Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2012 Mar 12;590(Pt 10):2317–2331. doi: 10.1113/jphysiol.2011.226605

Activation of glutamate transport evokes rapid glutamine release from perisynaptic astrocytes

Nneka M Uwechue 1, Mari-Carmen Marx 1, Quentin Chevy 2, Brian Billups 1
PMCID: PMC3424755  PMID: 22411007

Abstract

Stimulation of astrocytes by neuronal activity and the subsequent release of neuromodulators is thought to be an important regulator of synaptic communication. In this study we show that astrocytes juxtaposed to the glutamatergic calyx of Held synapse in the rat medial nucleus of the trapezoid body (MNTB) are stimulated by the activation of glutamate transporters and consequently release glutamine on a very rapid timescale. MNTB principal neurones express electrogenic system A glutamine transporters, and were exploited as glutamine sensors in this study. By simultaneous whole-cell voltage clamping astrocytes and neighbouring MNTB neurones in brainstem slices, we show that application of the excitatory amino acid transporter (EAAT) substrate d-aspartate stimulates astrocytes to rapidly release glutamine, which is detected by nearby MNTB neurones. This release is significantly reduced by the toxins l-methionine sulfoximine and fluoroacetate, which reduce glutamine concentrations specifically in glial cells. Similarly, glutamine release was also inhibited by localised inactivation of EAATs in individual astrocytes, using internal dl-threo-β-benzyloxyaspartic acid (TBOA) or dissipating the driving force by modifying the patch-pipette solution. These results demonstrate that astrocytes adjacent to glutamatergic synapses can release glutamine in a temporally precise, controlled manner in response to glial glutamate transporter activation. Since glutamine can be used by neurones as a precursor for glutamate and GABA synthesis, this represents a potential feedback mechanism by which astrocytes can respond to synaptic activation and react in a way that sustains or enhances further communication. This would therefore represent an additional manifestation of the tripartite relationship between synapses and astrocytes.


Key points

  • Following release of glutamate from excitatory synapses, excitatory amino acid transporters (EAATs) sequester this glutamate into neighbouring astrocytes.

  • The signalling effects on the astrocyte and the mechanisms by which this glutamate is recycled back to the synapse are currently unclear.

  • In this study we use electrophysiological recording from neurones and astrocytes to show that a surge of the neurotransmitter glutamate, as usually occurs during neuronal activity, activates astrocytes and causes them to rapidly release the amino acid glutamine.

  • This glutamine mediates a fast signal back to the neurones, where it is sequestered and is available for the biosynthesis of further neurotransmitters.

  • Our data demonstrate a novel feedback mechanism by which astrocytes can potentially modulate neuronal function, and pave the way for development of new therapeutic approaches to treat neurological disorders.

Introduction

Ultrastructual analysis of mammalian synapses has shown that the astrocytic membrane is located immediately adjacent to glutamatergic synapses in the cerebellum, cortex, hippocampus and brainstem (Spacek, 1985; Ventura & Harris, 1999; Satzler et al. 2002). This glial association is highly dynamic, with increases in synaptic activity due to long-term potentiation, kindling, sensory activity, motor skills learning and synaptic maturity causing hypertrophy of astrocytic processes, increased astrocytic volume and enhanced ensheathment of synapses (Wenzel et al. 1991; Hawrylak et al. 1993; Anderson et al. 1994; Jones & Greenough, 1996; Genoud et al. 2006; Witcher et al. 2007). More recently it has been established that astrocytes play a central role in regulating synaptic communication through their ability to sequester and release neurotransmitters and neuromodulators. This tripartite relationship between presynaptic, postsynaptic and astrocytic elements has been clearly demonstrated for neuroactive compounds such as adenosine, ATP, glutamate and d-serine (Perea & Araque, 2007; Bardoni et al. 2010; Henneberger et al. 2010). Accordingly, it is evident that astrocytes respond in a localised manner to specific neurotransmitters, in a rapid and controlled fashion (Perea et al. 2009).

Astrocytes are also proposed to be mediators of the glutamate–glutamine cycle, where synaptically released glutamate is recycled via conversion to glutamine in glial cells (van den Berg & Garfinkel, 1971; Hertz et al. 1999; Bak et al. 2006). Essential components of this cycle include: uptake by high-affinity excitatory amino acid transporters (EAATs; Danbolt, 2001), which are known to be concentrated on astrocytic membranes adjacent to synapses (Chaudhry et al. 1995); conversion of glutamate to glutamine by the astrocytic enzyme glutamine synthetase (Norenberg & Martinez-Hernandez, 1979); release of glutamine from astrocytes; sequestration of this glutamine into neurones (Chaudhry et al. 2002a); and subsequent conversion back to glutamate by enzymes such as phosphate-activated glutaminase (Kvamme et al. 2000). The importance of this cycle in maintaining neurotransmitter supply has been demonstrated in vitro (Pow & Robinson, 1994; Laake et al. 1995) and in vivo (Gibbs et al. 1996; Barnett et al. 2000; Sibson et al. 2001).

Consequently, the release of astrocytic glutamine is a vital step in the maintenance of neurotransmitter levels at glutamatergic synapses. Furthermore, since glutamate is the main precursor of GABA, the continued supply of glutamine has also been shown to be crucial for sustaining inhibitory neurotransmission (Sonnewald et al. 1993; Liang et al. 2006; Fricke et al. 2007). However, despite the physiological importance of this process, rapid release of glutamine from individually identified astrocytes in the vicinity of a synapse has not been directly demonstrated.

In this present study, we investigate astrocytic glutamine release at the calyx of Held synapse in the auditory brainstem. This glutamatergic synapse surrounds the principal cells of the medial nucleus of the trapezoid body (MNTB) and is large enough to be visually identified. In acutely isolated brain slices, the astrocyte immediately adjacent to the synapse can be easily located and the close proximity of the astrocytic cell body to the neuronal synapse is advantageous for patch-clamp recording. We have previously shown that the postsynaptic MNTB cells express system A glutamine transporters, the electrogenic nature of which results in a membrane current upon activation by glutamine. Taking advantage of this, we have used the MNTB principal neurones as glutamine detectors to sense glutamine release from glial cells immediately apposing the synapse. We find that stimulation of EAAT glutamate transporters on astrocytes results in a very rapid, tightly coupled release of glutamine, which would be available to the presynaptic terminal to replenish the glutamate supply for synaptic transmission.

Methods

Ethical approval

All procedures were carried out in accordance with the UK Animals (Scientific Procedures) Act 1986, and approved by the Local Ethical Review Panel at the University of Cambridge, UK.

Slice preparation

Wistar rats 10–15 days old were killed by decapitation and brains quickly removed and immediately transferred to an ice-cold, low-sodium solution, composed of (in mm): 250 sucrose, 2.5 KCl, 10 glucose, 1.25 NaH2PO4, 26 NaHCO3, 4 MgCl2, 0.1 CaCl2, and gassed with 95% O2–5% CO2 to pH 7.4. Transverse brain slices, 110 to 130 μm thick, containing MNTB were obtained, using an Integraslice 7550PSDS (Campden Instruments; Loughborough, UK). Slices were placed in an incubation chamber containing artificial cerebrospinal fluid (aCSF) with composition (in mm): 125 NaCl, 2.5 KCl, 10 glucose, 1.25 NaH2PO4, 26 NaHCO3, 1 MgCl2 and 2 CaCl2 which was gassed with 95% O2–5% CO2 (pH 7.4) and kept at 35°C for 1 h prior to storage at room temperature for up to 8 h.

Electrophysiological recording

For recording, brain slices were continually perfused with aCSF (composition as above; 31–35°C), at a rate of 1 ml min−1 and visualised with infrared differential interference contrast optics. MNTB principal cells were identified by their characteristically large (∼20 μm) spherical soma and astrocytes adjacent to principal cells were identified by their smaller soma (<10 μm). Astrocytes exhibited characteristic electrophysiological properties (membrane resistance <10 MΩ; resting membrane potential < −78 mV; lack of voltage-activated currents) and identification was confirmed in a separate series of experiments by labelling slices with 1 μm sulforhodamine 101 for 20 min at 35°C, immediately following slicing (Nimmerjahn et al. 2004; Kafitz et al. 2008). Furthermore, Lucifer yellow (<0.05%) was included in the patch pipette solution to enable fluorescent identification of cell type at the end of the experiment.

Membrane currents from MNTB principal neurones and astrocytes were obtained by whole-cell voltage clamping at −70 and −80 mV, respectively, using a HEKA EPC-10 double amplifier (HEKA Elektronik Dr Schulze GmbH; Lambrecht/Pfalz, Germany), filtered at 10 and 2.9 kHz, and digitized at 25 kHz with Patchmaster software (HEKA). Recordings were made using thick-walled glass pipettes (GC150F-7.5; Harvard Apparatus; Edenbridge, Kent, UK) with open-tip resistances of 5–8 MΩ for neurones and 6–9 MΩ for astrocytes. Whole-cell access resistances were <30 MΩ and series resistance compensation of 50% (100 μs lag time) was applied to neuronal recordings. For neuronal recordings, pipettes were filled with a solution containing (in mm): 110 caesium methanesulfonate, 40 HEPES, 10 TEA chloride, 5 Na2-phosphocreatine, 20 sucrose, 0.2 EGTA, 2 MgATP, 0.5 Na2GTP and 0.008 CaCl2 (pH 7.2 with CsOH). For astrocytes the patch pipette contained (in mm): 130 KCl, 2 NaCl, 4 glucose, 10 HEPES, 0.1 EGTA, 1 MgATP, 0.5 Na2GTP and 0.025 CaCl2 (pH 7.2 with KOH). In experiments where the effects of internal Cs+ or K+ removal were assessed, Cs+ and K+ were replaced by N-methyl-d-glucamine (NMDG).

All recordings were made in the presence of (in μm): 40 dl-2-amino-5-phospohonopentanoic acid (APV), 10 dizocilpine maleate (MK801), 10 (–)-bicuculline methochloride, 1 strychnine, 1 TTX, 20 2,3,dioxo-6-nitro-1,2,3,4-tetrahydrobenzoquinoxaline-7- sulfonamide (NBQX) and 10 mm TEA chloride. Transporter substrates (d-aspartate and l-glutamine) were dissolved in the external solution and applied by pressurized ejection (2–8 psi; 13.8-68.9 kPa) from a pipette (open tip resistance 4–6 MΩ) using a Picospritzer II (General Valve; Fairtrade, NJ, USA). In neurones, d-aspartate puffs were 2 s in duration and repeated every 30 s. In astrocytes, d-aspartate puffs were of 5 s duration and repeated every minute. Antagonists were applied by addition to the bathing solution and by inclusion in a second puff pipette, along with the agonist. The two puffer pipettes were of equal tip diameter (pulled from the same glass capillary), connected to the same pressure source and placed equidistant from the cells. For experiments investigating the proportional contribution of EAAT1 and EAAT2 to the glia d-aspartate response, 10 μm UCPH-101 and 450 μm dihydrokainate (DHK) were used. The percentage inhibitions were corrected for the fact that these competitive inhibitors do not block 100% of the EAAT1 or EAAT2 currents, respectively. UCPH-101 inhibition of EAAT1 was calculated to be 85%, using Ki = 0.41 μm (calculated from the data of Jensen et al. 2009) and a Km for d-aspartate of 60 μm (Arriza et al. 1994). DHK inhibition of EAAT2 was calculated to be 81%, using Ki = 23 μm and Km for d-aspartate of 54 μm (Arriza et al. 1994). Similarly 200 μm dl-threo-β-benzyloxyaspartic acid (TBOA) would be expected to inhibit EAAT1 by 52%, using Ki = 42 μm and inhibit EAAT2 by 88%, using Ki = 5.7 μm (Shimamoto et al. 1998). Using the relative proportions of EAAT1 and EAAT2 present (see Results), 200 μm TBOA would be expected to inhibit 64% of the total EAAT current in astrocytes. Furthermore, 20 μm of the higher affinity substrate inhibitor (3S)-3-[[3-[[4-(trifluoromethyl)benzoyl]amino]phenyl]methoxy]-l-aspartic acid (TFB-TBOA; Shimamoto et al. 2004) would be expected to inhibit 99.5% of the astrocytic EAAT current.

All chemicals were obtained from Sigma Aldrich (Gillingham, Dorset, UK) except TTX, bicuculline, NBQX, APV, MK801, UCPH-101 and DHK (Ascent Scientific; Bristol, UK) and TBOA and TFB-TBOA (Tocris BioScience; Bristol, UK).

Sodium imaging

For measuring sodium concentration changes in astrocytes, 0.5 mm SBFI (Minta & Tsien, 1989) was included in the patch-pipette solution, and NaCl omitted. Slices were imaged on a Nikon FN1 microscope using a 60× NA 1.0 fluorite lens (Nikon Corporation, Tokyo, Japan). Cells were illuminated at 350 and 380 nm for 100 ms at each wavelength (Optoscan monochromator, Carin Research, Faversham, UK) and emitted light separated by a 400 nm dichroic mirror and filtered through a 420 nm long-pass filter. Fluorescence was detected using a Cascade 512B electron-multiplying CCD camera (Photometrics, Tucson, AZ, USA), and 350:380 nm ratio images were analysed using Metafluor software (Molecular Devices, Sunnyvale, CA, USA). Calibration was performed by construction of a linear calibration curve using imaging with pipette solutions of known sodium concentrations. A 10% change in background-subtracted 350:380 nm ratio corresponded to a change in sodium concentration of 4.0 mm (R2 = 0.99).

Data analysis

Data are presented as mean ± SEM and regarded as statistically significant when P < 0.05 using one-way analysis of variance (ANOVA) with Dunnett's post hoc test (GraphPad Prism 5.01; GraphPad Software, San Diego, CA, USA), unless otherwise stated. Significance levels indicated on figures are 0.05 (*), 0.01 (**) and 0.001 (***).

Results

MNTB cells can act as glutamine sensors

Astrocytic glutamine release was assessed by recording system A glutamine transporter currents on adjacent neurones. To confirm the suitability of the postsynaptic MNTB cells as glutamine sensors in the region of the synapse, whole-cell patch-clamp recordings from these cells were made and their sensitivity to glutamine investigated. To ensure that trace contamination of glutamine with glutamate does not activate glutamate receptors on the MNTB cells and generate a current, all recordings were performed in the presence of ionotropic glutamate receptor antagonists APV, MK801 and NBQX. Puff-application of 10 mm glutamine from a glass pipette adjacent to the cell (Fig. 1A) induced a membrane current (IGln) of −20.8 ± 0.8 pA (Fig. 1B; mean ± SEM; n = 81). Removal of external Na+ (replaced with choline) abolished IGln (Fig. 1C; 99.5 ± 3.6% inhibition; n = 8; P < 0.001), demonstrating the sodium-dependent nature of the current. To further investigate the nature of IGln, the system A substrate α-(methylamino)isobutyric acid (MeAIB) was bath applied prior to and during the glutamine puff. MeAIB is a competitive substrate with specificity for system A transport over other glutamine transporters (Christensen et al. 1965), and 20 mm MeAIB reduced IGln by 92.9 ± 2.8% (Fig. 1C; n = 10; P < 0.001). These results show that glutamine induces a membrane current in MNTB neurones, which is mediated by the system A glutamine transporters (Blot et al. 2009). Any possible role of high-affinity glutamate transporters (EAATs) in generating this postsynaptic current was excluded by showing a lack of effect of the non-transportable EAAT inhibitor TBOA (200 μm) on the glutamine-evoked current (Fig. 1C; 2.6 ± 3.0% inhibition; n = 6; P > 0.05). We also tested the effects of the EAAT activator d-aspartate on IGln. Bath application of 200 μm d-aspartate did not have a significant effect on IGln (Fig. 1C; 110.4 ± 5.8% of control; n = 3; P > 0.05), confirming that this current is not mediated by EAATs and demonstrating that d-aspartate does not affect glutamine transport on the MNTB cell, which is an important control for later experiments.

Figure 1. The postsynaptic glutamine response is mediated by system A transporters.

Figure 1

A, a differential interference contrast (DIC) image of an MNTB neurone (asterisk) in a brainstem slice, which is whole-cell voltage-clamped at −70 mV via the patch pipette (arrow). Glutamine or glutamine + antagonist were applied by pressure ejection from one of two puffer pipettes placed equidistant from the cell (arrowheads). The scale bar is 20 μm. B, 1 s puff application of 10 mm glutamine induced a membrane current of −20.8 ± 0.8 pA (n = 81). C, the glutamine-induced current was abolished by the removal of external sodium ions and very significantly inhibited by the system A transporter substrate MeAIB (20 mm). An EAAT inhibitor (TBOA; 200 μm) or EAAT substrate (d-aspartate; 200 μm) had no effect on the glutamine-induced current.

Perisynaptic astrocytes express functional glutamate transporters

Astrocytes are known to express EAAT1 and EAAT2 transporters, which are responsible for removing glutamate from the synaptic cleft following neurotransmission (Bergles et al. 1999; Danbolt, 2001). To investigate EAAT-mediated glutamate transport on astrocytes in the MNTB, we whole-cell voltage-clamped astrocytes and puff-applied d-aspartate, an EAAT substrate that does not activate ionotropic glutamate receptors. Astrocytes adjacent to the calyx of Held synapse could be identified by bathing slices in sulforhodamine 101 (Fig. 2AC; Nimmerjahn et al. 2004; Kafitz et al. 2008). Additionally, electrical membrane properties were used to confirm astrocytic recordings (Muller et al. 2009): the resting membrane potential was <−78 mV, the current–voltage relationship was passive (Fig. 2E) and the membrane resistance was <10 MΩ. Puff-application of 200 μm glutamate (in the presence of glutamate receptor inhibitors APV, MK801 and NBQX) induced a current of –31.3 ± 10.5 pA (n = 5), which was inhibited 86.7 ± 11.8% by addition of the EAAT inhibitor TBOA (n = 5; P < 0.05; paired Student's two-tailed t test). To avoid complications resulting from activation of glutamate receptors, subsequent experiments used d-aspartate to stimulate EAATs. Puff-application of 200 μm d-aspartate induced a current (glial ID-asp) of −29.0 ± 1.5 pA (n = 46) which was inhibited by external application of TBOA (Fig. 2G; 72.5 ± 2.6% inhibition; n = 5; P < 0.001). In addition, glial ID-asp was eliminated by application of the high-affinity EAAT substrate inhibitor TFB-TBOA (Fig. 2FG; 99.1 ± 0.9% inhibition; n = 5; P < 0.001). Similarly, inclusion of 5 mm TBOA in the patch-pipette solution also significantly reduced glial ID-asp (Fig. 2G; 62.2 ± 12.4% inhibition; n = 3; P < 0.05), further indicating that it is mediated by astrocytic EAAT glutamate transporters. The subtypes of EAATs responsible for the astrocytic current were investigated using the isoform-specific inhibitors UCPH-101 and dihydrokainate (DHK). UCPH-101 exhibits a high degree of selectivity for EAAT1 (Jensen et al. 2009) and 10 μm inhibited glial ID-asp by 60.2 ± 4.6% (corrected for incomplete inhibition of EAAT1 – see Methods; n = 3). DHK is selective for EAAT2 (Arriza et al. 1994) and 450 μm inhibited glial ID-asp by 30.0 ± 3.9% (corrected; n = 3). A combination of UCPH-101 and DHK resulted in 9.7 ± 7.4% of glial ID-asp remaining uninhibited. To ensure that puff-application did not evoke a membrane current as the result of mechanical stimulation, external solution lacking any agonists was puff-applied to astrocytes. A very slight outward current was observed (0.96 ± 0.53 pA; n = 5) which, combined with the observation that TFB-TBOA eliminates glial ID-asp, demonstrates that our results are not significantly contaminated by puff artefacts.

Figure 2. Activation of perisynaptic astrocytes by d. aspartate application.

Figure 2

A, a DIC image of the MNTB showing astrocyte and neuronal somata. B, fluorescent image of sulforhodamine-101 labelled astrocytes in the same field of cells. C, overlay image of panels A and B showing the position of the astrocyte somata adjacent to the MNTB neurones. Scale bar in A, B and C is 10 μm. D, a fluorescent image of a whole-cell patch-clamped astrocyte (from a different slice to A) with Lucifer yellow included in the internal solution. The astrocyte processes can be seen surrounding the postsynaptic neurone (hash), in the same location as the presynaptic calyx of Held terminal (unlabelled). The scale bar is 20 μm. E, the linear current–voltage relationship of the astrocyte indicates a lack of significant voltage-activated currents, low membrane resistance and negative resting membrane potential. F, puff application of 200 μm d-aspartate at a holding potential of −80 mV induced an inward glutamate transport current in voltage-clamped astrocytes of −29.0 ± 1.5 pA (n = 46). Inclusion of 20 μm TFB-TBOA in the external solution eliminated the d-aspartate-induced current, indicating that it is mediated by activation of EAAT glutamate transporters. G, averaged data showing that the glial d-aspartate-induced current was inhibited by TBOA in the external and internal solutions and also by removal of K+ from the internal solution, demonstrating the involvement of EAATs. Inclusion of the system A substrate MeAIB (20 mm) had no effect on the current. H, fluorescent imaging of the astrocytic [Na+] by inclusion of SBFI in the patch-pipette solution. A 5 s puff application of 1 mm d-aspartate induced a [Na+] rise in the soma of 3.4 ± 0.3 mm (n = 7), which was inhibited by 50.9 ± 6.7% (n = 3) by the inclusion of 200 μm TBOA in the external solution, indicating that the [Na+] rise was mediated by EAAT transporters.

Glutamate transport is linked to the co-transport of three sodium ions, one proton and the counter transport of a potassium ion. Consistent with this stoichiometry, removal of intracellular K+ from the patch-pipette solution (replaced with NMDG) substantially inhibited the glutamate transporter current in the astrocyte (Barbour et al. 1988): glial ID-asp was reduced by 89.2 ± 3.9% of control values by removal of Inline graphic (Fig 2G; n = 5; P < 0.001). Co-transport of sodium would be expected to cause an increase in [Na+]i upon stimulation of glutamate transporters. This was investigated by fluorescent imaging of [Na+]i using 0.5 mm SBFI in the patch-pipette solution (Minta & Tsien, 1989). Application of d-aspartate induced a rapid rise in the somatic [Na+]i, by 3.4 ± 0.3 mm (n = 7), which was reduced by 50.9 ± 6.7% (Fig. 2H; n = 3; P < 0.05; paired Student's 2-tailed t test) by 200 μm TBOA. MeAIB (20 mm) had no significant effect on the glial ID-asp (0.5 ± 5.9% inhibition; n = 5; P > 0.05), indicating that system A glutamine transport is not involved in this astrocytic activation and that MeAIB does not affect EAATs. These data confirm that d-aspartate activates glutamate transporters on perisynaptic astrocytes, causing an inward membrane current and increase in [Na+]i.

Activation of astrocytic glutamate transporters causes glutamine release

Glutamine release into the region of the synapse was detected by whole-cell patch-clamping the MNTB cell. We find that activation of glial glutamate transporters using puff-application of 200 μm d-aspartate induces the release of glutamine, which subsequently stimulates the MNTB cell. This d-aspartate-induced glutamine transporter current recorded in MNTB cells (MNTB ID-asp) was observed in 113 of 125 MNTB cells (90.4%) and averaged −8.7 ± 0.6 pA (Fig. 3; n = 125). As these neurones express a high density of ionotropic glutamate receptors and exhibit glutamatergic EPSCs of >10 nA (Forsythe & Barnes-Davies, 1993; Borst et al. 1995), the orders of magnitude difference between the receptor and transporter currents would prevent adequate pharmacological separation of glutamate-induced currents and preclude the use of glutamate as an EAAT substrate in these experiments. Nevertheless, the role of glutamate transporters in initiating the d-aspartate-induced current is confirmed by the application of TBOA, which inhibited the MNTB ID-asp by 67.9 ± 9.1% (Fig. 3A; n = 6; P < 0.01). Furthermore, TFB-TBOA (20 μm) inhibited MNTB ID-asp by 94.2 ± 9.5% (Fig. 3D; n = 3; P < 0.01). It is possible that d-aspartate directly activates neuronal EAATs on MNTB principal cells, as immunohistochemical experiments reveal low levels of the neuronal EAAT3 transporter in the MNTB (Renden et al. 2005). However, to be certain that d-aspartate does not directly activate any putative neuronal EAATs, an intracellular solution lacking Cs+ (which substitutes for internal K+ on EAAT transporters; Barbour et al. 1991) was used. Upon removal of Inline graphic (substituted by NMDG), MNTB ID-asp was unchanged, at 113.0 ± 28.5% of control (Fig. 3B; n = 3; P > 0.05). In contrast to effects of removing internal K+ on the EAAT current in glial cells (Fig. 2G), this result demonstrates that EAATs are not functional in the postsynaptic MNTB cells and do not contribute to MNTB ID-asp. If glutamine acting on system A transporters is creating the observed MNTB current, then MeAIB or an excess of glutamine in the external solution would both be expected to abolish the current. Accordingly, 20 mm MeAIB inhibited MNTB ID-asp completely (Fig. 3C; n = 6; P < 0.001) as did 5 mm glutamine (Fig. 3D; n = 5; P < 0.001) consistent with MNTB ID-asp being mediated by released glutamine acting on MNTB cells.

Figure 3. Stimulating perisynaptic astrocytes results in glutamine release, detected by neighbouring postsynaptic neurones.

Figure 3

A, application of 200 μm d-aspartate to the brain slice resulted in a current in whole-cell voltage-clamped postsynaptic MNTB neurones, which was inhibited by adding 200 μm TBOA to the external solution, demonstrating the involvement of EAAT activation. B, the postsynaptic d-aspartate-evoked current was unaffected by the removal of Cs+ from the internal solution, highlighting the lack of involvement of postsynaptic EAATs in the d-aspartate response. C, application of the system A substrate MeAIB (20 mm) occluded the effect of d-aspartate application, indicating that system A is involved in the generation of the postsynaptic current. D, averaged data showing the effects of 200 μm TBOA, 20 μm TFB-TBOA, removal of Cs+, 20 mm MeAIB and 5 mm glutamine on the d-aspartate-induced neuronal current.

Release of glutamine is mediated by perisynaptic astrocytes

Two separate pharmacological approaches were used to demonstrate the involvement of astrocytes in releasing the glutamine that we detect. l-methionine sulfoximine (MSO) is an inhibitor of glutamine synthetase, and prevents the conversion of glutamate to glutamine in astrocytes. It specifically reduces glutamine production in comparison with other amino acids (Fonnum et al. 1997) and the subsequent reduction in the astrocytic glutamine concentration (Tani et al. 2010) would be expected to reduce glutamine release and thus inhibit MNTB ID-asp. Alternatively, fluoroacetate (FAc) is a glial-specific toxin that is taken up into astrocytes, converted to fluorocitrate and inhibits aconitase. The resulting inhibition of the astrocytic tricarboxylic acid (TCA) cycle causes astrocytic glutamate to be metabolised via α-ketoglutarate, and subsequently reduces glutamine production (Swanson & Graham, 1994; Fonnum et al. 1997), which would reduce MNTB ID-asp. Incubation of the brain slices in 10 mm MSO for at least 120 min prior to recording inhibited MNTB ID-asp by 64.0 ± 24.5% (Fig. 4A and C; n = 9; P < 0.01). Correspondingly, application of 10 mm FAc for 30–120 min inhibited MNTB ID-asp by 66.5 ± 7.4% (Fig. 4B and C; n = 8; P < 0.01) without any observable effects on the viability of the MNTB cells. These pharmacological data strongly implicate the astrocytes as being the source of the released glutamine. There are three times as many MNTB principal cells as astrocytes in this brain region (Reyes-Haro et al. 2010), suggesting that a single perisynaptic astrocyte may ensheathe more than one calyx of Held synapse and that each synapse may be ensheathed by only one astrocyte, which is coupled to neighbouring astrocytes via gap junctions (Muller et al. 2009). To investigate this relationship, we simultaneously whole-cell patch-clamped MNTB cells and neighbouring astrocytes, and activated glial glutamate transport (Fig. 5A). Puff-application of d-aspartate (Fig. 5B) demonstrated that MNTB ID-asp was activated very rapidly, within a few milliseconds of glial ID-asp. Two different techniques were used to investigate the role of an identified astrocyte in mediating the glutamine release. First, the astrocyte was whole-cell patch-clamped with a pipette containing extracellular solution and at a membrane potential of 0 mV, which would prevent astrocytic d-aspartate transport by removal of the driving force powering the EAATs. The glutamine release (as assessed by MNTB ID-asp) was recorded immediately before and after inactivating the astrocytic EAATs in this way. Our data show that the MNTB ID-asp was instantly inhibited by 81.6 ± 15.5% (Fig. 5C and D; n = 4; P < 0.01) by inactivation of the astrocyte. The second method was to include 5 mm TBOA in the astrocytic patch-pipette solution. This high concentration of TBOA was used to ensure that sufficient TBOA would diffuse down the astrocytic processes to inhibit glutamate transporters, which are located adjacent to the synapse. Internal TBOA reduced glial I D-asp by 62.2 ± 12.4% (n = 3; P < 0.05) and reduced MNTB ID-asp by 75.7 ± 10.3% (Fig. 5C; n = 3; P < 0.05). These data suggest that a perisynaptic astrocyte mediates the glutamate transporter-evoked glutamine release, with a high degree of temporal precision.

Figure 4. The stimulated release of glutamine is mediated by glial cells.

Figure 4

A, postsynaptic currents induced by the application of d-aspartate to the synapse are significantly inhibited by incubating the tissue in the glutamine synthetase inhibitor l-MSO (10 mm), which depletes astrocytes of glutamine. B, incubating the tissue with the glial-specific metabolic toxin fluoroacetate (FAc; 10 mm) also severely compromised the d-aspartate-induced postsynaptic current, further demonstrating the role of glial cells in the activated pathway. C, averaged data indicating the similar levels of inhibition of the postsynaptic d-aspartate-induced current by l-MSO (64%) and FAc (67%).

Figure 5. A perisynaptic astrocyte mediates the d. aspartate-induced release of glutamine.

Figure 5

A, a DIC image showing a voltage-clamped astrocyte (asterisk) adjacent to a voltage-clamped postsynaptic MNTB neurone (hash). A puffer pipette (arrowhead) is positioned nearby to apply d-aspartate. The scale bar is 20 μm. B, puff application of 200 μm d-aspartate induces an EAAT-mediated current in the astrocyte (black trace) and an almost instantaneous system A-mediated current in the adjacent neurone (grey trace). C, the neuronal current was significantly inhibited by inactivating the perisynaptic astrocyte by continual depolarization to 0 mV, or by including TBOA in the astrocyte patch-pipette solution. D, averaged data (from 4 cells) showing the time dependence of the reduction in postsynaptic d-aspartate-induced current when the astrocyte is inactivated by voltage-clamping the cell at 0 mV.

Discussion

The release of glutamine from astrocytes is a key component of the glutamate–glutamine cycle (Chaudhry et al. 2002a). However, the time scale on which this cycle may occur has not been established. Here, we directly demonstrate that activation of glutamate transport on perisynaptic astrocytes causes a rapid release of glutamine from the astrocyte, with tight temporal coupling. This suggests that astrocytic glutamine may be an immediate modulator of neuronal function.

We have studied glutamine release from astrocytes in the MNTB, adjacent to the calyx of Held synapse. This brain area has several key advantages for these experiments: the MNTB principal cells are spherical and have only one or two short dendrites (Leao et al. 2008), making them electrically compact and allowing recording of the relatively small transporter current without significant dendritic filtering; the main synaptic input to the MNTB principal cell is somatic and can be visually identified; the astrocyte adjacent to the synapse can be located; and the small number of astrocytes compared with principal cells (one astrocyte for three principal cells) suggests that only one astrocyte may be associated with the majority of the synaptic input to a single cell (Reyes-Haro et al. 2010). In addition, the ability to perform simultaneous patch-clamp experiments of the identified perisynaptic astrocyte and the postsynaptic neurone allows direct manipulation of this glial element in the synapse. Since we have demonstrated that the postsynaptic MNTB cell is sensitive to glutamine levels in the external solution, we have used the membrane current resulting from system A glutamine transporter activation as an indicator of glutamine release from the astrocyte when glutamate transport is stimulated in the immediate vicinity of the synapse.

Astrocytic glutamate uptake

Our data demonstrate that MNTB astrocytes possess functional glutamate transporters, making them likely candidates to take up synaptically released glutamate. Immunohistological studies reveal that MNTB astrocytes express both the EAAT1 (GLAST) and the EAAT2 (GLT-1) isoforms of glutamate transporters (Renden et al. 2005; Ford et al. 2009), and our data using isoform-specific antagonists, UCPH-101 and DHK, reveal that EAAT1 is the main functional isoform present, contributing approximately twice the current of EAAT2. This finding is consistent with the expression pattern observed in the cerebellum and in contrast to the forebrain where EAAT2 predominates (Chaudhry et al. 1995; Lehre & Danbolt, 1998). A negligible amount of current remains when both EAAT1 and EAAT2 are inhibited, which, in combination with the almost complete inhibition of current by removal of internal K+ and its elimination by TFB-TBOA, suggests an insignificant contribution of non-EAAT transport in mediating the d-aspartate current in astrocytes. Astrocytes are known to sequester synaptically released glutamate via EAATs and a direct activation of glial glutamate transporters following synaptic stimulation has been observed in the cerebellum (Clark & Barbour, 1997) and hippocampus (Bergles & Jahr, 1997), but not in the MNTB (Reyes-Haro et al. 2010). This difficulty in observing synaptically induced astrocytic transporter currents could be a result of the high conductance of the astrocyte membrane or as a consequence of the small number (<10%) of MNTB cells that retain active synaptic connections following the brainstem slice procedure (Billups et al. 2002). Moreover, inhibiting astrocytic glutamate transport influences the levels of glutamate in and around the synapse and alters the properties of synaptic transmission at the calyx of Held synapse (Renden et al. 2005), as well as at other calyceal synapses (Otis et al. 1996), in the cerebellum (Barbour et al. 1994; Overstreet et al. 1999; Carter & Regehr, 2000; Marcaggi et al. 2003) and in the hippocampus (Scanziani et al. 1997; Huang et al. 2004). This demonstrates that glutamate sequestration by astrocytic EAATs plays an important role in synaptic physiology. As location of EAATs adjacent to active synapses is a highly dynamic process and an increase in synaptic activity causes a rapid clustering of transporters adjacent to regions of higher glutamate release (Zhou & Sutherland, 2004; Yang et al. 2009; Benediktsson et al. 2012), astrocytic glutamate uptake mediates a key neuronal–glial communication pathway.

Astrocytic glutamine release

Our data show that stimulation of astrocytes by EAAT activation causes the release of glutamine. Glutamine release is detected by its subsequent activation of the system A amino acid transporter, and while it is possible that astrocytes release a system A substrate other than glutamine, this is not supported by our data. MSO is a relatively specific inhibitor of glutamine synthetase and inhibiting this pathway reduces astrocytic glutamine content but does not reduce glutamate, aspartate or alanine concentrations (Fonnum et al. 1997). Further support for glutamine mediating the glial–neuronal communication comes from the use of FAc. Inhibition of the glial TCA cycle by FAc (or its active metabolite fluorocitrate) has been shown to substantially reduce glutamine production, but not that of other amino acids or ATP, and does not inhibit glial glutamate or d-aspartate uptake (Paulsen et al. 1988; Swanson & Graham, 1994; Fonnum et al. 1997). The only other system A substrate thought capable of mediating transfer of neurotransmitter precursors from glia to neurones is alanine (Broer et al. 2007). However, both FAc and MSO have been shown to actually increase the alanine content of astrocytes by 2- to 3-fold (Fonnum et al. 1997), presumably as a consequence of diverting astrocytic glutamate metabolism away from glutamine production. As FAc and MSO inhibit the glial–neuronal communication in our experiments, it cannot be mediated by release of alanine or any of its metabolites from astrocytes. In addition to its action as a system A substrate, alanine also acts as an agonist at glycine receptors (Young & Snyder, 1973), which are expressed at the calyx of Held and have a depolarizing action (Price & Trussell, 2006; Huang & Trussell, 2008) that enhances neurotransmitter release (Turecek & Trussell, 2001; Awatramani et al. 2005). However, any contribution of alanine-activated currents to our results is eliminated by the strychnine present in the external solution.

It has been proposed that glutamine release from astrocytes is mediated by the system N transporter family, which is comprised of two main transporters: system N 1 (named SN1, NAT or SNAT3) and system N 2 (SN2 or SNAT5) (Kilberg et al. 1980; Chaudhry et al. 1999; Gu et al. 2000; Nakanishi et al. 2001). Immunohistochemical staining has shown that both isoforms are expressed in astrocytes of the MNTB, where they are located in astrocytic processes adjacent to presynaptic terminals (Boulland et al. 2002; Cubelos et al. 2005). The homologous transporter SNAT7 has also been ascribed to system N, although this transporter is expressed in neurons and not in astrocytes (Hagglund et al. 2011). Glutamine transport by system N is linked to the co-transport of a sodium ion and the counter-transport of a proton, resulting in electroneutral, pH-dependent transport (Broer et al. 2002; but see Fei et al. 2000). Compared with system A, which does not counter-transport a proton (Chaudhry et al. 2002b), system N operates near to thermodynamic equilibrium under physiological conditions and is therefore expected to reverse more readily to mediate glutamine efflux from astrocytes. In addition to systems A and N, astrocytes express systems L, y+L and ASC, which can also transport glutamine (Nagaraja & Brookes, 1996; Su et al. 1997; Broer et al. 1999; Heckel et al. 2003). These transporters are thought to be capable of mediating glutamine release and are obligate exchangers, requiring another amino acid to be transported in the opposite direction. Importantly, systems L and y+L are not sensitive to aspartate (Kanai et al. 1998; Segawa et al. 1999; Broer et al. 2000), and ASC is not influenced by d-aspartate (Heckel et al. 2003), excluding direct d-aspartate–glutamine exchange as a mechanism for glutamine release in our experiments.

Triggering glutamine release

Our data indicate that EAAT activation is required for glutamine release. However, it is important to establish that there is no direct activation of EAATs on the postsynaptic MNTB neurones themselves, and that the d-aspartate-induced currents were entirely mediated by astrocytic glutamine release. This was evident from their complete inhibition by MeAIB (Fig. 3C), which does not inhibit EAATs (Fig. 2G), and the lack of inhibition by removal of internal Cs+ (Fig. 3B). Therefore, we conclude that there is no EAAT transporter activity in postsynaptic MNTB neurones. While it is clear that d-aspartate application is activating astrocytes and not neurones, the exact mechanism linking glutamate transport to glutamine release is unclear. It has previously been shown that incubating cultured astrocytes with glutamate increases the internal glutamine concentration and can stimulate glutamine release by enhancing the affinity of system N transport for glutamine (Broer et al. 2004). However, this effect is only apparent after 10 min and the glutamine release that we observe occurs on the millisecond time scale. While we cannot rule out an effect of d-aspartate on the substrate affinity of the glutamine release mechanism, it is unlikely as d-aspartate is not metabolised to glutamine and would not be expected to have the same direct effects on system N transport as glutamate. A more likely mechanism underlying the incredibly rapid glutamate transporter-induced glutamine release we observe might be due to the properties of EAATs. As transport of d-aspartate by EAATs is driven by the co-transport of three sodium ions and a proton, and counter transport of a potassium ion (Zerangue & Kavanaugh, 1996), its application causes depolarization of the astrocytes. These ion and voltage changes may stimulate subsequent glutamine release. However, our experiments were performed under voltage-clamp, ruling out any possible effect of depolarization in mediating the stimulation of glutamine release. It is also unlikely that membrane voltage changes could drive the glutamine release physiologically, as (1) the high conductance of the astrocyte plasma membrane reduces any voltage change by a current-shunting mechanism, and (2) the electroneutrality of system N transport would suggest insensitivity to voltage changes. Similarly, the pH change associated with EAAT-mediated transport (internal acidification; Bouvier et al. 1992) would also not be expected to stimulate glutamine release, as the stoichiometry of system N transport indicates that internal acidification would inhibit release. In contrast, we show that activation of EAATs results in a significant and rapid rise in internal sodium ion concentration (Fig. 2H), consistent with previous reports of large sodium concentration rises in response to glutamate transporter activation in Bergmann glia (Kirischuk et al. 2007; Bennay et al. 2008) and astrocytes in cortical cultures (Chatton et al. 2000) and hippocampal slices (Langer & Rose, 2009; Langer et al. 2011). The low millimolar increase in sodium concentration that we observe in the astrocytic soma probably reflects a much larger increase in the fine processes (Langer & Rose, 2009; Langer et al. 2011), which will significantly alter the equilibrium of system N transport to one where a higher external glutamine concentration will be established, resulting in the release of glutamine to the extracellular space (Fig. 6). As there is a 3:1 relationship between sodium and glutamate influx on EAATs (Zerangue & Kavanaugh, 1996) and a 1:1 coupling of sodium and glutamine release by system N (Broer et al. 2002), amplification between glutamate uptake and glutamine release could occur. The concentration of glutamine in the extracellular space is thought to be 0.4 mm (Kanamori & Ross, 2004), which is below the level that would saturate the system A transporter isoforms thought to be responsible for the import of glutamine into neurones (Kms = 0.3–2.3 mm; Yao et al. 2000; Chaudhry et al. 2002b; Mackenzie et al. 2003). Therefore, any increase in glutamine release would enhance neuronal glutamine uptake.

Figure 6. A proposed mechanism for astrocytic glutamine release.

Figure 6

Activation of glial glutamate transporters (EAATs) causes a rise in glial [Na+]i, which favours the efflux of glutamine via glial system N transporters (SN). Released glutamine is subsequently detected by system A transporters (SA) located on adjacent MNTB neurones, resulting in a neuronal glutamine transport current.

Physiological implications

Glutamine is used by neurones in the generation of glutamate for excitatory neurotransmission (Lieth et al. 2001; Sibson et al. 2001; Rothman et al. 2003), and it is therefore likely that glutamine released from astrocytes is sequestered into presynaptic terminals for this purpose. The high levels of neurotransmission that are observed in brain-slice models of epilepsy are dependent on glutamine (Bacci et al. 2002; Tani et al. 2007, 2010), and the high firing frequencies of auditory synapses such as the calyx of Held (Hermann et al. 2007; Kopp-Scheinpflug et al. 2008) suggest that a similar mechanism also operates in this brain region under physiological conditions. The tight temporal coupling between sensing glutamate and releasing glutamine infers that astrocytic glutamine release may reflect a mechanism by which excitatory transmission may be rapidly modulated. In addition to glutamate production, glutamine is also required for the production of GABA (Liang et al. 2006; Fricke et al. 2007), indicating that astrocytic glutamine release may similarly enhance inhibitory neurotransmission. Hence, it is possible that glutamine release mediates the modulation of GABAergic transmission by astrocytes in the barrel cortex (Benedetti et al. 2011). These observations suggest that alteration of astrocytic glutamine release, such as by phosphorylation of system N transporters (Balkrishna et al. 2010; Nissen-Meyer et al. 2011) would be expected to have far-reaching effects on neuronal communication. Hence, our study demonstrates the potential for astrocytes to modulate neurotransmission on a very rapid time scale by the release of glutamine. This would represent a further manifestation of the tripartite relationship between astrocytes and synapses.

Acknowledgments

We thank Dr Daniela Billups for critical reading of the manuscript. The work was supported by the Royal Society.

Glossary

Abbreviations

aCSF

artificial cerebrospinal fluid

DHK

dihydrokainate

EAAT

excitatory amino acid transporter

FAc

fluoroacetate

MeAIB

α-(methylamino)isobutyric acid

MK801

dizocilpine maleate

MNTB

medial nucleus of the trapezoid body

MSO

l-methionine sulfoximine

Author contributions

N.M.U. and B.B. conceived and designed the experiments. N.M.U., M.-C.M., Q.C. and B.B. collected, analysed and interpreted the data. N.M.U. and B.B. drafted the manuscript and N.M.U., M.-C.M., Q.C. and B.B. critically revised the manuscript. All authors approved the final version of the manuscript. All experiments were performed in the Department of Pharmacology, University of Cambridge.

References

  1. Anderson BJ, Li X, Alcantara AA, Isaacs KR, Black JE, Greenough WT. Glial hypertrophy is associated with synaptogenesis following motor-skill learning, but not with angiogenesis following exercise. Glia. 1994;11:73–80. doi: 10.1002/glia.440110110. [DOI] [PubMed] [Google Scholar]
  2. Arriza JL, Fairman WA, Wadiche JI, Murdoch GH, Kavanaugh MP, Amara SG. Functional comparisons of three glutamate transporter subtypes cloned from human motor cortex. J Neurosci. 1994;14:5559–5569. doi: 10.1523/JNEUROSCI.14-09-05559.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Awatramani GB, Price GD, Trussell LO. Modulation of transmitter release by presynaptic resting potential and background calcium levels. Neuron. 2005;48:109–121. doi: 10.1016/j.neuron.2005.08.038. [DOI] [PubMed] [Google Scholar]
  4. Bacci A, Sancini G, Verderio C, Armano S, Pravettoni E, Fesce R, Franceschetti S, Matteoli M. Block of glutamate-glutamine cycle between astrocytes and neurons inhibits epileptiform activity in hippocampus. J Neurophysiol. 2002;88:2302–2310. doi: 10.1152/jn.00665.2001. [DOI] [PubMed] [Google Scholar]
  5. Bak LK, Schousboe A, Waagepetersen HS. The glutamate/GABA-glutamine cycle: aspects of transport, neurotransmitter homeostasis and ammonia transfer. J Neurochem. 2006;98:641–653. doi: 10.1111/j.1471-4159.2006.03913.x. [DOI] [PubMed] [Google Scholar]
  6. Balkrishna S, Broer A, Kingsland A, Broer S. Rapid downregulation of the rat glutamine transporter SNAT3 by a caveolin-dependent trafficking mechanism in Xenopus laevis oocytes. Am J Physiol Cell Physiol. 2010;299:C1047–C1057. doi: 10.1152/ajpcell.00209.2010. [DOI] [PubMed] [Google Scholar]
  7. Barbour B, Brew H, Attwell D. Electrogenic glutamate uptake in glial cells is activated by intracellular potassium. Nature. 1988;335:433–435. doi: 10.1038/335433a0. [DOI] [PubMed] [Google Scholar]
  8. Barbour B, Brew H, Attwell D. Electrogenic uptake of glutamate and aspartate into glial cells isolated from the salamander (Ambystoma) retina. J Physiol. 1991;436:169–193. doi: 10.1113/jphysiol.1991.sp018545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Barbour B, Keller BU, Llano I, Marty A. Prolonged presence of glutamate during excitatory synaptic transmission to cerebellar Purkinje cells. Neuron. 1994;12:1331–1343. doi: 10.1016/0896-6273(94)90448-0. [DOI] [PubMed] [Google Scholar]
  10. Bardoni R, Ghirri A, Zonta M, Betelli C, Vitale G, Ruggieri V, Sandrini M, Carmignoto G. Glutamate-mediated astrocyte-to-neuron signalling in the rat dorsal horn. J Physiol. 2010;588:831–846. doi: 10.1113/jphysiol.2009.180570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Barnett NL, Pow DV, Robinson SR. Inhibition of Muller cell glutamine synthetase rapidly impairs the retinal response to light. Glia. 2000;30:64–73. doi: 10.1002/(sici)1098-1136(200003)30:1<64::aid-glia7>3.0.co;2-i. [DOI] [PubMed] [Google Scholar]
  12. Benedetti B, Matyash V, Kettenmann H. Astrocytes control GABAergic inhibition of neurons in the mouse barrel cortex. J Physiol. 2011;589:1159–1172. doi: 10.1113/jphysiol.2010.203224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Benediktsson AM, Marrs GS, Tu JC, Worley PF, Rothstein JD, Bergles DE, Dailey ME. Neuronal activity regulates glutamate transporter dynamics in developing astrocytes. Glia. 2012;60:175–188. doi: 10.1002/glia.21249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bennay M, Langer J, Meier SD, Kafitz KW, Rose CR. Sodium signals in cerebellar Purkinje neurons and Bergmann glial cells evoked by glutamatergic synaptic transmission. Glia. 2008;56:1138–1149. doi: 10.1002/glia.20685. [DOI] [PubMed] [Google Scholar]
  15. Bergles DE, Diamond JS, Jahr CE. Clearance of glutamate inside the synapse and beyond. Curr Opin Neurobiol. 1999;9:293–298. doi: 10.1016/s0959-4388(99)80043-9. [DOI] [PubMed] [Google Scholar]
  16. Bergles DE, Jahr CE. Synaptic activation of glutamate transporters in hippocampal astrocytes. Neuron. 1997;19:1297–1308. doi: 10.1016/s0896-6273(00)80420-1. [DOI] [PubMed] [Google Scholar]
  17. Billups B, Wong AY, Forsythe ID. Detecting synaptic connections in the medial nucleus of the trapezoid body using calcium imaging. Pflugers Arch. 2002;444:663–669. doi: 10.1007/s00424-002-0861-6. [DOI] [PubMed] [Google Scholar]
  18. Blot A, Billups D, Bjorkmo M, Quazi AZ, Uwechue NM, Chaudhry FA, Billups B. Functional expression of two system A glutamine transporter isoforms in rat auditory brainstem neurons. Neuroscience. 2009;164:998–1008. doi: 10.1016/j.neuroscience.2009.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Borst JG, Helmchen F, Sakmann B. Pre- and postsynaptic whole-cell recordings in the medial nucleus of the trapezoid body of the rat. J Physiol. 1995;489:825–840. doi: 10.1113/jphysiol.1995.sp021095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Boulland JL, Osen KK, Levy LM, Danbolt NC, Edwards RH, Storm-Mathisen J, Chaudhry FA. Cell-specific expression of the glutamine transporter SN1 suggests differences in dependence on the glutamine cycle. Eur J Neurosci. 2002;15:1615–1631. doi: 10.1046/j.1460-9568.2002.01995.x. [DOI] [PubMed] [Google Scholar]
  21. Bouvier M, Szatkowski M, Amato A, Attwell D. The glial cell glutamate uptake carrier countertransports pH-changing anions. Nature. 1992;360:471–474. doi: 10.1038/360471a0. [DOI] [PubMed] [Google Scholar]
  22. Broer A, Albers A, Setiawan I, Edwards RH, Chaudhry FA, Lang F, Wagner CA, Broer S. Regulation of the glutamine transporter SN1 by extracellular pH and intracellular sodium ions. J Physiol. 2002;539:3–14. doi: 10.1113/jphysiol.2001.013303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Broer A, Brookes N, Ganapathy V, Dimmer KS, Wagner CA, Lang F, Broer S. The astroglial ASCT2 amino acid transporter as a mediator of glutamine efflux. J Neurochem. 1999;73:2184–2194. [PubMed] [Google Scholar]
  24. Broer A, Deitmer JW, Broer S. Astroglial glutamine transport by system N is upregulated by glutamate. Glia. 2004;48:298–310. doi: 10.1002/glia.20081. [DOI] [PubMed] [Google Scholar]
  25. Broer A, Wagner CA, Lang F, Broer S. The heterodimeric amino acid transporter 4F2hc/y+LAT2 mediates arginine efflux in exchange with glutamine. Biochem J. 2000;349:787–795. doi: 10.1042/bj3490787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Broer S, Broer A, Hansen JT, Bubb WA, Balcar VJ, Nasrallah FA, Garner B, Rae C. Alanine metabolism, transport, and cycling in the brain. J Neurochem. 2007;102:1758–1770. doi: 10.1111/j.1471-4159.2007.04654.x. [DOI] [PubMed] [Google Scholar]
  27. Carter AG, Regehr WG. Prolonged synaptic currents and glutamate spillover at the parallel fiber to stellate cell synapse. J Neurosci. 2000;20:4423–4434. doi: 10.1523/JNEUROSCI.20-12-04423.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Chatton JY, Marquet P, Magistretti PJ. A quantitative analysis of L-glutamate-regulated Na+ dynamics in mouse cortical astrocytes: implications for cellular bioenergetics. Eur J Neurosci. 2000;12:3843–3853. doi: 10.1046/j.1460-9568.2000.00269.x. [DOI] [PubMed] [Google Scholar]
  29. Chaudhry FA, Lehre KP, van Lookeren Campagne M, Ottersen OP, Danbolt NC, Storm-Mathisen J. Glutamate transporters in glial plasma membranes: highly differentiated localizations revealed by quantitative ultrastructural immunocytochemistry. Neuron. 1995;15:711–720. doi: 10.1016/0896-6273(95)90158-2. [DOI] [PubMed] [Google Scholar]
  30. Chaudhry FA, Reimer RJ, Edwards RH. The glutamine commute: take the N line and transfer to the A. J Cell Biol. 2002a;157:349–355. doi: 10.1083/jcb.200201070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Chaudhry FA, Reimer RJ, Krizaj D, Barber D, Storm-Mathisen J, Copenhagen DR, Edwards RH. Molecular analysis of system N suggests novel physiological roles in nitrogen metabolism and synaptic transmission. Cell. 1999;99:769–780. doi: 10.1016/s0092-8674(00)81674-8. [DOI] [PubMed] [Google Scholar]
  32. Chaudhry FA, Schmitz D, Reimer RJ, Larsson P, Gray AT, Nicoll R, Kavanaugh M, Edwards RH. Glutamine uptake by neurons: interaction of protons with system a transporters. J Neurosci. 2002b;22:62–72. doi: 10.1523/JNEUROSCI.22-01-00062.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Christensen HN, Oxender DL, Liang M, Vatz KA. The use of N-methylation to direct route of mediated transport of amino acids. J Biol Chem. 1965;240:3609–3616. [PubMed] [Google Scholar]
  34. Clark BA, Barbour B. Currents evoked in Bergmann glial cells by parallel fibre stimulation in rat cerebellar slices. J Physiol. 1997;502:335–350. doi: 10.1111/j.1469-7793.1997.335bk.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Cubelos B, Gonzalez-Gonzalez IM, Gimenez C, Zafra F. Amino acid transporter SNAT5 localizes to glial cells in the rat brain. Glia. 2005;49:230–244. doi: 10.1002/glia.20106. [DOI] [PubMed] [Google Scholar]
  36. Danbolt NC. Glutamate uptake. Prog Neurobiol. 2001;65:1–105. doi: 10.1016/s0301-0082(00)00067-8. [DOI] [PubMed] [Google Scholar]
  37. Fei YJ, Sugawara M, Nakanishi T, Huang W, Wang H, Prasad PD, Leibach FH, Ganapathy V. Primary structure, genomic organization, and functional and electrogenic characteristics of human system N 1, a Na+- and H+-coupled glutamine transporter. J Biol Chem. 2000;275:23707–23717. doi: 10.1074/jbc.M002282200. [DOI] [PubMed] [Google Scholar]
  38. Fonnum F, Johnsen A, Hassel B. Use of fluorocitrate and fluoroacetate in the study of brain metabolism. Glia. 1997;21:106–113. [PubMed] [Google Scholar]
  39. Ford MC, Grothe B, Klug A. Fenestration of the calyx of Held occurs sequentially along the tonotopic axis, is influenced by afferent activity, and facilitates glutamate clearance. J Comp Neurol. 2009;514:92–106. doi: 10.1002/cne.21998. [DOI] [PubMed] [Google Scholar]
  40. Forsythe ID, Barnes-Davies M. The binaural auditory pathway: excitatory amino acid receptors mediate dual timecourse excitatory postsynaptic currents in the rat medial nucleus of the trapezoid body. Proc Biol Sci. 1993;251:151–157. doi: 10.1098/rspb.1993.0022. [DOI] [PubMed] [Google Scholar]
  41. Fricke MN, Jones-Davis DM, Mathews GC. Glutamine uptake by System A transporters maintains neurotransmitter GABA synthesis and inhibitory synaptic transmission. J Neurochem. 2007;102:1895–1904. doi: 10.1111/j.1471-4159.2007.04649.x. [DOI] [PubMed] [Google Scholar]
  42. Genoud C, Quairiaux C, Steiner P, Hirling H, Welker E, Knott GW. Plasticity of astrocytic coverage and glutamate transporter expression in adult mouse cortex. PLoS Biol. 2006;4:e343. doi: 10.1371/journal.pbio.0040343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Gibbs ME, O’Dowd BS, Hertz L, Robinson SR, Sedman GL, Ng KT. Inhibition of glutamine synthetase activity prevents memory consolidation. Brain Res Cogn Brain Res. 1996;4:57–64. doi: 10.1016/0926-6410(96)00020-1. [DOI] [PubMed] [Google Scholar]
  44. Gu S, Roderick HL, Camacho P, Jiang JX. Identification and characterization of an amino acid transporter expressed differentially in liver. Proc Natl Acad Sci U S A. 2000;97:3230–3235. doi: 10.1073/pnas.050318197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Hagglund MG, Sreedharan S, Nilsson VC, Shaik JH, Almkvist IM, Backlin S, Wrange O, Fredriksson R. Identification of SLC38A7 (SNAT7) protein as a glutamine transporter expressed in neurons. J Biol Chem. 2011;286:20500–20511. doi: 10.1074/jbc.M110.162404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Hawrylak N, Chang FL, Greenough WT. Astrocytic and synaptic response to kindling in hippocampal subfield CA1. II. Synaptogenesis and astrocytic process increases to in vivo kindling. Brain Res. 1993;603:309–316. doi: 10.1016/0006-8993(93)91253-o. [DOI] [PubMed] [Google Scholar]
  47. Heckel T, Broer A, Wiesinger H, Lang F, Broer S. Asymmetry of glutamine transporters in cultured neural cells. Neurochem Int. 2003;43:289–298. doi: 10.1016/s0197-0186(03)00014-7. [DOI] [PubMed] [Google Scholar]
  48. Henneberger C, Papouin T, Oliet SH, Rusakov DA. Long-term potentiation depends on release of D-serine from astrocytes. Nature. 2010;463:232–236. doi: 10.1038/nature08673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Hermann J, Pecka M, von Gersdorff H, Grothe B, Klug A. Synaptic transmission at the calyx of Held under in vivo like activity levels. J Neurophysiol. 2007;98:807–820. doi: 10.1152/jn.00355.2007. [DOI] [PubMed] [Google Scholar]
  50. Hertz L, Dringen R, Schousboe A, Robinson SR. Astrocytes: glutamate producers for neurons. J Neurosci Res. 1999;57:417–428. [PubMed] [Google Scholar]
  51. Huang H, Trussell LO. Control of presynaptic function by a persistent Na+ current. Neuron. 2008;60:975–979. doi: 10.1016/j.neuron.2008.10.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Huang YH, Sinha SR, Tanaka K, Rothstein JD, Bergles DE. Astrocyte glutamate transporters regulate metabotropic glutamate receptor-mediated excitation of hippocampal interneurons. J Neurosci. 2004;24:4551–4559. doi: 10.1523/JNEUROSCI.5217-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Jensen AA, Erichsen MN, Nielsen CW, Stensbol TB, Kehler J, Bunch L. Discovery of the first selective inhibitor of excitatory amino acid transporter subtype 1. J Med Chem. 2009;52:912–915. doi: 10.1021/jm8013458. [DOI] [PubMed] [Google Scholar]
  54. Jones TA, Greenough WT. Ultrastructural evidence for increased contact between astrocytes and synapses in rats reared in a complex environment. Neurobiol Learn Mem. 1996;65:48–56. doi: 10.1006/nlme.1996.0005. [DOI] [PubMed] [Google Scholar]
  55. Kafitz KW, Meier SD, Stephan J, Rose CR. Developmental profile and properties of sulforhodamine 101-labeled glial cells in acute brain slices of rat hippocampus. J Neurosci Methods. 2008;169:84–92. doi: 10.1016/j.jneumeth.2007.11.022. [DOI] [PubMed] [Google Scholar]
  56. Kanai Y, Segawa H, Miyamoto K, Uchino H, Takeda E, Endou H. Expression cloning and characterization of a transporter for large neutral amino acids activated by the heavy chain of 4F2 antigen (CD98) J Biol Chem. 1998;273:23629–23632. doi: 10.1074/jbc.273.37.23629. [DOI] [PubMed] [Google Scholar]
  57. Kanamori K, Ross BD. Quantitative determination of extracellular glutamine concentration in rat brain, and its elevation in vivo by system A transport inhibitor, α-(methylamino)isobutyrate. J Neurochem. 2004;90:203–210. doi: 10.1111/j.1471-4159.2004.02478.x. [DOI] [PubMed] [Google Scholar]
  58. Kilberg MS, Handlogten ME, Christensen HN. Characteristics of an amino acid transport system in rat liver for glutamine, asparagine, histidine, and closely related analogs. J Biol Chem. 1980;255:4011–4019. [PubMed] [Google Scholar]
  59. Kirischuk S, Kettenmann H, Verkhratsky A. Membrane currents and cytoplasmic sodium transients generated by glutamate transport in Bergmann glial cells. Pflugers Arch. 2007;454:245–252. doi: 10.1007/s00424-007-0207-5. [DOI] [PubMed] [Google Scholar]
  60. Kopp-Scheinpflug C, Tolnai S, Malmierca MS, Rubsamen R. The medial nucleus of the trapezoid body: comparative physiology. Neuroscience. 2008;154:160–170. doi: 10.1016/j.neuroscience.2008.01.088. [DOI] [PubMed] [Google Scholar]
  61. Kvamme E, Roberg B, Torgner IA. Phosphate-activated glutaminase and mitochondrial glutamine transport in the brain. Neurochem Res. 2000;25:1407–1419. doi: 10.1023/a:1007668801570. [DOI] [PubMed] [Google Scholar]
  62. Laake JH, Slyngstad TA, Haug FM, Ottersen OP. Glutamine from glial cells is essential for the maintenance of the nerve terminal pool of glutamate: immunogold evidence from hippocampal slice cultures. J Neurochem. 1995;65:871–881. doi: 10.1046/j.1471-4159.1995.65020871.x. [DOI] [PubMed] [Google Scholar]
  63. Langer J, Rose CR. Synaptically induced sodium signals in hippocampal astrocytes in situ. J Physiol. 2009;587:5859–5877. doi: 10.1113/jphysiol.2009.182279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Langer J, Stephan J, Theis M, Rose CR. Gap junctions mediate intercellular spread of sodium between hippocampal astrocytes in situ. Glia. 2011;60:239–252. doi: 10.1002/glia.21259. [DOI] [PubMed] [Google Scholar]
  65. Leao RN, Leao RM, da Costa LF, Rock Levinson S, Walmsley B. A novel role for MNTB neuron dendrites in regulating action potential amplitude and cell excitability during repetitive firing. Eur J Neurosci. 2008;27:3095–3108. doi: 10.1111/j.1460-9568.2008.06297.x. [DOI] [PubMed] [Google Scholar]
  66. Lehre KP, Danbolt NC. The number of glutamate transporter subtype molecules at glutamatergic synapses: chemical and stereological quantification in young adult rat brain. J Neurosci. 1998;18:8751–8757. doi: 10.1523/JNEUROSCI.18-21-08751.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Liang SL, Carlson GC, Coulter DA. Dynamic regulation of synaptic GABA release by the glutamate-glutamine cycle in hippocampal area CA1. J Neurosci. 2006;26:8537–8548. doi: 10.1523/JNEUROSCI.0329-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Lieth E, LaNoue KF, Berkich DA, Xu B, Ratz M, Taylor C, Hutson SM. Nitrogen shuttling between neurons and glial cells during glutamate synthesis. J Neurochem. 2001;76:1712–1723. doi: 10.1046/j.1471-4159.2001.00156.x. [DOI] [PubMed] [Google Scholar]
  69. Mackenzie B, Schafer MK, Erickson JD, Hediger MA, Weihe E, Varoqui H. Functional properties and cellular distribution of the system A glutamine transporter SNAT1 support specialized roles in central neurons. J Biol Chem. 2003;278:23720–23730. doi: 10.1074/jbc.M212718200. [DOI] [PubMed] [Google Scholar]
  70. Marcaggi P, Billups D, Attwell D. The role of glial glutamate transporters in maintaining the independent operation of juvenile mouse cerebellar parallel fibre synapses. J Physiol. 2003;552:89–107. doi: 10.1113/jphysiol.2003.044263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Minta A, Tsien RY. Fluorescent indicators for cytosolic sodium. J Biol Chem. 1989;264:19449–19457. [PubMed] [Google Scholar]
  72. Muller J, Reyes-Haro D, Pivneva T, Nolte C, Schaette R, Lubke J, Kettenmann H. The principal neurons of the medial nucleus of the trapezoid body and NG2+ glial cells receive coordinated excitatory synaptic input. J Gen Physiol. 2009;134:115–127. doi: 10.1085/jgp.200910194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Nagaraja TN, Brookes N. Glutamine transport in mouse cerebral astrocytes. J Neurochem. 1996;66:1665–1674. doi: 10.1046/j.1471-4159.1996.66041665.x. [DOI] [PubMed] [Google Scholar]
  74. Nakanishi T, Kekuda R, Fei YJ, Hatanaka T, Sugawara M, Martindale RG, Leibach FH, Prasad PD, Ganapathy V. Cloning and functional characterization of a new subtype of the amino acid transport system N. Am J Physiol Cell Physiol. 2001;281:C1757–C1768. doi: 10.1152/ajpcell.2001.281.6.C1757. [DOI] [PubMed] [Google Scholar]
  75. Nimmerjahn A, Kirchhoff F, Kerr JN, Helmchen F. Sulforhodamine 101 as a specific marker of astroglia in the neocortex in vivo. Nat Methods. 2004;1:31–37. doi: 10.1038/nmeth706. [DOI] [PubMed] [Google Scholar]
  76. Nissen-Meyer LS, Popescu MC, Hamdani el H, Chaudhry FA. Protein kinase C-mediated phosphorylation of a single serine residue on the rat glial glutamine transporter SN1 governs its membrane trafficking. J Neurosci. 2011;31:6565–6575. doi: 10.1523/JNEUROSCI.3694-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Norenberg MD, Martinez-Hernandez A. Fine structural localization of glutamine synthetase in astrocytes of rat brain. Brain Res. 1979;161:303–310. doi: 10.1016/0006-8993(79)90071-4. [DOI] [PubMed] [Google Scholar]
  78. Otis TS, Wu YC, Trussell LO. Delayed clearance of transmitter and the role of glutamate transporters at synapses with multiple release sites. J Neurosci. 1996;16:1634–1644. doi: 10.1523/JNEUROSCI.16-05-01634.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Overstreet LS, Kinney GA, Liu YB, Billups D, Slater NT. Glutamate transporters contribute to the time course of synaptic transmission in cerebellar granule cells. J Neurosci. 1999;19:9663–9673. doi: 10.1523/JNEUROSCI.19-21-09663.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Paulsen RE, Contestabile A, Villani L, Fonnum F. The effect of fluorocitrate on transmitter amino acid release from rat striatal slices. Neurochem Res. 1988;13:637–641. doi: 10.1007/BF00973281. [DOI] [PubMed] [Google Scholar]
  81. Perea G, Araque A. Astrocytes potentiate transmitter release at single hippocampal synapses. Science. 2007;317:1083–1086. doi: 10.1126/science.1144640. [DOI] [PubMed] [Google Scholar]
  82. Perea G, Navarrete M, Araque A. Tripartite synapses: astrocytes process and control synaptic information. Trends Neurosci. 2009;32:421–431. doi: 10.1016/j.tins.2009.05.001. [DOI] [PubMed] [Google Scholar]
  83. Pow DV, Robinson SR. Glutamate in some retinal neurons is derived solely from glia. Neuroscience. 1994;60:355–366. doi: 10.1016/0306-4522(94)90249-6. [DOI] [PubMed] [Google Scholar]
  84. Price GD, Trussell LO. Estimate of the chloride concentration in a central glutamatergic terminal: a gramicidin perforated-patch study on the calyx of Held. J Neurosci. 2006;26:11432–11436. doi: 10.1523/JNEUROSCI.1660-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Renden R, Taschenberger H, Puente N, Rusakov DA, Duvoisin R, Wang LY, Lehre KP, von Gersdorff H. Glutamate transporter studies reveal the pruning of metabotropic glutamate receptors and absence of AMPA receptor desensitization at mature calyx of held synapses. J Neurosci. 2005;25:8482–8497. doi: 10.1523/JNEUROSCI.1848-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Reyes-Haro D, Muller J, Boresch M, Pivneva T, Benedetti B, Scheller A, Nolte C, Kettenmann H. Neuron-astrocyte interactions in the medial nucleus of the trapezoid body. J Gen Physiol. 2010;135:583–594. doi: 10.1085/jgp.200910354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Rothman DL, Behar KL, Hyder F, Shulman RG. In vivo NMR studies of the glutamate neurotransmitter flux and neuroenergetics: implications for brain function. Annu Rev Physiol. 2003;65:401–427. doi: 10.1146/annurev.physiol.65.092101.142131. [DOI] [PubMed] [Google Scholar]
  88. Satzler K, Sohl LF, Bollmann JH, Borst JG, Frotscher M, Sakmann B, Lubke JH. Three-dimensional reconstruction of a calyx of Held and its postsynaptic principal neuron in the medial nucleus of the trapezoid body. J Neurosci. 2002;22:10567–10579. doi: 10.1523/JNEUROSCI.22-24-10567.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Scanziani M, Salin PA, Vogt KE, Malenka RC, Nicoll RA. Use-dependent increases in glutamate concentration activate presynaptic metabotropic glutamate receptors. Nature. 1997;385:630–634. doi: 10.1038/385630a0. [DOI] [PubMed] [Google Scholar]
  90. Segawa H, Fukasawa Y, Miyamoto K, Takeda E, Endou H, Kanai Y. Identification and functional characterization of a Na+-independent neutral amino acid transporter with broad substrate selectivity. J Biol Chem. 1999;274:19745–19751. doi: 10.1074/jbc.274.28.19745. [DOI] [PubMed] [Google Scholar]
  91. Shimamoto K, Lebrun B, Yasuda-Kamatani Y, Sakaitani M, Shigeri Y, Yumoto N, Nakajima T. DL-threo-β-benzyloxyaspartate, a potent blocker of excitatory amino acid transporters. Mol Pharmacol. 1998;53:195–201. doi: 10.1124/mol.53.2.195. [DOI] [PubMed] [Google Scholar]
  92. Shimamoto K, Sakai R, Takaoka K, Yumoto N, Nakajima T, Amara SG, Shigeri Y. Characterization of novel L-threo-β-benzyloxyaspartate derivatives, potent blockers of the glutamate transporters. Mol Pharmacol. 2004;65:1008–1015. doi: 10.1124/mol.65.4.1008. [DOI] [PubMed] [Google Scholar]
  93. Sibson NR, Mason GF, Shen J, Cline GW, Herskovits AZ, Wall JE, Behar KL, Rothman DL, Shulman RG. In vivo13C NMR measurement of neurotransmitter glutamate cycling, anaplerosis and TCA cycle flux in rat brain during [2-13C] glucose infusion. J Neurochem. 2001;76:975–989. doi: 10.1046/j.1471-4159.2001.00074.x. [DOI] [PubMed] [Google Scholar]
  94. Sonnewald U, Westergaard N, Schousboe A, Svendsen JS, Unsgard G, Petersen SB. Direct demonstration by [13C]NMR spectroscopy that glutamine from astrocytes is a precursor for GABA synthesis in neurons. Neurochem Int. 1993;22:19–29. doi: 10.1016/0197-0186(93)90064-c. [DOI] [PubMed] [Google Scholar]
  95. Spacek J. Three-dimensional analysis of dendritic spines. III. Glial sheath. Anat Embryol (Berl) 1985;171:245–252. doi: 10.1007/BF00341419. [DOI] [PubMed] [Google Scholar]
  96. Su TZ, Campbell GW, Oxender DL. Glutamine transport in cerebellar granule cells in culture. Brain Res. 1997;757:69–78. doi: 10.1016/s0006-8993(97)00139-x. [DOI] [PubMed] [Google Scholar]
  97. Swanson RA, Graham SH. Fluorocitrate and fluoroacetate effects on astrocyte metabolism in vitro. Brain Res. 1994;664:94–100. doi: 10.1016/0006-8993(94)91958-5. [DOI] [PubMed] [Google Scholar]
  98. Tani H, Bandrowski AE, Parada I, Wynn M, Huguenard JR, Prince DA, Reimer RJ. Modulation of epileptiform activity by glutamine and system A transport in a model of post-traumatic epilepsy. Neurobiol Dis. 2007;25:230–238. doi: 10.1016/j.nbd.2006.08.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Tani H, Dulla CG, Huguenard JR, Reimer RJ. Glutamine is required for persistent epileptiform activity in the disinhibited neocortical brain slice. J Neurosci. 2010;30:1288–1300. doi: 10.1523/JNEUROSCI.0106-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Turecek R, Trussell LO. Presynaptic glycine receptors enhance transmitter release at a mammalian central synapse. Nature. 2001;411:587–590. doi: 10.1038/35079084. [DOI] [PubMed] [Google Scholar]
  101. van den Berg CJ, Garfinkel D. A stimulation study of brain compartments. Metabolism of glutamate and related substances in mouse brain. Biochem J. 1971;123:211–218. doi: 10.1042/bj1230211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Ventura R, Harris KM. Three-dimensional relationships between hippocampal synapses and astrocytes. J Neurosci. 1999;19:6897–6906. doi: 10.1523/JNEUROSCI.19-16-06897.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Wenzel J, Lammert G, Meyer U, Krug M. 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. 1991;560:122–131. doi: 10.1016/0006-8993(91)91222-m. [DOI] [PubMed] [Google Scholar]
  104. Witcher MR, Kirov SA, Harris KM. Plasticity of perisynaptic astroglia during synaptogenesis in the mature rat hippocampus. Glia. 2007;55:13–23. doi: 10.1002/glia.20415. [DOI] [PubMed] [Google Scholar]
  105. Yang Y, Gozen O, Watkins A, Lorenzini I, Lepore A, Gao Y, Vidensky S, Brennan J, Poulsen D, Won Park J, Li Jeon N, Robinson MB, Rothstein JD. Presynaptic regulation of astroglial excitatory neurotransmitter transporter GLT1. Neuron. 2009;61:880–894. doi: 10.1016/j.neuron.2009.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Yao D, Mackenzie B, Ming H, Varoqui H, Zhu H, Hediger MA, Erickson JD. A novel system A isoform mediating Na+/neutral amino acid cotransport. J Biol Chem. 2000;275:22790–22797. doi: 10.1074/jbc.M002965200. [DOI] [PubMed] [Google Scholar]
  107. Young AB, Snyder SH. Strychnine binding associated with glycine receptors of the central nervous system. Proc Natl Acad Sci U S A. 1973;70:2832–2836. doi: 10.1073/pnas.70.10.2832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Zerangue N, Kavanaugh MP. Flux coupling in a neuronal glutamate transporter. Nature. 1996;383:634–637. doi: 10.1038/383634a0. [DOI] [PubMed] [Google Scholar]
  109. Zhou J, Sutherland ML. Glutamate transporter cluster formation in astrocytic processes regulates glutamate uptake activity. J Neurosci. 2004;24:6301–6306. doi: 10.1523/JNEUROSCI.1404-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES