Abstract
Embryonic rat ventral spinal cord neurons were dissociated at day 15 and grown on: (i) poly-D-lysine (PDL); (ii) a confluent monolayer of type I astrocytes; or (iii) PDL in astrocyte-conditioned medium (ACM) to examine the influence of astroglia on the regulation of GABAA receptor/Cl− channel properties.
Potentiometric oxonol dye recordings of intact cells indicated that embryonic neurons were uniformly depolarized by muscimol. The depolarizing effects disappeared in cells dissociated during the early postnatal period and recovered in culture for 24 h. Similar recordings using the calcium-imaging dye fura-2 AM revealed that GABA or muscimol triggered a sustained rise in cytosolic Ca2+ () in embryonic neurons that was dependent on extracellular Ca2+, blocked by bicuculline and nifedipine and sensitive to changes in extracellular chloride. The incidence and amplitude of the Ca2+ response decreased with time in vitro and was accelerated in neurons cultured on astrocytes compared with those on PDL.
Perforated patch-clamp recordings revealed that GABA depolarized neurons in a Cl−-dependent and bicuculline-sensitive manner. Both the resting membrane potential and the GABA equilibrium potential became more hyperpolarized with time in vitro.
Astrocytes and ACM accelerated the transformation of GABAergic potential responses from depolarizing to hyperpolarizing. The change occurred over the first 4 days in co-culture or in ACM but took more than 2 weeks in neurons cultured on PDL alone.
The intrinsic, elementary properties of GABAA receptor/Cl− channels including open time and unitary conductance changed independently of the presence of astrocytes or ACM. Mean open time of the dominant kinetic component decreased and conductance increased with time in vitro.
In sum, astrocytes accelerate the developmental change in the Cl− ion gradient extrinsic to GABAA receptor/Cl− channels, which is critical for triggering Ca2+ entry, without influencing parallel changes in the intrinsic properties of the channels.
Recent evidence indicates that amino acids like GABA and glutamate and their receptors are expressed throughout the embryonic mammalian central nervous system (CNS) where they are considered to play morphogenic roles in development (for review, see Lauder, 1993) in addition to their classic roles as fast-acting neurotransmitters. In this regard, the molecular components of a putative GABAergic signalling system, including transcripts encoding GABA-synthesizing enzymes and specific GABAA receptor subunits have been detected in the embryonic rat spinal cord along with the corresponding gene products and GABA (Behar, Schaffner, Laing, Hudson, Komoly & Barker, 1993; Ma, Saunders, Somogyi, Poulter & Barker, 1993). The depolarizing effects of GABA and muscimol, an agonist at GABAA receptor/Cl− channels, on embryonic spinal cord neurons has been well established using potentiometric dye recordings (Walton, Schaffner & Barker, 1993) as well as perforated-patch (Wang, Reichling, Kyrozis & MacDermott, 1994) and on-cell patch techniques (Serafini, Valeyev, Barker & Poulter, 1995). The depolarizing effects of GABA and muscimol are blocked by bicuculline, which selectively antagonizes the activation of GABAA receptor/Cl− channels. Ca2+-indicator dye recordings of embryonic spinal cord cells dissociated from the dorsal horn reveal that both GABA and muscimol trigger bicuculline-sensitive, extracellular Ca2+-dependent elevations in cytosolic Ca2+ () (Reichling, Kyrozis, Wang & MacDermott, 1994; Wang et al. 1994). These are blocked by nifedipine, leading to the conclusion that the sustained elevation of by GABA involves Cl−-dependent depolarization and activation of voltage-dependent Ca2+ channels. In vivo (Wu, Ziskind-Conhaim & Sweet, 1992) and in vitro (Reichling et al. 1994; Wang et al. 1994) the Cl−-dependent depolarization of spinal cord neurons and the elevation in their levels gradually disappear as the Cl− gradient across the membrane increases. The mechanisms involved in these developmental transformations in the polarity of GABAA receptor-coupled, Cl−-dependent signals have yet to be elucidated.
Spinal astrocytes expressing glial fibrillary acid protein (GFAP) proliferate and differentiate during the embryonic and postnatal period (Yang, Lieska, Shao, Kriho & Pappas, 1993) when Cl−-dependent GABAergic signals switch polarity in vivo (Wu et al. 1992). Here we have used dissociated cell-culture techniques to study the effects of GFAP+ astrocytes on the differentiation of GABAergic signal polarity and associated channel properties expressed by embryonic ventral horn neurons maintained in vitro for several weeks. The results indicate that in vitro, astrocytes accelerate the time-dependent switch in the polarity of GABAergic signals independently of developmental changes in the elementary properties of GABAA receptor/Cl− channels and these astrocyte-induced effects can be reproduced using astrocyte-conditioned medium (ACM). Some of these results have been reported in abstract form (Li, Schaffner, Walton & Barker, 1995).
METHODS
Cell culture
Astrocyte monolayer
Rat pups (3-day-old) were rapidly decapitated. Cortices, free of hippocampi and striata and cleaned of meninges, were removed and placed in 10 ml of L-15 medium with 50 U ml−1 gentamicin. The tissue was triturated through a 5 ml pipette followed by mechanical dissociation through a series of small bore needles (3 × 19 g, 3 × 22 g, 1 × 25 g) and filtered through 62 μm nylon mesh (Nitex, TETKO, Inc., Kansas City, MO, USA). The dissociated cells were centrifuged and resuspended in Dulbecco's modified Eagle's medium (DMEM) supplemented with 10 % v/v fetal calf serum (FCS) and 50 U ml−1 gentamicin and plated in 75 cm2 flasks at the equivalent of two brains per flask (all tissue culture supplies from Gibco, Grand Island, NY, USA). The medium was changed after 72 h and twice each week thereafter. When a confluent monolayer was present (after ∼1 week) the flasks were tightly capped and placed overnight on a rotary shaker at 180 r.p.m. in an incubator at 37°C. The following morning the supernatant, containing microglia, loosely adherent O-2A progenitor cells and debris, was rapidly removed. Cultures were rinsed once with DMEM and re-fed with plating medium. To eliminate any surviving neurons and O-2A progenitor cells the flasks were then subjected to complement-mediated lysis, as described by Armstrong, Dorn, Kufta, Friedman & Dubois-Dalcq (1992). Briefly, the cultures were incubated with a 1:50 dilution of A2B5 ascites in DMEM with 1 % FCS or full strength A2B5 culture supernatant for 1 h at 37°C. Cultures were rinsed twice in DMEM-1 % FCS and treated with rabbit complement diluted 1:8 in DMEM-1 % FCS for 1 h at 37°C. To reduce the amount of antibody and expose a greater surface area for antibody binding the cells could also be trypsinized off the flask and resuspended in a small (1-2 ml) volume of antibody solution since A2B5 is a trypsin-resistant surface antigen. After cytolysis, the cultures were rinsed twice in DMEM-1 % FCS and held in DMEM-5 % FCS. The cultures could be maintained in flasks for an indefinite period or trypsinized and transferred to 35 mm2 dishes precoated with 5 μg ml−1 poly-D-lysine (PDL; 53 K, Sigma, St Louis, MO, USA). When cells reached confluency in the dishes they were exposed to 10 μM cytosine arabinoside for 2 days. If flasks or dishes were to be maintained for several weeks the medium was changed to DMEM-2.5 % FCS or Minimum Essential Medium (MEM)-5 % horse serum. Cultures prepared in this way contained ≤ 95 % type-1 astrocytes as determined by GFAP, S100β, and A2B5 immunocytochemistry and morphological examination. Greater than 95 % of the cells on the monolayer were GFAP/S100β positive and A2B5 negative (data not shown).
Astrocyte-conditioned medium
ACM was prepared by the addition of MEM-5 % horse serum to flasks or dishes of ‘purified’ astrocytes for 24 h; 5 % FCS was added to ACM for incubation with newly dissociated neurons but was not added to ACM intended for neuronal cultures on or after day 4.
Neurons
Pregnant Sprague-Dawley rats were narcotized by CO2 inhalation followed by cervical dislocation. E15 embryos were removed from the uteri by Caesarean section, rapidly decapitated and placed in phosphate-buffered saline at room temperature (20-22°C). Ventral spinal cords were dissociated and placed into a solution of Earle's Balanced Salt Solution (EBSS) containing 20 U ml−1 papain (Worthington Biochemical, Freehold, NJ, USA), 0.01 % w/v DNase (Boehringer Mannheim, Indianapolis, IN, USA), 0.5 mM EGTA, and 1 mM L-cysteine for 45 min at 37°C (Huettner & Baughman, 1986) to initiate dissociation of the cells. After trituration the cells were spun at 300 g for 5 min, and resuspended in EBSS with 1 mg ml−1 bovine serum albumin (BSA; Sigma) and 1 mg ml−1 ovomucoid trypsin inhibitor (Sigma). The cell suspension was layered over 5 ml of EBSS with 10 mg ml−1 each of BSA and ovomucoid and centrifuged at 80 g for 7 min. The neurons were plated in 35 mm diameter plastic dishes (NUNC, Thomas Sci., Swedesboro, NJ, USA) that had been previously coated with high molecular weight PDL or on a confluent layer of astrocytes. Plating medium consisted of modified Eagle's medium (MEM) with 3.7 g l−1 sodium bicarbonate, 6 g l−1 glucose (Gibco), 5 % fetal calf serum and 5 % v/v horse serum. In some experiments medium conditioned by astrocytes for 24 h was used as the culture medium. Cultures were kept at 36°C in a CO2 incubator. The medium was changed twice weekly and the cells were maintained in MEM and 5 % horse serum.
Cultures used for digital videomicroscopy were in 35 mm diameter glass-bottom microwell dishes (MatTek Corp., Ashland, MA, USA).
Microelectrode recording
Cultures were removed from the incubator, medium was removed and replaced with recording saline. Recordings were made on the stage of an inverted phase microscope (Nikon). Patch-clamp recordings were carried out at room temperature either in the whole-cell configuration or with the perforated-patch technique using gramicidin. Ionic currents and voltages were measured with a patch-clamp amplifier (Axopatch 200A), band-pass filtered at 0.1 Hz to 2 kHz and monitored on a storage oscilloscope and displayed on a pen recorder. The electrical signals were recorded at 9 kHz on a four-channel VCR PCM system (Instrutech VR-100B, Great Neck, NY, USA). For fluctuation analysis of GABA-activated current responses, the currents were low-pass filtered at 0.4-250 Hz, digitized (12 bits, National Instrument LAB-PC card) at 2 kHz, stored on a 386 PC microcomputer and analysed using SPAN (software courtesy of J. Dempster, Strathclyde University, Glasgow, UK).
The patch pipettes for whole-cell recordings were filled with (mM): 145 CsCl, 5 Hepes, 2 Mg-ATP, 5 BAPTA, 5 sodium phosphocreatine, pH 7.2, and the osmolarity was adjusted to 310 mosmol l−1 with sucrose. The K2SO4 pipette solution contained (mM): 10 Hepes, 70 K2SO4, 10 KCl, 5 EGTA, 0.5 CaCl2, pH 7.2, with osmolarity at 310 mosmol l−1. Electrode resistance measured in the recording saline was between 4 and 6 MΩ. Series resistance was ∼15 MΩ and was compensated for by 80-90 %. The standard external solution contained (mM): 145 NaCl, 10 Hepes, 10 D-glucose, 5.4 KCl, 1.8 CaCl2, 0.8 MgCl2, titrated to pH 7.4 with NaOH, and osmolarity adjusted to 330 mosmol l−1 with sucrose. The low Cl− solution was made by equimolar replacement of NaCl with sodium gluconate.
For perforated-patch recordings, the patch pipette solution contained (mM): 140 potassium gluconate, 10 KCl, 10 Hepes, titrated to pH 7.2 with NaOH, and osmolarity adjusted to 320 mosmol l−1 with sucrose. Gramicidin (Sigma) was initially dissolved in dimethyl sulphoxide (DMSO, 50 mg ml−1). The pipettes were first filled with gramicidin-free pipette solution by brief immersion of the tip. The remainder of the pipette was then backfilled with the solution containing gramicidin diluted to a final concentration of 100 μg ml−1. Series resistance was ≤ 20 MΩ and was compensated for by 80-90 %.
Dishes were superperfused continuously with recording medium at a rate of ∼1.2 ml min−1. Drugs were prepared and stored at -20°C in 10 mM aliquots and were applied from a puffer pipette.
To estimate intracellular Cl− concentration ([Cl−]i) a version of the Nernst equation was used:
where 155.6 is the concentration of extracellular Cl− and EGABA is the reversal potential for GABA-induced responses.
Digital video microscopy
Oxonol dye potentiometry
Details are described in Walton et al. (1993). Briefly, a Nikon Diaphot microscope was equipped with a xenon source and epi-illumination with a rhodamine filter was set appropriate for the oxonol dye DiBaC4(5) (Molecular Probes, Eugene, OR, USA). Dye (50 nM) was added to all superfusion media which consisted of a physiological saline with and without various test substances. Two glass inserts were placed inside the culture dish to decrease the volume to approximately 0.2 ml thus ensuring a rapid change in solutions. An image intensifier coupled to a CCD video camera was attached to the video output port of the microscope and fluorescent images were captured using a frame-grabber at the rate of one per minute. The anionic oxonol dyes diffuse across the membrane and achieve a Nernstian equilibrium (Apell & Bersch, 1987). Depolarization is recorded as an increase in fluorescence. Since fluorescence is directly related to the intracellular dye concentration and altered via diffusion, the fluorescence changes are quantitatively continuous rather than exhibiting a threshold. However, due to intrinsic electronic noise, the analysis procedure accepted only changes of at least 0.1 log units as an unequivocal change in membrane potential. Manipulation of extracellular [K+] indicated that a fluorescence change in the order of 0.1 log units corresponded to a potential difference of ∼8 mV, evidence that the technique is sensitive to discrete changes in membrane potential (Walton et al. 1993).
Calcium imaging
Neurons were loaded with the calcium indicator dye fura-2 by exposure to 4 μM fura-2 AM (Molecular Probes) in standard bath solution for 30 min and then washed and maintained for 45 min for ester hydrolysis at 37°C. As with oxonol recordings, rapid solution changes were obtained by decreasing the dish volume to 0.2 ml with glass inserts. Digital video imaging fluorescence microscopy was used for measuring the fluorescence of a chosen field of cells (× 40, using a Nikon Diaphot inverted microscope), and images using excitation wavelengths of 340 and 380 nm were captured and stored. The ratio of fluorescence at the two exciting wavelengths was calculated for each pixel within a cell boundary. Calibration of the ratio in terms of Ca2+ was carried out by adding ionomycin to the dish and recording images in calcium-free medium (no Ca2+ plus 5 mM EGTA) and in normal medium (1.8 mM Ca2+). The calcium concentration was derived from:
where KD is the fura-Ca2+ binding constant (∼220 nM), R is a ratio of fluorescence at two wavelengths and Rmin and Rmax are values of R in calcium-free and normal medium, respectively. F0/F0 is the ratio of fluorescence at 380 nm in calcium-free medium versus 1.8 mM Ca2+.
Statistical tests
A two-tail t test was used to assess statistical significance in Figs 5 and 7. Differences were considered significant if P < 0.05 and have been indicated by *, while P < 0.01 are denoted by ** in these figures. The data points reflect means ± standard error of the mean. For electrophysiological studies n represents the number of neurons tested; for optical recordings n is the number of fields tested.
Figure 5. EGABA hyperpolarization in long-term culture occurs more rapidly in the presence of astrocytes or astrocyte-conditioned medium (ACM).

A, the resting membrane potentials (Vm) recorded with perforated-patch techniques become more negative over time in culture independently of the presence of astrocytes. All the shift in EGABA occurs during the first week. B, EGABA hyperpolarizes during long-term culture and the rate and extent is greater on astrocytes or in ACM relative to that recorded in neurons on PDL. C, the differences between the absolute values of EGABA and Vm are plotted, with positive values representing the depolarizing responses and the negative values representing hyperpolarizing responses. D, intracellular chloride ([Cl−]i) derived from measurements of EGABA, changes over time in culture more rapidly in neurons on astrocytes and in neurons exposed to ACM compared with changes occurring on PDL alone. The change in [Cl−]i occurs over the first week with differences becoming statistically significant at 2 days. The data points represent the means ± s.e.m. (n = 5-7 cells). *P significant at the 0.05 level, **P significant at the 0.01 level. Significant differences were present between cells on PDL versus astrocytes and ACM in B, C and D.
Figure 7. Unitary properties of GABA-activated Cl− channels change in culture independently of astrocytes.

Unitary properties of channels were estimated at different times in neurons cultured under three different experimental conditions: on PDL; on astrocytes; or in astrocyte-conditioned medium (ACM). Each data point reflects the mean ± s.e.m. of 5-7 cells. Biophysical properties were very similar in all three conditions. A, the estimated mean open time of the long-lasting component (τlong) shortens from ≈75 to ≈50 ms in 7 days, then remains stable. B, the contribution of power in this component to the fluctuating signal remains relatively constant at ≈ 80 % in all three conditions. C, the unitary conductance increases from ≈16 to ≈21 pS in 7 days, then remains stable. D, the estimated mean open time of the short-lasting component (τshort) remains constant at 3-4 ms throughout the culture period.
A Fisher exact probability test using 2 × 2 contingency tables was used to determine significance between responders and non-responders on the two surfaces in Fig. 3A;* denotes P < 0.05.
Figure 3. GABA-evoked [Ca2+]c responses disappear more rapidly in neurons co-cultured on astrocytes.

A, the percentage of neurons responding to 10 μM GABA with an increase in [Ca2+]c of at least 15 nM is plotted as a function of days in culture. The number of cells responding decreases with time in culture on both PDL (▪) and astrocytes (
) but astrocytes accelerate this change. The numbers above each column are the number of responding cells/number of cells tested. * denotes P < 0.05. B, the mean amplitude of the response evoked by 10 μM GABA decreases with time in culture on both surfaces, but the decrease is magnified in the presence of astrocytes. *P significant at the 0.05 level, **P significant at the 0.01 level.
Animal procedures
All animals procedures were done in accordance with Public Health Service Policy on Humane Care and Use of Laboratory Animals and the Guide for the Care and Use of Laboratory Animals in the USA.
RESULTS
Muscimol-induced depolarization disappears in vivo during the early postnatal period
We used oxonol dye potentiometry to record from intact ventral spinal cord cells dissociated at late embryonic and early postnatal days to assess the time course over which the GABAA agonist muscimol is able to depolarize neurons. We found that about 80 % of embryonic ventral horn cells dissociated and recorded within 2-3 h of plating on PDL (79 ± 2 %; n = 3 fields/129 cells) or after 24 h in culture (75 ± 6 %; n = 6 fields/201 cells) depolarized to 2 μM muscimol (Fig. 1A). All the responding cells exhibited processes and were TuJ1 or tetanus-toxin positive (data not shown), identifying them as neurons (Koulakoff, Bizzini & Berwald-Netter, 1983; Moody, Quigg & Frankfurter, 1989). At this age and subsequent ages it is unlikely that non-responders are dead or traumatized. These states should result in a very high level of baseline oxonol fluorescence due to leaky plasma membranes and this was not found to be the case. A similar percentage of neurons dissociated at birth or at different postnatal days depolarized to muscimol when the cells were recorded within 2-3 h of plating. However, progressively fewer neurons depolarized after recovery in culture for ∼24 h (Fig. 1A). Thus, by postnatal day 14 few neurons recovered in culture for 24 h depolarized to muscimol (4 ± 3 %; n = 7 fields/98 cells). This was not due to selective loss of muscimol-responsive neurons. When the same neuron was recorded after several hours in culture and again after 24 h, the disappearance of a depolarizing response was demonstrated in the same cell. In contrast, the percentage of neurons that depolarized to 50 μM kainate remained high at all times whether they were derived from the embryonic or postnatal period or whether they were recorded several hours or 24 h after dissociation (Fig. 1A). A similar percentage of embryonic neurons depolarizing at ∼2 and ∼24 h in response to kainate strongly suggests that the depolarizing effects are not due to the trauma of cell dissociation, but reflect the physiological state of the cells that exists in vivo. These potentiometric results recorded in vitro on ventral spinal cord neurons indicate that in vivo the depolarizing effects of muscimol and, by inference GABA, disappear during the early postnatal period. The results with acutely recorded postnatal neurons also imply that the trauma of cell dissociation leads to Cl− loaded cells whose intracellular Cl− decreases dramatically during ∼24 h recovery in vitro.
Figure 1. GABAA agonists elicit depolarization and an increase in intracellular Ca2+.

A, depolarizing responses to muscimol, examined with the potentiometric dye oxonol, disappear in postnatal dissociates recovered for 1 day in vitro (1 Div). Eighty percent of the embryonic cells depolarize to muscimol after 2-3 h in vitro and after a 1 day recovery in vitro. Progressively fewer postnatal cells depolarize to muscimol after 24 h with only rare cells from postnatal day (PN) 14 responding. The percentage of neurons that depolarize to kainate remains high at all times independent of age at dissociation or time after dissociation. Each data point represents the mean ± s.e.m. for 3-7 fields (11-54 cells per field). B-D, GABA induces Ca2+-dependent increases in [Ca2+]c in ventral spinal cord neurons dissociated at embryonic day (E) 15 and recovered overnight in culture. B, 10 μM GABA induces a sustained increase in that is completely and reversibly blocked by co-application of 100 μM bicuculline. C, the increase is completely eliminated in Ca2+-free saline and recovers within 10 min. Note that Ca2+-free saline also lowers the steady-state level in a reversible manner. D, 5 μM nifedipine blocks the increase induced by 10 μM GABA in a reversible manner. The left-hand panels illustrate representative recordings and the right-hand panels are plots of mean values and their standard errors (s.e.m.) from multiple recordings (n = 31-36 cells).
GABA triggers a sustained rise in in embryonic ventral spinal cord neurons
We recorded the effects of GABA on ventral spinal cord neurons dissociated at E15 and recovered for ∼24 h in culture on PDL. GABA triggered a rise in from about 50 nM to about 200 nM (Fig. 1B-D) in the same percentage of neurons (80 %) as depolarized to muscimol. The response to GABA was blocked almost completely and in an entirely reversible manner by co-application of bicuculline, which by itself had no effect (Fig. 1B). The steady-state level of was noticeably decreased in -free medium while the response to GABA was completely eliminated in a reversible manner (Fig. 1C). The rise to GABA was also markedly, but not completely, blocked by nifedipine in a reversible manner (Fig. 1D). Virtually identical effects were recorded regarding bicuculline and nifedipine sensitivity and dependency when muscimol was used instead of GABA (data not shown). Co-application of 5 μM ω-conotoxin GVIA, 3 μM ω-conotoxin MVIIC or 100 nM tetrodotoxin with muscimol did not affect the response (n = 4; data not shown). These results implicate bicuculline-sensitive GABAA receptor/Cl− channels, which, when activated, depolarize neurons and trigger extracellular Ca2+ entry via nifedipine-sensitive, voltage-dependent Ca2+ channels.
response is sensitive to changes in extracellular chloride
Since our results implicated Cl−-dependent depolarization in the GABA- and muscimol-evoked Ca2+ responses, we changed extracellular Cl− () and recorded responses. Virtually all neurons that exhibited responses of ∼100 nM to muscimol in normal saline responded to muscimol co-incidentally delivered with low saline with about a 50 % increase in the amplitude (∼150 nM) of the sustained signal (Fig. 2A,B and D). When neurons were allowed to equilibrate in low saline and then exposed to muscimol delivered in normal saline, most neurons (∼75 %) no longer responded (Fig. 2A and B (▪)). Those that did respond manifested markedly reduced responses (Fig. 2B (
)). Neurons were also recorded that did not respond to muscimol in normal saline but upon switching to low did express quite detectable responses indistinguishable from neuronal responses in normal saline (Fig. 2C). These results are consistent with the idea that the observed Ca2+ responses are triggered via voltage-dependent Ca2+ channels activated by the depolarization of the cell in response to opening of GABAA receptor/Cl− channels.
Figure 2. Muscimol-induced changes in [Ca2+]c are sensitive to extracellular Cl−.

Embryonic neurons were plated on poly-D-lysine (PDL), then recorded for their responses to 10 μM muscimol under control conditions and in salines with altered [Cl−]. Representative recordings are shown on the left and plots summarizing all the results from multiple recordings are depicted on the right (means ± s.e.m.). A and B, neurons that exhibited increases in of ≈100 nM in response to muscimol in normal saline (150 mM) responded to muscimol coincidentally delivered in low saline (50 mM) with ≈50 % increase in the amplitude of the sustained signal (≈150 nM). Neurons equilibrated in low saline, and then exposed to muscimol delivered in normal saline, no longer responded (A or ▪in B) or exhibited responses of markedly reduced amplitude (
, B). C, neurons were also recorded that did not respond to muscimol in normal saline but did so when exposed to muscimol delivered coincidentally in low saline. These responses were indistinguishable from those typically recorded in normal saline. D, all the cells that responded to muscimol delivered in standard saline with an increase in were defined as the starting population (100 %). The vast majority of these cells responded similarly when the saline was changed to one with low Cl−. However, when neurons were allowed to equilibrate in low saline a much smaller percentage responded with an increase in when muscimol was subsequently delivered in normal saline. In A and C, smaller ticks are drawn at 5 s intervals, taller ticks at 1 min intervals.
Astrocytes accelerate time-dependent change in GABA-induced responses occurring in vitro
We recorded responses to GABA in E15 neurons cultured on a commonly used surface (PDL) in a standard medium and compared them to responses of neurons plated on confluent cortical astrocytes in the same medium. After one day in culture, similar percentages of neurons (∼80 %) differentiating on PDL or on astrocytes exhibited responses of identical amplitude (Fig. 3A). At 4 days, both the percentage of cells exhibiting GABA-induced responses and their mean amplitude had decreased considerably more in neurons differentiating on astrocytes compared with neurons growing on PDL. At 1 week, more than 60 % of neurons cultured on PDL still exhibited responses to GABA. In those neurons that responded the mean elevation of was greater than 90 nM. Less than 20 % of neurons on astrocytes expressed signals and in those neurons the mean response was less than 60 nM (Fig. 3A and B). By 2 weeks in culture, neurons on astrocytes did not respond to GABA with an elevation in while one-third of those grown on PDL still manifested signals with a mean about 60 nM. Thus, relative to cultivation of ventral spinal neurons on PDL, co-culture of neurons on astrocytes accelerated the developmental disappearance in the cellular distribution of GABA-induced responses.
Astrocytes enhance developmental changes in EGABA that involve diffusible factors
We used gramicidin-perforated patch recordings to record the resting membrane potentials and equilibrium potential for GABA-induced conductance and polarization (EGABA). GABA depolarized neurons cultured for 1 day on either PDL or astrocytes in a bicuculline-sensitive manner (n = 7) (Fig. 4A). Bicuculline by itself did not alter resting membrane potential. Depolarizing responses evoked under these experimental conditions usually did not provoke regenerative action potential activity (Fig. 4A) except in several neurons cultured on PDL for 1-2 weeks (data not shown). We recorded brief pulses to GABA under voltage clamp in normal saline and low -containing saline. In the cell illustrated (Fig. 4B and C), GABA-induced current reversed polarity at approximately -38 mV in normal saline and -2 mV in low , consistent with a Nernstian shift in EGABA. Similar results were obtained in six out of six cells tested.
Figure 4. GABA depolarizes cells in gramicidin-perforated patch recordings via activation of GABAA receptor/Cl− channels.

A, a depolarizing response to GABA was recorded under current clamp (I = 0) and could be reversibly blocked by 100 μM bicuculline. Note that neither depolarizing response triggers an action potential. B, the reversal in polarity of superimposed current responses activated by 10 μM GABA (bar at bottom) occurs at different potentials in the two salines. C, current-voltage curves constructed from the data shown in B(same cell) reveal the Cl−dependency of the reversal potential in current polarity.
We recorded steady-state membrane potential and GABA-induced polarization of neurons cultured on PDL or on astrocytes using gramicidin-perforated patch-recording techniques. Resting membrane potential recorded in this manner progressively increased from an initial level of about -40 mV within hours of plating to about -55 mV at 1 week in vitro in all neurons studied (Fig. 5A; PDL, -40 ± 0.6 to -56 ± 1.0 mV; astrocytes, -42 ± 1.6 to -55 ± 1.6 mV). EGABA was about -30 mV at 1 day in culture in all neurons recorded (Fig. 5B). Astrocytes accelerated the time course of changes in EGABA, which hyperpolarized to -60 ± 2.5 mV by 1 week in culture, eventually reaching -64 ± 1.0 mV by 3 weeks. In contrast, EGABA recorded in neurons on PDL, was between -48 ± 0.8 mV by 1 week, eventually reaching -53 ± 1.8 mV by 3 weeks (Fig. 5B). Medium conditioned by astrocytes (ACM) completely mimicked the initial rate of change in EGABA recorded in co-cultures of neurons with astrocytes (Fig. 5B). However, the effectiveness of ACM in promoting hyperpolarization became limiting such that EGABA levelled off at -58 ± 1.4 mV, a value intermediate between those recorded in the two other conditions. When the data were plotted relative to the resting potential, both astrocytes and ACM stimulated changes in EGABA that, by 4 days in vitro, led to no potential response or hyperpolarization relative to the resting potential. Hyperpolarizing responses took 3 weeks to manifest in neurons on PDL alone (Fig. 5C). Intracellular chloride, [Cl−]i (estimated from EGABA), was calculated to change from ∼45-50 mM at the time of plating to ∼12-15 mM after 4-7 days in culture in neurons on astrocytes or in ACM. A similar decrease in [Cl−]i was evident in neurons on PDL but took 3 weeks to drop to ∼20 mM (Fig. 5D).
GABAA receptor/Cl− channel properties change in neurons independently of astrocytes
We used the whole-cell patch-clamp recording technique with Cl−-filled pipettes to set ECl near or at 0 mV and clamped cells at -80 mV to optimize 10 μM GABA-induced Cl−-dependent current signals. Macroscopic GABA-induced currents were always superimposed with microscopic fluctuations characteristic of underlying Cl− channel activity (Fig. 6A and B). We used fluctuation analysis methods to estimate the elementary properties of the Cl− channels activated by GABA in the whole cell (Neher & Stevens, 1977). At the recording bandwidth used under our experimental conditions we consistently calculated two components in every spectrum of current fluctuations in all neurons (Figs 6 and 7). We computed the relative areas of the two components and found that the long-lasting component accounted for about 80-85 % of the spectrum in all neurons at all time points (Fig. 7B). At day 1 for example, PDL = 82 ± 3.5 %, astrocytes = 79 ± 4.3 %, ACM = 79 ± 1.7 %; and by day 21, PDL = 77 ± 3.0 %, astrocytes = 81 ± 2.1 %, ACM = 80 ± 2.0 %; n = 6-8 cells. Since the estimated mean channel open time, or burst duration, of the greater component (τlong) was considerably longer than the shorter component at all times in culture, more than 95 % of the pharmacologically induced current reflects Cl− channel activity with these kinetics. τlong decreased progressively in all neurons from ∼75 ms in acutely recorded cells to ∼50 ms in cells cultured for 3 weeks (Fig. 7A). In contrast, a short-lasting component averaging 3-4 ms was recorded in all neurons at all times (Fig. 7D). The estimated elementary conductance progressively increased from ∼16 to ∼22 pS over the first week in culture in all neurons recorded (Fig. 7C). These results demonstrate that unitary GABAA receptor/Cl− channel properties inferred from analysis of fluctuations in GABA-evoked Cl− current recorded from whole cells change in a concerted manner primarily during the first week in vitro and that these changes occur independently of astrocytes.
Figure 6. The long-lasting component of GABA-activated Cl− channel activity shortens over time in culture.

Cells were recorded within hours of plating (0 Div) or after 7 days in culture (7 Div). At both times GABA macroscopic currents superimposed with microscopic fluctuations. Representative traces reveal the presence of low frequency fluctuations at 0 Div (A), which are less apparent at 7 Div (B). Spectral analysis of the fluctuations show that both spectra can be adequately explained with two components indicated by the corner frequencies, fc1 and fc2. The mean open times (τ) of the underlying channel activity estimated from τ = (2πfc)−1 shorten from ≈70 ms to ≈50 ms for the long-lasting openings, but remain at about 3 ms for the short-lasting component.
DISCUSSION
The present study shows that: (1) pharmacological activation of GABAA receptor/Cl− channels depolarizes cultured ventral spinal cord neurons and triggers sustained elevation of ; (2) both responses disappear in long-term culture, with GABAA receptor/Cl− channel activation eventually hyperpolarizing cells relative to the resting potential; (3) astrocytes in co-culture and medium conditioned by them for 24 h accelerate these changes relative to the rate recorded in neurons cultured on PDL; and (4) the unitary properties of GABAA receptor/Cl− channels change independently of astrocytes, as channels shorten in open time and increase in conductance. Thus, in vitro, astrocytes regulate changes in the Cl− gradient extrinsic to GABAA receptor/Cl− channels without modulating their intrinsic properties, which change in parallel.
GABA depolarizes embryonic neurons
Depolarization via GABAA receptor/Cl− channels has been recorded in cells dissociated from the cervical region of the embryonic rat spinal cord using dye potentiometry (Walton et al. 1993). Depolarizing responses have also been recorded electrically using perforated-patch techniques in the majority of dorsal horn neurons plated at E15-16 and studied during the first week in culture (Wang et al. 1994). Embryonic, rat lumbar motoneurons in slices of spinal cord prepared from E16 and then recorded at PN2 with high resistance microelectrodes also depolarized in response to GABA (Wu et al. 1992). In the latter preparation, depolarizing synaptic potentials mediated by GABA can be evoked in motoneurons by electrical stimulation of primary afferents in dorsal roots. These depolarizing GABAergic potentials did not trigger action potentials. This may be due to the peak of the depolarizing potential lying subthreshold to action potential activation as well as to the associated conductance increase to Cl− ions. Depolarizations evoked pharmacologically in perforated-patch recordings of cultured dorsal horn neurons (Wang et al. 1994) or ventral cervical cord neurons (this study) only rarely triggered action potentials. However, in on-cell patch-clamp recordings of spinal neurons, GABA's activation of GABAA receptor/Cl− channels enclosed in the patch triggered action potentials and promoted their occurrence in the remainder of the membrane (Serafini et al. 1995). These excitatory effects of single-channel openings may result from charge transfer throughout the intact membrane that effectively depolarizes the cell outside the patch. Thus, differences in experimental conditions and/or electrical recording configurations may account for the variability in recording action potentials.
The Cl−-dependent depolarization of embryonic and early postnatal spinal cord neurons, and presumably embryonic neurons from brain as well, is due to high intracellular [Cl−]i and disappears with age as the Cl− gradient across the membrane increases. In spinal motoneurons GABA generates membrane depolarization until at least 1-2 days after birth but it has not been determined when the response becomes hyperpolarizing (Wu et al. 1992). CA3 and CA1 neurons of the hippocampus show the transition from Cl−-dependent depolarizations to hyperpolarizations at the beginning of the second postnatal week (Ben-Ari, Cherubini, Corradetti & Gaiarsa, 1989). In the present study the number of ventral spinal cord neurons exhibiting depolarizations to muscimol (after a 24 h recovery) dropped from ∼80 % at E17 to ∼40 % at PN0/PN7 and to 4 % at PN14. Neurons dissociated at E15 and grown on astrocytes or PDL exhibit the transition to GABA-induced hyperpolarizations after 4 days in vitro or 3 weeks in vitro, respectively. These results indicate that the transition occurring in vitro follows roughly the same time course as has been seen for spinal and hippocampal neurons in vivo.
Cl−-dependent depolarization activates Ca2+ entry
Previous studies on cultured embryonic dorsal horn neurons have revealed that GABA evokes a sustained elevation in , which has been attributed to voltage-dependent Ca2+ channels activated by depolarization (Reichling et al. 1994). In this study, we have found similar effects of GABAA receptor/Cl− channel activation on levels in embryonic ventral cervical spinal cord neurons. The sensitivity of the response to nifedipine in the present study implicates voltage-dependent L-type Ca2+ channels as the primary pathway for Ca2+ entry. Thus, GABA induces responses indirectly via its Cl−-dependent depolarizing actions. Quite similar conclusions regarding L-type Ca2+ channel involvement in GABA-evoked responses have been reported in cultured embryonic hypothalamic neurons (Obrietan & van den Pol, 1995). Perforated-patch recordings of hypothalamic neurons cultured for several days show that in about half of the cells GABA evoked depolarizing responses that triggered one or several action potentials. Bicuculline blocked many of these action potentials occurring spontaneously together with most, but not all of the depolarizing synaptic activity (Chen, Trombley & van den Pol, 1996). The coupling of bicuculline-sensitive, GABA-mediated depolarization and elevation has also been reported in recordings of cultured rat hippocampal neurons (Segal, 1993), pyramidal neurons in early postnatal hippocampal slice preparations (Leinekugel, Tseeb, Ben-Ari & Bregestovski, 1995) and in tissue print preparations of embryonic neocortex (Owens, Boyce, Davis & Kriegstein, 1996). In the hippocampal slice preparation, the coupling between GABAA receptor/Cl− conductance, depolarization and activation of voltage-dependent Ca2+ channels disappears during the postnatal period coincident with a progressive shift in ECl to more hyperpolarized potentials.
Astrocyte effects
Changes in the functional effects of GABAA receptor activation recorded in slice preparations during the postnatal period parallel the developmental appearance of astrocytes in the hippocampus (Nixdorf-Bergweiler, Albrecht & Heinemann, 1994). In the developing rat spinal cord, GFAP+ elements appear at E16 and their numbers increase during the late embryonic and early postnatal period (Yang et al. 1993). In the present study, accelerated changes in the polarity of the GABA response and the progressive loss of a concomitant elevation in occur in ventral spinal cord neurons dissociated at E15 and cultured over a 3 week period in the presence of astrocytes. This is the same time period during which these neurons would be subject to the influence(s) of proliferating astrocytes in vivo. Hence, in vivo astrocytes could play a role in regulating these developmental changes. There have been numerous studies on the profound effects of astrocytes on neuronal proliferation, morphology, electrophysiological properties and homeostasis (Silver, Lorenz, Wahlsten & Coughline, 1982; Sivron, Eitan, Schreyer & Schwartz, 1993; Travis, 1994; Wu & Barish, 1994; Barish, 1995; Sontheimer, 1995; Tsacopoulos & Magistretti, 1996).
The effects of astrocytes on GABA-induced, chloride-dependent potential in vitro were mimicked by exposure to medium conditioned for 24 h by astrocytes. Thus, confluent astrocytes derived from postnatal cortex secrete substances independently of the presence of embryonic neurons that serve to signal developmental changes in ECl and, in turn, responses. There are several possible targets for astrocyte factor regulation that could influence levels of intracellular Cl−. These include a Cl−/HCO3− exchanger driven by the intracellular production of CO2/HCO3− (Kobayashi, Morgans, Casey & Kopito, 1994), a Cl−/cation transporter driven by the transmembrane gradient for Na+ (Misgeld, Deisz, Dodt & Lux, 1986; Rohrbough & Spitzer, 1996), or an ATP-dependent Cl− pump (Inagaki, Hara & Inoue, 1992). Interestingly, brain astrocytes and their conditioned media have recently been shown to stimulate Na+–K+-Cl− cotransporter activity in brain endothelial cells (Sun, Lytle & O'Donnell, 1997). In the latter study the cytokine, IL-6 appeared to be the stimulatory factor. Further work should help to identify which transporter, if any, is a target for astrocyte regulation in developing spinal cord neurons.
Developmental changes in GABAA receptor/Cl− channel properties
We found that the unitary properties of GABAA receptor/Cl− channels inferred from fluctuation analyses of Cl− current responses changed in parallel with the hyperpolarization of ECl and loss of GABA-induced responses. However, these changes occurred with the same time course whether or not neurons were cultured with astrocytes. Thus, in vitro, astrocytes do not regulate the intrinsic biophysical properties of GABAA receptor/Cl− channels. The developmental changes in Cl− channel properties could involve a switch in the subunit composition of the putative pentameric GABAA receptors. In this regard, we have carried out a preliminary study of GABAA receptor subunit transcript expression using in situ hybridization applied to cells dissociated at E17 and P12 (data not shown). We found cells expressing transcripts for α2, α3, α4, α5, β1, β2, β3, γ1 and γ2 in E17 dissociates, and in P12 dissociates, transcripts for α2, α3, α4, β2, β3 and γ2, consistent with a previous study in vivo (Ma et al. 1993), were found.
It is significant that GABAA receptor subtypes have been shown to change during development (Poulter, Barker, O'Carroll, Lolait & Mahan, 1992; Fritschy, Paysan, Enna & Mohler, 1994; Chang, Luntz-Leybman, Evans, Rotter & Frostholm, 1995; Ma et al. 1993) and different combinations of subunits are associated with different physiological properties (Mathews et al. 1994; Saxena & Macdonald, 1994). Hence, the shortening in the open time derived from the major component in the spectral analyses of GABA-induced fluctuations and the coincident increase in estimated unitary conductance occurring over time in vitro may reflect changes in subunit composition. This remains to be determined.
Conclusion
In summary, astrocytes facilitate developmental changes in the Cl− gradient of spinal neurons extrinsic to GABAA receptor/Cl− channels that profoundly influence the functional effects of GABAA receptor/Cl− channel activation. These astrocyte effects are mediated by diffusible substances and occur independently of parallel changes in the intrinsic properties of the receptor.
References
- Apell H-J, Bersch B. Oxonol VI as an optical indicator for membrane potentials in lipid vesicles. Biochimica et Biophysica Acta. 1987;903:480–494. doi: 10.1016/0005-2736(87)90055-1. [DOI] [PubMed] [Google Scholar]
- Armstrong RC, Dorn HH, Kufta CV, Friedman E, Dubois-Dalcq ME. Pre-oligodendrocytes from adult human CNS. Journal of Neuroscience. 1992;12:1538–1547. doi: 10.1523/JNEUROSCI.12-04-01538.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barish ME. Modulation of the electrical differentiation of neurons by interactions with glia and other non-neuronal cells. Perspectives on Developmental Neurobiology. 1995;2:357–370. [PubMed] [Google Scholar]
- Behar T, Schaffner A, Laing P, Hudson L, Komoly S, Barker JL. Many spinal cord cells transiently express low molecular weight forms of glutamic acid decarboxylase during embryonic development. Developmental Brain Research. 1993;72:203–218. doi: 10.1016/0165-3806(93)90185-d. [DOI] [PubMed] [Google Scholar]
- Ben-Ari Y, Cherubini E, Corradetti R, Gaiarsa J-L. Giant synaptic potentials in immature rat CA3 hippocampal neurones. The Journal of Physiology. 1989;416:303–325. doi: 10.1113/jphysiol.1989.sp017762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang CCA, Luntz-Leybman V, Evans JE, Rotter A, Frostholm A. Developmental changes in the expression of γ-aminobutyric acidA/benzodiazepine receptor subunit mRNAs in the murine inferior olivary complex. Journal of Comparative Neurology. 1995;356:615–628. doi: 10.1002/cne.903560410. [DOI] [PubMed] [Google Scholar]
- Chen G, Trombley PQ, van den Pol AN. Excitatory actions of GABA in developing rat hypothalamic neurones. The Journal of Physiology. 1996;494:451–464. doi: 10.1113/jphysiol.1996.sp021505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fritschy J-M, Paysan J, Enna A, Mohler H. Switch in the expression of rat GABAA-receptor subtypes during postnatal development: an immunohistochemical study. Journal of Neuroscience. 1994;14:5302–5324. doi: 10.1523/JNEUROSCI.14-09-05302.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huettner JE, Baughman RW. Primary culture of identified neurons from the visual cortex. Journal of Neuroscience. 1986;6:3044–3060. doi: 10.1523/JNEUROSCI.06-10-03044.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Inagaki C, Hara M, Inoue M. Neuronal intracellular chloride ion concentration and their regulatory mechanisms. Nippon Yakurigaku Zasshi. 1992;99:307–315. doi: 10.1254/fpj.99.307. [DOI] [PubMed] [Google Scholar]
- Kobayashi S, Morgans CW, Casey JR, Kopito RR. AE3 anion exchanger isoforms in the vertebrate retina: developmental regulation and differential expression in neurons and glia. Journal of Neuroscience. 1994;14:6266–6279. doi: 10.1523/JNEUROSCI.14-10-06266.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koulakoff A, Bizzini B, Berwald-Netter Y. Neuronal acquisition of tetanus toxin binding sites: relationship with the last mitotic cycle. Developmental Biology. 1983;100:350–357. doi: 10.1016/0012-1606(83)90229-4. [DOI] [PubMed] [Google Scholar]
- Lauder JM. Neurotransmitters as growth regulatory signals: role of receptors and second messengers. Trends in Neurosciences. 1993;16:233–240. doi: 10.1016/0166-2236(93)90162-f. 10.1016/0166-2236(93)90162-F. [DOI] [PubMed] [Google Scholar]
- Leinekugel X, Tseeb V, Ben-Ari Y, Bregestovski P. Synaptic GABAA activation induces Ca2+ rise in pyramidal cells and interneurons from rat neonatal hippocampal slices. The Journal of Physiology. 1995;487:319–329. doi: 10.1113/jphysiol.1995.sp020882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y-X, Schaffner AE, Walton MK, Barker JL. Cortical astrocytes affect the differentiation of cultured embryonic rat spinal cord neurons. Journal of Neuroscience. 1995;21:A1298. [Google Scholar]
- Ma W, Saunders PA, Somogyi R, Poulter MO, Barker JL. Ontogeny of GABAA receptor subunit mRNAs in rat spinal cord and dorsal root ganglia. Journal of Comparative Neurology. 1993;338:337–359. doi: 10.1002/cne.903380303. [DOI] [PubMed] [Google Scholar]
- Mathews G, Bolossy AM, Holland KD, Isenberg KE, Covey DF, Ferrendelli JA, Rothman SM. Developmental alteration in GABAA receptor structure and physiological properties in cultured cerebellar granule neurons. Neuron. 1994;13:149–158. doi: 10.1016/0896-6273(94)90465-0. 10.1016/0896-6273(94)90465-0. [DOI] [PubMed] [Google Scholar]
- Misgeld U, Deisz RA, Dodt HU, Lux HD. The role of chloride transport in postsynaptic inhibition of hippocampal neurons. Science. 1986;232:1413–1415. doi: 10.1126/science.2424084. [DOI] [PubMed] [Google Scholar]
- Moody SA, Quigg MS, Frankfurter A. Development of the peripheral trigeminal system in the chick revealed by an isotype-specific anti-beta-tubulin monoclonal antibody. Journal of Comparative Neurology. 1989;279:567–580. doi: 10.1002/cne.902790406. [DOI] [PubMed] [Google Scholar]
- Neher E, Stevens CF. Conductance fluctuations and ionic pores in membranes. Annual Reviews of Biophysics and Bioengineering. 1977;6:345–381. doi: 10.1146/annurev.bb.06.060177.002021. 10.1146/annurev.bb.06.060177.002021. [DOI] [PubMed] [Google Scholar]
- Nixdorf-Bergweiler BE, Albrecht D, Heinemann U. Developmental changes in the number, size and orientation of GFAP-positive cells in the CA1 region of rat hippocampus. Glia. 1994;12:180–195. doi: 10.1002/glia.440120304. [DOI] [PubMed] [Google Scholar]
- Obrietan K, van den Pol AN. GABA activity mediating cytosolic Ca2+ rises in developing neurons is modulated by cAMP-dependent signal transduction. Journal of Neuroscience. 1995;17:4785–4799. doi: 10.1523/JNEUROSCI.17-12-04785.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Owens DF, Boyce LH, Davis MB, Kriegstein AR. Excitatory GABA responses in embryonic and neonatal cortical slices demonstrated by gramicidin perforated-patch recordings and calcium imaging. Journal of Neuroscience. 1996;16:6414–6423. doi: 10.1523/JNEUROSCI.16-20-06414.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poulter MO, Barker JL, O'Carroll AM, Lolait SJ, Mahan LC. Differential and transient expression of GABAA receptor α-subunit mRNAs in the developing rat CNS. Journal of Neuroscience. 1992;12:2888–2900. doi: 10.1523/JNEUROSCI.12-08-02888.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reichling DB, Kyrozis A, Wang J, MacDermott AB. Mechanisms of GABA and glycine depolarization-induced calcium transients in rat dorsal horn neurons. The Journal of Physiology. 1994;476:411–421. doi: 10.1113/jphysiol.1994.sp020142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rohrbough J, Spitzer NC. Regulation of intracellular Cl− levels by Na(+)-dependent Cl− cotransport distinguishes depolarizing from hyperpolarizing GABAA receptor-mediated responses in spinal neurons. Journal of Neuroscience. 1996;16:82–91. doi: 10.1523/JNEUROSCI.16-01-00082.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saxena NC, Macdonald RL. Assembly of GABAA receptor subunits: role of delta subunit. Journal of Neuroscience. 1994;14:7077–7086. doi: 10.1523/JNEUROSCI.14-11-07077.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Segal M. GABA induces a unique rise of [Ca2+]i in cultured rat hippocampal neurons. Hippocampus. 1993;3:229–238. doi: 10.1002/hipo.450030214. [DOI] [PubMed] [Google Scholar]
- Serafini R, Valeyev AY, Barker JL, Poulter MO. Depolarizing GABA-activated Cl− channels in embryonic rat spinal and olfactory bulb cells. The Journal of Physiology. 1995;488:371–386. doi: 10.1113/jphysiol.1995.sp020973. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silver J, Lorenz SE, Wahlsten D, Coughline J. Axon guidance during development of the great cerebral commissures: descriptive and experimental studies, in vivo, on the role of preformed glial pathways. Journal of Comparative Neurology. 1982;210:10–29. doi: 10.1002/cne.902100103. [DOI] [PubMed] [Google Scholar]
- Sivron T, Eitan S, Schreyer DJ, Schwartz M. Astrocytes play a major role in the control of neuronal proliferation in vitro. Brain Research. 1993;629:199–208. doi: 10.1016/0006-8993(93)91321-i. 10.1016/0006-8993(93)91321-I. [DOI] [PubMed] [Google Scholar]
- Sontheimer H. Glial neuronal interactions: a physiological perspective. The Neuroscientist. 1995;1:328–337. [Google Scholar]
- Sun D, Lytle C, O'Donnell ME. IL-6 secreted by astroglial cells regulates Na-K-Cl cotransport in brain microvessel endothelial cells. American Journal of Physiology. 1997;272:C1829–1835. doi: 10.1152/ajpcell.1997.272.6.C1829. [DOI] [PubMed] [Google Scholar]
- Travis J. Glia: the brain's other cells. Science. 1994;266:970–972. doi: 10.1126/science.7973679. [DOI] [PubMed] [Google Scholar]
- Tsacopoulos M, Magistretti PJ. Metabolic coupling between glia and neurons. Journal of Neuroscience. 1996;16:877–885. doi: 10.1523/JNEUROSCI.16-03-00877.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Walton MK, Schaffner AE, Barker JL. Sodium channels, GABAA receptors, glutamate receptors develop sequentially on embryonic rat spinal cord cells. Journal of Neuroscience. 1993;13:2068–2084. doi: 10.1523/JNEUROSCI.13-05-02068.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J, Reichling DB, Kyrozis A, MacDermott AB. Developmental loss of GABA- and glycine-induced depolarization and Ca2+ transients in embryonic rat dorsal horn neurons in culture. European Journal of Neuroscience. 1994;6:1275–1280. doi: 10.1111/j.1460-9568.1994.tb00317.x. [DOI] [PubMed] [Google Scholar]
- Wu RL, Barish ME. Astroglia modulation of transient potassium current development in cultured mouse hippocampal neurons. Journal of Neuroscience. 1994;14:1677–1687. doi: 10.1523/JNEUROSCI.14-03-01677.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu WL, Ziskind-Conhaim L, Sweet MA. Early development of glycine and GABA-mediated synapses in rat spinal cord. Journal of Neuroscience. 1992;12:3935–3945. doi: 10.1523/JNEUROSCI.12-10-03935.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang HY, Lieska N, Shao D, Kriho V, Pappas GD. Immunotyping of radial glia and their glial derivatives during development of the rat spinal cord. Journal of Neurocytology. 1993;22:558–571. doi: 10.1007/BF01189043. [DOI] [PubMed] [Google Scholar]
