Heterogeneous distribution and utilization of ... - Semantic Scholar

4 downloads 0 Views 680KB Size Report
Feb 1, 2015 - Animal Physiology Group, Department of Biology, University of Kaiserslautern Erwin Schroedinger-Strasse 13, D-67663 .... Ventral respiratory.
Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491; © 2015 by Jonathan Stephan http://www.smartscitech.com/index.php/nt

REVIEW

Heterogeneous distribution and utilization of inhibitory neurotransmitter transporters Jonathan Stephan Animal Physiology Group, Department of Biology, University of Kaiserslautern Erwin Schroedinger-Strasse 13, D-67663 Kaiserslautern, Germany Correspondence: Jonathan Stephan E-mail: [email protected] Received: December 23, 2014 Published online: February 1, 2015

Neurotransmitter homeostasis is important for proper synaptic signal transmission. Synaptically released neurotransmitters are recycled by direct reuptake into the presynaptic terminal and/or uptake into perisynaptic astrocyte processes. In the latter case, most neurotransmitters are metabolized and shuttled back to the presynaptic terminal to restore the neurotransmitter content in that compartment. In various regions of the central nervous system (CNS), the neurotransmitters glycine and GABA are heterogeneously utilized for inhibition of neuronal activity – glycine predominantly in caudal regions, GABA predominantly in rostral ones. Accordingly, respective transporters are used in neurons and astrocytes. Both cell types utilize different members of the neurotransmitter transporter family in a region-dependent manner. While glycine transporter (GlyT) 1 is predominantly utilized in astrocytes, GlyT2 is prevailing in neurons. The GABA transporters (GAT)-1 and -3 are not distinctly attributed to neurons or astrocytes, and transporter localization depends on the CNS region. Additionally, there is a complex interplay of different neurotransmitters and respective transporters as the uptake of one neurotransmitter into neurons as well as astrocytes can trigger the release of other transmitters from both cell types. This review summarizes the heterogeneous distribution and utilization of the four inhibitory neurotransmitter transporters GlyT1, GlyT2, GAT-1, and GAT-3 in olfactory bulb, retina, cortex, hippocampus, cerebellum, brainstem, thalamus, and spinal cord. Furthermore, it elaborates possible functional consequences of the transporter interplay for neurotransmitter homeostasis. Keywords: GlyT1; GlyT2; GAT-1; GAT-3; neurons; astrocytes To cite this article: Jonathan Stephan. Heterogeneous distribution and utilization of inhibitory neurotransmitter transporters. Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491.

GABA via inhibitory neurotransmitter transporters causes an inward current [3, 5-9] and concomitant depolarization [7].

Introduction The inhibitory neurotransmitter transporters GlyT1, GlyT2, GAT-1, and GAT-3 belong to the solute carrier family 6 (Slc6). They are composed of 12 transmembrane segments and transporter-specific N- and C-terminal domains (Figure 1A). Each transporter is electrogenic and three of them share the same stoichiometry (Na+ : Cl- : Glycine/GABA): GlyT1 (Slc1A9; 2 : 1 : 1), GlyT2 (Slc1A5; 3 : 1 : 1), GAT-1 (Slc1A1; 2 : 1 : 1), and GAT-3 (Slc1A11; 2 : 1 : 1; Figure 1B) [1-4]. Thus, the uptake of glycine or

Glycine-mediated inhibition is predominant in brainstem and spinal cord, whereas GABA-mediated inhibition takes is prevalent in rostral brain regions. Nevertheless, the composition of inhibitory neurotransmitter transporters and their specificity for distinct cell types depends on the brain region [10-12]. GlyT1 and GAT-3 have long been considered to be located in astrocytes, whereas GlyT2 and GAT-3 were postulated to be utilized in neurons. However, results from in

Page 1 of 7

Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491; © 2015 by Jonathan Stephan http://www.smartscitech.com/index.php/nt

Figure 1. Topology and stoichiometry of inhibitory neurotransmitter transporters. A: Inhibitory neurotransmitter transporters exhibit 12 membrane spanning helices. N- and C-termini are located intracellularly. Differences between N- and C-termini of different transporters, which give rise to multiple regulatory domains, are neglected. B: All four inhibitory neurotransmitter transporters are secondary active and their transport activity mainly depends on the electrochemical Na+ gradient. GlyT1, GAT-1, and GAT-3 are electrogenic and exhibit the same stoichiometry (Na+ : Cl- : Glycine/GABA): 2 : 1 : 1. GlyT2 more strongly depends on the Na+ gradient (3 : 1 : 1).

situ hybridization and immunohistochemistry levels did not reveal conclusive evidence for cell type specificity. Meanwhile, the initially proposed distribution of inhibitory neurotransmitter transporters has been challenged, and a high heterogeneity regarding cell type specificity of GlyTs and GATs was frequently shown (Table 1). For a detailed analysis of cell type-specific utilization of respective transporters, different techniques are used: Whole-cell patch-clamp, immunohistochemistry, electron microscopy, or uptake/release assays using [3H]-labeled glycine or GABA. Functional analysis is done in acute tissue slices, freshly isolated cells, primary neuronal and astrocytic cultures, synaptosomes, gliosomes, and heterologous expression systems (Table 1). Properties of inhibitory neurotransmitter transporters The transport activity of the four secondary active inhibitory neurotransmitter transporters mainly depends on the Na+ gradient across the membrane (Figure 1). Taking the different stoichiometry of the transporters into account, only GlyT1, GAT-1, and GAT-3 are able to work bidirectionally under physiological conditions, whereas GlyT2 exclusively mediates glycine uptake [13, 14]. In addition to Na+, the transport requires all co-substrates. Consequently, GlyT1 activity depends on [Na+]i/o, [Cl-]i/o, and [Gly]i/o as well as the membrane potential EM [2, 5, 7]. Because of the similar stoichiometry, GAT activity similarly depends on [Na+]i/o, [Cl-]i/o, [GABA]i/o, and EM [3, 7-9, 15-19]. In comparison to neurons, astrocytes exhibit elevated [Na+]i and [Cl-]i of about 17mM and 30mM, respectively [20]. Together with [Gly]i and [GABA]i of about 2mM, respectively, astrocytes promote GlyT1 and GAT reversal under physiological conditions [14, 20, 21] . Interference of neurotransmitter transporters

inhibitory synaptic transmission and postsynaptic signal integration [22]. Transporter regulation not only depends on ion and neurotransmitter gradients [18], but can also be achieved via intracellular signaling [23, 24] through interaction with N- and C-termini of the transporters [25, 26]. Furthermore, as the reversal potential of GlyT1 and GATs is close to the resting EM of astrocytes under physiological conditions those transporters can be easily forced into reverse mode by increasing [Na+]i, [Cl-]i, [transmitter]i, or depolarization of the cell (cf. Figure 2) [17, 20]. Thus, astrocytes can directly serve tonic inhibition of neuronal network activity [27, 28].For reviews on tonic and phasic inhibition, see [29, 30]. Furthermore, as glycine is a co-agonist of NMDA receptors, it can directly modulate glutamatergic synaptic transmission. GlyT1-mediated release of glycine has been shown to alter NMDA receptor-mediated signaling [13, 31]. Inhibition of GlyT1-mediated glycine uptake was shown to facilitate NMDA receptor-mediated currents and long-term potentiation as well as axonal sprouting [32, 33]. Regarding GABA release, not only activation of ionotropic GABAA receptors is of interest, but also activation of the high-affinity metabotropic GABAB receptors. Indeed, regulation of GAT activity modulates GABAB receptor function [34]. Besides regulated neurotransmitter transporter-mediated uptake and release, the interplay of different neurotransmitters can have manifold effects. The most obvious interplay of two neurotransmitters takes place at mixed inhibitory synapses, which co-release glycine and GABA. Such synapses are present in various CNS regions, e.g. cerebellum, brainstem, and spinal cord [35-42]. Here, two mechanisms can lead to neurotransmitter release: 1) vesicle fusion and 2) transporter reversal. In presynaptic terminals, glycine and GABA concur on the binding site of the vesicular inhibitory amino acid transporter (VIAAT), which mediates the filling of vesicles with both neurotransmitters, depending on the ratio of available neurotransmitters [42] .

The regulation of neurotransmitter uptake modulates

Page 2 of 7

Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491; © 2015 by Jonathan Stephan http://www.smartscitech.com/index.php/nt Table 1. Distribution of inhibitory neurotransmitter transporters within the CNS. Region Olfactory bulb

Cell type A

Transporter / Method / Reference GlyT1 GlyT2

N

Retina

Cortex

MC

Fixed tissue, culture; IHC

[49, 50]

N

Fixed tissue; IHC, EM Fixed slices, cultures; IHC, EM

[51, 52]

A

N

Hippocampus

A

N

Cerebellum

Thalamus

Fixed slices; IHC

BG

Fixed slices, synaptosomes; IHC, EM, [3H]Glycine Fixed slices; IHC, EM, in situ hybridization Fixed slices, Synaptosomes; IHC, EM, [3H]Glycine

Acute slices, PC

[52, 53]

Ventral respiratory group(Ventrolateral medulla) Spinal cord

Acute slices; PC

[54]

[54]

Fixed slices; IHC, EM Fixed slices; IHC, EM

[55-58]

Acute slices; PC Fixed slices; IHC, EM Fixed slices; IHC, EM

Acute slices; PC

[9, 63]

Acute slices, PC (?) Fixed slices; IHC, EM

[8]

Acute slices; PC Fixed slices; IHC, EM

[68]

Acute slices; PC

[7]

[53]

[52, 64]

Fixed slices, synaptosomes; IHC, [3H]Glycine

[48, 65]

Acute slices; PC

[9, 63]

[19]

[5]

Cultures; [3H]GABA Acute slices, PC

[5]

Fixed slices; IHC

[10]

Acute slices, PC (?)

Fixed slices, synaptosoms; IHC, [3H]Glycine

Fixed slices; IHC Acute slices; PC

[7]

Acute slices; PC Fixed slices; IHC

[70]

N

N

[15]

Fixed slices; IHC

A

Fixed slices; IHC, EM Fixed slices synaptosomes; IHC, EM, [3H]GABA

Fresh isolation; IHC, PC

[52, 53, 62]

[61]

A

[15]

[48]

Synaptosomes; [3H]Glycine

N A

[53]

[47]

Fixed slices; IHC

N

A

Cultures, fixed slices; IHC

GAT-3 [47]

[52, 61]

N Brainstem Auditory BS

[48]

GAT-1 Fixed slices; IHC, in situ hybridization Fixed slices; IHC, in situ hybridization Fresh isolation; IHC, PC

[52] [52, 73, 74]

[45, 48, 61, 67]

[55-58]

[8]

Fixed slices, synaptosoms, IHC, EM Acute slices; PC

[45, 66]

Fixed slices; IHC, EM

[55]

Acute slices; PC

[7]

[68]

[55, 58-60] [58, 59]

[66]

[55]

[10, 48]

Fixed slices; IHC

[48, 69]

Fixed slices; IHC,in situ hybridization Gliosomes; IHC

[71, 72]

Fixed slices, synaptosomes; IHC, [3H]GABA

[48, 74]

[73]

Cell types: A: Astrocytes; BG: Bergmann glia; MC: Müller cells; N: Neurons; (?): Speculative results. Methods: EM: Electron microscopy; IHC: immunohistochemistry; PC: patch-clamp.

Glutamate uptake into these mixed inhibitory presynaptic terminals via neuron-specific glutamate transporters leads to production of GABA by degradation of glutamate via the glutamic acid decarboxylase (GAD) [37, 42, 43]. Consecutively, GABA can replace glycine at VIAAT, thus giving the synapse a more GABAergic phenotype, which alters the kinetics of inhibitory postsynaptic currents [37]. Besides altering the vesicular content, electrogenic glutamate uptake

(stoichiometry: 3Na+ : 1 glutamate : 1H+ against 1K+) [44] causing reduction of the electrical and Na+ gradient and synthesis of GABA - induces GAT reversal and consecutive GABA release [37, 43]. Additionally, GlyT2 activity in neurons can trigger GABA release via GAT reversal [45]. Thus, release of glycine and/or GABA depends on the distribution and amount of both neurotransmitters as well as the interplay of different transporters.

Page 3 of 7

Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491; © 2015 by Jonathan Stephan http://www.smartscitech.com/index.php/nt

Figure 2. Neurotransmitter dynamics. The scheme illustrates interference between released inhibitory and excitatory neurotransmitters. Release of glycine and GABA from inhibitory projections activates GlyRs and GABAARs at the postsynapse (1). Glycine and GABA are then taken up by GlyT1 and GATs into neighboring astrocytes (2), which causes an increase of [Gly]i, [GABA]i, [Na+]i, and [Cl-]i and a concomitant depolarization (3) due to the electrogenic nature of the transporters (cf. Figure 1). These intracellular alterations shift the reversal potential of GlyT1 causing a reversal of transport direction and glycine release at sites of low resting [gly]o (4) and may thereby modulate NMDARs at a nearby excitatory synapse (5) (cf. [13]). Excitatory terminals release glutamate activating AMPARs and NMDARs at the postsynapse (6). Glutamate is then taken up by astrocytic Glast (7), causing [Na+]i elevation and depolarization (8). Within the astrocyte, glutamate can be converted into GABA via GAD (9). GABA synthesis in combination with [Na+]i and [Cl-]i increase and concomitant depolarization (3, 8) after GlyT, GAT, and Glast activity (2, 7) can induce GAT reversal (10). Released GABA can now activate GABABRs at both inhibitory and excitatory presynaptic nerve terminals (11) and may thus modulate synaptic transmission. Direct recycling of released glycine and GABA by re-uptake, glycine shuttling via astrocytic GlyT1 and neuronal GlyT2, vesicle refilling via VIAAT, glutamate re-uptake, recycling via the glutamate-glutamine-cycle, and vesicle refilling via VGluT have been neglected for didactic reasons (cf. [13, 44]).

Conclusion Astrocytes in many brain regions simultaneously utilize GlyTs and GATs (cf. Table 1). As both neurotransmitter transporters are linked to the chemical gradient of Na+ and Cl- as well as astrocytic EM they can reciprocally reduce their activity [7]. Furthermore, uptake of glutamate into astrocytes after excitatory synaptic transmission can lead to production and release of GABA via GAT reversal (Figure 2) [43, 46]. Additionally, glycinergic synaptic transmission may trigger GABA release via GAT reversal as well (Figure 2). Independent from the possible effect of GAT activity on neuronal signaling, inhibitory glycinergic transmission will cause an increase of [Na+]i and [Cl-]i and depolarization of astrocytes, which subsequently enables glycine release via transporter reversal (Figure 2). If the release site is located near excitatory glutamatergic terminals, released glycine can act as a co-agonist on postsynaptic NMDA receptors and modulate synaptic transmission [13].

Both, the transport activity and the direction of neurotransmitter transport depend on the chemical (Na+, Cl-, glycine/GABA) and electrical gradient (EM) as well as transporter stoichiometry. Especially in astrocytes, which comprise higher [Na+]i and [Cl-]I than neurons [20], neurotransmitter transporters can reverse under physiological conditions and mediate transmitter release [14, 20]. The transporters are heterogeneously distributed within the CNS and utilized by both neurons and astrocytes (cf. Table 1) causing a brain region-dependent interplay. Conflict of interest The author declares that there are no conflicting interests. Acknowledgements Grant sponsor: Institutional funding

Page 4 of 7

Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491; © 2015 by Jonathan Stephan http://www.smartscitech.com/index.php/nt

The author thanks Dr. Eckhard Friauf for critical comments on an earlier version of the manuscript. References 1.

Roux MJ, Supplisson S. Neuronal and glial glycine transporters have different stoichiometries. Neuron 2000; 25:373-383.

2.

Aragon MC, Gimenez C, Mayor F. Stoichiometry of sodium- and chloride-coupled glycine transport in synaptic plasma membrane vesicles derived from rat brain. FEBS Lett 1987; 212:87-90.

3.

4.

5.

Karakossian MH, Spencer SR, Gomez AQ, Padilla OR, Sacher A, Loo DD, et al. Novel properties of a mouse gamma-aminobutyric acid transporter (GAT4). J MembrBiol 2005; 203:65-82. Loo DD, Eskandari S, Boorer KJ, Sarkar HK, Wright EM. Role of Cl- in electrogenic Na+-coupled cotransporters GAT1 and SGLT1. J BiolChem 2000; 275:37414-37422. Huang H, Barakat L, Wang D, Bordey A. Bergmann glial GlyT1 mediates glycine uptake and release in mouse cerebellar slices. J Physiol 2004; 560:721-736.

6.

Gonzales AL, Lee W, Spencer SR, Oropeza RA, Chapman JV, Ku JY, et al. Turnover rate of the gamma-aminobutyric acid transporter GAT1. J MembrBiol 2007; 220:33-51.

7.

Stephan J, Friauf E. Functional analysis of the inhibitory neurotransmitter transporters GlyT1, GAT-1, and GAT-3 in astrocytes of the lateral superior olive. Glia 2014; 62:1992-2003.

8.

Barakat L, Bordey A. GAT-1 and reversible GABA transport in Bergmann glia in slices. J Neurophysiol 2002; 88:1407-1419.

9.

Egawa K, Yamada J, Furukawa T, Yanagawa Y, Fukuda A. Clhomeodynamics in gap junction-coupled astrocytic networks on activation of GABAergic synapses. J Physiol 2013; 591:3901-3917.

10. Zafra F, Aragon C, Olivares L, Danbolt NC, Gimenez C, Storm-Mathisen J. Glycine transporters are differentially expressed among CNS cells. J Neurosci 1995; 15:3952-3969. 11. Zafra F, Gomeza J, Olivares L, Aragon C, Gimenez C. Regional distribution and developmental variation of the glycine transporters GLYT1 and GLYT2 in the rat CNS. Eur J Neurosci 1995; 7:1342-1352. 12. Borden LA. GABA transporter heterogeneity: pharmacology and cellular localization. NeurochemInt 1996; 29:335-356. 13. Supplisson S, Roux MJ. Why glycine transporters have different stoichiometries. FEBS Lett 2002; 529:93-101. 14. Richerson GB, Wu Y. Dynamic equilibrium of neurotransmitter transporters: not just for reuptake anymore. J Neurophysiol 2003; 90:1363-1374. 15. Biedermann B, Bringmann A, Reichenbach A. High-affinity GABA uptake in retinal glial (Muller) cells of the guinea pig: electrophysiological characterization, immunohistochemical localization, and modeling of efficiency. Glia 2002; 39:217-228. 16. Bossi E, Giovannardi S, Binda F, Forlani G, Peres A. Role of anion-cation interactions on the pre-steady-state currents of the rat Na(+)-Cl(-)-dependent GABA cotransporter rGAT1. J Physiol 2002; 541:343-350. 17. Wu Y, Wang W, Diez-Sampedro A, Richerson GB. Nonvesicular inhibitory neurotransmission via reversal of the GABA transporter

GAT-1. Neuron 2007; 56:851-865. 18. Wu Y, Wang W, Richerson GB. The transmembrane sodium gradient influences ambient GABA concentration by altering the equilibrium of GABA transporters. J Neurophysiol 2006; 96:2425-2436. 19. Bernstein EM, Quick MW. Regulation of gamma-aminobutyric acid (GABA) transporters by extracellular GABA.J BiolChem 1999; 274:889-895. 20. Kirischuk S, Parpura V, Verkhratsky A. Sodium dynamics: another key to astroglial excitability? Trends Neurosci 2012; 35:497-506. 21. Berger SJ, Carter JC, Lowry OH. The distribution of glycine, GABA, glutamate and aspartate in rabbit spinal cord, cerebellum and hippocampus.J Neurochem 1977; 28:149-158. 22. Eulenburg V, Gomeza J. Neurotransmitter transporters expressed in glial cells as regulators of synapse function. Brain Res Rev 2010; 63:103-112. 23. Cristovao-Ferreira S, Navarro G, Brugarolas M, Perez-Capote K, Vaz SH, Fattorini G, et al. A1R-A2AR heteromers coupled to Gs and G i/0 proteins modulate GABA transport into astrocytes. Purinergic Signal 2013; 9:433-449. 24. Jacob PF, Vaz SH, Ribeiro JA, Sebastiao AM. P2Y1 receptor inhibits GABA transport through a calcium signalling-dependent mechanism in rat cortical astrocytes. Glia 2014; 62:1211-1226. 25. Pramod AB, Foster J, Carvelli L, Henry LK. SLC6 transporters: structure, function, regulation, disease association and therapeutics. Mol Aspects Med 2013; 34:197-219. 26. Deken SL, Beckman ML, Boos L, Quick MW. Transport rates of GABA transporters: regulation by the N-terminal domain and syntaxin 1A. Nat Neurosci 2000; 3:998-1003. 27. Sipila ST, Spoljaric A, Virtanen MA, Hiironniemi I, Kaila K. Glycine transporter-1 controls nonsynaptic inhibitory actions of glycine receptors in the neonatal rat hippocampus. J Neurosci 2014; 34:10003-10009. 28. Zhang LH, Gong N, Fei D, Xu L, Xu TL. Glycine uptake regulates hippocampal network activity via glycine receptor-mediated tonic inhibition. Neuropsychopharmacology 2008; 33:701-711. 29. Brickley SG, Mody I. ExtrasynapticGABA(A) receptors: their function in the CNS and implications for disease. Neuron 2011; 73:23-34. 30. Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABA(A) receptors. Nat Rev Neurosci 2005; 6:215-229. 31. Harsing LG, Jr., Matyus P. Mechanisms of glycine release, which build up synaptic and extrasynaptic glycine levels: the role of synaptic and non-synaptic glycine transporters. Brain Res Bull 2013; 93:110-119. 32. Martina M, Gorfinkel Y, Halman S, Lowe JA, Periyalwar P, Schmidt CJ, et al. Glycine transporter type 1 blockade changes NMDA receptor-mediated responses and LTP in hippocampal CA1 pyramidal cells by altering extracellular glycine levels. J Physiol 2004; 557:489-500. 33. Schmitz Y, Castagna C, Mrejeru A, Lizardi-Ortiz JE, Klein Z, Lindsley CW, et al. Glycine Transporter-1 Inhibition Promotes Striatal Axon Sprouting via NMDA Receptors in Dopamine Neurons. J Neurosci 2013; 33:16778-16789.

Page 5 of 7

Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491; © 2015 by Jonathan Stephan http://www.smartscitech.com/index.php/nt 34. Beenhakker MP, Huguenard JR. Astrocytes as gatekeepers of GABAB receptor function. J Neurosci 2010; 30:15262-15276. 35. Dufour A, Tell F, Kessler JP, Baude A. Mixed GABA-glycine synapses delineate a specific topography in the nucleus tractussolitarii of adult rat. J Physiol 2010; 588:1097-1115.

52. Cubelos B, Gimenez C, Zafra F. Localization of the GLYT1 glycine transporter at glutamatergic synapses in the rat brain. Cereb Cortex 2005; 15:448-459. 53. Aroeira RI, Sebastiao AM, Valente CA. GlyT1 and GlyT2 in brain astrocytes: expression, distribution and function. Brain StructFunct 2013; 219:817-830.

36. Dumoulin A, Triller A, Dieudonne S. IPSC kinetics at identified GABAergic and mixed GABAergic and glycinergic synapses onto cerebellar Golgi cells. J Neurosci 2001; 21:6045-6057.

54. Kinney GA, Spain WJ. Synaptically evoked GABA transporter currents in neocortical glia. J Neurophysiol 2002; 88:2899-2908.

37. Ishibashi H, Yamaguchi J, Nakahata Y, Nabekura J. Dynamic regulation of glycine-GABA co-transmission at spinal inhibitory synapses by neuronal glutamate transporter. J Physiol 2013; 591:3821-3832.

55. Vitellaro-Zuccarello L, Calvaresi N, De Biasi S. Expression of GABA transporters, GAT-1 and GAT-3, in the cerebral cortex and thalamus of the rat during postnatal development. Cell Tissue Res 2003; 313:245-257.

38. Jonas P, Bischofberger J, Sandkuhler J. Corelease of two fast neurotransmitters at a central synapse. Science 1998; 281:419-424.

56. Conti F, Melone M, De Biasi S, Minelli A, Brecha NC, Ducati A. Neuronal and glial localization of GAT-1, a high-affinity gamma-aminobutyric acid plasma membrane transporter, in human cerebral cortex: with a note on its distribution in monkey cortex. J Comp Neurol 1998; 396:51-63.

39. Kotak VC, Korada S, Schwartz IR, Sanes DH. A developmental shift from GABAergic to glycinergic transmission in the central auditory system.J Neurosci 1998; 18:4646-4655. 40. Nabekura J, Katsurabayashi S, Kakazu Y, Shibata S, Matsubara A, Jinno S, et al. Developmental switch from GABA to glycine release in single central synaptic terminals. Nat Neurosci 2004; 7:17-23. 41. Rahman J, Latal AT, Besser S, Hirrlinger J, Hulsmann S. Mixed miniature postsynaptic currents resulting from co-release of glycine and GABA recorded from glycinergic neurons in the neonatal respiratory network. Eur J Neurosci 2013; 37:1229-1241. 42. Apostolides PF, Trussell LO.Rapid, activity-independent turnover of vesicular transmitter content at a mixed glycine/GABA synapse. J Neurosci 2013; 33:4768-4781.

57. Minelli A, Brecha NC, Karschin C, DeBiasi S, Conti F. GAT-1, a high-affinity GABA plasma membrane transporter, is localized to neurons and astroglia in the cerebral cortex. J Neurosci 1995; 15:7734-7746. 58. Melone M, Ciappelloni S, Conti F. Plasma membrane transporters GAT-1 and GAT-3 contribute to heterogeneity of GABAergic synapses in neocortex. Front Neuroanat 2014; 8:72. 59. Minelli A, Barbaresi P, Conti F. Postnatal development of high-affinity plasma membrane GABA transporters GAT-2 and GAT-3 in the rat cerebral cortex. Brain Res Dev Brain Res 2003; 142:7-18.

44. Danbolt NC. Glutamate uptake.ProgNeurobiol 2001; 65:1-105.

60. Minelli A, DeBiasi S, Brecha NC, Zuccarello LV, Conti F. GAT-3, a high-affinity GABA plasma membrane transporter, is localized to astrocytic processes, and it is not confined to the vicinity of GABAergic synapses in the cerebral cortex. J Neurosci 1996; 16:6255-6264.

45. Milanese M, Romei C, Usai C, Oliveri M, Raiteri L. A new function for glycine GlyT2 transporters: Stimulation of gamma-aminobutyric acid release from cerebellar nerve terminals through GAT1 transporter reversal and Ca -dependent anion channels. J Neurosci Res 2013; 92:398-408.

61. Herdon HJ, Godfrey FM, Brown AM, Coulton S, Evans JR, Cairns WJ. Pharmacological assessment of the role of the glycine transporter GlyT-1 in mediating high-affinity glycine uptake by rat cerebral cortex and cerebellum synaptosomes. Neuropharmacology 2001; 41:88-96.

46. Lee M, McGeer EG, McGeer PL. Mechanisms of GABA release from human astrocytes. Glia 2011; 59:1600-1611.

62. Fujita H, Sato K, Wen TC, Peng Y, Sakanaka M. Differential expressions of glycine transporter 1 and three glutamate transporter mRNA in the hippocampus of gerbils with transient forebrain ischemia. J Cereb Blood Flow Metab 1999; 19:604-615.

43. Heja L, Barabas P, Nyitrai G, Kekesi KA, Lasztoczi B, Toke O, et al. Glutamate uptake triggers transporter-mediated GABA release from astrocytes. PLoS One 2009; 4:e7153.

47. Nishimura M, Sato K, Mizuno M, Yoshiya I, Shimada S, Saito N, et al. Differential expression patterns of GABA transporters (GAT1-3) in the rat olfactory bulb. Brain Res Mol Brain Res 1997; 45:268-274. 48. Zeilhofer HU, Studler B, Arabadzisz D, Schweizer C, Ahmadi S, Layh B, et al. Glycinergic neurons expressing enhanced green fluorescent protein in bacterial artificial chromosome transgenic mice. J Comp Neurol 2005; 482:123-141. 49. Lee SC, Zhong YM, Yang XL. Expression of glycine receptor and transporter on bullfrog retinal Muller cells.Neurosci Lett 2005; 387:75-79. 50. Reye P, Penfold P, Pow DV. Glyt-1 expression in cultured human Muller cells and intact retinae. Glia 2001; 34:311-315. 51. Pow DV, Hendrickson AE. Expression of glycine and the glycine transporter Glyt-1 in the developing rat retina. Vis Neurosci 2000; 17:1R-9R.

63. Kersante F, Rowley SC, Pavlov I, Gutierrez-Mecinas M, Semyanov A, Reul JM, et al. A functional role for both -aminobutyric acid (GABA) transporter-1 and GABA transporter-3 in the modulation of extracellular GABA and GABAergic tonic conductances in the rat hippocampus. J Physiol 2013; 591:2429-2441. 64. Musante V, Summa M, Cunha RA, Raiteri M, Pittaluga A. Pre-synaptic glycine GlyT1 transporter--NMDA receptor interaction: relevance to NMDA autoreceptor activation in the presence of Mg2+ ions. J Neurochem 2011; 117:516-527. 65. Luccini E, Romei C, Raiteri L. Glycinergic nerve endings in hippocampus and spinal cord release glycine by different mechanisms in response to identical depolarizing stimuli. J Neurochem 2008; 105:2179-2189.

Page 6 of 7

Neurotransmitter 2015; 2: e491. doi: 10.14800/nt.491; © 2015 by Jonathan Stephan http://www.smartscitech.com/index.php/nt 66. Takayama C, Inoue Y. Developmental expression of GABA transporter-1 and 3 during formation of the GABAergic synapses in the mouse cerebellar cortex. Brain Res Dev Brain Res 2005; 158:41-49. 67. Romei C, Raiteri M, Raiteri L. GABA transporters mediate glycine release from cerebellum nerve endings: roles of Ca(2+)channels, mitochondrial Na(+)/Ca(2+) exchangers, vesicular GABA/glycine transporters and anion channels. NeurochemInt 2012; 61:133-140. 68. Pirttimaki T, Parri HR, Crunelli V. Astrocytic GABA transporter GAT-1 dysfunction in experimental absence seizures. J Physiol 2013; 591:823-833. 69. Friauf E, Aragon C, Lohrke S, Westenfelder B, Zafra F. Developmental expression of the glycine transporter GLYT2 in the auditory system of rats suggests involvement in synapse maturation. J Comp Neurol 1999; 412:17-37. 70. Szoke K, Hartel K, Grass D, Hirrlinger PG, Hirrlinger J, Hulsmann S. Glycine transporter 1 expression in the ventral

respiratory group is restricted to protoplasmic astrocytes. Brain Res 2006; 1119:182-189. 71. Schreihofer AM, Stornetta RL, Guyenet PG. Evidence for glycinergic respiratory neurons: Botzinger neurons express mRNA for glycinergic transporter 2. J Comp Neurol 1999; 407:583-597. 72. Tanaka I, Ezure K, Kondo M. Distribution of glycine transporter 2 mRNA-containing neurons in relation to glutamic acid decarboxylase mRNA-containing neurons in rat medulla. Neurosci Res 2003; 47:139-151. 73. Raiteri L, Stigliani S, Usai C, Diaspro A, Paluzzi S, Milanese M, et al. Functional expression of release-regulating glycine transporters GLYT1 on GABAergic neurons and GLYT2 on astrocytes in mouse spinal cord. NeurochemInt 2008; 52:103-112. 74. Raiteri L, Raiteri M, Bonanno G. Glycine is taken up through GLYT1 and GLYT2 transporters into mouse spinal cord axon terminals and causes vesicular and carrier-mediated release of its proposed co-transmitter GABA. J Neurochem 2001; 76:1823-1832.

Page 7 of 7