi: 10.1021/bi050987n

PubMed Abstract | CrossRef Full Text | Google Scholar

Grewer, C., Madani Mobarekeh, S. A., Watzke, N., Rauen, T., and Schaper, K. (2001). Substrate translocation kinetics of excitatory amino acid carrier 1 probed with laser-pulse photolysis of a new photolabile precursor of D-aspartic acid. Biochemistry 40, 232–240. doi: 10.1021/bi0015919

PubMed Abstract | CrossRef Full Text | Google Scholar

Grewer, C., and Rauen, T. (2005). Electrogenic glutamate transporters in the CNS: molecular mechanism, pre-steady-state kinetics, and their impact on synaptic signaling. J. Membr. Biol. 203, 1–20. doi: 10.1007/s00232-004-0731-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Grewer, C., Watzke, N., Wiessner, M., and Rauen, T. (2000). Glutamate translocation of the neuronal glutamate transporter EAAC1 occurs within milliseconds. Proc. Natl. Acad. Sci. U.S.A. 97, 9706–9711. doi: 10.1073/pnas.160170397

PubMed Abstract | CrossRef Full Text | Google Scholar

Gu, H., Wall, S. C., and Rudnick, G. (1994). Stable expression of biogenic amine transporters reveals differences in inhibitor sensitivity, kinetics, and ion dependence. J. Biol. Chem. 269, 7124–7130. doi: 10.1016/s0021-9258(17)37256-3

CrossRef Full Text | Google Scholar

Guillet, B. A., Velly, L. J., Canolle, B., Masmejean, F. M., Nieoullon, A. L., and Pisano, P. (2005). Differential regulation by protein kinases of activity and cell surface expression of glutamate transporters in neuron-enriched cultures. Neurochem. Int. 46, 337–346. doi: 10.1016/j.neuint.2004.10.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Gulley, J. M., Doolen, S., and Zahniser, N. R. (2002). Brief, repeated exposure to substrates down-regulates dopamine transporter function in Xenopus oocytes in vitro and rat dorsal striatum in vivo. J. Neurochem. 83, 400–411. doi: 10.1046/j.1471-4159.2002.01133.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Guskov, A., Jensen, S., Faustino, I., Marrink, S. J., and Slotboom, D. J. (2016). Coupled binding mechanism of three sodium ions and aspartate in the glutamate transporter homologue GltTk. Nat. Commun. 7:13420.

Google Scholar

Hagiwara, T., Tanaka, K., Takai, S., Maeno-Hikichi, Y., Mukainaka, Y., and Wada, K. (1996). Genomic organization, promoter analysis, and chromosomal localization of the gene for the mouse glial high-affinity glutamate transporter Slc1a3. Genomics 33, 508–515. doi: 10.1006/geno.1996.0226

PubMed Abstract | CrossRef Full Text | Google Scholar

Hancock, J. F. (2006). Lipid rafts: contentious only from simplistic standpoints. Nat. Rev. Mol. Cell Biol. 7, 456–462. doi: 10.1038/nrm1925

PubMed Abstract | CrossRef Full Text | Google Scholar

Hansen, F. H., Skjørringe, T., Yasmeen, S., Arends, N. V., Sahai, M. A., Erreger, K., et al. (2014). Missense dopamine transporter mutations associate with adult parkinsonism and ADHD. J. Clin. Invest. 124, 3107–3120. doi: 10.1172/jci73778

PubMed Abstract | CrossRef Full Text | Google Scholar

Harsing, L. G. Jr., Solyom, S., and Salamon, C. (2001). The role of glycineB binding site and glycine transporter (GlyT1) in the regulation of [3H]GABA and [3H]glycine release in the rat brain. Neurochem. Res. 26, 915–923.

Google Scholar

Haugeto, O., Ullensvang, K., Levy, L. M., Chaudhry, F. A., Honore, T., Nielsen, M., et al. (1996). Brain glutamate transporter proteins form homomultimers. J. Biol. Chem. 271, 27715–27722. doi: 10.1074/jbc.271.44.27715

PubMed Abstract | CrossRef Full Text | Google Scholar

Hefendehl, J. K., LeDue, J., Ko, R. W., Mahler, J., Murphy, T. H., and MacVicar, B. A. (2016). Mapping synaptic glutamate transporter dysfunction in vivo to regions surrounding Abeta plaques by iGluSnFR two-photon imaging. Nat. Commun. 7:13441.

Google Scholar

Hepp, R., Perraut, M., Chasserot-Golaz, S., Galli, T., Aunis, D., Langley, K., et al. (1999). Cultured glial cells express the SNAP-25 analogue SNAP-23. Glia 27, 181–187. doi: 10.1002/(sici)1098-1136(199908)27:2<181::aid-glia8>3.0.co;2-9

CrossRef Full Text | Google Scholar

Herborg, F., Andreassen, T. F., Berlin, F., Loland, C. J., and Gether, U. (2018). Neuropsychiatric disease-associated genetic variants of the dopamine transporter display heterogeneous molecular phenotypes. J. Biol. Chem. 293, 7250–7262. doi: 10.1074/jbc.RA118.001753

PubMed Abstract | CrossRef Full Text | Google Scholar

Herde, M. K., Bohmbach, K., Domingos, C., Vana, N., Komorowska-Muller, J. A., Passlick, S., et al. (2020). Local efficacy of glutamate uptake decreases with synapse size. Cell Rep. 32:108182. doi: 10.1016/j.celrep.2020.108182

PubMed Abstract | CrossRef Full Text | Google Scholar

Hertz, L., Bock, E., and Schousboe, A. (1978). GFA content, glutamate uptake and activity of glutamate metabolizing enzymes in differentiating mouse astrocytes in primary cultures. Dev. Neurosci. 1, 226–238. doi: 10.1159/000112577

CrossRef Full Text | Google Scholar

Heshmati, M., Christoffel, D. J., LeClair, K., Cathomas, F., Golden, S. A., Aleyasin, H., et al. (2020). Depression and social defeat stress are associated with inhibitory synaptic changes in the nucleus accumbens. J. Neurosci. 40, 6228–6233. doi: 10.1523/jneurosci.2568-19.2020

PubMed Abstract | CrossRef Full Text | Google Scholar

Higuera-Matas, A., Miguens, M., Coria, S. M., Assis, M. A., Borcel, E., del Olmo, N., et al. (2012). Sex-specific disturbances of the glutamate/GABA balance in the hippocampus of adult rats subjected to adolescent cannabinoid exposure. Neuropharmacology 62, 1975–1984. doi: 10.1016/j.neuropharm.2011.12.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoffman, A. F., Lupica, C. R., and Gerhardt, G. A. (1998). Dopamine transporter activity in the substantia nigra and striatum assessed by high-speed chronoamperometric recordings in brain slices. J. Pharmacol. Exp. Ther. 287, 487–496.

Google Scholar

Hofmann, K., Duker, M., Fink, T., Lichter, P., and Stoffel, W. (1994). Human neutral amino acid transporter ASCT1: structure of the gene (SLC1A4) and localization to chromosome 2p13-p15. Genomics 24, 20–26. doi: 10.1006/geno.1994.1577

PubMed Abstract | CrossRef Full Text | Google Scholar

Holmseth, S., Scott, H. A., Real, K., Lehre, K. P., Leergaard, T. B., Bjaalie, J. G., et al. (2009). The concentrations and distributions of three C-terminal variants of the GLT1 (EAAT2; slc1a2) glutamate transporter protein in rat brain tissue suggest differential regulation. Neuroscience 162, 1055–1071. doi: 10.1016/j.neuroscience.2009.03.048

PubMed Abstract | CrossRef Full Text | Google Scholar

Holton, K. L., Loder, M. K., and Melikian, H. E. (2005). Nonclassical, distinct endocytic signals dictate constitutive and PKC-regulated neurotransmitter transporter internalization. Nat. Neurosci. 8, 881–888. doi: 10.1038/nn1478

PubMed Abstract | CrossRef Full Text | Google Scholar

Hong, W. C., and Amara, S. G. (2010). Membrane cholesterol modulates the outward facing conformation of the dopamine transporter and alters cocaine binding. J. Biol. Chem. 285, 32616–32626. doi: 10.1074/jbc.m110.150565

PubMed Abstract | CrossRef Full Text | Google Scholar

Hong, W. C., and Amara, S. G. (2013). Differential targeting of the dopamine transporter to recycling or degradative pathways during amphetamine- or PKC-regulated endocytosis in dopamine neurons. FASEB J. 27, 2995–3007. doi: 10.1096/fj.12-218727

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoover, B. R., Everett, C. V., Sorkin, A., and Zahniser, N. R. (2007). Rapid regulation of dopamine transporters by tyrosine kinases in rat neuronal preparations. J. Neurochem. 101, 1258–1271. doi: 10.1111/j.1471-4159.2007.04522.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Horton, N., and Quick, M. W. (2001). Syntaxin 1A up-regulates GABA transporter expression by subcellular redistribution. Mol. Membr. Biol. 18, 39–44. doi: 10.1080/09687680010029383

CrossRef Full Text | Google Scholar

Hu, J., Fei, J., Reutter, W., and Fan, H. (2011). Involvement of sialic acid in the regulation of gamma–aminobutyric acid uptake activity of gamma-aminobutyric acid transporter 1. Glycobiology 21, 329–339. doi: 10.1093/glycob/cwq166

PubMed Abstract | CrossRef Full Text | Google Scholar

Hu, J., and Quick, M. W. (2008). Substrate-mediated regulation of gamma-aminobutyric acid transporter 1 in rat brain. Neuropharmacology 54, 309–318. doi: 10.1016/j.neuropharm.2007.09.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, H. T., Liao, C. K., Chiu, W. T., and Tzeng, S. F. (2017). Ligands of peroxisome proliferator-activated receptor-alpha promote glutamate transporter-1 endocytosis in astrocytes. Int. J. Biochem. Cell Biol. 86, 42–53. doi: 10.1016/j.biocel.2017.03.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Hudson, B. D., Hebert, T. E., and Kelly, M. E. (2010). Physical and functional interaction between CB1 cannabinoid receptors and beta2-adrenoceptors. Br. J. Pharmacol. 160, 627–642. doi: 10.1111/j.1476-5381.2010.00681.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Huff, R. A., Vaughan, R. A., Kuhar, M. J., and Uhl, G. R. (1997). Phorbol esters increase dopamine transporter phosphorylation and decrease transport Vmax. J. Neurochem. 68, 225–232. doi: 10.1046/j.1471-4159.1997.68010225.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Imoukhuede, P. I., Moss, F. J., Michael, D. J., Chow, R. H., and Lester, H. A. (2009). Ezrin mediates tethering of the gamma-aminobutyric acid transporter GAT1 to actin filaments via a C-terminal PDZ-interacting domain. Biophys. J. 96, 2949–2960. doi: 10.1016/j.bpj.2008.11.070

PubMed Abstract | CrossRef Full Text | Google Scholar

Ingram, S. L., Prasad, B. M., and Amara, S. G. (2002). Dopamine transporter-mediated conductances increase excitability of midbrain dopamine neurons. Nat. Neurosci. 5, 971–978. doi: 10.1038/nn920

PubMed Abstract | CrossRef Full Text | Google Scholar

Jacob, C. P., Koutsilieri, E., Bartl, J., Neuen-Jacob, E., Arzberger, T., Zander, N., et al. (2007). Alterations in expression of glutamatergic transporters and receptors in sporadic Alzheimer’s disease. J. Alzheimers Dis. 11, 97–116. doi: 10.3233/jad-2007-11113

PubMed Abstract | CrossRef Full Text | Google Scholar

Jansson, L. C., Louhivuori, L., Wigren, H. K., Nordstrom, T., Louhivuori, V., Castren, M. L., et al. (2013). Effect of glutamate receptor antagonists on migrating neural progenitor cells. Eur. J. Neurosci. 37, 1369–1382. doi: 10.1111/ejn.12152

PubMed Abstract | CrossRef Full Text | Google Scholar

Jayanthi, L. D., Vargas, G., and DeFelice, L. J. (2002). Characterization of cocaine and antidepressant-sensitive norepinephrine transporters in rat placental trophoblasts. Br. J. Pharmacol. 135, 1927–1934. doi: 10.1038/sj.bjp.0704658

PubMed Abstract | CrossRef Full Text | Google Scholar

Ji, Y. F., Xu, S. M., Zhu, J., Wang, X. X., and Shen, Y. (2011). Insulin increases glutamate transporter GLT1 in cultured astrocytes. Biochem. Biophys. Res. Commun. 405, 691–696. doi: 10.1016/j.bbrc.2011.01.105

PubMed Abstract | CrossRef Full Text | Google Scholar

Ji, Y. F., Zhou, L., Xie, Y. J., Xu, S. M., Zhu, J., Teng, P., et al. (2013). Upregulation of glutamate transporter GLT-1 by mTOR-Akt-NF-small ka, CyrillicB cascade in astrocytic oxygen-glucose deprivation. Glia 61, 1959–1975. doi: 10.1002/glia.22566

PubMed Abstract | CrossRef Full Text | Google Scholar

Jiang, N. W., Wang, D. J., Xie, Y. J., Zhou, L., Su, L. D., Li, H., et al. (2016). Downregulation of glutamate transporter EAAT4 by conditional knockout of rheb1 in Cerebellar Purkinje cells. Cerebellum 15, 314–321. doi: 10.1007/s12311-015-0701-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Johnson, L. A., Furman, C. A., Zhang, M., Guptaroy, B., and Gnegy, M. E. (2005a). Rapid delivery of the dopamine transporter to the plasmalemmal membrane upon amphetamine stimulation. Neuropharmacology 49, 750–758. doi: 10.1016/j.neuropharm.2005.08.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Johnson, L. A., Guptaroy, B., Lund, D., Shamban, S., and Gnegy, M. E. (2005b). Regulation of amphetamine-stimulated dopamine efflux by protein kinase C beta. J. Biol. Chem. 280, 10914–10919. doi: 10.1074/jbc.m413887200

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, K. T., Zhen, J., and Reith, M. E. (2012). Importance of cholesterol in dopamine transporter function. J. Neurochem. 123, 700–715. doi: 10.1111/jnc.12007

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, S. R., Gainetdinov, R. R., Wightman, R. M., and Caron, M. G. (1998). Mechanisms of amphetamine action revealed in mice lacking the dopamine transporter. J. Neurosci. 18, 1979–1986. doi: 10.1523/jneurosci.18-06-01979.1998

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, S. R., Joseph, J. D., Barak, L. S., Caron, M. G., and Wightman, R. M. (1999). Dopamine neuronal transport kinetics and effects of amphetamine. J. Neurochem. 73, 2406–2414. doi: 10.1046/j.1471-4159.1999.0732406.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Kahlig, K. M., Binda, F., Khoshbouei, H., Blakely, R. D., McMahon, D. G., Javitch, J. A., et al. (2005). Amphetamine induces dopamine efflux through a dopamine transporter channel. Proc. Natl. Acad. Sci. U.S.A. 102, 3495–3500. doi: 10.1073/pnas.0407737102

PubMed Abstract | CrossRef Full Text | Google Scholar

Kahlig, K. M., and Galli, A. (2003). Regulation of dopamine transporter function and plasma membrane expression by dopamine, amphetamine, and cocaine. Eur. J. Pharmacol. 479, 153–158. doi: 10.1016/j.ejphar.2003.08.065

PubMed Abstract | CrossRef Full Text | Google Scholar

Kahlig, K. M., Javitch, J. A., and Galli, A. (2004). Amphetamine regulation of dopamine transport. Combined measurements of transporter currents and transporter imaging support the endocytosis of an active carrier. J. Biol. Chem. 279, 8966–8975.

Google Scholar

Kalandadze, A., Wu, Y., and Robinson, M. B. (2002). Protein kinase C activation decreases cell surface expression of the GLT-1 subtype of glutamate transporter. Requirement of a carboxyl-terminal domain and partial dependence on serine 486. J. Biol. Chem. 277, 45741–45750. doi: 10.1074/jbc.m203771200

PubMed Abstract | CrossRef Full Text | Google Scholar

Kang, M., Ryu, J., Kim, J. H., Na, H., Zuo, Z., and Do, S. H. (2010). Corticosterone decreases the activity of rat glutamate transporter type 3 expressed in Xenopus oocytes. Steroids 75, 1113–1118. doi: 10.1016/j.steroids.2010.07.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Kanner, B. I. (2003). Transmembrane domain I of the gamma-aminobutyric acid transporter GAT-1 plays a crucial role in the transition between cation leak and transport modes. J. Biol. Chem. 278, 3705–3712. doi: 10.1074/jbc.m210525200

PubMed Abstract | CrossRef Full Text | Google Scholar

Karam, C. S., and Javitch, J. A. (2018). Phosphorylation of the amino terminus of the dopamine transporter: regulatory mechanisms and implications for amphetamine action. Adv. Pharmacol. 82, 205–234. doi: 10.1016/bs.apha.2017.09.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Karatas-Wulf, U., Koepsell, H., Bergert, M., Sonnekes, S., and Kugler, P. (2009). Protein kinase C-dependent trafficking of glutamate transporters excitatory amino acid carrier 1 and glutamate transporter 1b in cultured cerebellar granule cells. Neuroscience 161, 794–805. doi: 10.1016/j.neuroscience.2009.04.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Karki, P., Webb, A., Smith, K., Johnson, J. Jr., Lee, K., Son, D. S., et al. (2014). Yin Yang 1 is a repressor of glutamate transporter EAAT2, and it mediates manganese-induced decrease of EAAT2 expression in astrocytes. Mol. Cell Biol. 34, 1280–1289. doi: 10.1128/mcb.01176-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Karylowski, O., Zeigerer, A., Cohen, A., and McGraw, T. E. (2004). GLUT4 is retained by an intracellular cycle of vesicle formation and fusion with endosomes. Mol. Biol. Cell 15, 870–882. doi: 10.1091/mbc.e03-07-0517

PubMed Abstract | CrossRef Full Text | Google Scholar

Kerkerian, L., Dusticier, N., and Nieoullon, A. (1987). Modulatory effect of dopamine on high-affinity glutamate uptake in the rat striatum. J. Neurochem. 48, 1301–1306. doi: 10.1111/j.1471-4159.1987.tb05661.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Keynan, S., and Kanner, B. I. (1988). gamma-Aminobutyric acid transport in reconstituted preparations from rat brain: coupled sodium and chloride fluxes. Biochemistry 27, 12–17. doi: 10.1021/bi00401a003

PubMed Abstract | CrossRef Full Text | Google Scholar

Khoshbouei, H., Sen, N., Guptaroy, B., Johnson, L., Lund, D., Gnegy, M. E., et al. (2004). N-terminal phosphorylation of the dopamine transporter is required for amphetamine-induced efflux. PLoS Biol. 2:E78. doi: 10.1371/journal.pbio.0020078

PubMed Abstract | CrossRef Full Text | Google Scholar

Kilty, J. E., Lorang, D., and Amara, S. G. (1991). Cloning and expression of a cocaine-sensitive rat dopamine transporter. Science 254, 578–579. doi: 10.1126/science.1948035

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, D. U., Kim, M. K., Cho, Y. W., Kim, Y. S., Kim, W. J., Lee, M. G., et al. (2011). Association of a synonymous GAT3 polymorphism with antiepileptic drug pharmacoresistance. J. Hum. Genet. 56, 640–646. doi: 10.1038/jhg.2011.73

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, S., Westphalen, R., Callahan, B., Hatzidimitriou, G., Yuan, J., and Ricaurte, G. A. (2000). Toward development of an in vitro model of methamphetamine-induced dopamine nerve terminal toxicity. J. Pharmacol. Exp. Ther. 293, 625–633.

Google Scholar

Kimmel, H. L., Carroll, F. I., and Kuhar, M. J. (2000). Dopamine transporter synthesis and degradation rate in rat striatum and nucleus accumbens using RTI-76. Neuropharmacology 39, 578–585. doi: 10.1016/s0028-3908(99)00160-4

CrossRef Full Text | Google Scholar

Kinney, G. A. (2005). GAT-3 transporters regulate inhibition in the neocortex. J, Neurophysiol. 94, 4533–4537. doi: 10.1152/jn.00420.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Kitayama, S., Dohi, T., and Uhl, G. R. (1994). Phorbol esters alter functions of the expressed dopamine transporter. Eur. J. Pharmacol. 268, 115–119. doi: 10.1016/0922-4106(94)90180-5

CrossRef Full Text | Google Scholar

Kivell, B., Uzelac, Z., Sundaramurthy, S., Rajamanickam, J., Ewald, A., Chefer, V., et al. (2014). Salvinorin A regulates dopamine transporter function via a kappa opioid receptor and ERK1/2-dependent mechanism. Neuropharmacology 86, 228–240. doi: 10.1016/j.neuropharm.2014.07.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Koch, H. P., Brown, R. L., and Larsson, H. P. (2007). The glutamate-activated anion conductance in excitatory amino acid transporters is gated independently by the individual subunits. J. Neurosci. 27, 2943–2947. doi: 10.1523/jneurosci.0118-07.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Kolen, B., Kortzak, D., Franzen, A., and Fahlke, C. (2020). An amino-terminal point mutation increases EAAT2 anion currents without affecting glutamate transport rates. J. Biol. Chem. 295, 14936–14947. doi: 10.1074/jbc.ra120.013704

PubMed Abstract | CrossRef Full Text | Google Scholar

Kovtun, O., Sakrikar, D., Tomlinson, I. D., Chang, J. C., Arzeta-Ferrer, X., Blakely, R. D., et al. (2015). Single-quantum-dot tracking reveals altered membrane dynamics of an attention-deficit/hyperactivity-disorder-derived dopamine transporter coding variant. ACS Chem. Neurosci. 6, 526–534. doi: 10.1021/cn500202c

PubMed Abstract | CrossRef Full Text | Google Scholar

Kristensen, A. S., Andersen, J., Jorgensen, T. N., Sorensen, L., Eriksen, J., Loland, C. J., et al. (2011). SLC6 neurotransmitter transporters: structure, function, and regulation. Pharmacol. Rev. 63, 585–640.

Google Scholar

Krueger, B. K. (1990). Kinetics and block of dopamine uptake in synaptosomes from rat caudate nucleus. J. Neurochem. 55, 260–267. doi: 10.1111/j.1471-4159.1990.tb08847.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Laughlin, T. M., Tram, K. V., Wilcox, G. L., and Birnbaum, A. K. (2002). Comparison of antiepileptic drugs tiagabine, lamotrigine, and gabapentin in mouse models of acute, prolonged, and chronic nociception. J. Pharmacol. Exp. Ther. 302, 1168–1175. doi: 10.1124/jpet.302.3.1168

PubMed Abstract | CrossRef Full Text | Google Scholar

Law, R. M., Stafford, A., and Quick, M. W. (2000). Functional regulation of gamma-aminobutyric acid transporters by direct tyrosine phosphorylation. J. Biol. Chem. 275, 23986–23991. doi: 10.1074/jbc.m910283199

PubMed Abstract | CrossRef Full Text | Google Scholar

Leary, G. P., Holley, D. C., Stone, E. F., Lyda, B. R., Kalachev, L. V., and Kavanaugh, M. P. (2011). The central cavity in trimeric glutamate transporters restricts ligand diffusion. Proc. Natl. Acad. Sci. U.S.A. 108, 14980–14985. doi: 10.1073/pnas.1108785108

PubMed Abstract | CrossRef Full Text | Google Scholar

Leary, G. P., Stone, E. F., Holley, D. C., and Kavanaugh, M. P. (2007). The glutamate and chloride permeation pathways are colocalized in individual neuronal glutamate transporter subunits. J. Neurosci. 27, 2938–2942. doi: 10.1523/jneurosci.4851-06.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, F. J., Liu, F., Pristupa, Z. B., and Niznik, H. B. (2001). Direct binding and functional coupling of alpha-synuclein to the dopamine transporters accelerate dopamine-induced apoptosis. FASEB J. 15, 916–926. doi: 10.1096/fsb2fj000334com

CrossRef Full Text | Google Scholar

Lee, F. J., Pei, L., Moszczynska, A., Vukusic, B., Fletcher, P. J., and Liu, F. (2007). Dopamine transporter cell surface localization facilitated by a direct interaction with the dopamine D2 receptor. EMBO J. 26, 2127–2136. doi: 10.1038/sj.emboj.7601656

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, G., Huang, Y., Washington, J. M., Briggs, N. W., and Zuo, Z. (2005). Carbamazepine enhances the activity of glutamate transporter type 3 via phosphatidylinositol 3-kinase. Epilepsy Res. 66, 145–153. doi: 10.1016/j.eplepsyres.2005.08.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, H. S., Bae, E. J., Yi, S. H., Shim, J. W., Jo, A. Y., Kang, J. S., et al. (2010). Foxa2 and Nurr1 synergistically yield A9 nigral dopamine neurons exhibiting improved differentiation, function, and cell survival. Stem Cells 28, 501–512.

Google Scholar

Lee, K. H., Kim, M. Y., Kim, D. H., and Lee, Y. S. (2004). Syntaxin 1A and receptor for activated C kinase interact with the N-terminal region of human dopamine transporter. Neurochem. Res. 29, 1405–1409. doi: 10.1023/b:nere.0000026404.08779.43

CrossRef Full Text | Google Scholar

Lehre, K. P., and Danbolt, N. C. (1998). The number of glutamate transporter subtype molecules at glutamatergic synapses: chemical and stereological quantification in young adult rat brain. J. Neurosci. 18, 8751–8757. doi: 10.1523/jneurosci.18-21-08751.1998

PubMed Abstract | CrossRef Full Text | Google Scholar

Lehre, K. P., Levy, L. M., Ottersen, O. P., Storm-Mathisen, J., and Danbolt, N. C. (1995). Differential expression of two glial glutamate transporters in the rat brain: quantitative and immunocytochemical observations. J. Neurosci. 15, 1835–1853. doi: 10.1523/jneurosci.15-03-01835.1995

PubMed Abstract | CrossRef Full Text | Google Scholar

Leng, K., Li, E., Eser, R., Piergies, A., Sit, R., Tan, M., et al. (2021). Molecular characterization of selectively vulnerable neurons in Alzheimer’s disease. Nat. Neurosci. 24, 276–287.

Google Scholar

Leonova, J., Thorlin, T., Aberg, N. D., Eriksson, P. S., Ronnback, L., and Hansson, E. (2001). Endothelin-1 decreases glutamate uptake in primary cultured rat astrocytes. Am. J. Physiol. Cell Physiol. 281, C1495–C1503.

Google Scholar

Levy, L., Lehre, K., Walaas, S., Storm-Mathisen, J., and Danbolt, N. (1995). Down-regulation of glial glutamate transporters after glutamatergic denervation in the rat brain. Eur. J. Neurosci. 7, 2036–2041. doi: 10.1111/j.1460-9568.1995.tb00626.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Levy, L. M., Lehre, K. P., Rolstad, B., and Danbolt, N. C. (1993). A monoclonal antibody raised against an [Na(+)+K+]coupled L-glutamate transporter purified from rat brain confirms glial cell localization. FEBS Lett. 317, 79–84. doi: 10.1016/0014-5793(93)81495-l

CrossRef Full Text | Google Scholar

Levy, L. M., Warr, O., and Attwell, D. (1998). Stoichiometry of the glial glutamate transporter GLT-1 expressed inducibly in a Chinese hamster ovary cell line selected for low endogenous Na+-dependent glutamate uptake. J. Neurosci. 18, 9620–9628. doi: 10.1523/jneurosci.18-23-09620.1998

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, L. B., Toan, S. V., Zelenaia, O., Watson, D. J., Wolfe, J. H., Rothstein, J. D., et al. (2006). Regulation of astrocytic glutamate transporter expression by Akt: evidence for a selective transcriptional effect on the GLT-1/EAAT2 subtype. J. Neurochem. 97, 759–771. doi: 10.1111/j.1471-4159.2006.03743.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, M. H., Underhill, S. M., Reed, C., Phillips, T. J., Amara, S. G., and Ingram, S. L. (2017). Amphetamine and methamphetamine increase NMDAR-GluN2B synaptic currents in midbrain dopamine neurons. Neuropsychopharmacology 42, 1539–1547. doi: 10.1038/npp.2016.278

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, S., Mallory, M., Alford, M., Tanaka, S., and Masliah, E. (1997). Glutamate transporter alterations in Alzheimer disease are possibly associated with abnormal APP expression. J. Neuropathol. Exp. Neurol. 56, 901–911. doi: 10.1097/00005072-199708000-00008

PubMed Abstract | CrossRef Full Text | Google Scholar

Lim, S. K., Park, M. J., Jung, H. K., Park, A. Y., Kim, D. I., Kim, J. C., et al. (2008). Bradykinin stimulates glutamate uptake via both B1R and B2R activation in a human retinal pigment epithelial cells. Life Sci. 83, 761–770. doi: 10.1016/j.lfs.2008.09.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Lin, C. L., Tzingounis, A. V., Jin, L., Furuta, A., Kavanaugh, M. P., and Rothstein, J. D. (1998). Molecular cloning and expression of the rat EAAT4 glutamate transporter subtype. Brain Res. Mol. Brain Res. 63, 174–179. doi: 10.1016/s0169-328x(98)00256-3

CrossRef Full Text | Google Scholar

Loder, M. K., and Melikian, H. E. (2003). The dopamine transporter constitutively internalizes and recycles in a protein kinase C-regulated manner in stably transfected PC12 cell lines. J. Biol. Chem. 278, 22168–22174. doi: 10.1074/jbc.m301845200

PubMed Abstract | CrossRef Full Text | Google Scholar

Lu, C.-C., and Hilgemann, D. W. (1999). GAT1 (GABA: Na+: Cl–) cotransport function: steady state studies in giant Xenopus oocyte membrane patches. J. Gen. Physiol. 114, 429–444. doi: 10.1085/jgp.114.3.429

PubMed Abstract | CrossRef Full Text | Google Scholar

Lushnikova, I., Skibo, G., Muller, D., and Nikonenko, I. (2009). Synaptic potentiation induces increased glial coverage of excitatory synapses in CA1 hippocampus. Hippocampus 19, 753–762. doi: 10.1002/hipo.20551

PubMed Abstract | CrossRef Full Text | Google Scholar

Lute, B. J., Khoshbouei, H., Saunders, C., Sen, N., Lin, R. Z., Javitch, J. A., et al. (2008). PI3K signaling supports amphetamine-induced dopamine efflux. Biochem. Biophys. Res. Commun. 372, 656–661. doi: 10.1016/j.bbrc.2008.05.091

PubMed Abstract | CrossRef Full Text | Google Scholar

MacAulay, N., Zeuthen, T., and Gether, U. (2002). Conformational basis for the Li(+)-induced leak current in the rat gamma-aminobutyric acid (GABA) transporter-1. J. Physiol. 544, 447–458. doi: 10.1113/jphysiol.2002.022897

PubMed Abstract | CrossRef Full Text | Google Scholar

Mager, S., Min, C., Henry, D. J., Chavkin, C., Hoffman, B. J., Davidson, N., et al. (1994). Conducting states of a mammalian serotonin transporter. Neuron 12, 845–859. doi: 10.1016/0896-6273(94)90337-9

CrossRef Full Text | Google Scholar

Mager, S., Naeve, J., Quick, M., Labarca, C., Davidson, N., and Lester, H. A. (1993). Steady states, charge movements, and rates for a cloned GABA transporter expressed in Xenopus oocytes. Neuron 10, 177–188. doi: 10.1016/0896-6273(93)90309-f

CrossRef Full Text | Google Scholar

Malik, A. R., and Willnow, T. E. (2019). Excitatory amino acid transporters in physiology and disorders of the central nervous system. Int. J. Mol. Sci. 20:5671. doi: 10.3390/ijms20225671

PubMed Abstract | CrossRef Full Text | Google Scholar

Maragakis, N. J., Dykes-Hoberg, M., and Rothstein, J. D. (2004). Altered expression of the glutamate transporter EAAT2b in neurological disease. Ann. Neurol. 55, 469–477. doi: 10.1002/ana.20003

PubMed Abstract | CrossRef Full Text | Google Scholar

Martinez-Lozada, Z., Guillem, A. M., and Robinson, M. B. (2016). Transcriptional regulation of glutamate transporters: from extracellular signals to transcription factors. Adv. Pharmacol. 76, 103–145.

Google Scholar

Martins, R. S., de Freitas, I. G., Sathler, M. F., Martins, V., Schitine, C. S., da Silva Sampaio, L., et al. (2018). Beta-adrenergic receptor activation increases GABA uptake in adolescent mice frontal cortex: Modulation by cannabinoid receptor agonist WIN55,212-2. Neurochem. Int. 120, 182–190. doi: 10.1016/j.neuint.2018.08.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Mason, J. N., Farmer, H., Tomlinson, I. D., Schwartz, J. W., Savchenko, V., DeFelice, L. J., et al. (2005). Novel fluorescence-based approaches for the study of biogenic amine transporter localization, activity, and regulation. J. Neurosci. Methods 143, 3–25. doi: 10.1016/j.jneumeth.2004.09.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Matsusaki, M., Kishida, A., Stainton, N., Ansell, C. W. G., and Akashi, M. (2001). Synthesis and characterization of novel biodegradable polymers composed of hydroxycinnamic acid and D,L-lactic acid. J. Appl. Polym. Sci. 82, 2357–2364. doi: 10.1002/app.2085

CrossRef Full Text | Google Scholar

Matthews, E. Jr., Rahnama-Vaghef, A., and Eskandari, S. (2009). Inhibitors of the γ-aminobutyric acid transporter 1 (GAT1) do not reveal a channel mode of conduction. Neurochem. Intern. 55, 732–740. doi: 10.1016/j.neuint.2009.07.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Mazei-Robison, M. S., Couch, R. S., Shelton, R. C., Stein, M. A., and Blakely, R. D. (2005). Sequence variation in the human dopamine transporter gene in children with attention deficit hyperactivity disorder. Neuropharmacology 49, 724–736. doi: 10.1016/j.neuropharm.2005.08.003

PubMed Abstract | CrossRef Full Text | Google Scholar

McElvain, J. S., and Schenk, J. O. (1992). A multisubstrate mechanism of striatal dopamine uptake and its inhibition by cocaine. Biochem. Pharmacol. 43, 2189–2199. doi: 10.1016/0006-2952(92)90178-l

CrossRef Full Text | Google Scholar

McGeer, P., McGeer, E., Scherer, U., and Singh, K. (1977). A glutamatergic corticostriatal path? Brain Res. 128, 369–373. doi: 10.1016/0006-8993(77)91003-4

CrossRef Full Text | Google Scholar

McHugh, E. M., Zhu, W., Milgram, S., and Mager, S. (2004). The GABA transporter GAT1 and the MAGUK protein Pals1: interaction, uptake modulation, and coexpression in the brain. Mol. Cell Neurosci. 26, 406–417. doi: 10.1016/j.mcn.2004.03.006

PubMed Abstract | CrossRef Full Text | Google Scholar

McNair, L. F., Andersen, J. V., Aldana, B. I., Hohnholt, M. C., Nissen, J. D., Sun, Y., et al. (2019). Deletion of neuronal GLT-1 in mice reveals its role in synaptic glutamate homeostasis and mitochondrial function. J. Neurosci. 39, 4847–4863. doi: 10.1523/jneurosci.0894-18.2019

PubMed Abstract | CrossRef Full Text | Google Scholar

McNair, L. F., Andersen, J. V., Nissen, J. D., Sun, Y., Fischer, K. D., Hodgson, N. W., et al. (2020). Conditional knockout of GLT-1 in neurons leads to alterations in aspartate homeostasis and synaptic mitochondrial metabolism in Striatum and Hippocampus. Neurochem. Res. 45, 1420–1437. doi: 10.1007/s11064-020-03000-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Medvedev, N., Popov, V., Henneberger, C., Kraev, I., Rusakov, D. A., and Stewart, M. G. (2014). Glia selectively approach synapses on thin dendritic spines. Philos. Trans. R. Soc. Lond. B Biol. Sci. 369:20140047. doi: 10.1098/rstb.2014.0047

PubMed Abstract | CrossRef Full Text | Google Scholar

Melikian, H. E., and Buckley, K. M. (1999). Membrane trafficking regulates the activity of the human dopamine transporter. J. Neurosci. 19, 7699–7710. doi: 10.1523/jneurosci.19-18-07699.1999

PubMed Abstract | CrossRef Full Text | Google Scholar

Melone, M., Barbaresi, P., Fattorini, G., and Conti, F. (2005). Neuronal localization of the GABA transporter GAT-3 in human cerebral cortex: a procedural artifact? J. Chem. Neuroanat. 30, 45–54. doi: 10.1016/j.jchemneu.2005.04.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Melone, M., Bellesi, M., and Conti, F. (2009). Synaptic localization of GLT-1a in the rat somatic sensory cortex. Glia 57, 108–117. doi: 10.1002/glia.20744

PubMed Abstract | CrossRef Full Text | Google Scholar

Melone, M., Bellesi, M., Ducati, A., Iacoangeli, M., and Conti, F. (2011). Cellular and synaptic localization of EAAT2a in human cerebral cortex. Front. Neuroanat. 4:151. doi: 10.3389/fnana.2010.00151

PubMed Abstract | CrossRef Full Text | Google Scholar

Melone, M., Ciappelloni, S., and Conti, F. (2015). A quantitative analysis of cellular and synaptic localization of GAT-1 and GAT-3 in rat neocortex. Brain Struct. Funct. 220, 885–897. doi: 10.1007/s00429-013-0690-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Melone, M., Ciriachi, C., Pietrobon, D., and Conti, F. (2019). Heterogeneity of Astrocytic and neuronal GLT-1 at cortical excitatory synapses, as revealed by its colocalization with Na+/K+-ATPase alpha isoforms. Cereb. Cortex 29, 3331–3350. doi: 10.1093/cercor/bhy203

PubMed Abstract | CrossRef Full Text | Google Scholar

Melone, M., Cozzi, A., Pellegrini-Giampietro, D. E., and Conti, F. (2003). Transient focal ischemia triggers neuronal expression of GAT-3 in the rat perilesional cortex. Neurobiol. Dis. 14, 120–132. doi: 10.1016/s0969-9961(03)00042-1

CrossRef Full Text | Google Scholar

Mergy, M. A., Gowrishankar, R., Davis, G. L., Jessen, T. N., Wright, J., Stanwood, G. D., et al. (2014). Genetic targeting of the amphetamine and methylphenidate-sensitive dopamine transporter: on the path to an animal model of attention-deficit hyperactivity disorder. Neurochem. Int. 73, 56–70. doi: 10.1016/j.neuint.2013.11.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Metzger, R. R., Haughey, H. M., Wilkins, D. G., Gibb, J. W., Hanson, G. R., and Fleckenstein, A. E. (2000). Methamphetamine-induced rapid decrease in dopamine transporter function: role of dopamine and hyperthermia. J. Pharmacol. Exp. Ther. 295, 1077–1085.

Google Scholar

Meyer, T., Speer, A., Meyer, B., Sitte, W., Kuther, G., and Ludolph, A. C. (1996). The glial glutamate transporter complementary DNA in patients with amyotrophic lateral sclerosis. Ann. Neurol. 40, 456–459. doi: 10.1002/ana.410400317

PubMed Abstract | CrossRef Full Text | Google Scholar

Michaluk, P. H., Heller, J., and Rusakov, D. A. (2020). Rapid recycling of glutamate transporters on the astroglial surface. bioRxiv [Preprint], doi: 10.1101/2020.11.08.373233v1

CrossRef Full Text | Google Scholar

Mick, E., and Faraone, S. V. (2009). Family and genetic association studies of bipolar disorder in children. Child Adolesc. Psychiatr. Clin. N. Am. 18, 441–453. doi: 10.1016/j.chc.2008.11.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Mim, C., Balani, P., Rauen, T., and Grewer, C. (2005). The glutamate transporter subtypes EAAT4 and EAATs 1-3 transport glutamate with dramatically different kinetics and voltage dependence but share a common uptake mechanism. J. Gen. Physiol. 126, 571–589. doi: 10.1085/jgp.200509365

PubMed Abstract | CrossRef Full Text | Google Scholar

Minelli, A., Barbaresi, P., Reimer, R. J., Edwards, R. H., and Conti, F. (2001). The glial glutamate transporter GLT-1 is localized both in the vicinity of and at distance from axon terminals in the rat cerebral cortex. Neuroscience 108, 51–59. doi: 10.1016/s0306-4522(01)00375-x

CrossRef Full Text | Google Scholar

Miranda, M., Dionne, K. R., Sorkina, T., and Sorkin, A. (2007). Three ubiquitin conjugation sites in the amino terminus of the dopamine transporter mediate protein kinase C-dependent endocytosis of the transporter. Mol. Biol. Cell 18, 313–323. doi: 10.1091/mbc.e06-08-0704

PubMed Abstract | CrossRef Full Text | Google Scholar

Mitrovic, A. D., Plesko, F., and Vandenberg, R. J. (2001). Zn(2+) inhibits the anion conductance of the glutamate transporter EEAT4. J. Biol. Chem. 276, 26071–26076. doi: 10.1074/jbc.m011318200

PubMed Abstract | CrossRef Full Text | Google Scholar

Mookherjee, P., Green, P. S., Watson, G. S., Marques, M. A., Tanaka, K., Meeker, K. D., et al. (2011). GLT-1 loss accelerates cognitive deficit onset in an Alzheimer’s disease animal model. J. Alzheimers Dis. 26, 447–455. doi: 10.3233/jad-2011-110503

PubMed Abstract | CrossRef Full Text | Google Scholar

Moritz, A. E., Foster, J. D., Gorentla, B. K., Mazei-Robison, M. S., Yang, J. W., Sitte, H. H., et al. (2013). Phosphorylation of dopamine transporter serine 7 modulates cocaine analog binding. J. Biol. Chem. 288, 20–32. doi: 10.1074/jbc.m112.407874

PubMed Abstract | CrossRef Full Text | Google Scholar

Moron, J. A., Zakharova, I., Ferrer, J. V., Merrill, G. A., Hope, B., Lafer, E. M., et al. (2003). Mitogen-activated protein kinase regulates dopamine transporter surface expression and dopamine transport capacity. J. Neurosci. 23, 8480–8488. doi: 10.1523/jneurosci.23-24-08480.2003

PubMed Abstract | CrossRef Full Text | Google Scholar

Moss, F. J., Imoukhuede, P. I., Scott, K., Hu, J., Jankowsky, J. L., Quick, M. W., et al. (2009). GABA transporter function, oligomerization state, and anchoring: correlates with subcellularly resolved FRET. J. Gen. Physiol. 134, 489–521. doi: 10.1085/jgp.200910314

PubMed Abstract | CrossRef Full Text | Google Scholar

Moszczynska, A., Saleh, J., Zhang, H., Vukusic, B., Lee, F. J., and Liu, F. (2007). Parkin disrupts the alpha-synuclein/dopamine transporter interaction: consequences toward dopamine-induced toxicity. J. Mol. Neurosci. 32, 217–227. doi: 10.1007/s12031-007-0037-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Mukainaka, Y., Tanaka, K., Hagiwara, T., and Wada, K. (1995). Molecular cloning of two glutamate transporter subtypes from mouse brain. Biochim. Biophys. Acta 1244, 233–237. doi: 10.1016/0304-4165(95)00062-g

CrossRef Full Text | Google Scholar

Mundorf, M. L., Hochstetler, S. E., and Wightman, R. M. (1999). Amine weak bases disrupt vesicular storage and promote exocytosis in chromaffin cells. J. Neurochem. 73, 2397–2405. doi: 10.1046/j.1471-4159.1999.0732397.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Murphy-Royal, C., Dupuis, J. P., Varela, J. A., Panatier, A., Pinson, B., Baufreton, J., et al. (2015). Surface diffusion of astrocytic glutamate transporters shapes synaptic transmission. Nat. Neurosci. 18, 219–226. doi: 10.1038/nn.3901

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakayama, T., Kawakami, H., Tanaka, K., and Nakamura, S. (1996). Expression of three glutamate transporter subtype mRNAs in human brain regions and peripheral tissues. Brain Res. Mol. Brain Res. 36, 189–192. doi: 10.1016/0169-328x(95)00297-6

CrossRef Full Text | Google Scholar

Navaroli, D. M., Stevens, Z. H., Uzelac, Z., Gabriel, L., King, M. J., Lifshitz, L. M., et al. (2011). The plasma membrane-associated GTPase Rin interacts with the dopamine transporter and is required for protein kinase C-regulated dopamine transporter trafficking. J. Neurosci. 31, 13758–13770. doi: 10.1523/jneurosci.2649-11.2011

PubMed Abstract | CrossRef Full Text | Google Scholar

Nirenberg, M. J., Chan, J., Pohorille, A., Vaughan, R. A., Uhl, G. R., Kuhar, M. J., et al. (1997). The dopamine transporter: comparative ultrastructure of dopaminergic axons in limbic and motor compartments of the nucleus accumbens. J. Neurosci. 17, 6899–6907. doi: 10.1523/jneurosci.17-18-06899.1997

PubMed Abstract | CrossRef Full Text | Google Scholar

Norregaard, L., and Gether, U. (2001). The monoamine neurotransmitter transporters: structure, conformational changes and molecular gating. Curr. Opin. Drug Discov. Dev. 4, 591–601.

Google Scholar

O’Donovan, S. M., Sullivan, C. R., and McCullumsmith, R. E. (2017). The role of glutamate transporters in the pathophysiology of neuropsychiatric disorders. NPJ Schizophr. 3:32.

Google Scholar

Otis, T. S., and Kavanaugh, M. P. (2000). Isolation of current components and partial reaction cycles in the glial glutamate transporter EAAT2. J. Neurosci. 20, 2749–2757. doi: 10.1523/jneurosci.20-08-02749.2000

PubMed Abstract | CrossRef Full Text | Google Scholar

Padovano, V., Massari, S., Mazzucchelli, S., and Pietrini, G. (2009). PKC induces internalization and retention of the EAAC1 glutamate transporter in recycling endosomes of MDCK cells. Am. J. Physiol. Cell Physiol. 297, C835–C844.

Google Scholar

Page, G., Barc-Pain, S., Pontcharraud, R., Cante, A., Piriou, A., and Barrier, L. (2004). The up-regulation of the striatal dopamine transporter’s activity by cAMP is PKA-, CaMK II- and phosphatase-dependent. Neurochem. Int. 45, 627–632. doi: 10.1016/j.neuint.2004.04.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Pakulska, W. (2007). Influence of tiagabine on the antinociceptive action of morphine, metamizole and indomethacin in mice. Acta Pol. Pharm. 64, 263–270.

Google Scholar

Palos, T. P., Ramachandran, B., Boado, R., and Howard, B. D. (1996). Rat C6 and human astrocytic tumor cells express a neuronal type of glutamate transporter. Brain Res. Mol. Brain Res. 37, 297–303. doi: 10.1016/0169-328x(95)00331-l

CrossRef Full Text | Google Scholar

Parinejad, N., Peco, E., Ferreira, T., Stacey, S. M., and van Meyel, D. J. (2016). Disruption of an EAAT-mediated chloride channel in a Drosophila model of Ataxia. J. Neurosci. 36, 7640–7647. doi: 10.1523/jneurosci.0197-16.2016

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, H. Y., Kim, J. H., Zuo, Z., and Do, S. H. (2008). Ethanol increases the activity of rat excitatory amino acid transporter type 4 expressed in Xenopus oocytes: role of protein kinase C and phosphatidylinositol 3-kinase. Alcohol. Clin. Exp. Res. 32, 348–354. doi: 10.1111/j.1530-0277.2007.00577.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, S. J., Shin, H. J., Gu, B. W., Woo, K. I., Zuo, Z., Do, S. H., et al. (2015). Desflurane increased the activity of excitatory amino-acid carrier 1 (EAAC1) expressed in Xenopus oocytes. Eur. J. Pharmacol. 757, 84–89. doi: 10.1016/j.ejphar.2015.03.058

PubMed Abstract | CrossRef Full Text | Google Scholar

Parker, E. M., and Cubeddu, L. X. (1988). Comparative effects of amphetamine, phenylethylamine and related drugs on dopamine efflux, dopamine uptake and mazindol binding. J. Pharmacol. Exp. Ther. 245, 199–210.

Google Scholar

Parpura, V., Fang, Y., Basarsky, T., Jahn, R., and Haydon, P. G. (1995). Expression of synaptobrevin II, cellubrevin and syntaxin but not SNAP-25 in cultured astrocytes. FEBS Lett. 377, 489–492. doi: 10.1016/0014-5793(95)01401-2

CrossRef Full Text | Google Scholar

Patrushev, I., Gavrilov, N., Turlapov, V., and Semyanov, A. (2013). Subcellular location of astrocytic calcium stores favors extrasynaptic neuron-astrocyte communication. Cell Calc. 54, 343–349. doi: 10.1016/j.ceca.2013.08.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Peacey, E., Miller, C. C., Dunlop, J., and Rattray, M. (2009). The four major N- and C-terminal splice variants of the excitatory amino acid transporter GLT-1 form cell surface homomeric and heteromeric assemblies. Mol. Pharmacol. 75, 1062–1073. doi: 10.1124/mol.108.052829

PubMed Abstract | CrossRef Full Text | Google Scholar

Peghini, P., Janzen, J., and Stoffel, W. (1997). Glutamate transporter EAAC-1-deficient mice develop dicarboxylic aminoaciduria and behavioral abnormalities but no neurodegeneration. EMBO J. 16, 3822–3832. doi: 10.1093/emboj/16.13.3822

PubMed Abstract | CrossRef Full Text | Google Scholar

Penmatsa, A., Wang, K. H., and Gouaux, E. (2013). X-ray structure of dopamine transporter elucidates antidepressant mechanism. Nature 503, 85–90. doi: 10.1038/nature12533

PubMed Abstract | CrossRef Full Text | Google Scholar

Penmatsa, A., Wang, K. H., and Gouaux, E. (2015). X-ray structures of Drosophila dopamine transporter in complex with Nisoxetine and Reboxetine. Nat. Struct. Mol. Biol. 22, 506–508. doi: 10.1038/nsmb.3029

PubMed Abstract | CrossRef Full Text | Google Scholar

Perego, C., Di Cairano, E. S., Ballabio, M., and Magnaghi, V. (2012). Neurosteroid allopregnanolone regulates EAAC1-mediated glutamate uptake and triggers actin changes in Schwann cells. J. Cell Physiol. 227, 1740–1751. doi: 10.1002/jcp.22898

PubMed Abstract | CrossRef Full Text | Google Scholar

Perez-Alvarez, A., Navarrete, M., Covelo, A., Martin, E. D., and Araque, A. (2014). Structural and functional plasticity of astrocyte processes and dendritic spine interactions. J. Neurosci. 34, 12738–12744. doi: 10.1523/jneurosci.2401-14.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

Petr, G. T., Sun, Y., Frederick, N. M., Zhou, Y., Dhamne, S. C., Hameed, M. Q., et al. (2015). Conditional deletion of the glutamate transporter GLT-1 reveals that astrocytic GLT-1 protects against fatal epilepsy while neuronal GLT-1 contributes significantly to glutamate uptake into synaptosomes. J. Neurosci. 35, 5187–5201. doi: 10.1523/jneurosci.4255-14.2015

PubMed Abstract | CrossRef Full Text | Google Scholar

Piani, D., Frei, K., Pfister, H. W., and Fontana, A. (1993). Glutamate uptake by astrocytes is inhibited by reactive oxygen intermediates but not by other macrophage-derived molecules including cytokines, leukotrienes or platelet-activating factor. J. Neuroimmunol. 48, 99–104. doi: 10.1016/0165-5728(93)90063-5

CrossRef Full Text | Google Scholar

Picaud, S., Larsson, H. P., Wellis, D. P., Lecar, H., and Werblin, F. (1995). Cone photoreceptors respond to their own glutamate release in the tiger salamander. Proc. Natl. Acad. Sci. U.S.A. 92, 9417–9421. doi: 10.1073/pnas.92.20.9417

PubMed Abstract | CrossRef Full Text | Google Scholar

Pines, G., Danbolt, N. C., Bjoras, M., Zhang, Y., Bendahan, A., Eide, L., et al. (1992). Cloning and expression of a rat brain L-glutamate transporter. Nature 360, 464–467.

Google Scholar

Plachez, C., Danbolt, N. C., and Recasens, M. (2000). Transient expression of the glial glutamate transporters GLAST and GLT in hippocampal neurons in primary culture. J. Neurosci. Res. 59, 587–593. doi: 10.1002/(sici)1097-4547(20000301)59:5<587::aid-jnr1>3.0.co;2-l

CrossRef Full Text | Google Scholar

Pogun, S., Dawson, V., and Kuhar, M. J. (1994). Nitric oxide inhibits 3H-glutamate transport in synaptosomes. Synapse 18, 21–26. doi: 10.1002/syn.890180104

PubMed Abstract | CrossRef Full Text | Google Scholar

Pogun, S., and Kuhar, M. J. (1994). Regulation of neurotransmitter reuptake by nitric oxide. Ann. N.Y. Acad. Sci. 738, 305–315. doi: 10.1111/j.1749-6632.1994.tb21816.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Porton, B., Greenberg, B. D., Askland, K., Serra, L. M., Gesmonde, J., Rudnick, G., et al. (2013). Isoforms of the neuronal glutamate transporter gene, SLC1A1/EAAC1, negatively modulate glutamate uptake: relevance to obsessive-compulsive disorder. Transl. Psychiatry 3:e259. doi: 10.1038/tp.2013.35

PubMed Abstract | CrossRef Full Text | Google Scholar

Pristupa, Z. B., McConkey, F., Liu, F., Man, H. Y., Lee, F. J., Wang, Y. T., et al. (1998). Protein kinase-mediated bidirectional trafficking and functional regulation of the human dopamine transporter. Synapse 30, 79–87. doi: 10.1002/(sici)1098-2396(199809)30:1<79::aid-syn10>3.0.co;2-k

CrossRef Full Text | Google Scholar

Quick, M. W., Corey, J. L., Davidson, N., and Lester, H. A. (1997). Second messengers, trafficking-related proteins, and amino acid residues that contribute to the functional regulation of the rat brain GABA transporter GAT1. J. Neurosci. 17, 2967–2979. doi: 10.1523/jneurosci.17-09-02967.1997

PubMed Abstract | CrossRef Full Text | Google Scholar

Radian, R., and Kanner, B. I. (1983). Stoichiometry of sodium- and chloride-coupled gamma-aminobutyric acid transport by synaptic plasma membrane vesicles isolated from rat brain. Biochemistry 22, 1236–1241. doi: 10.1021/bi00274a038

PubMed Abstract | CrossRef Full Text | Google Scholar

Raiteri, L., Stigliani, S., Zedda, L., Raiteri, M., and Bonanno, G. (2002). Multiple mechanisms of transmitter release evoked by “pathologically” elevated extracellular [K+]: involvement of transporter reversal and mitochondrial calcium. J. Neurochem. 80, 706–714. doi: 10.1046/j.0022-3042.2001.00750.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Raiteri, M., Cerrito, F., Cervoni, A. M., and Levi, G. (1979). Dopamine can be released by two mechanisms differentially affected by the dopamine transport inhibitor nomifensine. J. Pharmacol. Exp. Ther. 208, 195–202.

Google Scholar

Rauen, T., and Kanner, B. I. (1994). Localization of the glutamate transporter GLT-1 in rat and macaque monkey retinae. Neurosci. Lett. 169, 137–140. doi: 10.1016/0304-3940(94)90375-1

CrossRef Full Text | Google Scholar

Rauen, T., Wiessner, M., Sullivan, R., Lee, A., and Pow, D. V. (2004). A new GLT1 splice variant: cloning and immunolocalization of GLT1c in the mammalian retina and brain. Neurochem. Int. 45, 1095–1106. doi: 10.1016/j.neuint.2004.04.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Reyes, N., Ginter, C., and Boudker, O. (2009). Transport mechanism of a bacterial homologue of glutamate transporters. Nature 462, 880–885. doi: 10.1038/nature08616

PubMed Abstract | CrossRef Full Text | Google Scholar

Richerson, G. B., and Wu, Y. (2004). Role of the GABA transporter in epilepsy. Adv. Exp. Med. Biol. 548, 76–91. doi: 10.1007/978-1-4757-6376-8_6

CrossRef Full Text | Google Scholar

Rimmele, T. S., and Rosenberg, P. A. (2016). GLT-1: the elusive presynaptic glutamate transporter. Neurochem. Int. 98, 19–28. doi: 10.1016/j.neuint.2016.04.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Rivera, C., Voipio, J., Payne, J. A., Ruusuvuori, E., Lahtinen, H., Lamsa, K., et al. (1999). The K+/Cl- co-transporter KCC2 renders GABA hyperpolarizing during neuronal maturation. Nature 397, 251–255. doi: 10.1038/16697

PubMed Abstract | CrossRef Full Text | Google Scholar

Roberts, B. M., Doig, N. M., Brimblecombe, K. R., Lopes, E. F., Siddorn, R. E., Threlfell, S., et al. (2020). GABA uptake transporters support dopamine release in dorsal striatum with maladaptive downregulation in a parkinsonism model. Nat. Commun. 11:4958.

Google Scholar

Roberts, J. G., Lugo-Morales, L. Z., Loziuk, P. L., and Sombers, L. A. (2013). Real-time chemical measurements of dopamine release in the brain. Methods Mol. Biol. 964, 275–294. doi: 10.1007/978-1-62703-251-3_16

CrossRef Full Text | Google Scholar

Robinson, M. B. (2002). Regulated trafficking of neurotransmitter transporters: common notes but different melodies. J. Neurochem. 80, 1–11. doi: 10.1046/j.0022-3042.2001.00698.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Rodriguez-Traver, E., Solis, O., Diaz-Guerra, E., Ortiz, O., Vergano-Vera, E., Mendez-Gomez, H. R., et al. (2016). Role of Nurr1 in the generation and differentiation of dopaminergic neurons from stem cells. Neurotox Res. 30, 14–31. doi: 10.1007/s12640-015-9586-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Roginski, R., Choudhury, K., Marone, M., and Geller, H. (1993). Molecular characterization and expression of a rat-brain glutamate transporter CDNA. Anesthesiology 79:A787.

Google Scholar

Romero, J., de Miguel, R., Ramos, J. A., and Fernandez-Ruiz, J. J. (1998). The activation of cannabinoid receptors in striatonigral GABAergic neurons inhibited GABA uptake. Life Sci. 62, 351–363. doi: 10.1016/s0024-3205(97)01117-x

CrossRef Full Text | Google Scholar

Rosenblum, L. T., and Trotti, D. (2017). EAAT2 and the molecular signature of amyotrophic lateral sclerosis. Adv. Neurobiol. 16, 117–136. doi: 10.1007/978-3-319-55769-4_6

CrossRef Full Text | Google Scholar

Rothstein, J. D., Dykes-Hoberg, M., Pardo, C. A., Bristol, L. A., Jin, L., Kuncl, R. W., et al. (1996). Knockout of glutamate transporters reveals a major role for astroglial transport in excitotoxicity and clearance of glutamate. Neuron 16, 675–686. doi: 10.1016/s0896-6273(00)80086-0

CrossRef Full Text | Google Scholar

Rothstein, J. D., Martin, L., Levey, A. I., Dykes-Hoberg, M., Jin, L., Wu, D., et al. (1994). Localization of neuronal and glial glutamate transporters. Neuron 13, 713–725. doi: 10.1016/0896-6273(94)90038-8

CrossRef Full Text | Google Scholar

Rothstein, J. D., Van Kammen, M., Levey, A. I., Martin, L. J., and Kuncl, R. W. (1995). Selective loss of glial glutamate transporter GLT-1 in amyotrophic lateral sclerosis. Ann. Neurol. 38, 73–84. doi: 10.1002/ana.410380114

PubMed Abstract | CrossRef Full Text | Google Scholar

Ryan, R. M., and Mindell, J. A. (2007). The uncoupled chloride conductance of a bacterial glutamate transporter homolog. Nat. Struct. Mol. Biol. 14, 365–371. doi: 10.1038/nsmb1230

PubMed Abstract | CrossRef Full Text | Google Scholar

Ryan, R. M., Mitrovic, A. D., and Vandenberg, R. J. (2004). The chloride permeation pathway of a glutamate transporter and its proximity to the glutamate translocation pathway. J. Biol. Chem. 279, 20742–20751. doi: 10.1074/jbc.m304433200

PubMed Abstract | CrossRef Full Text | Google Scholar

Ryan, R. M., and Vandenberg, R. J. (2016). Elevating the alternating-access model. Nat. Struct. Mol. Biol. 23, 187–189. doi: 10.1038/nsmb.3179

PubMed Abstract | CrossRef Full Text | Google Scholar

Sacchetti, P., Brownschidle, L. A., Granneman, J. G., and Bannon, M. J. (1999). Characterization of the 5′-flanking region of the human dopamine transporter gene. Brain Res. Mol. Brain Res. 74, 167–174. doi: 10.1016/s0169-328x(99)00275-2

CrossRef Full Text | Google Scholar

Sakrikar, D., Mazei-Robison, M. S., Mergy, M. A., Richtand, N. W., Han, Q., Hamilton, P. J., et al. (2012). Attention deficit/hyperactivity disorder-derived coding variation in the dopamine transporter disrupts microdomain targeting and trafficking regulation. J. Neurosci. 32, 5385–5397. doi: 10.1523/jneurosci.6033-11.2012

PubMed Abstract | CrossRef Full Text | Google Scholar

Saksela, K., and Permi, P. (2012). SH3 domain ligand binding: what’s the consensus and where’s the specificity? FEBS Lett. 586, 2609–2614. doi: 10.1016/j.febslet.2012.04.042

PubMed Abstract | CrossRef Full Text | Google Scholar

Salat, K., and Kulig, K. (2011). GABA transporters as targets for new drugs. Future Med. Chem. 3, 211–222. doi: 10.4155/fmc.10.298

PubMed Abstract | CrossRef Full Text | Google Scholar

Salat, K., Podkowa, A., Malikowska, N., Kern, F., Pabel, J., Wojcieszak, E., et al. (2017). Novel, highly potent and in vivo active inhibitor of GABA transporter subtype 1 with anticonvulsant, anxiolytic, antidepressant and antinociceptive properties. Neuropharmacology 113, 331–342. doi: 10.1016/j.neuropharm.2016.10.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Salemi, S., Baktash, P., Rajaei, B., Noori, M., Amini, H., Shamsara, M., et al. (2016). Efficient generation of dopaminergic-like neurons by overexpression of Nurr1 and Pitx3 in mouse induced pluripotent stem cells. Neurosci. Lett. 626, 126–134. doi: 10.1016/j.neulet.2016.05.032

PubMed Abstract | CrossRef Full Text | Google Scholar

Sandoval, V., Riddle, E. L., Ugarte, Y. V., Hanson, G. R., and Fleckenstein, A. E. (2001). Methamphetamine-induced rapid and reversible changes in dopamine transporter function: an in vitro model. J. Neurosci. 21, 1413–1419. doi: 10.1523/jneurosci.21-04-01413.2001

PubMed Abstract | CrossRef Full Text | Google Scholar

Saucedo-Cardenas, O., Quintana-Hau, J. D., Le, W. D., Smidt, M. P., Cox, J. J., De Mayo, F., et al. (1998). Nurr1 is essential for the induction of the dopaminergic phenotype and the survival of ventral mesencephalic late dopaminergic precursor neurons. Proc. Natl. Acad. Sci. U.S.A. 95, 4013–4018. doi: 10.1073/pnas.95.7.4013

PubMed Abstract | CrossRef Full Text | Google Scholar

Saunders, C., Ferrer, J. V., Shi, L., Chen, J., Merrill, G., Lamb, M. E., et al. (2000). Amphetamine-induced loss of human dopamine transporter activity: an internalization-dependent and cocaine-sensitive mechanism. Proc. Natl. Acad. Sci. U.S.A. 97, 6850–6855. doi: 10.1073/pnas.110035297

PubMed Abstract | CrossRef Full Text | Google Scholar

Schallier, A., Smolders, I., Van Dam, D., Loyens, E., De Deyn, P. P., Michotte, A., et al. (2011). Region- and age-specific changes in glutamate transport in the AbetaPP23 mouse model for Alzheimer’s disease. J. Alzheimers Dis. 24, 287–300. doi: 10.3233/jad-2011-101005

PubMed Abstract | CrossRef Full Text | Google Scholar

Schikorski, T., and Stevens, C. F. (1997). Quantitative ultrastructural analysis of hippocampal excitatory synapses. J. Neurosci. 17, 5858–5867. doi: 10.1523/jneurosci.17-15-05858.1997

PubMed Abstract | CrossRef Full Text | Google Scholar

Schitine, C. S., Mendez-Flores, O. G., Santos, L. E., Ornelas, I., Calaza, K. C., Perez-Toledo, K., et al. (2015). Functional plasticity of GAT-3 in avian Muller cells is regulated by neurons via a glutamatergic input. Neurochem. Int. 82, 42–51. doi: 10.1016/j.neuint.2015.02.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Schmitt, A., Asan, E., Lesch, K. P., and Kugler, P. (2002). A splice variant of glutamate transporter GLT1/EAAT2 expressed in neurons: cloning and localization in rat nervous system. Neuroscience 109, 45–61. doi: 10.1016/s0306-4522(01)00451-1

CrossRef Full Text | Google Scholar

Schmitt, A., Asan, E., Puschel, B., Jons, T., and Kugler, P. (1996). Expression of the glutamate transporter GLT1 in neural cells of the rat central nervous system: non-radioactive in situ hybridization and comparative immunocytochemistry. Neuroscience 71, 989–1004. doi: 10.1016/0306-4522(95)00477-7

CrossRef Full Text | Google Scholar

Schmitt, K. C., Rothman, R. B., and Reith, M. E. (2013). Nonclassical pharmacology of the dopamine transporter: atypical inhibitors, allosteric modulators, and partial substrates. J. Pharmacol. Exp. Ther. 346, 2–10. doi: 10.1124/jpet.111.191056

PubMed Abstract | CrossRef Full Text | Google Scholar

Schoffelmeer, A. N., Drukarch, B., De Vries, T. J., Hogenboom, F., Schetters, D., and Pattij, T. (2011). Insulin modulates cocaine-sensitive monoamine transporter function and impulsive behavior. J. Neurosci. 31, 1284–1291. doi: 10.1523/jneurosci.3779-10.2011

PubMed Abstract | CrossRef Full Text | Google Scholar

Schrödinger, L. L. C. (2010). The PyMOL molecular graphics system. Version 1:5.

Google Scholar

Schwartz, M. D., Canales, J. J., Zucchi, R., Espinoza, S., Sukhanov, I., and Gainetdinov, R. R. (2018). Trace amine-associated receptor 1: a multimodal therapeutic target for neuropsychiatric diseases. Expert Opin. Ther. Targets 22, 513–526. doi: 10.1080/14728222.2018.1480723

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwerdt, H. N., Zhang, E., Kim, M. J., Yoshida, T., Stanwicks, L., Amemori, S., et al. (2018). Cellular-scale probes enable stable chronic subsecond monitoring of dopamine neurochemicals in a rodent model. Commun. Biol. 1:144.

Google Scholar

Scimemi, A. (2014a). Plasticity of GABA transporters: an unconventional route to shape inhibitory synaptic transmission. Front. Cell Neurosci. 8:128. doi: 10.3389/fncel.2014.00128

PubMed Abstract | CrossRef Full Text | Google Scholar

Scimemi, A. (2014b). Structure, function, and plasticity of GABA transporters. Front. Cell. Neurosci. 8:161. doi: 10.3389/fncel.2014.00161

PubMed Abstract | CrossRef Full Text | Google Scholar

Scimemi, A., Fine, A., Kullmann, D. M., and Rusakov, D. A. (2004). NR2B-containing receptors mediate cross talk among hippocampal synapses. J. Neurosci. 24, 4767–4777. doi: 10.1523/jneurosci.0364-04.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Scimemi, A., Meabon, J. S., Woltjer, R. L., Sullivan, J. M., Diamond, J. S., and Cook, D. G. (2013). Amyloid-beta1-42 slows clearance of synaptically released glutamate by mislocalizing astrocytic GLT-1. J. Neurosci. 33, 5312–5318. doi: 10.1523/jneurosci.5274-12.2013

PubMed Abstract | CrossRef Full Text | Google Scholar

Scimemi, A., Semyanov, A., Sperk, G., Kullmann, D. M., and Walker, M. C. (2005). Multiple and plastic receptors mediate tonic GABA(A) receptor currents in the hippocampus. J. Neurosci. 25, 10016–10024. doi: 10.1523/jneurosci.2520-05.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Scimemi, A., Tian, H., and Diamond, J. S. (2009). Neuronal transporters regulate glutamate clearance, NMDA receptor activation, and synaptic plasticity in the hippocampus. J. Neurosci. 29, 14581–14595. doi: 10.1523/jneurosci.4845-09.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Scopelliti, A. J., Font, J., Vandenberg, R. J., Boudker, O., and Ryan, R. M. (2018). Structural characterisation reveals insights into substrate recognition by the glutamine transporter ASCT2/SLC1A5. Nat. Commun. 9:38.

Google Scholar

Scott, H. A., Gebhardt, F. M., Mitrovic, A. D., Vandenberg, R. J., and Dodd, P. R. (2011). Glutamate transporter variants reduce glutamate uptake in Alzheimer’s disease. Neurobiol. Aging 32, 553.e551–e511.

Google Scholar

Seal, R. P., and Amara, S. G. (1999). Excitatory amino acid transporters: a family in flux. Annu. Rev. Pharmacol. Toxicol. 39, 431–456. doi: 10.1146/annurev.pharmtox.39.1.431

PubMed Abstract | CrossRef Full Text | Google Scholar

Seal, R. P., Shigeri, Y., Eliasof, S., Leighton, B. H., and Amara, S. G. (2001). Sulfhydryl modification of V449C in the glutamate transporter EAAT1 abolishes substrate transport but not the substrate-gated anion conductance. Proc. Natl. Acad. Sci. U.S.A. 98, 15324–15329. doi: 10.1073/pnas.011400198

PubMed Abstract | CrossRef Full Text | Google Scholar

Serretti, A., and Mandelli, L. (2008). The genetics of bipolar disorder: genome ‘hot regions,’ genes, new potential candidates and future directions. Mol. Psychiatry 13, 742–771. doi: 10.1038/mp.2008.29

PubMed Abstract | CrossRef Full Text | Google Scholar

Sharma, A., Kazim, S. F., Larson, C. S., Ramakrishnan, A., Gray, J. D., McEwen, B. S., et al. (2019). Divergent roles of astrocytic versus neuronal EAAT2 deficiency on cognition and overlap with aging and Alzheimer’s molecular signatures. Proc. Natl. Acad. Sci. U.S.A. 116, 21800–21811. doi: 10.1073/pnas.1903566116

PubMed Abstract | CrossRef Full Text | Google Scholar

Sheldon, A. L., Gonzalez, M. I., Krizman-Genda, E. N., Susarla, B. T., and Robinson, M. B. (2008). Ubiquitination-mediated internalization and degradation of the astroglial glutamate transporter, GLT-1. Neurochem. Int. 53, 296–308. doi: 10.1016/j.neuint.2008.07.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Shifman, M. (1991). The effect of gangliosides upon recovery of aspartate/glutamatergic synapses in striatum after lesions of the rat sensorimotor cortex. Brain Res. 568, 323–324. doi: 10.1016/0006-8993(91)91419-2

CrossRef Full Text | Google Scholar

Shin, J. W., Nguyen, K. T., Pow, D. V., Knight, T., Buljan, V., Bennett, M. R., et al. (2009). Distribution of glutamate transporter GLAST in membranes of cultured astrocytes in the presence of glutamate transport substrates and ATP. Neurochem. Res. 34, 1758–1766. doi: 10.1007/s11064-009-9982-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Sieber, S. A., and Marahiel, M. A. (2005). Molecular mechanisms underlying nonribosomal peptide synthesis: approaches to new antibiotics. Chem. Rev. 105, 715–738. doi: 10.1021/cr0301191

PubMed Abstract | CrossRef Full Text | Google Scholar

Sims, K. D., Straff, D. J., and Robinson, M. B. (2000). Platelet-derived growth factor rapidly increases activity and cell surface expression of the EAAC1 subtype of glutamate transporter through activation of phosphatidylinositol 3-kinase. J. Biol. Chem. 275, 5228–5237. doi: 10.1074/jbc.275.7.5228

PubMed Abstract | CrossRef Full Text | Google Scholar

Sitcheran, R., Gupta, P., Fisher, P. B., and Baldwin, A. S. (2005). Positive and negative regulation of EAAT2 by NF-kappaB: a role for N-myc in TNFalpha-controlled repression. EMBO J. 24, 510–520. doi: 10.1038/sj.emboj.7600555

PubMed Abstract | CrossRef Full Text | Google Scholar

Sitte, H. H., Huck, S., Reither, H., Boehm, S., Singer, E. A., and Pifl, C. (1998). Carrier-mediated release, transport rates, and charge transfer induced by amphetamine, tyramine, and dopamine in mammalian cells transfected with the human dopamine transporter. J. Neurochem. 71, 1289–1297. doi: 10.1046/j.1471-4159.1998.71031289.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Sogaard, R., Borre, L., Braunstein, T. H., Madsen, K. L., and MacAulay, N. (2013). Functional modulation of the glutamate transporter variant GLT1b by the PDZ domain protein PICK1. J. Biol. Chem. 288, 20195–20207. doi: 10.1074/jbc.m113.471128

PubMed Abstract | CrossRef Full Text | Google Scholar

Sonders, M. S., and Amara, S. G. (1996). Channels in transporters. Curr. Opin. Neurobiol. 6, 294–302.

Google Scholar

Sonders, M. S., Zhu, S. J., Zahniser, N. R., Kavanaugh, M. P., and Amara, S. G. (1997). Multiple ionic conductances of the human dopamine transporter: the actions of dopamine and psychostimulants. J. Neurosci. 17, 960–974. doi: 10.1523/jneurosci.17-03-00960.1997

PubMed Abstract | CrossRef Full Text | Google Scholar

Sorkina, T., Doolen, S., Galperin, E., Zahniser, N. R., and Sorkin, A. (2003). Oligomerization of dopamine transporters visualized in living cells by fluorescence resonance energy transfer microscopy. J. Biol. Chem. 278, 28274–28283. doi: 10.1074/jbc.m210652200

PubMed Abstract | CrossRef Full Text | Google Scholar

Sorkina, T., Hoover, B. R., Zahniser, N. R., and Sorkin, A. (2005). Constitutive and protein kinase C-induced internalization of the dopamine transporter is mediated by a clathrin-dependent mechanism. Traffic 6, 157–170. doi: 10.1111/j.1600-0854.2005.00259.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Stenovec, M., Kreft, M., Grilc, S., Pangrsic, T., and Zorec, R. (2008). EAAT2 density at the astrocyte plasma membrane and Ca(2 +)-regulated exocytosis. Mol. Membr. Biol. 25, 203–215. doi: 10.1080/09687680701790925

PubMed Abstract | CrossRef Full Text | Google Scholar

Stoffel, W., Sasse, J., Duker, M., Muller, R., Hofmann, K., Fink, T., et al. (1996). Human high affinity, Na(+)-dependent L-glutamate/L-aspartate transporter GLAST-1 (EAAT-1): gene structure and localization to chromosome 5p11-p12. FEBS Lett. 386, 189–193. doi: 10.1016/0014-5793(96)00424-3

CrossRef Full Text | Google Scholar

Su, Z. Z., Leszczyniecka, M., Kang, D. C., Sarkar, D., Chao, W., Volsky, D. J., et al. (2003). Insights into glutamate transport regulation in human astrocytes: cloning of the promoter for excitatory amino acid transporter 2 (EAAT2). Proc. Natl. Acad. Sci. U.S.A. 100, 1955–1960. doi: 10.1073/pnas.0136555100

PubMed Abstract | CrossRef Full Text | Google Scholar

Sullivan, R., Rauen, T., Fischer, F., Wiessner, M., Grewer, C., Bicho, A., et al. (2004). Cloning, transport properties, and differential localization of two splice variants of GLT-1 in the rat CNS: implications for CNS glutamate homeostasis. Glia 45, 155–169. doi: 10.1002/glia.10317

PubMed Abstract | CrossRef Full Text | Google Scholar

Sulzer, D., Chen, T. K., Lau, Y. Y., Kristensen, H., Rayport, S., and Ewing, A. (1995). Amphetamine redistributes dopamine from synaptic vesicles to the cytosol and promotes reverse transport. J. Neurosci. 15, 4102–4108. doi: 10.1523/jneurosci.15-05-04102.1995

PubMed Abstract | CrossRef Full Text | Google Scholar

Sulzer, D., Maidment, N. T., and Rayport, S. (1993). Amphetamine and other weak bases act to promote reverse transport of dopamine in ventral midbrain neurons. J. Neurochem. 60, 527–535. doi: 10.1111/j.1471-4159.1993.tb03181.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Susarla, B. T., and Robinson, M. B. (2003). Rottlerin, an inhibitor of protein kinase Cdelta (PKCdelta), inhibits astrocytic glutamate transport activity and reduces GLAST immunoreactivity by a mechanism that appears to be PKCdelta-independent. J. Neurochem. 86, 635–645. doi: 10.1046/j.1471-4159.2003.01886.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Susarla, B. T., and Robinson, M. B. (2008). Internalization and degradation of the glutamate transporter GLT-1 in response to phorbol ester. Neurochem. Int. 52, 709–722. doi: 10.1016/j.neuint.2007.08.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Susarla, B. T., Seal, R. P., Zelenaia, O., Watson, D. J., Wolfe, J. H., Amara, S. G., et al. (2004). Differential regulation of GLAST immunoreactivity and activity by protein kinase C: evidence for modification of amino and carboxyl termini. J. Neurochem. 91, 1151–1163. doi: 10.1111/j.1471-4159.2004.02791.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Swanson, R. A., Liu, J., Miller, J. W., Rothstein, J. D., Farrell, K., Stein, B. A., et al. (1997). Neuronal regulation of glutamate transporter subtype expression in astrocytes. J. Neurosci. 17, 932–940. doi: 10.1523/jneurosci.17-03-00932.1997

PubMed Abstract | CrossRef Full Text | Google Scholar

Sylantyev, S., Savtchenko, L. P., O’Neill, N., and Rusakov, D. A. (2020). Extracellular GABA waves regulate coincidence detection in excitatory circuits. J. Physiol. 598, 4047–4062. doi: 10.1113/jp279744

PubMed Abstract | CrossRef Full Text | Google Scholar

Tai, Y. H., Wang, Y. H., Tsai, R. Y., Wang, J. J., Tao, P. L., Liu, T. M., et al. (2007). Amitriptyline preserves morphine’s antinociceptive effect by regulating the glutamate transporter GLAST and GLT-1 trafficking and excitatory amino acids concentration in morphine-tolerant rats. Pain 129, 343–354. doi: 10.1016/j.pain.2007.01.031

PubMed Abstract | CrossRef Full Text | Google Scholar

Tan, J., Zelenaia, O., Correale, D., Rothstein, J. D., and Robinson, M. B. (1999). Expression of the GLT-1 subtype of Na+-dependent glutamate transporter: pharmacological characterization and lack of regulation by protein kinase C. J. Pharmacol. Exp. Ther. 289, 1600–1610.

Google Scholar

Tanaka, K. (1993). Cloning and expression of a glutamate transporter from mouse brain. Neurosci. Lett. 159, 183–186. doi: 10.1016/0304-3940(93)90829-a

CrossRef Full Text | Google Scholar

Tebano, M. T., Martire, A., Potenza, R. L., Gro, C., Pepponi, R., Armida, M., et al. (2008). Adenosine A(2A) receptors are required for normal BDNF levels and BDNF-induced potentiation of synaptic transmission in the mouse hippocampus. J. Neurochem. 104, 279–286.

Google Scholar

Tian, Y., Kapatos, G., Granneman, J. G., and Bannon, M. J. (1994). Dopamine and gamma-aminobutyric acid transporters: differential regulation by agents that promote phosphorylation. Neurosci. Lett. 173, 143–146. doi: 10.1016/0304-3940(94)90169-4

CrossRef Full Text | Google Scholar

Todorov, A. A., Kolchev, C. B., and Todorov, A. B. (2005). Tiagabine and gabapentin for the management of chronic pain. Clin. J. Pain 21, 358–361. doi: 10.1097/01.ajp.0000110637.14355.77

CrossRef Full Text | Google Scholar

Tonsfeldt, K. J., Suchland, K. L., Beeson, K. A., Lowe, J. D., Li, M. H., and Ingram, S. L. (2016). Sex Differences in GABAA signaling in the periaqueductal gray induced by persistent inflammation. J. Neurosci. 36, 1669–1681. doi: 10.1523/jneurosci.1928-15.2016

PubMed Abstract | CrossRef Full Text | Google Scholar

Torp, R., Danbolt, N. C., Babaie, E., Bjoras, M., Seeberg, E., Storm-Mathisen, J., et al. (1994). Differential expression of two glial glutamate transporters in the rat brain: an in situ hybridization study. Eur. J. Neurosci. 6, 936–942. doi: 10.1111/j.1460-9568.1994.tb00587.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Torp, R., Hoover, F., Danbolt, N. C., Storm-Mathisen, J., and Ottersen, O. P. (1997). Differential distribution of the glutamate transporters GLT1 and rEAAC1 in rat cerebral cortex and thalamus: an in situ hybridization analysis. Anat. Embryol. 195, 317–326. doi: 10.1007/s004290050051

PubMed Abstract | CrossRef Full Text | Google Scholar

Torres, G. E., Gainetdinov, R. R., and Caron, M. G. (2003). Plasma membrane monoamine transporters: structure, regulation and function. Nat. Rev. Neurosci. 4, 13–25. doi: 10.1038/nrn1008

PubMed Abstract | CrossRef Full Text | Google Scholar

Torres, G. E., Yao, W. D., Mohn, A. R., Quan, H., Kim, K. M., Levey, A. I., et al. (2001). Functional interaction between monoamine plasma membrane transporters and the synaptic PDZ domain-containing protein PICK1. Neuron 30, 121–134. doi: 10.1016/s0896-6273(01)00267-7

CrossRef Full Text | Google Scholar

Torres-Salazar, D., and Fahlke, C. (2007). Neuronal glutamate transporters vary in substrate transport rate but not in unitary anion channel conductance. J. Biol. Chem. 282, 34719–34726. doi: 10.1074/jbc.m704118200

PubMed Abstract | CrossRef Full Text | Google Scholar

Trotti, D., Peng, J. B., Dunlop, J., and Hediger, M. A. (2001). Inhibition of the glutamate transporter EAAC1 expressed in Xenopus oocytes by phorbol esters. Brain Res. 914, 196–203. doi: 10.1016/s0006-8993(01)02802-5

CrossRef Full Text | Google Scholar

Trotti, D., Volterra, A., Lehre, K. P., Rossi, D., Gjesdal, O., Racagni, G., et al. (1995). Arachidonic acid inhibits a purified and reconstituted glutamate transporter directly from the water phase and not via the phospholipid membrane. J. Biol. Chem. 270, 9890–9895. doi: 10.1074/jbc.270.17.9890

PubMed Abstract | CrossRef Full Text | Google Scholar

Tzingounis, A. V., and Wadiche, J. I. (2007). Glutamate transporters: confining runaway excitation by shaping synaptic transmission. Nat. Rev. Neurosci. 8, 935–947. doi: 10.1038/nrn2274

PubMed Abstract | CrossRef Full Text | Google Scholar

Underhill, S. M., Ingram, S. L., Ahmari, S. E., Veenstra-VanderWeele, J., and Amara, S. G. (2019). Neuronal excitatory amino acid transporter EAAT3: emerging functions in health and disease. Neurochem. Int. 123, 69–76. doi: 10.1016/j.neuint.2018.05.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Underhill, S. M., Wheeler, D. S., and Amara, S. G. (2015). Differential regulation of two isoforms of the glial glutamate transporter EAAT2 by DLG1 and CaMKII. J. Neurosci. 35, 5260–5270. doi: 10.1523/jneurosci.4365-14.2015

PubMed Abstract | CrossRef Full Text | Google Scholar

Underhill, S. M., Wheeler, D. S., Li, M., Watts, S. D., Ingram, S. L., and Amara, S. G. (2014). Amphetamine modulates excitatory neurotransmission through endocytosis of the glutamate transporter EAAT3 in dopamine neurons. Neuron 83, 404–416. doi: 10.1016/j.neuron.2014.05.043

PubMed Abstract | CrossRef Full Text | Google Scholar

Untiet, V., Kovermann, P., Gerkau, N. J., Gensch, T., Rose, C. R., and Fahlke, C. (2017). Glutamate transporter-associated anion channels adjust intracellular chloride concentrations during glial maturation. Glia 65, 388–400. doi: 10.1002/glia.23098

PubMed Abstract | CrossRef Full Text | Google Scholar

Vandenbergh, D. J., Persico, A. M., Hawkins, A. L., Griffin, C. A., Li, X., Jabs, E. W., et al. (1992a). Human dopamine transporter gene (DAT1) maps to chromosome 5p15.3 and displays a VNTR. Genomics 14, 1104–1106. doi: 10.1016/s0888-7543(05)80138-7

CrossRef Full Text | Google Scholar

Vandenbergh, D. J., Persico, A. M., and Uhl, G. R. (1992b). A human dopamine transporter cDNA predicts reduced glycosylation, displays a novel repetitive element and provides racially-dimorphic TaqI RFLPs. Brain Res. Mol. Brain Res. 15, 161–166. doi: 10.1016/0169-328x(92)90165-8

CrossRef Full Text | Google Scholar

Vaughan, R. A., Huff, R. A., Uhl, G. R., and Kuhar, M. J. (1997). Protein kinase C-mediated phosphorylation and functional regulation of dopamine transporters in striatal synaptosomes. J. Biol. Chem. 272, 15541–15546. doi: 10.1074/jbc.272.24.15541

PubMed Abstract | CrossRef Full Text | Google Scholar

Vaz, S. H., Cristovao-Ferreira, S., Ribeiro, J. A., and Sebastiao, A. M. (2008). Brain-derived neurotrophic factor inhibits GABA uptake by the rat hippocampal nerve terminals. Brain Res. 1219, 19–25. doi: 10.1016/j.brainres.2008.04.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Vaz, S. H., Jorgensen, T. N., Cristovao-Ferreira, S., Duflot, S., Ribeiro, J. A., Gether, U., et al. (2011). Brain-derived neurotrophic factor (BDNF) enhances GABA transport by modulating the trafficking of GABA transporter-1 (GAT-1) from the plasma membrane of rat cortical astrocytes. J. Biol. Chem. 286, 40464–40476. doi: 10.1074/jbc.m111.232009

PubMed Abstract | CrossRef Full Text | Google Scholar

Venderova, K., Brown, T. M., and Brotchie, J. M. (2005). Differential effects of endocannabinoids on [(3)H]-GABA uptake in the rat Globus pallidus. Exp. Neurol. 194, 284–287. doi: 10.1016/j.expneurol.2005.02.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Ventura, R., and Harris, K. M. (1999). Three-dimensional relationships between hippocampal synapses and astrocytes. J. Neurosci. 19, 6897–6906. doi: 10.1523/jneurosci.19-16-06897.1999

PubMed Abstract | CrossRef Full Text | Google Scholar

Verdon, G., and Boudker, O. (2012). Crystal structure of an asymmetric trimer of a bacterial glutamate transporter homolog. Nat. Struct. Mol. Biol. 19, 355–357. doi: 10.1038/nsmb.2233

PubMed Abstract | CrossRef Full Text | Google Scholar

Veruki, M. L., Morkve, S. H., and Hartveit, E. (2006). Activation of a presynaptic glutamate transporter regulates synaptic transmission through electrical signaling. Nat. Neurosci. 9, 1388–1396. doi: 10.1038/nn1793

PubMed Abstract | CrossRef Full Text | Google Scholar

Vina-Vilaseca, A., and Sorkin, A. (2010). Lysine 63-linked polyubiquitination of the dopamine transporter requires WW3 and WW4 domains of Nedd4-2 and UBE2D ubiquitin-conjugating enzymes. J. Biol. Chem. 285, 7645–7656. doi: 10.1074/jbc.m109.058990

PubMed Abstract | CrossRef Full Text | Google Scholar

Voisin, P., Viratelle, O., Girault, J. M., Morrison-Bogorad, M., and Labouesse, J. (1993). Plasticity of astroglial glutamate and γ-Aminobutyric acid uptake in cell cultures derived from postnatal mouse cerebellum. J. Neurochem. 60, 114–127. doi: 10.1111/j.1471-4159.1993.tb05829.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Volterra, A., Trotti, D., Cassutti, P., Tromba, C., Galimberti, R., Lecchi, P., et al. (1992). A role for the arachidonic acid cascade in fast synaptic modulation: ion channels and transmitter uptake systems as target proteins. Adv. Exp. Med. Biol. 318, 147–158. doi: 10.1007/978-1-4615-3426-6_13

CrossRef Full Text | Google Scholar

Volterra, A., Trotti, D., and Racagni, G. (1994a). Glutamate uptake is inhibited by arachidonic acid and oxygen radicals via two distinct and additive mechanisms. Mol. Pharmacol. 46, 986–992.

Google Scholar

Volterra, A., Trotti, D., Tromba, C., Floridi, S., and Racagni, G. (1994b). Glutamate uptake inhibition by oxygen free radicals in rat cortical astrocytes. J. Neurosci. 14, 2924–2932. doi: 10.1523/jneurosci.14-05-02924.1994

PubMed Abstract | CrossRef Full Text | Google Scholar

Voutsinos, B., Dutuit, M., Reboul, A., Fevre-Montange, M., Bernard, A., Trouillas, P., et al. (1998). Serotoninergic control of the activity and expression of glial GABA transporters in the rat cerebellum. Glia 23, 45–60. doi: 10.1002/(sici)1098-1136(199805)23:1<45::aid-glia5>3.0.co;2-3

CrossRef Full Text | Google Scholar

Wadiche, J. I., Amara, S. G., and Kavanaugh, M. P. (1995a). Ion fluxes associated with excitatory amino acid transport. Neuron 15, 721–728. doi: 10.1016/0896-6273(95)90159-0

CrossRef Full Text | Google Scholar

Wadiche, J. I., Arriza, J. L., Amara, S. G., and Kavanaugh, M. P. (1995b). Kinetics of a human glutamate transporter. Neuron 14, 1019–1027. doi: 10.1016/0896-6273(95)90340-2

CrossRef Full Text | Google Scholar

Wadiche, J. I., and Jahr, C. E. (2005). Patterned expression of Purkinje cell glutamate transporters controls synaptic plasticity. Nat. Neurosci. 8, 1329–1334. doi: 10.1038/nn1539

PubMed Abstract | CrossRef Full Text | Google Scholar

Wadiche, J. I., and Kavanaugh, M. P. (1998). Macroscopic and microscopic properties of a cloned glutamate transporter/chloride channel. J. Neurosci. 18, 7650–7661. doi: 10.1523/jneurosci.18-19-07650.1998

PubMed Abstract | CrossRef Full Text | Google Scholar

Walker, M. C., and van der Donk, W. A. (2016). The many roles of glutamate in metabolism. J. Ind. Microbiol. Biotechnol. 43, 419–430. doi: 10.1007/s10295-015-1665-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, D., Deken, S. L., Whitworth, T. L., and Quick, M. W. (2003). Syntaxin 1A inhibits GABA flux, efflux, and exchange mediated by the rat brain GABA transporter GAT1. Mol. Pharmacol. 64, 905–913. doi: 10.1124/mol.64.4.905

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Z., Li, W., Mitchell, C. K., and Carter-Dawson, L. (2003). Activation of protein kinase C reduces GLAST in the plasma membrane of rat Muller cells in primary culture. Vis. Neurosci. 20, 611–619. doi: 10.1017/s0952523803206039

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, D., and Quick, M. W. (2005). Trafficking of the plasma membrane gamma-aminobutyric acid transporter GAT1. Size and rates of an acutely recycling pool. J. Biol. Chem. 280, 18703–18709. doi: 10.1074/jbc.m500381200

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, J., Michelhaugh, S. K., and Bannon, M. J. (2007). Valproate robustly increases Sp transcription factor-mediated expression of the dopamine transporter gene within dopamine cells. Eur. J. Neurosci. 25, 1982–1986. doi: 10.1111/j.1460-9568.2007.05460.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, K. H., Penmatsa, A., and Gouaux, E. (2015). Neurotransmitter and psychostimulant recognition by the dopamine transporter. Nature 521, 322–327. doi: 10.1038/nature14431

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, X., and Boudker, O. (2020). Large domain movements through the lipid bilayer mediate substrate release and inhibition of glutamate transporters. eLife 9:e58417.

Google Scholar

Wang, Y., Lu, S., Qu, Z., Wu, L., and Wang, Y. (2017). Sonic hedgehog induces GLT-1 degradation via PKC delta to suppress its transporter activities. Neuroscience 365, 217–225. doi: 10.1016/j.neuroscience.2017.09.051

PubMed Abstract | CrossRef Full Text | Google Scholar

Watzke, N., Bamberg, E., and Grewer, C. (2001). Early intermediates in the transport cycle of the neuronal excitatory amino acid carrier EAAC1. J. Gen. Physiol. 117, 547–562. doi: 10.1085/jgp.117.6.547

PubMed Abstract | CrossRef Full Text | Google Scholar

Watzke, N., and Grewer, C. (2001). The anion conductance of the glutamate transporter EAAC1 depends on the direction of glutamate transport. FEBS Lett. 503, 121–125. doi: 10.1016/s0014-5793(01)02715-6

CrossRef Full Text | Google Scholar

Wenzel, J., Lammert, G., Meyer, U., and Krug, M. (1991). The influence of long-term potentiation on the spatial relationship between astrocyte processes and potentiated synapses in the dentate gyrus neuropil of rat brain. Brain Res. 560, 122–131. doi: 10.1016/0006-8993(91)91222-m

CrossRef Full Text | Google Scholar

Wheeler, D. S., Ebben, A. L., Kurtoglu, B., Lovell, M. E., Bohn, A. T., Jasek, I. A., et al. (2017). Corticosterone regulates both naturally occurring and cocaine-induced dopamine signaling by selectively decreasing dopamine uptake. Eur. J. Neurosci. 46, 2638–2646. doi: 10.1111/ejn.13730

PubMed Abstract | CrossRef Full Text | Google Scholar

Wheeler, D. S., Underhill, S. M., Stolz, D. B., Murdoch, G. H., Thiels, E., Romero, G., et al. (2015). Amphetamine activates Rho GTPase signaling to mediate dopamine transporter internalization and acute behavioral effects of amphetamine. Proc. Natl. Acad. Sci. U.S.A. 112, E7138–E7147.

Google Scholar

Whitworth, T. L., and Quick, M. W. (2001). Upregulation of gamma-aminobutyric acid transporter expression: role of alkylated gamma-aminobutyric acid derivatives. Biochem. Soc. Trans. 29, 736–741. doi: 10.1042/bst0290736

CrossRef Full Text | Google Scholar

Wieczorek, W. J., and Kruk, Z. L. (1994). Differential action of (+)-amphetamine on electrically evoked dopamine overflow in rat brain slices containing corpus striatum and nucleus accumbens. Br. J. Pharmacol. 111, 829–836. doi: 10.1111/j.1476-5381.1994.tb14813.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Williams, J. M., Owens, W. A., Turner, G. H., Saunders, C., Dipace, C., Blakely, R. D., et al. (2007). Hypoinsulinemia regulates amphetamine-induced reverse transport of dopamine. PLoS Biol. 5:e274. doi: 10.1371/journal.pbio.0050274

PubMed Abstract | CrossRef Full Text | Google Scholar

Wilson, J. M., Khabazian, I., Pow, D. V., Craig, U. K., and Shaw, C. A. (2003). Decrease in glial glutamate transporter variants and excitatory amino acid receptor down-regulation in a murine model of ALS-PDC. Neuromolecul. Med. 3, 105–118. doi: 10.1385/nmm:3:2:105

CrossRef Full Text | Google Scholar

Winter, N., Kovermann, P., and Fahlke, C. (2012). A point mutation associated with episodic ataxia 6 increases glutamate transporter anion currents. Brain 135, 3416–3425. doi: 10.1093/brain/aws255

PubMed Abstract | CrossRef Full Text | Google Scholar

Witcher, M. R., Kirov, S. A., and Harris, K. M. (2007). Plasticity of perisynaptic astroglia during synaptogenesis in the mature rat hippocampus. Glia 55, 13–23. doi: 10.1002/glia.20415

PubMed Abstract | CrossRef Full Text | Google Scholar

Witcher, M. R., Park, Y. D., Lee, M. R., Sharma, S., Harris, K. M., and Kirov, S. A. (2010). Three-dimensional relationships between perisynaptic astroglia and human hippocampal synapses. Glia 58, 572–587.

Google Scholar

Woo, T. U., Whitehead, R. E., Melchitzky, D. S., and Lewis, D. A. (1998). A subclass of prefrontal gamma-aminobutyric acid axon terminals are selectively altered in schizophrenia. Proc. Natl. Acad. Sci. U.S.A. 95, 5341–5346. doi: 10.1073/pnas.95.9.5341

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, X., Kihara, T., Akaike, A., Niidome, T., and Sugimoto, H. (2010). PI3K/Akt/mTOR signaling regulates glutamate transporter 1 in astrocytes. Biochem. Biophys. Res. Commun. 393, 514–518. doi: 10.1016/j.bbrc.2010.02.038

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, Y., Wang, W., Diez-Sampedro, A., and Richerson, G. B. (2007). Nonvesicular inhibitory neurotransmission via reversal of the GABA transporter GAT-1. Neuron 56, 851–865. doi: 10.1016/j.neuron.2007.10.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, Y., Wang, W., and Richerson, G. B. (2003). Vigabatrin induces tonic inhibition via GABA transporter reversal without increasing vesicular GABA release. J. Neurophysiol. 89, 2021–2034. doi: 10.1152/jn.00856.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, Y. Y., Qu, J., Zhou, L., Lv, N., Gong, J. E., Cao, Y. Z., et al. (2017). Lack of association between SLC6A11 genetic polymorphisms and drug resistant epilepsy in chinese han population. Clin. Lab. 63, 1113–1120.

Google Scholar

Xu, Y. F., Cai, Y. Q., Cai, G. Q., Jiang, J., Sheng, Z. J., Wang, Z. G., et al. (2008). Hypoalgesia in mice lacking GABA transporter subtype 1. J. Neurosci. Res. 86, 465–470. doi: 10.1002/jnr.21499

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamashita, A., Singh, S. K., Kawate, T., Jin, Y., and Gouaux, E. (2005). Crystal structure of a bacterial homologue of Na+/Cl–dependent neurotransmitter transporters. Nature 437, 215–223. doi: 10.1038/nature03978

PubMed Abstract | CrossRef Full Text | Google Scholar

Yan, X., Yadav, R., Gao, M., and Weng, H. R. (2014). Interleukin-1 beta enhances endocytosis of glial glutamate transporters in the spinal dorsal horn through activating protein kinase C. Glia 62, 1093–1109. doi: 10.1002/glia.22665

PubMed Abstract | CrossRef Full Text | Google Scholar

Yan, X. X., Cariaga, W. A., and Ribak, C. E. (1997). Immunoreactivity for GABA plasma membrane transporter, GAT-1, in the developing rat cerebral cortex: transient presence in the somata of neocortical and hippocampal neurons. Brain Res. Dev. Brain Res. 99, 1–19. doi: 10.1016/s0165-3806(96)00192-7

CrossRef Full Text | Google Scholar

Yan, X. X., and Ribak, C. E. (1998). Developmental expression of gamma-aminobutyric acid transporters (GAT-1 and GAT-3) in the rat cerebellum: evidence for a transient presence of GAT-1 in Purkinje cells. Brain Res. Dev. Brain Res. 111, 253–269. doi: 10.1016/s0165-3806(98)00144-8

CrossRef Full Text | Google Scholar

Yang, Y., Kinney, G. A., Spain, W. J., Breitner, J. C., and Cook, D. G. (2004). Presenilin-1 and intracellular calcium stores regulate neuronal glutamate uptake. J. Neurochem. 88, 1361–1372. doi: 10.1046/j.1471-4159.2003.02279.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Yernool, D., Boudker, O., Jin, Y., and Gouaux, E. (2004). Structure of a glutamate transporter homologue from Pyrococcus horikoshii. Nature 431, 811–818. doi: 10.1038/nature03018

PubMed Abstract | CrossRef Full Text | Google Scholar

Yoo, S. Y., Kim, J. H., Do, S. H., and Zuo, Z. (2008). Inhibition of the activity of excitatory amino acid transporter 4 expressed in Xenopus oocytes after chronic exposure to ethanol. Alcohol. Clin. Exp. Res. 32, 1292–1298. doi: 10.1111/j.1530-0277.2008.00697.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, X., Plotnikova, O., Bonin, P. D., Subashi, T. A., McLellan, T. J., Dumlao, D., et al. (2019). Cryo-EM structures of the human glutamine transporter SLC1A5 (ASCT2) in the outward-facing conformation. eLife 8:e48120.

Google Scholar

Zeigerer, A., McBrayer, M. K., and McGraw, T. E. (2004). Insulin stimulation of GLUT4 exocytosis, but not its inhibition of endocytosis, is dependent on RabGAP AS160. Mol. Biol. Cell 15, 4406–4415. doi: 10.1091/mbc.e04-04-0333

PubMed Abstract | CrossRef Full Text | Google Scholar

Zelenaia, O., Schlag, B. D., Gochenauer, G. E., Ganel, R., Song, W., Beesley, J. S., et al. (2000). Epidermal growth factor receptor agonists increase expression of glutamate transporter GLT-1 in astrocytes through pathways dependent on phosphatidylinositol 3-kinase and transcription factor NF-kappaB. Mol. Pharmacol. 57, 667–678. doi: 10.1124/mol.57.4.667

PubMed Abstract | CrossRef Full Text | Google Scholar

Zerangue, N., Arriza, J. L., Amara, S. G., and Kavanaugh, M. P. (1995). Differential modulation of human glutamate transporter subtypes by Arachidonic acid. J. Biol. Chem. 270, 6433–6435. doi: 10.1074/jbc.270.12.6433

PubMed Abstract | CrossRef Full Text | Google Scholar

Zerangue, N., and Kavanaugh, M. P. (1996). Flux coupling in a neuronal glutamate transporter. Nature 383, 634–637. doi: 10.1038/383634a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Zestos, A. G., Mikelman, S. R., Kennedy, R. T., and Gnegy, M. E. (2016). PKCbeta inhibitors attenuate amphetamine-stimulated dopamine efflux. ACS Chem. Neurosci. 7, 757–766. doi: 10.1021/acschemneuro.6b00028

PubMed Abstract | CrossRef Full Text | Google Scholar

Zetterstrom, R. H., Solomin, L., Jansson, L., Hoffer, B. J., Olson, L., and Perlmann, T. (1997). Dopamine neuron agenesis in Nurr1-deficient mice. Science 276, 248–250. doi: 10.1126/science.276.5310.248

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, H., Li, S., Wang, M., Vukusic, B., Pristupa, Z. B., and Liu, F. (2009). Regulation of dopamine transporter activity by carboxypeptidase E. Mol. Brain 2:10. doi: 10.1186/1756-6606-2-10

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, L., Coffey, L. L., and Reith, M. E. (1997). Regulation of the functional activity of the human dopamine transporter by protein kinase C. Biochem. Pharmacol. 53, 677–688. doi: 10.1016/s0006-2952(96)00898-2

CrossRef Full Text | Google Scholar

Zhang, L., Li, L., Wang, B., Qian, D. M., Song, X. X., and Hu, M. (2014). HCMV induces dysregulation of glutamate uptake and transporter expression in human fetal astrocytes. Neurochem. Res. 39, 2407–2418. doi: 10.1007/s11064-014-1445-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Q., Fukuda, M., Van Bockstaele, E., Pascual, O., and Haydon, P. G. (2004). Synaptotagmin IV regulates glial glutamate release. Proc. Natl. Acad. Sci. U.S.A. 101, 9441–9446. doi: 10.1073/pnas.0401960101

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, J., and Sutherland, M. L. (2004). Glutamate transporter cluster formation in astrocytic processes regulates glutamate uptake activity. J. Neurosci. 24, 6301–6306. doi: 10.1523/jneurosci.1404-04.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhu, S. J., Kavanaugh, M. P., Sonders, M. S., Amara, S. G., and Zahniser, N. R. (1997). Activation of protein kinase C inhibits uptake, currents and binding associated with the human dopamine transporter expressed in Xenopus oocytes. J. Pharmacol. Exp. Ther. 282, 1358–1365.

Google Scholar

Zike, I. D., Chohan, M. O., Kopelman, J. M., Krasnow, E. N., Flicker, D., Nautiyal, K. M., et al. (2017). OCD candidate gene SLC1A1/EAAT3 impacts basal ganglia-mediated activity and stereotypic behavior. Proc. Natl. Acad. Sci. U.S.A. 114, 5719–5724. doi: 10.1073/pnas.1701736114

PubMed Abstract | CrossRef Full Text | Google Scholar

Zschocke, J., Bayatti, N., Clement, A. M., Witan, H., Figiel, M., Engele, J., et al. (2005). Differential promotion of glutamate transporter expression and function by glucocorticoids in astrocytes from various brain regions. J. Biol. Chem. 280, 34924–34932. doi: 10.1074/jbc.m502581200

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: glutamate, GABA, dopamine, transporter, uptake, surface mobility

Citation: Ryan RM, Ingram SL and Scimemi A (2021) Regulation of Glutamate, GABA and Dopamine Transporter Uptake, Surface Mobility and Expression. Front. Cell. Neurosci. 15:670346. doi: 10.3389/fncel.2021.670346

Received: 21 February 2021; Accepted: 15 March 2021;
Published: 13 April 2021.

Edited by:

Fiorenzo Conti, Marche Polytechnic University, Italy

Reviewed by:

Janosch P. Heller, Dublin City University, Ireland
Michael B. Robinson, Children’s Hospital of Philadelphia Research Institute, United States

Copyright © 2021 Ryan, Ingram and Scimemi. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Annalisa Scimemi, scimemia@gmail.com; ascimemi@albany.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.