Glutamate modulation of GABA transport in ... - Wiley Online Library

2 downloads 0 Views 348KB Size Report
Yazulla, 1986; Marc, 1992; Wu, 1992; Kamermans &. Spekreijse, 1999 for ... were obtained from the Marine Biological Laboratory in Woods. Hole (MA, USA).
Journal of Physiology

J Physiol (2003), 546.3, pp. 717–731 © The Physiological Society 2002

DOI: 10.1113/jphysiol.2002.034421 www.jphysiol.org

Glutamate modulation of GABA transport in retinal horizontal cells of the skate Matthew A. Kreitzer *, Kristen A. Andersen * and Robert Paul Malchow *† Departments of *Biological Sciences and † Ophthalmology and Visual Science, University of Illinois at Chicago, Chicago, IL 60607, USA

Transport of the amino acid GABA into neurons and glia plays a key role in regulating the effects of GABA in the vertebrate retina. We have examined the modulation of GABA-elicited transport currents of retinal horizontal cells by glutamate, the likely neurotransmitter of vertebrate photoreceptors. Enzymatically isolated external horizontal cells of skate were examined using whole-cell voltage-clamp techniques. GABA (1 mM ) elicited an inward current that was completely suppressed by the GABA transport inhibitors tiagabine (10 mM) and SKF89976-A (100 mM), but was unaffected by 100 mM picrotoxin. Prior application of 100 mM glutamate significantly reduced the GABA-elicited current. Glutamate depressed the GABA dose–response curve without shifting the curve laterally or altering the voltage dependence of the current. The ionotropic glutamate receptor agonists kainate and AMPA also reduced the GABA-elicited current, and the effects of glutamate and kainate were abolished by the ionotropic glutamate receptor antagonist 6-cyano-7-nitroquinoxaline. NMDA neither elicited a current nor modified the GABA-induced current, and metabotropic glutamate analogues were also without effect. Inhibition of the GABA-elicited current by glutamate and kainate was reduced when extracellular calcium was removed and when recording pipettes contained high concentrations of the calcium chelator BAPTA. Caffeine (5 mM) and thapsigargin (2 nM), agents known to alter intracellular calcium levels, also reduced the GABAelicited current, but increases in calcium induced by depolarization alone did not. Our data suggest that glutamate regulates GABA transport in retinal horizontal cells through a calcium-dependent process, and imply a close physical relationship between calcium-permeable glutamate receptors and GABA transporters in these cells. (Resubmitted 14 October 2002; accepted after revision 8 November 2002; first published online 13 December 2002) Corresponding author R. P. Malchow: M/C 067, 840 West Taylor Street, Chicago, IL 60607, USA. Email: [email protected]

The amino acid g-aminobutyric acid (GABA) is believed to be the most widely used inhibitory neurotransmitter in the vertebrate nervous system. In the vertebrate retina, there is compelling evidence to suggest that certain classes of horizontal cells use GABA as the neurotransmitter in such processes as the establishment of the surround portion of the centre-surround receptive fields of retinal neurons (cf. Yazulla, 1986; Marc, 1992; Wu, 1992; Kamermans & Spekreijse, 1999 for review). The postsynaptic effects of this neurotransmitter are thought to be terminated primarily by the transport of GABA into the neurons and glia surrounding the release site (Iversen & Kelly, 1975). Agents that can alter the transport process thus have the potential to significantly alter the postsynaptic effects of GABA in the nervous system, and the receptive field properties of retinal cells specifically. Retinal horizontal cells have proved to be a useful model system with which to study the properties of GABA transport. The large size of catfish and skate horizontal cells in particular have greatly facilitated the ease with which the electrical currents associated with the transport of GABA can be examined. Horizontal cells from these

species have been used to characterize the ionic dependence of the transport current, its voltage dependence and its pharmacology (Malchow & Ripps, 1990; Cammack & Schwartz, 1993). The electrical currents associated with the transport process in these cells require the presence of sodium and chloride, are not affected by typical GABAreceptor blockers such as bicuculline, picrotoxin and phaclofen, and are abolished by GABA-transport blockers such as tiagabine, NO-711 and SKF 89976-A. Retinal horizontal cells receive direct input from photoreceptors, which are believed to use glutamate as their neurotransmitter (Copenhagen & Jahr, 1989; Barnstable, 1993). When dark-adapted, the photoreceptors are believed to be tonically depolarized and to release glutamate continually into the synaptic cleft; light causes a hyperpolarization of the photoreceptors and a decrease in the release of glutamate (Dowling & Ripps, 1973; Ayoub & Dorst, 1998; Ayoub et al. 1998). In the present work, we have used electrophysiological techniques to examine the effects of glutamate on the GABA-elicited current of enzymatically isolated skate horizontal cells. The electrical current induced by GABA in these cells is believed to result exclusively from the transport

Journal of Physiology

718

M. A. Kreitzer, K. A. Andersen and R. P. Malchow

of GABA into the cells (Malchow & Ripps, 1990). We found that glutamate downregulates the GABA-elicited current in skate horizontal cells. Our data implicate the activation of ionotropic glutamate receptors in this modulation and further suggest that calcium entering the cell through these channels plays a key role in this process.

METHODS The skate used for these studies (Raja erinacea and R. ocellata) were obtained from the Marine Biological Laboratory in Woods Hole (MA, USA). They were maintained for up to 3 weeks in a tank of circulating artificial seawater kept at 14 °C. Prior to enucleation of the eyes, animals were anaesthetized for 5 min by placing them in approximately 1 gallon (3.79 l) of water containing 1 g gallon_1 (~0.26 g l–1) MS 222 (tricaine), then cervically transected and double pithed. Cell dissociation Eyes were hemisected and the anterior portion containing the lens was discarded. The remaining eyecup was cut into four pieces, being careful to excise the optic disc. The pieces of eyecup were placed in a skate-modified L-15 solution containing 2 mg ml_1 papain (Calbiochem-Novabiochem, San Diego, CA, USA) and 1 mg ml_1 L-cysteine. The L-15 medium was adjusted to meet the ionic and osmotic requirements for this marine elasmobranch by adding 3.0 g NaCl, 10.5 g urea, 2.5 g glucose, 0.5 g Hepes and 1.0 ml penicillin–streptomycin solution (Sigma, P0781) to a volume of 500 ml; the pH was adjusted to 7.6 with NaOH. The retinal pieces were bubbled in the solution with air for 45 min and then rinsed eight times with skate-modified L-15 solution lacking papain or cysteine. The retinae were divided into two aliquots of 2 ml of modified L-15 and were triturated through a 5 ml glass graduated sterile pipette until the tissue broke apart. Drops of the retinal solution were placed in 40 Falcon-3001 35 mm tissue culture dishes (Becton Dickinson; Franklin Lakes, NJ, USA) each containing ~2 ml of skate-modified L-15 solution. In some instances, dishes were coated with protamine and concanavalin A to ensure rapid adhesion of the cells to the dishes. For such coatings, 2 ml of a solution containing 5 mg protamine sulphate (ml H2O)_1 was added per dish and allowed to sit overnight. Dishes were then washed three times with distilled water, and 2 ml of a solution containing 1 mg concanavalin A (ml H2O)_1 again added overnight; dishes were then washed again three times with distilled water prior to adding modified L-15. No differences were noted in coated as compared with uncoated dishes with respect to the magnitude of the GABA-elicited current, the magnitude of the glutamate-elicited current, or the degree of modulation of the GABA-elicited current by glutamate. Cells were kept for up to 3 days at 4 °C. A Nikon inverted microscope equipped with Hoffman modulation contrast optics was used to visualize the cells during the electrical recording procedure. The microscope, perfusion system, electrode manipulator and suction system were mounted on an air isolation table (to minimize vibrations), which was enclosed in a Faraday cage to reduce electrical noise. Electrophysiology Whole-cell, patch-clamp recordings (Hamill et al. 1981) were made from external horizontal cells from the skate retina. Microelectrodes with tip diameters of 2–4 mm and typical resistances of 2–4 MV were pulled from Kovar capillary tubing (Garner Glass; Claremont, CA, USA) using a P-97 Micropipette Puller (Sutter

J Physiol 546.3

Instruments) and were used without flame polishing; pipettes were bent using a micro torch so as to permit the electrodes to approach the cell from a more vertical angle. Each micropipette was filled with an intracellular solution containing (mM) 204 CsCl, 0.05 CaCl2, 0.5 EGTA, 2 MgCl2 and 10 Hepes; 100 ml of 0.5 % phenol red was added to 100 ml of the above solution, which was then adjusted to a pH of 7.6 with caesium hydroxide. The electrode was connected via a silver chloride wire to a HEKA EPC-9 patch-clamp amplifier. The return pathway went through a silver–silver chloride wire contacting the media in the culture dish. No adjustment was made for series resistance except in experiments in which the voltage dependence of GABA transport was examined (where 50 % of the series resistance was compensated); the measured series resistance was typically 15–35 MV. Series resistance was checked periodically during the course of all trials; if the value for this resistance changed by more than 50 %, results from the trial were discarded. We also conducted a separate series of experiments on six cells to examine specifically alterations in series resistance before and after a 50 s application of 20 mM kainate. Our results showed no significant alteration in series resistance under these conditions: the average series resistance was 18.3 ± 0.6 MV prior to application of kainate and 18.2 ± 0.6 MV after application. A three-dimensional remote-controlled Newport manipulator was used to guide the electrode onto the cell membrane. A giga-Ohm seal was formed by applying gentle suction. After about 30 s, additional light suction produced the whole-cell-recording configuration. Recordings were performed with cells voltage clamped at _70 mV, except in those experiments examining the voltage dependence of the GABA-induced currents. In the latter experiments, the Ringer solution was supplemented with 10 mM caesium chloride and 100 mM nimodipine (gift from Dr Simon Alford, University of Illinois at Chicago, USA) to reduce voltage-dependent contributions from the inwardly rectifying current and the calcium conductance known to be present in these cells (Malchow et al. 1990). In addition, cells were initially voltage clamped at _40 mV in these latter experiments to inactivate the potassiumdependent A current and the TTX-blockable sodium conductance also known to be present in these cells; the voltage was then jumped to +40 mV and then ramped at a rate of 0.52 V s_1 to _110 mV. Drugs were applied using a Warner Instrument three-barrel perfusion system that dispensed solution at a rate of about 2 ml min_1, and solution was removed via a suction pipette located several millimetres from the cell. Exchange of solutions was complete within 50 ms. Ringer solution consisted of (mM): 270 NaCl, 6 KCl, 1 MgCl2, 4 CaCl2, 360 urea, 1 glucose and 5 Hepes at pH 7.6. The urea that is included is common to all elasmobranch Ringer solutions, and such concentrations are typically detected in the plasma of these species, including the skate used here. Fresh solutions were prepared on the mornings of experiments. L(+)-2-amino-4-phosphonobutyrate (L-AP4) and (±)-1-aminocyclopentane-trans-1,3-dicarboxylic acid (transACPD) were purchased from Tocris (Ballwin, MO, USA). K-BAPTA was a gift from Dr Jonathan Art (University of Illinois at Chicago, USA) and NMDA was a gift from Dr John Leonard (University of Illinois at Chicago, USA). Except where stated, all other chemicals were obtained from Sigma Chemical. Student’s t tests were used to examine statistical significance, with a criterion of 0.01 being chosen as indicating significantly different distributions. Data are presented throughout the paper as the mean ± S.E.M. Parameters for dose–response relationships were determined using an iterative non-linear regression statistical

J Physiol 546.3

Glutamate modulation of GABA transport

Journal of Physiology

package incorporated in the SigmaPlot 4.0 program (Jandel Scientific, San Rafael, California, USA). Calcium fluorescence measurements Fluorescence measurements were made using the membranepermeable dye Oregon Green 488 BAPTA-1 AM (Molecular Probes, O-6807). A 50 mg aliquot of the dye was reconstituted into 20 ml of dimethyl sulphoxide (DMSO) containing 20 % Pluronic F-127 detergent (Molecular Probes). The mixture was added to 7.95 ml of Ringer solution to make a final concentration of 5 mM Oregon Green, 0.25 % DMSO and 0.05 % Pluronic. Cells were incubated in Oregon Green for 30 min at 14 °C and subsequently rinsed in Ringer solution. Cells were viewed with the aid of a Nikon Diaphot inverted microscope. Light from a mercury lamp passed through a XF22 filter set (Omega Optical), with a bandpass excitation filter having peak transmittance at 485 nm (halfbandwith ±22 nm) and an emission filter with maximum transmittance at 530 ± 30 nm. Emitted light from the viewing area seen through a w 40 long-working-distance Hoffman Modulation Contrast objective was measured using a R-3896 Hamamatsu photomultiplier tube (PMT) with the signal amplified, integrated and digitized using electronics and V-clamp software provided by Prairie Technologies (Madison, WI, USA). The software also controlled the opening and closing of a Uniblitz Vincent Associates shutter, illuminating the cell with 485 nm light for 0.1 s at intervals of every 2 s for the length of the recording. Fluorescence measurements were conducted while cells were whole-cell voltage clamped; the holding potential was maintained at _70 mV to prevent activation of voltage-gated calcium channels. Care was taken to ensure that only the field under view in the microscope was illuminated with 485 nm light, and that the field contained only a single horizontal cell.

719

as bicuculline or phaclofen, nor evoked by the GABAA receptor agonist muscimol nor the GABAB receptor agonist baclofen (Malchow & Ripps, 1990). Moreover, the current requires the presence of extracellular sodium and chloride, ions that are necessary for the electrogenic transport of GABA into the cells (Malchow & Ripps, 1990; Malchow & Andersen, 2001). Thus, the electrical current recorded here probably reflects solely the transport of GABA into the cells and not the activation of GABAA, GABAB or GABAC receptors. Finally, as reported by Malchow & Andersen (2001), the GABA-elicited currents were also stable with time; the GABA-induced current elicited some 14 min after establishment of a whole-cell recording configuration was 98% as large as the current obtained at the beginning of the experiment (n = 5). The third trace in Fig. 2A shows that prior application of glutamate decreased the magnitude of the GABA-elicited inward current. In this and most of the figures that follow, the standard protocol was to first obtain a control response to 1 mM GABA, then to superfuse 100 mM glutamate for 50 s, and then to wash away the glutamate with normal skate Ringer solution. Glutamate initiates its own inward electrical current in these cells due to the activation of ionotropic glutamate receptors (not shown). When the electrical current had returned to baseline, 1 mM GABA was again applied to the cell for 10 s and the peak inward current measured. The final trace shows the virtually full recovery of the GABA-elicited current that occurred after super-

RESULTS An example of the isolated external horizontal cells used in the present experiments is shown in Fig. 1. These cells are large, possessing cell somas extending 50–100 mm in length and have multiple thick processes typically extending 10–50 mm from the soma. The large size and characteristic morphology of these cells permitted ready and sure identification of cell type. Figure 2A shows GABA-elicited currents obtained from a single horizontal cell. In this and most of the subsequent figures, whole-cell recordings were obtained from isolated skate external horizontal cells that were voltage clamped at _70 mV. The left-most trace shows the control response induced by a 10 s application of 1 mM GABA in normal skate Ringer solution. The second trace demonstrates that this current was not altered by the presence of 100 mM picrotoxin, a blocker of GABAA and GABAC receptor-mediated currents (cf. Malchow et al. 1989; Qian & Dowling, 1993). In separate experiments (cf. Malchow & Andersen, 2001), we found that 10 mM tiagabine and 100 mM SKF 89976-A, compounds known to block the transport of GABA into cells (Larsson et al. 1988; Krogsgaard-Larsen et al. 2000), completely eliminated the GABA-elicited current in a reversible fashion. Previous work from this laboratory has also shown that the GABA-initiated current in these cells is neither reduced by other GABA receptor antagonists such

Figure 1. Photomicrograph of an isolated external horizontal cell typical of those used in the course of this study The glass electrode used to whole-cell voltage clamp the neuron can be seen in the left portion of the picture. The cell is characterized by its large size relative to other neurons in the skate retina. The scale bar is 50 mm. The photomicrograph was taken on a Zeiss Axiovert 25 inverted microscope equipped for variable relief contrast.

Journal of Physiology

720

M. A. Kreitzer, K. A. Andersen and R. P. Malchow

J Physiol 546.3

Figure 2. Effects of picrotoxin and glutamate on the GABA-elicited currents of skate horizontal cells A, electrical current from a single horizontal cell to 10 s applications of 1 mM GABA. The cell was whole-cell voltage clamped at _70 mV. The first trace shows the inward current elicited by 1 mM GABA applied in normal skate Ringer solution. The Ringer solution was then switched to one containing 100 mM picrotoxin; the second trace shows that 1 mM GABA applied in the picrotoxin-containing solution produced a current of essentially the same size. Glutamate (100 mM) mixed in the picrotoxin-containing Ringer solution was then superfused over the cell for 50 s (the glutamate-induced current is not shown in the figure) and then washed away. After return of the current to baseline, 100 mM GABA was again applied to the cell, and a substantial decrease in the GABA-elicited current was seen (third trace). The fourth trace shows the recovery of the GABA-induced current following 5 min further superfusion in the picrotoxin-containing Ringer solution. B, averaged responses to 1 mM GABA from seven cells; responses are normalized to those obtained in normal Ringer solution.

fusion with Ringer for 5 min. Figure 2B shows normalized responses obtained from seven cells, and demonstrates: firstly, that picrotoxin had no effect on the GABA-initiated current; secondly, that prior application of 100 mM glutamate for 50 s significantly decreased the response to 74 ± 2 % of control levels and thirdly, that the GABA current recovered substantially after approximately 5 min. Since picrotoxin failed to alter the GABA-elicited currents, the remainder of the experiments presented in this paper were performed without this compound in the bath.

The dose–response relationship for the GABA-elicited current obtained in normal Ringer solution and following the application of glutamate is shown in Fig. 3. Recordings from a single cell are shown in Fig. 3A, and depict the responses to 100 mM, 1 mM and 10 mM GABA before and after the application of 100 mM glutamate. The responses to all three concentrations of GABA were decreased following the glutamate application. A more complete range of responses to GABA concentrations ranging from 10 mM to 10 mM is shown in Fig. 3B. The data before and

Figure 3. Glutamate depresses the concentration–response relationship for the GABA-induced current A, responses of a single cell to 100 mM, 1 mM and 10 mM GABA applications before (control) and after application of 100 mM glutamate (Glu). Glutamate was applied for 50 s and then washed away; the second applications of GABA commenced once the electrical current elicited by glutamate had returned to baseline levels. B, concentration–response relationship for the GABA-elicited current before (continuous line, •) and after (dashed line, 8) the application of 100 mM glutamate. Curves are derived fits using eqn (1). GABA depressed the overall size of the concentration–response relationship but did not alter its shape or shift it laterally.

J Physiol 546.3

Glutamate modulation of GABA transport

after glutamate application were both well fitted to the Hill equation of the form:

Journal of Physiology

R = Rmax In/(In + sn),

(1)

where R is the response induced by a GABA concentration of [I], Rmax is the maximum response, here taken to be the current elicited by 10 mM GABA, s is the concentration producing a half-maximal response, and n is a coefficient reflecting potential co-operativity in the reaction scheme. For the present experiments, the data were well fitted with a coefficient near 1 (0.91 ± 0.09 for GABA in normal Ringer solution prior to glutamate application, 0.95 ± 0.11 after glutamate), and s values of 67 ± 8 mM in normal Ringer solution and 76 ± 11 mM following glutamate. After glutamate application, the value for Rmax declined to 76% of control values. The lack of any significant alteration in the values for s or n indicate that the effect of glutamate was to reduce the size of the GABA-elicited currents without altering the affinity of the transporter for GABA. We next investigated the dose–response relationship for glutamate-gated currents and the glutamate-induced reduction of the GABA transport current. Figure 4A shows electrical currents elicited by glutamate in one horizontal cell at concentrations of 1, 3, 10 and 100 mM. Concentrations of 1 and 3 mM glutamate did not elicit detectable currents; glutamate-induced inward currents were first clearly observable at 10 mM. Figure 4B shows the dose–response relationship for the glutamate-induced current obtained from 14 cells; each cell was exposed to four different glutamate concentrations. The dose–response curve for the glutamate-induced inward current was fitted to a Hill equation with a half-maximal concentration at 35 ± 2.3 mM and a Hill coefficient of 3.2 ± 0.98. Figure 4C shows the dose–response relationship for modulation by glutamate of the current elicited by 1 mM GABA. In these experiments, the size of the GABA-elicited current was first examined without glutamate, and that value was taken as 100 %. One concentration of glutamate was then applied for 50 s then washed away, and the GABA-elicited current was examined several seconds after the glutamateinduced current had subsided. As the graph makes clear, the concentration dependence of glutamate modulation of the GABA-elicited current matched closely with the inward current elicited by glutamate itself: a half-maximal reduction in the GABA-elicited current was induced by 32 ± 3.5 mM, and the inhibition curve was best fitted with a Hill coefficient of 3.2 ± 1.3. We next used specific pharmacological agents to distinguish whether the effects of glutamate were due to the activation of ionotropic or metabotropic glutamate receptors. We initially examined the effects of 6-cyano7-nitroquinoxaline-2,3-dione (CNQX), a kainate/AMPA glutamate receptor antagonist, on the glutamate-

721

evoked currents of horizontal cells. Figure 5A and B shows that 100 m M CNQX reversibly eliminated the large inward electrical current elicited by 100 mM glutamate. Figure 5C and D further demonstrates that 20 m M kainate, a potent activator of kainate/AMPA glutamate receptor channels, also elicited a sizeable inward electrical current in the skate horizontal cells and that this current was also completely and reversibly

Figure 4. Dose–response relationship for glutamateelicited currents and glutamate-induced inhibition of the GABA-elicited current A, responses from a single cell to 5 s applications of 1, 3 , 10 and 100 mM glutamate. B, dose–response relationship for the glutamate-induced inward currents of skate horizontal cells. Responses were obtained from 14 cells, each exposed to four different glutamate concentrations. C, dose–response relationship for the inhibition of the GABA-elicited current by glutamate. Control responses to 1 mM GABA were first obtained from a cell, and then glutamate applied at a given concentration for 50 s. The glutamate was then washed away and the cell stimulated with 1 mM GABA again once the glutamate-induced current had subsided. Points reflect average responses for 1mM, n = 7; 10 mM, n = 8; 30 mM, n = 6; 60 mM, n = 6; 100 mM, n = 11 and 300 mM glutamate, n = 7 cells.

722

M. A. Kreitzer, K. A. Andersen and R. P. Malchow

Journal of Physiology

blocked by 100 m M CNQX. We also found that 50 m M of the ionotropic glutamate receptor agonist AMPA was capable of eliciting similar large inward currents (data not shown). The ability of glutamate to reduce the GABA-elicited current was eliminated when the ionotropic glutamate receptor blocker CNQX was present in the bath. The average GABA-elicited current following a 50 s application of 100 mM glutamate in the presence of 100 mM CNQX was 106 ± 3% of responses obtained in control Ringer for seven cells. Following washout of the CNQX, 100 mM glutamate was able to reduce the GABA-elicited current in these same cells to 76 ± 2 % of control values. Figure 6 shows that application of CNQX also eliminated the ability of kainate to depress the GABA-elicited current. The top trace shows the control current elicited by 1 mM GABA with 100 mM CNQX in the bath, and the second trace the response following 50 s with 20 mM kainate and CNQX also present; no decline in the GABA-elicited current was detected under these conditions, and no change was seen after an additional 5 min (third trace). CNQX was then removed from the bath and the cell once again challenged with 20 mM kainate for 50 s. As shown in the fourth trace, kainate was now able to reduce the GABA-elicited current. The fifth trace shows the recovery of the GABA-elicited

Figure 5. Glutamate and kainate induced inward currents are eliminated by the ionotropic glutamate receptor blocker 6-cyano-7-nitroquinoxaline (CNQX) A, responses from a single cell to 100 mM glutamate in normal Ringer solution (top trace), 100 mM CNQX (middle trace) and following washout of CNQX (bottom trace). CNQX completely blocked the glutamate-elicited current. B, average responses obtained from five cells. C, responses from a single cell to 20 mM kainate (KA) applied in normal Ringer solution (top trace), Ringer solution containing 100 mM CNQX (middle trace) and following washout of the CNQX (bottom trace). D, average responses to 20 mM kainate, n = 7 cells.

J Physiol 546.3

current that took place 5 min later. Figure 6B shows averaged results obtained from seven cells and demonstrates that in the presence of 100 mM CNQX, kainate (20 mM) was unable to reduce the GABA transport current. Again, following washout of CNQX, kainate was able to reduce the GABA-initiated current to 50 ± 2 % of control values. In additional experiments (seven cells, not shown), we also observed that 50 s applications of 50 mM AMPA reduced the GABA-elicited currents by 71.3 ± 3.3 %; again, following a 5 min washout, the current induced by GABA returned to 99.1 % of control levels. We also examined whether NMDA receptors might be involved in this process, since the horizontal cells of catfish have been reported to possess NMDA-activatable glutamate receptors (O’Dell & Christensen, 1989b; Linn & Christensen, 1992). We conducted these studies in Ringer solution containing 1 mM glycine and 0 mM magnesium to create an environment conducive to the opening of NMDAglutamate receptor channels. The application of 100 mM NMDA under these conditions failed to elicit any detectable currents when cells were voltage clamped at _70, 0 or +30 mV. In addition, 100 mM D-2-amino-5-phosphonopentanoate, an NMDA-glutamate receptor antagonist, failed to reduce the current elicited by 300 mM glutamate. This suggests that NMDA-sensitive channels do not contribute to the glutamate-induced current in these cells, a finding in concert with studies demonstrating a lack of NMDA receptors in the horizontal cells of the closely related stingray (O’Dell & Christensen, 1989a). We also examined currents elicited by 1 mM GABA before and after a 50 s application of 100 mM NMDA. NMDA failed to reduce the GABA current (100 ± 1 % of control levels), while 100 mM glutamate reduced the GABA-elicited current in these same cells to 74 ± 2 % of control values. These data suggest that activation of ionotropic glutamate receptors of the kainate/AMPA subclass is sufficient to produce modulation of the GABA-elicited current. To determine whether metabotropic glutamate receptors might also be involved, we examined the effects of the metabotropic glutamate receptor agonists trans-ACPD and L-AP4. Neither of these agents was effective in reducing the GABA-elicited current. The average GABAelicited current from seven cells following a 50 s application of 100 mM ACPD was 100 ± 2 % as compared to control responses to GABA in Ringer solution; for L-AP4, the GABA-elicited current was 104 ± 3 % from an additional seven cells. Control experiments for the transACPD and L-AP4 studies demonstrated that 100 mM glutamate by itself still reduced the GABA-elicited currents to 76.4 ± 1.8 and 61.5 ± 6.2 %, respectively, in these same cells. Electrophysiological and calcium-imaging data obtained from the horizontal cells of catfish and carp indicate that the ionotropic non-NMDA glutamate receptors present in

Journal of Physiology

J Physiol 546.3

Glutamate modulation of GABA transport

723

Figure 6. CNQX eliminates the ability of the glutamate analogue kainate to reduce the GABA-elicited current A, traces from a single cell to applications of 1 mM GABA. Top trace: response in a Ringer solution containing 100 mM CNQX. Second trace: response following a 50 s application of 20 mM kainate in Ringer solution containing 100 mM CNQX. Third trace: response 5 min following washout of kainate and 100 mM CNQX. Fourth trace: response after 50 s of 20 mM kainate applied in normal Ringer solution. Fifth trace: recovery of the GABA current 5 min after kainate application. B, average responses from seven cells; currents normalized to the response initially obtained in Ringer solution containing 100 mM CNQX.

these cells are permeable to calcium (Linn & Christensen, 1992; Okada et al. 1999). This raises the possibility that calcium entering the cell through the glutamate-gated channels might be important in the modulation of the GABA transport current we have observed in the present experiments. To assess calcium influx through the

glutamate receptors present in the horizontal cells of the skate, we measured changes in calcium fluorescence using the calcium indicator dye Oregon Green, monitoring fluorescence with a photomultiplier tube while voltage clamping cells at _70 mV. Figure 7A shows changes in dye fluorescence due to extracellular application of 100 mM

Figure 7. Changes in fluorescence in cells filled with the calcium indicator dye Oregon Green A, recordings from a single cell showing that 100 mM glutamate application produced an increase in fluorescence both in normal Ringer solution (left trace) and when the solution contained 100 mM of the calcium channel blocker nifedipine. B, traces from the same cell as shown in A demonstrating a fluorescence increase upon depolarization of the cell to 0 mV for 50 s in normal Ringer solution (left trace); depolarization in the presence of 100 mM nifedipine failed to induce a significant increase in fluorescence (right trace). C, recording from another horizontal cell showing the increase in dye fluorescence induced by 20 mM kainic acid (left trace) and the block of this fluorescence change when the bath contained in addition 100 mM of the ionotropic glutamate receptor blocker CNQX. D, traces from two horizontal cells showing the typical increase in dye fluorescence induced by 20 mM kainic acid and the elimination of this increased fluorescence when kainic acid was applied in a 0 mM calcium Ringer solution.

Journal of Physiology

724

M. A. Kreitzer, K. A. Andersen and R. P. Malchow

glutamate alone for 50 s in a single cell (left trace). These changes persisted in the presence of 100 mM nifedipine, a blocker of voltage-gated calcium channels (Fig. 7A, right trace). The average DF/F value for seven cells was 0.20 ± 0.02 with glutamate present and 0.17 ± 0.01 when both glutamate and nifedipine were added. Depolarization to 0 mV also produced a robust increase in fluorescence. The increase produced by depolarization to 0 mV could be greatly reduced when the cell was superfused with 100 mM nifedipine (Fig. 7B), confirming the presence of voltagegated calcium channels. In approximately half of the cells tested, nifedipine completely inhibited changes in fluorescence, while in the remainder, a small transient response, which quickly declined to baseline while the cell was still depolarized, was noted. In seven cells tested, depolarization to 0 mV for 50 s resulted in an increase in the peak value of DF/F to 0.50 ± 0.05, while in the presence of 100 mM nifedipine, the value for peak DF/F was only 0.06 ± 0.02. The glutamate analogue kainate was also able to induce significant increases in calcium in the skate horizontal cells. The left trace of Fig. 7C shows the increase in calciuminduced fluorescence upon the application of 20 mM

Figure 8. Effect of alterations in extracellular calcium on glutamate modulation of the GABA-induced current A, traces from a single cell showing the response to GABA in Ringer solution (top trace), following a 50 s application of 100 mM glutamate (second trace), and 5 min after continuing to wash with Ringer solution (bottom trace). B, average results from eight cells, showing that glutamate reduced the GABA-elicited current to 66 % of control levels. C, traces from a single horizontal cell bathed continuously in a Ringer solution lacking any added extracellular calcium. The top trace shows the control current elicited by GABA; the second trace shows the response after 50 s of 100 mM glutamate, and the bottom trace shows the recovery of the response after 5 min. D, averaged responses from eight cells, showing that glutamate was now only able to reduce the GABA elicited current by 11 %.

J Physiol 546.3

kainate to one cell voltage clamped at _70 mV. Bathing the same cell in 100 mM CNQX completely abolished the kainate-induced increase in fluorescence (Fig. 7C, right trace). In seven cells similarly tested, kainate induced an increase in DF/F of 0.12 ± 0.02, while the change in fluorescence induced by kainate in the presence of CNQX was only 0.01 ± 0.01. Moreover, when calcium was removed from the Ringer solution and replaced with 4 mM magnesium, changes in calcium fluorescence with kainate application were abolished. Figure 7D shows traces from two different cells from the same retinal dissociation. When kainate was applied in normal Ringer solution (4 mM calcium) changes in calcium fluorescence were readily detected (left trace). However, when kainate was applied in a zero-calcium Ringer solution, no change in calcium fluorescence was observed (right trace). Removal of calcium from the extracellular Ringer solution did not greatly reduce the magnitude of glutamate-induced currents. Currents elicited by 100 mM glutamate in zerocalcium Ringer solution in six cells voltage clamped at _70 mV were 95.3 ± 4.1 % of control currents obtained from the same six cells bathed in normal extracellular Ringer solution containing 4 mM calcium. Thus, like the horizontal cells of catfish and carp, skate horizontal cells possess ionotropic glutamate receptors that permit the influx of calcium. In an attempt to see whether extracellular calcium entering the cells through these receptors was necessary to obtain modulation of the GABA transporter current by glutamate, we examined modulation when extracellular calcium was removed from the Ringer solution. Figure 8A shows the reduction of the GABA transport current by 100 mM glutamate from a single cell bathed in normal calcium, and in Fig. 8B the averaged responses from eight cells are shown; glutamate reduced the GABA transporter current to 66 ± 4 % of control levels. Figure 8C and D shows that the ability of glutamate to reduce the GABA-elicited current in cells from the same dissociation was markedly reduced when cells were bathed in a Ringer solution containing nominally 0 mM calcium. Under these conditions, 100 mM glutamate reduced the GABA-elicited current by only 11 ± 2 % of control values, significantly less than that obtained when the extracellular Ringer solution contained the normal 4 mM calcium. Figure 9 shows that extracellular calcium was similarly required in order for the glutamate analogue kainate to exert its full depressive effect. Responses from cells bathed in the normal 4 mM external calcium are shown in Fig. 9A (single cell) and B (average of seven cells); prior treatment with 20 mM kainate reduced the GABA-elicited current to 55 ± 3 % of control levels. Data in Fig. 9C and D were obtained from cells from the same dissociation bathed in nominally 0 mM calcium (single cell traces in C, averaged responses from seven cells in D). In zero-calcium Ringer

J Physiol 546.3

Glutamate modulation of GABA transport

Journal of Physiology

solution, kainate was now able to reduce the GABAelicited current by only 12 ± 1 %. To further examine the calcium dependence of the glutamate-induced reduction of GABA transport, we used intracellular solutions containing a high concentration of the calcium chelator BAPTA (10 mM). We reasoned that if increases in the intracellular calcium concentration were important in the modulation of the GABA transport current, then the higher concentration of BAPTA might be expected to reduce the extent of modulation induced by glutamate. A series of trials was performed with cells obtained from the same dissociation, with some cells examined with the typical 0.5 mM EGTA intracellular solution, and others using a solution containing 10 mM BAPTA. With 0.5 mM EGTA in the pipette solution, 100 mM glutamate reduced the GABA current to 67 ± 2 % of control values in eight cells, while in an additional seven cells, recorded with 10 mM BAPTA in the pipette, the reduction induced by prior application of glutamate was only to 82 ± 2 % of control values. Further evidence supporting a role for calcium in the modulation of the GABA-elicited current came from experiments using caffeine. This agent has been shown to promote an increase in intracellular calcium in skate horizontal cells via release of calcium from intracellular stores (Haugh-Scheidt et al. 1995). In these experiments, 1 mM GABA was first applied in normal

Figure 9. Modulation of the GABA-elicited current by kainate depends upon extracellular calcium A, responses from a single cell in normal Ringer solution to a challenge by 1 mM GABA before (top trace), immediately following a 50 s application of 20 mM kainate (middle trace) and 5 min after the application of kainate (bottom trace). B, averaged responses from seven cells showing that kainate reduced the GABA-elicited current to 55 % of control levels. C, traces from a single cell using the identical protocol in A but in a Ringer solution lacking calcium. D, averaged responses from seven cells, showing that kainate now reduced the current by only 12 %.

725

Ringer solution; then the solution was switched to one containing 1 mM GABA and 5 mM caffeine (Fig. 10). These trials were performed in two types of solutions: a normal Ringer solution and a block Ringer solution that was designed to eliminate voltage-gated conductances including contributions from voltage-gated calcium channels (see Methods); these two solutions produced similar results. With pipettes containing 0.5 mM EGTA, the addition of caffeine resulted in a reduction of the GABA current to 69 ± 4 % of control levels in seven cells (Fig. 10A and B). The caffeine-induced reduction of the GABA-transport current was significantly less when the intracellular solution contained 10 mM BAPTA: caffeine reduced the GABA-uptake current to 88 ± 1 % of control levels in eight cells (Fig. 10C). This suggests that release of intracellular calcium from intracellular stores can indeed reduce the GABA-elicited current. We also conducted experiments to examine whether prior release of intracellular calcium from internal stores would reduce the effects of calcium on the GABA-elicited current. In these experiments, cells were bathed in 20 mM ryanodine for 5–6 min prior to running a trial. In these circumstances, the application of ryanodine followed by addition of caffeine resulted in a reduction of the GABA-induced current to 82 ± 2 % of control levels, while in cells not pretreated with ryanodine, caffeine by itself reduced GABA-elicited currents to 60 ± 3 % of control levels. Finally, we note that thapsigargin, an agent that acts to increase intracellular calcium levels by blocking the activity of ATP-dependent calcium transport

Figure 10. Caffeine reduces the GABA-elicited current A, trace from a single cell showing the response to 1 mM GABA first in normal Ringer solution and then during the application of 5 mM caffeine. B, averaged responses from 10 cells using an intracellular pipette containing 0.5 mM of the calcium chelator EGTA. C, effects of caffeine on the GABA-elicited current using pipettes containing 10 mM BAPTA.

Journal of Physiology

726

M. A. Kreitzer, K. A. Andersen and R. P. Malchow

mechanisms in the endoplasmic reticulum (Treiman et al. 1998), was also able to reduce the size of the GABA-elicited current. Thapsigargin (2 nM) superfused onto the cells for 5 min depressed the GABA-elicited current to 53.9 ± 2.2 % of control values in five cells tested. Given our data showing the dependence of modulation of the GABA transport current on calcium, we anticipated that simple depolarization of the cells, with consequent calcium influx through voltage-gated calcium channels, would also promote a reduction in the GABA transport current. However, to our surprise we found that the increase in intracellular calcium induced by depolarization was not sufficient to modulate GABA transport. Figure 11 gives a sample result from a single cell loaded with Oregon Green to measure calcium levels while simultaneously examining the GABA-transport current; the top traces show the GABA-elicited currents, while the bottom traces show the changes in Oregon Green fluorescence for the same cell. We first obtained control responses to 1 mM GABA with the cell voltage clamped at _70 mV (Fig. 11A, left-most trace). The cell was then depolarized to 0 mV for 50 s to open the voltage-sensitive calcium channels; the left-most trace of Fig. 11B shows the rise in Oregon Green fluorescence that was induced. The cell was then returned to a holding potential of _70 mV and challenged with 1 mM GABA a second time. Interestingly, no reduction in

Figure 11. Simultaneous measurements of whole-cell transmembrane currents and Oregon Green fluorescence from a single isolated horizontal cell The holding potential of the cell was _70 mV. A, currents induced by 1 mM GABA in ordinary Ringer solution (left trace), following a depolarization to 0 mV for 50 s (second trace), after a 50 s application of 20 mM kainate (third trace) and 5 min after the application of the kainate (last trace). B, concurrent fluorescence measurements showing the increase in fluorescence that occurred during the 50 s, 0 mV depolarization (left trace) and that resulting from the 50 s application of kainic acid (second trace).

J Physiol 546.3

the GABA-elicited currents was found (second trace in A). Kainate (20 mM) was then applied to the cell for 50 s, producing an increase in Oregon Green fluorescence of similar magnitude (right trace of Fig. 11B). After washing away the kainate, the typical decrease in the amplitude of the GABA-elicited current was observed (third trace, Fig. 11A). The right-most trace in Fig. 11A shows the return of the GABA-induced current to control levels after a 5 min period. In eight cells so examined, GABA-elicited currents were 100.5 ± 3.9 % of control values following 50 s of depolarization, despite an increase in the DF/F of 0.40 ± 0.05 for calcium fluorescence during depolarization. GABA-elicited currents in the same eight cells following a 50 s challenge with kainate were reduced to 58.5 ± 4.8 % of control values, with DF/F values for calcium fluorescence during kainate application increasing to 0.34 ± 0.05. The activation of ionotropic glutamate channels permits the inward flux of both calcium and sodium into horizontal cells (Linn & Christensen, 1992; Okada et al. 1999). The sodium gradient across the cell membrane is a key factor in the transport of GABA (Kanner, 1994; Mager et al. 1996), and alterations in intracellular sodium might play a partial role in the modulation induced by glutamate. The following experiments, however, demonstrate that glutamate could decrease GABA-elicited currents even when no sodium was likely to flow into the cell through glutamate-gated channels. In these experiments, the GABA-elicited current was first obtained in a Ringer solution containing normal concentrations of sodium. Glutamate (100 mM) was then applied for 50 s but in a Ringer solution in which all the extracellular sodium had been replaced with N-methyl-D-glucamine (NMDG). Under these conditions, the average current elicited by

Figure 12. Removal of extracellular sodium during the application of glutamate did not alter the ability of glutamate to reduce the GABA-elicited current A, responses from a single cell showing the electrical current elicited by 1 mM GABA in normal Ringer solution (top trace). Glutamate (100 mM) in 0 mM sodium Ringer solution was then bathed over the cell for 50 s. The solution was then switched back to normal Ringer solution and the GABA-elicited current examined 15 s (middle trace) and 5 min (bottom trace) after superfusion with glutamate. B, average responses from seven cells after applying glutamate in a 0 mM sodium solution.

Journal of Physiology

J Physiol 546.3

Glutamate modulation of GABA transport

glutamate was _13.5 ± 8.1 pA (six cells). Following the glutamate application, normal Ringer solution containing 270 mM sodium was superfused over the cell and GABA currents once again elicited. In these trials, application of 100 mM glutamate in 0 mM sodium for 50 s was still able to significantly reduce the size of the GABA-elicited current, with glutamate depressing the GABA-elicited current to 64 ± 5 % of control values in seven cells (Fig. 12). In additional control experiments, exposing horizontal cells to 0 mM external sodium by itself for 50 s (that is, without the co-application of 100 mM glutamate) did not alter GABAelicited currents: GABA currents after 50 s of 0 mM external sodium alone were 102 ± 3 % of initial values in seven cells tested. These data suggest that alterations in intracellular sodium levels are not likely to be the primary mechanism by which glutamate exerts its downregulation of the GABAtransport current, and also argue against the possibility that modulation of GABA transport currents might result from reversal of sodium–calcium exchange activity. Finally, we looked to see whether glutamate altered the voltage dependence of the GABA-elicited current. The transport of GABA into cells is known to be highly voltage dependent, with hyperpolarizing potentials greatly facilitating the transport of GABA into the cells (Malchow & Ripps, 1990; Mager et al. 1996). Alteration of the voltage dependence of GABA transport could thus be an

727

important means of regulation of GABA uptake. Figure 13 shows, however, that glutamate did not alter the voltage dependence of the GABA-elicited current in skate horizontal cells. Figure 13A shows the GABA-elicited current as a function of voltage from a single horizontal cell before glutamate application (continuous line), immediately after the application of 100 mM glutamate (dotted line), and 5 min after the challenge with glutamate (dashed line). The response to GABA was depressed at all voltages examined following the application of glutamate. The ramp data were then normalized to more easily detect lateral shifts in the voltage dependence of the response; the resultant normalized curves are plotted in Fig. 13B. In this case, the response obtained at _70 mV was normalized so as to produce a maximum response (_1 on the graph) for all three curves. The curves lie directly on top of one another, indicating that glutamate did not exert its effects by altering the voltage dependence of the GABA-elicited current. Figure 13C shows the averaged current–voltage relationship for the GABA-induced current obtained from seven cells before and after the application of glutamate. Responses were measured from _90 mV to +20 mV in 10 mV increments. The close alignment of the curves before and after the addition of glutamate further supports the notion that glutamate does not influence the voltage dependence of the GABA-elicited current.

Figure 13. Voltage dependence of the GABA-elicited current is not altered by glutamate A, GABA-elicited current as a function of voltage from a single horizontal cell before glutamate application (continuous line), immediately after the application of 100 mM glutamate (dotted line) and 5 min after the challenge with glutamate (dashed line). The data were obtained by applying a voltage ramp to the cell from +20 to _90 mV at a rate of 0.52 V s_1. In all cases, the Ringer solution contained 10 mM caesium chloride and 100 mM nimodipine to reduce contributions from voltage-gated conductances. Electrical currents were first obtained in Ringer solution and then in a solution containing 1 mM GABA. The electrical currents obtained in the Ringer solution were then subtracted from those obtained in GABA; the resultant current reflects that elicited by GABA alone (continuous line). The cell was then challenged with 100 mM glutamate for 50 s and then the glutamate was washed away. Another ramp was obtained once the current of the cell had returned to baseline, followed by recordings when 1 mM GABA was present; subtraction of the second Ringer solution ramp from that obtained in GABA gave the resultant GABA current (dotted line). After a 5 min recovery, the procedure was repeated (dashed curve). B, normalized current–voltage (I–V) curves for the GABA-elicited current from the same cell. Responses were normalized to that obtained at _70 mV. No lateral shift in the curve was seen as a function of the application of glutamate. C, normalized I–V curves averaged from seven cells. Responses were measured from _90 to +20 mV in 10 mV increments.

728

M. A. Kreitzer, K. A. Andersen and R. P. Malchow

Journal of Physiology

DISCUSSION Our data demonstrate that extracellular glutamate downregulates the current elicited by extracellular GABA in skate horizontal cells. Glutamate modulation depends upon the activation of ionotropic kainate/AMPA-sensitive receptors: kainate and AMPA mimic the modulation of GABA-elicited currents, and CNQX, an antagonist of ionotropic glutamate receptors, abolishes the glutamate effect. We found no evidence for the involvement of NMDA- or metabotropic-glutamate receptors in this modulation: application of NMDA or metabotropic glutamate receptor agonists failed to modulate GABAelicited currents. The downregulation requires the presence of calcium, since removal of extracellular calcium reduced the modulation by glutamate and kainate. A role for calcium was also supported by findings that caffeine and thapsigargin were able to reduce the size of the GABAelicited current. Glutamate has previously been implicated as a modulator of several other properties of retinal horizontal cells. In addition to its well-known activation of ionotropic receptors present in the horizontal cells of a wide number of species, glutamate has been shown to modulate voltagesensitive potassium and calcium conductances in these cells (Akopian & Witkovsky, 1994; Dixon & Copenhagen, 1997; Linn & Gafka, 1999). A key difference between our present findings and those cited above is in the nature of the glutamate receptor involved. Modulation of the inward rectifier conductance by glutamate in catfish conedriven horizontal cells occurs via activation of metabotropic glutamate receptors through a cGMP-coupled mechanism (Dixon & Copenhagen, 1997). Similarly, Linn & Gafka (1999) have shown that the glutamate upregulation of voltage-gated sustained calcium currents in catfish horizontal cells appears to be driven by metabotropic glutamate receptors through the activation of diacylglycerol and inositol trisphosphate pathways (Linn, 2000). Our own data strongly suggest that the modulation of the GABA-transport current examined in the skate horizontal cells is mediated by ionotropic, rather than metabotropic, glutamate receptors. Previous calcium imaging studies in the horizontal cells of catfish, carp and rat indicate that calcium can directly permeate through the kainate/AMPA ionotropic glutamate receptors on these cells (Linn & Christensen, 1992; Okada et al. 1999; Rivera et al. 2001). Our data examining fluorescence of the calcium indicator dye Oregon Green suggest strongly that the kainate/AMPA receptors of skate horizontal cells are similarly permeable to calcium: removal of extracellular calcium abolishes the increase in fluorescence upon application of glutamate or kainate, and changes in Oregon Green fluorescence are also prevented by prior application of the ionotropic glutamate blocker CNQX. This is a striking conservation of function

J Physiol 546.3

over such a wide range of evolutionarily distant species, ranging from the relatively primitive elasmobranchs used in the present study all the way to mammals. Calcium influx through the glutamate receptors of retinal horizontal cells has also been implicated in controlling synaptic structure at the level of the outer plexiform layer (Weiler et al. 1988; Weiler & Schultz, 1993). In the retinae of teleost fish such as carp and goldfish, horizontal cells extend complex finger-like extrusions, referred to as spinules, into the synaptic complexes of cone photoreceptor synaptic pedicles (cf. Wagner & Djamgoz, 1993 for review). The extent and number of these appendages depends upon the state of light adaptation of the animal: the number of spinules is greatest when the fish is light adapted and significantly reduced when the animal is dark adapted. These spinules have been correlated with an increase in lateral inhibition and the processing of colourdependent information (Weiler & Wagner, 1984). Glutamate promotes the retraction of spinules, a finding consistent with its probable continual release in the dark and the relative paucity of spinules in dark-adapted animals. The retraction of spinules occurs upon activation of AMPA/kainate-sensitive receptors; moreover, the influx of calcium through these glutamate-activated ionotropic channels appears to be the key regulatory event (Weiler et al. 1988; Weiler & Schultz, 1993; Okada et al. 1999). In these studies, removal of extracellular calcium greatly reduced the ability of glutamate and glutamate analogues to induce the retraction of spinules, a finding reminiscent of our observation that the glutamate-induced modulation of the GABA transport current is greatly reduced when extracellular calcium is removed from the bath. Moreover, spinule retraction by glutamate occurred even when voltage-gated calcium channels were blocked with cobalt, implicating calcium flux specifically through the glutamate channels themselves as a key regulatory element. This finding is particularly interesting in the light of our own observations that simple depolarization, which permits significant entry of calcium into the cell through voltageactivated channels, does not by itself alter the GABAelicited current of skate horizontal cells. The observation that the GABA transport current is modulated by elevations in intracellular calcium through glutamate-gated receptors but not through voltage-gated calcium channels, even when of comparable magnitude (cf. Fig. 11), suggests a close spatial relationship between the ionotropic glutamate receptors and GABA transporters. Moreover, the downregulatory effects on GABA transport of caffeine, which releases calcium from intracellular stores in these cells (Haugh-Scheidt et al. 1995), implies a close relationship of transporters with, at least some, sites of intracellular calcium storage. Calcium entering the cell through glutamate receptors may cause the release in turn of additional calcium from local intracellular stores, amplifying the ability of glutamate to downregulate GABA transport.

Journal of Physiology

J Physiol 546.3

Glutamate modulation of GABA transport

Activation of glutamate-gated channels will permit the flux of sodium as well as calcium into horizontal cells, and it is possible that alterations in the concentration of internal sodium might contribute to changes in the transport of GABA into the cells. Indeed, in experiments conducted in nominally 0 mM extracellular calcium, glutamate and kainate were still able to reduce the size of the GABA-elicited current by 11 and 12 %, respectively. The following calculations suggest that relatively small changes in internal sodium concentration are indeed likely to have occurred under our experimental conditions. The average glutamateinduced current recorded from 30 cells voltage clamped at _70 mV was 2.08 nA, and glutamate was applied for 50 s. If this inward current was considered to result entirely from the influx of sodium, it would result in the flux of approximately 1.1 w 10_12 mol of sodium into the cell. As is clear from the photomicrograph shown in Fig. 1, the external horizontal cells of the skate are quite large. The total surface area of these cells can be estimated from measurements of total membrane capacitance, which in 17 cells examined in this study averaged 171 ± 12 pF. Taking a value of 1.77 mF cm_2 for the specific membrane capacitance (Newman, 1985) leads to an estimated average surface area of 9.6 w 103 mm2. The cells possess thick somas, often 15 mm or more at the thickest part, with this depth extending for a considerable portion of the cell’s entire length (Malchow et al. 1990). Taking a conservative estimate of 10 mm for the thickness of the cell to account for the decrease in thickness at the extremities of the cell’s processes, we estimate a total intracellular volume of approximately 3.6 w 10_11 l; a similar value was obtained using morphological measurements of cells examined in the present studies. A 50 s application of 100 mM glutamate would thus be expected to introduce about 30 mM of sodium into the cells. Unstimulated horizontal cells have a resting internal sodium concentration of 50 mM (Andersen & Malchow, 1998) and 270 mM sodium remains on the outside of the cell. The addition of 30 mM sodium to the cell interior will thus alter the sodium gradient across the cell membrane only modestly, changing the calculated value for the Nernst potential for sodium from an initial value of +42 to +30 mV at the temperatures used in the present experiments. This small decline in driving force might partially account for the decrease in GABA-elicited current observed in calcium-free conditions. The rough estimation above does not include consideration of Na–KATPase activity, nor does it take into account possible inhomogeneities in the location of GABA transporters or ionotropic glutamate receptors, which might be expected to cluster in select portions of the cell. Such inhomogeneities could increase the local changes in sodium and have a more significant impact on GABA transport activity and other sodium-dependent processes. However, it is also possible that the reason horizontal cells of the skate retina are so large is to minimize changes in intracellular ion concentrations during the course of their normal function.

729

Modulation of GABA uptake by metabotropic glutamate receptors has been reported in cultured hippocampal neurons and glial cells. Modulation of GABA transport in these cells can result from the activation of several different signal transduction pathways. One pathway appears to involve protein kinase C (PKC) and its activation of a syntaxin-dependent mechanism (Beckman et al. 1999). GABA transport in hippocampal neurons has also been shown to be sensitive to a tyrosine receptor kinase pathway (Law et al. 2000). It has been suggested that these signal transduction cascades ultimately modulate the transport of GABA by reducing the number of active transporters present in the plasma membrane, by reducing the rate at which the GABA transporters operate, or both (Corey et al. 1994; Quick et al. 1997; Deken et al. 2000). Our data show that glutamate induces a depression of the GABA dose–response curve in skate horizontal cells without any apparent alteration in the affinity of the transporter for GABA, since no lateral shift in the dose–response relationship is seen. These data are consistent with either a reduction in the number of functional transporters present in the cell membrane or a reduction in the rate at which each transporter can ferry GABA across the membrane. Our examination of the voltage dependence of GABA transport before and after the addition of glutamate eliminates the possibility that glutamate acts by shifting the voltage-activation curve for GABA uptake. The precise signal transduction mechanisms responsible for the calcium-dependent reduction of the GABA-elicited currents remain undetermined and are currently under active investigation. Our data demonstrate that an increase in extracellular glutamate, as would be expected to occur under darkadapted conditions or following the cessation of a light stimulus, can depress the uptake of GABA into skate horizontal cells. Reducing the uptake of GABA into the horizontal cells by this means clearly has the potential to leave more GABA in the synaptic cleft, thus potentially enhancing and prolonging the inhibitory effects of this compound in the outer plexiform layer. On the other hand, it has been suggested that the GABA transporter of retinal horizontal cells also plays a key role in extruding GABA from the horizontal cell into the synaptic cleft, by means of reverse transport (Schwartz, 1982, 1987; Yazulla & Kleinschmidt, 1983; Yazulla, 1985; Yazulla et al. 1985; Schwartz, 1987; Cammack & Schwartz, 1993; Yang et al. 1999). It is not yet known whether glutamate also modulates the efflux of GABA mediated by reverse transport, but this would be expected if glutamate decreases the number of functional transporters present in the cell membrane. In summary, our results suggest an additional mechanism whereby glutamate can regulate signal processing within the neural circuits of the vertebrate retina. In addition to activating non-selective cation channels via ionotropic

Journal of Physiology

730

M. A. Kreitzer, K. A. Andersen and R. P. Malchow

receptors and modulating voltage-gated conductances by activation of metabotropic receptors, glutamate can also modulate the transport of GABA, thus potentially regulating the extent of inhibition provided by this widely used neurotransmitter.

REFERENCES Akopian A & Witkovsky P (1994). Modulation of transient outward potassium current by GTP, calcium, and glutamate in horizontal cells of the Xenopus retina. J Neurophysiol 71, 1661–1671. Andersen KA & Malchow RP (1998). Fluorometric analysis of intracellular sodium concentrations in isolated retinal horizontal cells. Biol Bull 195, 204–205. Ayoub GS & Dorst K (1998). Imaging of glutamate release from the goldfish retinal slice. Vision Res 38, 2909–2912. Ayoub GS, Grutsis S & Simko H (1998). Imaging of endogenous neurotransmitter release. J Neurosci Methods 81, 113–119. Barnstable CJ (1993). Glutamate and GABA in retinal circuitry. Curr Opin Neurobiol 3, 520–525. Beckman ML, Bernstein EM & Quick MW (1999). Multiple G protein-coupled receptors initiate protein kinase C redistribution of GABA transporters in hippocampal neurons. J Neurosci 19, RC9. Cammack JN & Schwartz EA (1993). Ions required for the electrogenic transport of GABA by horizontal cells of the catfish retina. J Physiol 472, 81–102. Copenhagen DR & Jahr CE (1989). Release of endogenous excitatory amino acids from turtle photoreceptors. Nature 341, 536–539. Corey JL, Davidson N, Lester HA, Brecha N & Quick MW (1994). Protein kinase C modulates the activity of a cloned gammaaminobutyric acid transporter expressed in Xenopus oocytes via regulated subcellular redistribution of the transporter. J Biol Chem 269, 14759–14767. Deken SL, Beckman ML, Boos L & Quick MW (2000). Transport rates of GABA transporters: regulation by the N-terminal domain and syntaxin 1A. Nature Neurosci 3, 998–1003. Dixon DB & Copenhagen DR (1997). Metabotropic glutamate receptor-mediated suppression of an inward rectifier current is linked via a cGMP cascade. J Neurosci 17, 8945–8954. Dowling JE & Ripps H. (1973). Effect of magnesium on horizontal cell activity in the skate retina. Nature 242, 101–103. Haugh-Scheidt L, Malchow RP & Ripps H (1995). GABA transport and calcium dynamics in horizontal cells from the skate retina. J Physiol 488, 565–576. Iversen LL & Kelly JS. (1975). Uptake and metabolism of gammaaminobutyric acid by neurons and glial cells. Biochem Pharmacol 24, 933–938. Kamermans M & Spekreijse H (1999). The feedback pathway from horizontal cells to cones. A mini review with a look ahead. Vision Res 39, 2449–2468. Kanner BI (1994). Sodium-coupled neurotransmitter transport: structure, function and regulation. J Exp Biol 196, 237–249. Krogsgaard-Larsen P, Frolund B & Frydenvang K (2000). GABA uptake inhibitors. Design, molecular pharmacology and therapeutic aspects. Curr Pharm Des 6, 1193–1209. Larsson OM, Falch E, Krogsgaard-Larsen P & Schousboe A (1988). Kinetic characterization of inhibition of gamma-aminobutyric acid uptake into cultured neurons and astrocytes by 4, 4-diphenyl3-butenyl derivatives of nipecotic acid and guvacine. J Neurochem 50, 818–823.

J Physiol 546.3

Law RM, Stafford A & Quick MW (2000). Functional regulation of gamma-aminobutyric acid transporters by direct tyrosine phosphorylation. J Biol Chem 275, 23986–23991. Linn CL (2000). Second messenger pathways involved in upregulation of an L-type calcium channel. Vis Neurosci 17, 473–482. Linn CL & Gafka AC (1999). Activation of metabotropic glutamate receptors modulates the voltage-gated sustained calcium current in a teleost horizontal cell. J Neurophysiol 81, 425–434. Linn CP & Christensen BN (1992). Excitatory amino acid regulation of intracellular Ca2+ in isolated catfish cone horizontal cells measured under voltage- and concentration-clamp conditions. J Neurosci 12, 2156–2164. Mager S, Kleinberger-Doron N, Keshet GI, Davidson N, Kanner BI & Lester HA (1996). Ion binding and permeation at the GABA transporter GAT1. J Neurosci 16, 5405–5414. Malchow RP & Andersen KA (2001). GABA transporter function in the horizontal cells of the skate. Prog Brain Res 131, 267–275. Malchow RP, Qian HH & Ripps H (1989). g-Aminobutyric acid (GABA)-induced currents of skate Muller (glial) cells are mediated by neuronal-like GABAA receptors. Proc Natl Acad Sci U S A 86, 4226–4230. Malchow RP, Qian HH, Ripps H & Dowling JE (1990). Structural and functional properties of two types of horizontal cell in the skate retina. J Gen Physiol 95, 177–198. Malchow RP & Ripps H (1990). Effects of gamma-aminobutyric acid on skate retinal horizontal cells: evidence for an electrogenic uptake mechanism. Proc Natl Acad Sci U S A 87, 8945–8949. Marc RE (1992). Structural organization of GABAergic circuitry in ectotherm retinas. Prog Brain Res 90, 61–92. Newman EA (1985). Membrane physiology of retinal glial (Muller) cells. J Neurosci 5, 2225-2239. O’Dell TJ & Christensen BN (1989a). A voltage-clamp study of isolated stingray horizontal cell non-NMDA excitatory amino acid receptors. J Neurophysiol 61, 162–172. O’Dell TJ & Christensen BN (1989b). Horizontal cells isolated from catfish retina contain two types of excitatory amino acid receptors. J Neurophysiol 61, 1097–1109. Okada T, Schultz K, Geurtz W, Hatt H & Weiler R (1999). AMPApreferring receptors with high Ca2+ permeability mediate dendritic plasticity of retinal horizontal cells. Eur J Neurosci 11, 1085–1095. Qian H & Dowling JE (1993). Novel GABA responses from roddriven retinal horizontal cells. Nature 361, 162–164. Quick MW, Corey JL, Davidson N & Lester HA (1997). Second messengers, trafficking-related proteins, and amino acid residues that contribute to the functional regulation of the rat brain GABA transporter GAT1. J Neurosci 17, 2967–2979. Rivera L, Blanco R & De La Villa P (2001). Calcium-permeable glutamate receptors in horizontal cells of the mammalian retina. Visual Neurosci 18, 995–1002. Schwartz EA (1982). Calcium-independent release of GABA from isolated horizontal cells of the toad retina. J Physiol 323, 211–227. Schwartz EA (1987). Depolarization without calcium can release gamma-aminobutyric acid from a retinal neuron. Science 238, 350–355. Treiman M, Caspersen C & Christensen SB (1998). A tool coming of age: thapsigargin as an inhibitor of sarco-endoplasmic reticulum Ca(2+)-ATPases. Trends Pharmacol Sci 19, 131–135. Wagner HJ & Djamgoz MB (1993). Spinules: a case for retinal synaptic plasticity. Trends Neurosci 16, 201–206. Weiler R, Kohler K, Kirsch M & Wagner HJ (1988). Glutamate and dopamine modulate synaptic plasticity in horizontal cell dendrites of fish retina. Neurosci Lett 87, 205–209.

Journal of Physiology

J Physiol 546.3

Glutamate modulation of GABA transport

Weiler R & Schultz K (1993). Ionotropic non-N-methyl-D-aspartate agonists induce retraction of dendritic spinules from retinal horizontal cells. Proc Natl Acad Sci U S A 90, 6533–6537. Weiler R & Wagner HJ (1984). Light-dependent change of conehorizontal cell interactions in carp retina. Brain Res 298, 1–9. Wu SM (1992). Functional organization of GABAergic circuitry in ectotherm retinas. Prog Brain Res 90, 93–106. Yang XL, Gao F & Wu SM (1999). Modulation of horizontal cell function by GABA(A) and GABA(C). receptors in dark- and lightadapted tiger salamander retina. Vis Neurosci 16, 967–979. Yazulla S (1985). Evoked efflux of [3H]GABA from goldfish retina in the dark. Brain Res 325, 171–180. Yazulla S (1986). GABAergic mechanisms in the retina. Prog Retin Eye Res 5, 1–52. Yazulla S, Cunningham JR & Neal M (1985). Stimulated release of endogenous GABA and glycine from the goldfish retina. Brain Res 345, 384–388. Yazulla S & Kleinschmidt J (1983). Carrier-mediated release of GABA from retinal horizontal cells. Brain Res 263, 63–75.

731

Acknowledgements This work was supported by NIH grant EYO9411 (R.P.M.), NSF grant 009–1240 (R.P.M.), core grant NIH EY01792 and an unrestricted grant from the Research to Prevent Blindness, Inc. to the Department of Ophthalmology. We thank Drs John Leonard (UIC Dept. Biology) and Haohua Qian (UIC Dept. Ophthalmology) for their comments on the manuscript, and Mr Anthony Molina (UIC Dept. Biology) for helpful discussions and criticisms. We are grateful to Mr Michael Szulczewski and his colleagues at Prairie Technologies (Madison, WI, USA) for their help and guidance in developing and using the calcium-imaging system employed in the present experiments and for providing the engineering drawings required to fabricate a light-tight PMT housing to connect the photomultiplier tube to the side port of our microscope. We also thank Mr Marek Mori (UIC Dept. Ophthalmology) for outstanding technical support. Author’s present address K. A. Andersen: Department of Medicine, University of Chicago, 5841 S. Maryland Avenue, MC 6094, Chicago, Illinois 60637, USA.