Skip to main content
Advertisement

Main menu

  • Home
  • Articles
    • Current Issue
    • Fast Forward
    • Latest Articles
    • Special Sections
    • Archive
  • Information
    • Instructions to Authors
    • Submit a Manuscript
    • FAQs
    • For Subscribers
    • Terms & Conditions of Use
    • Permissions
  • Editorial Board
  • Alerts
    • Alerts
    • RSS Feeds
  • Virtual Issues
  • Feedback
  • Submit
  • Other Publications
    • Drug Metabolism and Disposition
    • Journal of Pharmacology and Experimental Therapeutics
    • Molecular Pharmacology
    • Pharmacological Reviews
    • Pharmacology Research & Perspectives
    • ASPET

User menu

  • My alerts
  • Log in
  • Log out
  • My Cart

Search

  • Advanced search
Journal of Pharmacology and Experimental Therapeutics
  • Other Publications
    • Drug Metabolism and Disposition
    • Journal of Pharmacology and Experimental Therapeutics
    • Molecular Pharmacology
    • Pharmacological Reviews
    • Pharmacology Research & Perspectives
    • ASPET
  • My alerts
  • Log in
  • Log out
  • My Cart
Journal of Pharmacology and Experimental Therapeutics

Advanced Search

  • Home
  • Articles
    • Current Issue
    • Fast Forward
    • Latest Articles
    • Special Sections
    • Archive
  • Information
    • Instructions to Authors
    • Submit a Manuscript
    • FAQs
    • For Subscribers
    • Terms & Conditions of Use
    • Permissions
  • Editorial Board
  • Alerts
    • Alerts
    • RSS Feeds
  • Virtual Issues
  • Feedback
  • Submit
  • Visit jpet on Facebook
  • Follow jpet on Twitter
  • Follow jpet on LinkedIn
Research ArticleNEUROPHARMACOLOGY

Selective Depression by General Anesthetics of Glutamate Versus GABA Release from Isolated Cortical Nerve Terminals

Robert I. Westphalen and Hugh C. Hemmings Jr.
Journal of Pharmacology and Experimental Therapeutics March 2003, 304 (3) 1188-1196; DOI: https://doi.org/10.1124/jpet.102.044685
Robert I. Westphalen
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Hugh C. Hemmings Jr.
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

The role of presynaptic mechanisms in general anesthetic depression of excitatory glutamatergic neurotransmission and facilitation of GABA-mediated inhibitory neurotransmission is unclear. A dual isotope method allowed simultaneous comparisons of the effects of a representative volatile (isoflurane) and intravenous (propofol) anesthetic on the release of glutamate and GABA from isolated rat cerebrocortical nerve terminals (synaptosomes). Synaptosomes were prelabeled with l-[3H]glutamate and [14C]GABA, and release was determined by superfusion with pulses of 30 mM K+ or 1 mM 4-aminopyridine (4AP) in the absence or presence of 1.9 mM free Ca2+. Isoflurane maximally inhibited Ca2+-dependent 4AP-evokedl-[3H]glutamate release (99 ± 8% inhibition) to a greater extent than [14C]GABA release (74 ± 6% inhibition; P = 0.023). Greater inhibition of l-[3H]glutamate versus [14C]GABA release was also observed for the Na+ channel antagonists tetrodotoxin (99 ± 4 versus 63 ± 5% inhibition; P < 0.001) and riluzole (84 ± 5 versus 52 ± 12% inhibition; P= 0.041). Propofol did not differ in its maximum inhibition of Ca2+-dependent 4AP-evokedl-[3H]glutamate release (76 ± 12% inhibition) compared with [14C]GABA (84 ± 31% inhibition; P = 0.99) release. Neither isoflurane (1 mM) nor propofol (15 μM) affected K+-evoked release, consistent with a molecular target upstream of the synaptic vesicle exocytotic machinery or voltage-gated Ca2+ channels coupled to transmitter release. These findings support selective presynaptic depression of excitatory versus inhibitory neurotransmission by clinical concentrations of isoflurane, probably as a result of Na+ channel blockade.

General anesthetics, like most other neuroactive drugs, modulate synaptic transmission in the central nervous system. The synaptic actions of general anesthetics are agent-specific, but considerable evidence indicates that they depress fast excitatory synaptic transmission mediated by glutamate and/or enhance fast inhibitory synaptic transmission mediated by GABA and glycine (MacIver, 1997). The cellular (e.g., presynaptic versus postsynaptic) and molecular (e.g., ion channel, receptor, signal transduction pathway, and fusion machinery) mechanisms of these synaptic effects remain to be elucidated. Moreover, the relative importance of the effects of various anesthetics on excitatory versus inhibitory synaptic transmission remains unclear (Perouansky and Hemmings, 2003).

Electrophysiological evidence supports both presynaptic (effects on neurotransmitter release) and postsynaptic (receptor interactions) mechanisms for the synaptic actions of general anesthetics. Prolongation of synaptic inhibition by positive modulation of postsynaptic GABAA receptor function at GABAergic synapses is an important component of the depressant effects of volatile anesthetics and the primary action of several chemically distinct intravenous anesthetics at clinical concentrations (Hales and Lambert, 1991; Franks and Lieb, 1994; MacIver, 1997; Wakasugi et al., 1999; Buggy et al., 2000). Depression of excitatory transmission is also observed at clinical concentrations of many general anesthetics, particularly volatile anesthetics (Perouansky et al., 1995; MacIver et al., 1996; Ouanounou et al., 1998; Wakasugi et al., 1999). The mechanisms of these effects are unclear. Intravenous and volatile anesthetics decrease excitatory postsynaptic potentials in hippocampal neurons, which has been indirectly attributed to a presynaptic mechanism (Perouansky et al., 1995; MacIver et al., 1996).

Presynaptic effects of general anesthetics on glutamate release have been demonstrated directly at the neurochemical level using isolated nerve terminals (synaptosomes). Synaptosomes provide a superior experimental system for investigating presynaptic effects of drugs on synaptic transmission in isolation of indirect effects present in intact neural networks, such as brain slices. Chemical depolarization by superfusion with K+ channel blockers or elevated K+ stimulates neurotransmitter release from synaptosomes with comparable pharmacological properties to intact nerve terminals (Tibbs et al., 1989), while minimizing the effects of released transmitter (Garcia-Sanz et al., 2001). Volatile and intravenous anesthetics inhibit depolarization-induced glutamate release from isolated nerve terminals (Miao et al., 1995; Schlame and Hemmings, 1995; Ratnakumari and Hemmings, 1998). Considerable evidence indicates that this effect is due to suppression of presynaptic voltage-gated Na+ channels coupled to glutamate release (Ratnakumari and Hemmings, 2000).

Elucidation of the effects of general anesthetics on release of the major excitatory transmitter glutamate and major inhibitory transmitter GABA is essential to understanding the neurophysiological outcomes of presynaptic anesthetic actions. Effects on GABA release are of particular interest given the postsynaptic facilitation of GABAergic transmission by most anesthetics (Franks and Lieb, 1994). The apparent conservation in the fundamental machinery involved in mediating transmitter release among various nerve terminals (Scheller, 1995) suggests that the release of many neurotransmitters, in addition to glutamate, should also be inhibited by general anesthetics. Previous studies have failed to clarify the effects of general anesthetics on GABA release. We therefore compared the effects of a prototypical volatile (isoflurane) and intravenous (propofol) anesthetic on the release of glutamate and GABA from isolated rat cerebrocortical nerve terminals using a dual radiolabel superfusion technique.

Materials and Methods

Materials.

Isoflurane was obtained from Abbott Laboratories (North Chicago, IL) and propofol (2,6-diisopropylphenol) from Sigma-Aldrich (St. Louis, MO). Tetrodotoxin, riluzole, and 4-aminopyridine (4AP) were from Sigma-Aldrich,l-[3H]glutamate (42 Ci/mmol) fromAmersham Biosciences UK, Ltd. (Buckinghamshire, UK), and [14C]GABA (0.24 Ci/mmol) from PerkinElmer Life Sciences (Boston, MA).

Synaptosome Preparation.

Experiments were done in accordance with the National Institutes of Health Guidelines for the Care and Use of Laboratory Animals as approved by the Weill Medical College of Cornell University Institutional Animal Care and Use Committee. Male Sprague-Dawley rats (250–350 g) were anesthetized with 80% CO2/20% O2 and sacrificed by cervical dislocation. The brain was rapidly removed and placed on ice. The cerebral cortex was removed and homogenized in 10 volumes of ice-cold 0.32 M sucrose with a motor-driven (500 rpm) Teflon-glass homogenizer for 10 strokes, and the homogenate was centrifuged for 2 min at 4,000g. Crude rat cerebrocortical synaptosomes (supernatant) were demyelinated by centrifugation through 0.8 M sucrose (10 ml of supernatant layered onto 10 ml of 0.8 M sucrose) for 30 min at 36,000g (Dodd et al., 1981). The pellet containing isolated nerve terminals was resuspended in ice-cold 0.32 M sucrose for use within 2 h.

Glutamate and GABA Release.

Demyelinated rat cerebrocortical synaptosomes were loaded with 8 nMl-[3H]glutamate and 440 nM [14C]GABA for 15 min at 30°C in Krebs-HEPES buffer (composition: 140 mM NaCl, 5 mM KCl, 20 mM HEPES, 1 mM MgCl2, 1.2 mM Na2HPO4, 5 mM NaHCO3, 0.1 mM EGTA, and 10 mMd-glucose, pH 7.4 with NaOH), pelleted by centrifugation (10 min at 20,000g at 4°C), and resuspended in ice-cold 0.32 M sucrose. For concentration-effect experiments, the prelabeling procedure was followed by a 5-min incubation with 0.2 mMd-aspartate to reduce cytoplasmicl-[3H]glutamate and, thus, Ca2+-independentl-[3H]glutamate release via reverse transport (Lester et al., 1994). Synaptosomes were confined by Whatman GF/B glass fiber filter disks (Maidstone, UK) and superfused at 0.5 ml/min with Krebs-HEPES buffer (initially bubbled with 95% O2/5% CO2 for 10 min) at 36°C using a customized (Diagram FD1) Brandel SF12 superfusion apparatus (Gaithersburg, MD) set to collect 1-min fractions. Release was induced by pulses of either 30 mM K+ (with KCl replacing equivalent NaCl) for 3 min or 1 mM 4AP for 2 min. All pulses were induced in the presence (addition of 2 mM CaCl2 yielding final free [Ca2+] of 1.9 mM; MaxChelator v2.10;http://www.stanford.edu/∼cpatton/maxc.html) or absence of Ca2+. At the end of each experiment, synaptosomes were lysed with 0.2 M perchloric acid, and radioactivity in the synaptosomes and each fraction was quantified by liquid scintillation spectrometry with dual isotope quench correction (Beckman Coulter LS 6000IC; Beckman Coulter, Inc., Fullerton, CA) using BioSafe II scintillation cocktail (RPI, Mt. Prospect, IL). The accuracy of the dual radiolabel assay has been verified previously in detecting changes in the release of each amino acid independently of the other (Westphalen and Hemmings, 2003).

Figure FD1
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure FD1

Modified superfusion system integrating a “closed” design for use with volatile anesthetics. Isoflurane or isoflurane with secretagogue are contained in separate valved glass syringes such that the superfusate can be switched between three solutions.

Application of Anesthetics.

Isoflurane was added to glass syringes by dilution of saturated solutions (10–12 mM) in Krebs-HEPES buffer prepared 12 to 24 h before use. Propofol was added from concentrated solutions prepared in dimethyl sulfoxide and diluted in Krebs-HEPES buffer in glass tubes before addition to assay [maximum final [dimethyl sulfoxide] < 0.5% (v/v)]. The aqueous isoflurane concentration for initial experiments (1 mM) was approximately 3 times minimum alveolar concentration, the EC50for suppression of movement in response to a painful stimulus (Taheri et al., 1991). The propofol concentration (>12 μM) was approximately 5 times the EC50 for loss of righting reflex (Tonner et al., 1992). Isoflurane or propofol were introduced through Teflon tubing from glass syringes (closed system) or glass tubes (open system), respectively. Anesthetics were added 3 min before, during, and 1 min after stimulation to insure the presence of anesthetic throughout the concurrent release pulse. In parallel experiments, isoflurane and propofol concentrations exiting the synaptosome chamber in the superfusate were determined by gas chromatography (Ratnakumari and Hemmings, 1998) or high-performance liquid chromatography (Lingamaneni et al., 2001), respectively.

Data Analysis.

Release in each fraction was expressed as a fraction of synaptosomal content of labeled transmitter before each fraction collected (fractional release). The magnitude of release pulses was determined by subtracting baseline release (average of basal release before and after pulse) from cumulative fractional release values of the release pulse (sum FR; Garcia-Sanz et al., 2001). For analysis, sum FR data from each experiment were normalized to mean control release in the presence or absence of Ca2+. Concentration-effect data are expressed as a percentage of the within experiment control in the presence of Ca2+. Ca2+-dependent release is defined as release in the presence of Ca2+ minus release in the absence of Ca2+.

Mean sum FR values all followed Gaussian distributions, with some significantly differing in variance, as determined by a modification of the method of Kolmogorov and Smirnov and Bartlett's test, respectively. Concentration-effect data were fitted by least-squares analysis to estimate Imax, IC50, and Hill slope with standard errors (Prism v. 3.02; GraphPad Software, San Diego CA). Significant differences between mean sum FR values and between inhibition curve parameters were determined by an unpaired t test with Welch correction for variances that were not assumed to be equal (Instat v. 3.0a; GraphPad Software).

Results

4-Aminopyridine-Evoked Release.

Stimulating preloaded rat cerebrocortical synaptosomes with 1 mM 4AP (Tibbs et al., 1989) induced Ca2+-independent release ofl-[3H]glutamate and [14C]GABA. In the presence of 1.9 mM free Ca2+, total release increased 2.4-fold (P < 0.001) and 3.7-fold (P < 0.001) over Ca2+-independent release, respectively (Fig.1A). In the absence of Ca2+, the release of glutamate and GABA occurs primarily by carrier-mediated reverse transport of cytoplasmic transmitter (Lester et al., 1994); the addition of Ca2+ triggers vesicular release, significantly increasing total release (Südhof, 1995). In the presence of Ca2+, 4AP evoked significantly (P< 0.001) more [14C]GABA thanl-[3H]glutamate.

Figure 1
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 1

Release of l-[3H]glutamate and [14C]GABA from rat cerebrocortical synaptosomes preloaded with l-[3H]glutamate and [14C]GABA in the presence (filled symbols) or absence (open symbols) of 1.9 mM free Ca2+. A, 2-min pulse of 1 mM 4AP (n = 12–18). B, 3-min pulse of 30 mM K+ (n = 20–26). C, 2-min pulse of 1 mM 4AP with (n = 7) or without (n= 7) isoflurane in the presence of 1.9 mM free Ca2+. Concentrations of isoflurane were measured from the synaptosome chamber. Application bars indicate secretagogue or anesthetic entry into synaptosome chamber; lag in measurements represents chamber dead space. Anesthetic concentrations reported under Resultsare the average of values at 28 and 29 min. Data are shown as the mean ± S.E.M.

Neither isoflurane (1.07 ± 0.04 mM) nor propofol (12.6 ± 3.0 μM) affected Ca2+-independent 4AP-evoked release of l-[3H]glutamate and [14C]GABA (Fig.2A). In the presence of Ca2+, isoflurane inhibited 4AP-evokedl-[3H]glutamate (P< 0.001) and [14C]GABA (P = 0.002) release (Fig. 2A). Propofol also inhibited 4AP-evokedl-[3H]glutamate (P = 0.010) and [14C]GABA (P = 0.024) release in the presence of Ca2+ (Fig. 2A).

Figure 2
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 2

Effects of isoflurane, propofol, and tetrodotoxin on 4AP-evoked (A), K+-evoked (B), and basal (C) release ofl-[3H]glutamate and [14C]GABA. Isoflurane (1.07 ± 0.04, 1.08 ± 0.04, 1.03 ± 0.13 mM, respectively), propofol (12.6 ± 3.0, 17.0 ± 0.5, 13.5 ± 1.2 μM, respectively), or tetrodotoxin (2 μM) were present during stimulation of release by 1 mM 4AP or 30 mM K+. Data are shown as the mean ± S.E.M. Statistical comparisons between experimental groups (★, P < 0.05; ★★,P < 0.01; ★★★, P < 0.001) or between amino acids (†, P <0.05; ††,P <0.01; †††, P <0.001) employed unpaired t tests (see Results for definedP values).

In the presence of Ca2+, isoflurane maximally inhibited (Imax)l-[3H]glutamate release to a significantly greater extent than [14C]GABA release (P = 0.023), with no significant difference in IC50 values (Fig. 3A; Table1). Isoflurane completely inhibited Ca2+-dependentl-[3H]glutamate release but inhibited Ca2+-dependent [14C]GABA release by only 74%, as determined by the subtraction from maximum inhibition in the presence of Ca2+ of that in the absence of Ca2+. In the absence of Ca2+, there was no concentration-dependent effect of isoflurane onl-[3H]glutamate or [14C]GABA release (Fig. 3A).

Figure 3
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 3

Concentration-dependent effects of isoflurane (A), propofol (B), tetrodotoxin (C), and riluzole (D) on 1 mM 4AP-evoked release of l-[3H]glutamate and [14C]GABA. Data are shown as the percentage of control in the presence of Ca2+. Ca2+-independent controls are shown as the mean ± S.E.M. (n = 3–5). Parameters of fitted curves are presented in Table 1.

View this table:
  • View inline
  • View popup
Table 1

Parameters for inhibition of 4-aminopyridine-evokedl-[3H]glutamate and [14C]GABA release

An apparent greater potency of propofol inhibition of Ca2+-dependentl-[3H]glutamate release than [14C]GABA release was not significant (P = 0.21), with no difference in efficacy (Fig. 3B). Inhibition of Ca2+-independent release of bothl-[3H]glutamate and [14C]GABA by propofol was also not significant (Fig. 3B).

Inhibition of 4AP-evoked release ofl-[3H]glutamate and [14C]GABA by tetrodotoxin or riluzole (Figs. 3, C and D) paralleled the effects of isoflurane. That is, tetrodotoxin (P < 0.001) or riluzole (P = 0.041) inhibited Ca2+-dependentl-[3H]glutamate release to a greater extent than [14C]GABA release, with no difference in IC50 values (P = 0.26 and P = 0.71, respectively). Tetrodotoxin also inhibited Ca2+-independent release of bothl-[3H]glutamate and [14C]GABA (Figs. 2A and 3C), a finding reproduced by riluzole (Fig. 3D).

Potency for inhibition of Ca2+-independent versus Ca2+-dependent amino acid release differed for some drugs (Table 1). This disparity in potency probably reflects the different mechanisms associated with Ca2+-independent (reverse transport) and Ca2+-dependent (vesicular) release.

K+-Evoked Release.

In the absence of Ca2+, depolarization of preloaded synaptosomes by 30 mM K+ evoked the release ofl-[3H]glutamate and [14C]GABA. In the presence of 1.9 mM free Ca2+, release increased by 2.0-fold (P < 0.001) and 2.2-fold (P < 0.001) over Ca2+-independent release, respectively (Fig.1B). Neither isoflurane (1.08 ± 0.04 mM) nor propofol (17.0 ± 1.5 μM) affected K+-evoked release in the absence or presence of Ca2+ (Fig. 2B). Tetrodotoxin partially inhibited Ca2+-independent K+-evoked release ofl-[3H]glutamate (P = 0.010) and [14C]GABA (P = 0.033) and [14C]GABA release (P = 0.012) in the presence of Ca2+ (Fig. 2B).

Basal Release.

Addition of Ca2+ did not affect basal (unstimulated)l-[3H]glutamate release (P = 0.29) but increased basal [14C]GABA release (P = 0.017; Fig. 2C). Isoflurane (1.03 ± 0.13 mM) inhibited basall-[3H]glutamate release in the absence (P = 0.020) or presence (P = 0.14) of Ca2+ but significantly enhanced basal [14C]GABA release in the absence (P < 0.001) or presence (P = 0.017) of Ca2+ (Fig. 2C). Propofol enhanced basall-[3H]glutamate release (P = 0.031) but not basal [14C]GABA release (P = 0.71) in the presence of Ca2+ (Fig. 2C). Tetrodotoxin inhibited basall-[3H]glutamate release in the absence (P = 0.16) or presence (P = 0.18) of Ca2+ and significantly inhibited basal [14C]GABA release in the absence (P = 0.010) or presence (P = 0.006) of Ca2+ (Fig. 2C).

Equating Secretagogue Intensities.

The sum FR of [14C]GABA released by 1 mM 4AP in the presence of Ca2+ (0.11) was greater (P < 0.001) than that ofl-[3H]glutamate (0.055; Fig. 2A). To test the possibility that the lower sensitivity of [14C]GABA release to isoflurane was due to the greater stimulus intensity, the magnitude of [14C]GABA release was reduced to that approximatingl-[3H]glutamate release by using 0.1 mM 4AP (Fig. 4). The profile of inhibition of 0.1 mM 4AP-evoked [14C]GABA release by isoflurane was similar to that observed with 1 mM 4AP (Fig.5B; Table 1); maximum inhibition remained significantly lower (P < 0.001) than for 1 mM 4AP-evokedl-[3H]glutamate release (Fig. 3A), with no significant change in IC50.

Figure 4
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 4

Concentration-dependent effects of 4AP and K+ on l-[3H]glutamate and [14C]GABA release in the presence of Ca2+. Data are shown as the mean ± S.E.M. (n = 4).

Figure 5
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 5

A, effect of isoflurane (1.15 ± 0.04 mM) on the release of l-[3H]glutamate and [14C]GABA evoked by a reduced K+ stimulus (15 mM) in the absence or presence of Ca2+. Data are shown as the mean ± S.E.M. Statistical comparisons between experimental groups used unpaired t tests (★★,P < 0.01; see Results for definedP values). B, concentration-effect curves for inhibition by isoflurane of the release ofl-[3H]glutamate and [14C]GABA evoked by a reduced 4AP stimulus (0.1 mM) compared with data for 1 mM 4AP (from Fig. 3A).

The sum FR of controll-[3H]glutamate released by 1 mM 4AP (0.055; Fig. 2A) was also significantly less (P < 0.001) than that released by 30 mM K+ (0.12; Fig.2B) in the presence of Ca2+. This difference between K+-evoked and 4AP-evoked release was less marked for [14C]GABA release (0.13 versus 0.11;P = 0.056). To test the possibility that the lower sensitivity to isoflurane of K+-evokedl-[3H]glutamate release was due to the greater stimulus intensity, the magnitude of K+-evokedl-[3H]glutamate release was reduced to that released by 1 mM 4AP by using 15 mM K+ (Fig. 4). Isoflurane (1.15 ± 0.04 mM) also failed to inhibit 15 mM K+-evokedl-[3H]glutamate release (Fig. 5A). The reduced intensity of the K+stimulus also equalized [14C]GABA release with that evoked by 1 mM 4AP (Fig. 4), which was also unaffected by isoflurane in the presence of Ca2+ (Fig. 5A). In the absence of Ca2+, isoflurane slightly enhanced 15 mM K+-evoked [14C]GABA release (P < 0.001; Fig. 5A).

Isoflurane-Evoked Release.

Isoflurane (1.03 ± 0.13 mM) alone inhibited basal release ofl-[3H]glutamate but enhanced basal [14C]GABA release in the absence or presence of Ca2+ (Fig. 2C). At concentrations above 0.8 mM, isoflurane evoked [14C]GABA release (Fig.6). Above a threshold concentration of about 2 mM, isoflurane evokedl-[3H]glutamate release (Fig. 6). Statistical comparisons in the absence (2.09 ± 0.1 mM isoflurane;n = 6; Fig. 6A) or presence (2.07 ± 0.09 mM isoflurane; n = 6; Fig. 6B) of Ca2+ showed that isoflurane-evoked release of both l-[3H]glutamate (P = 0.079) and [14C]GABA (P = 0.68) release was Ca2+-independent. Isoflurane (2 mM) preferentially released [14C]GABA overl-[3H]glutamate in the absence (P = 0.0013) or presence (P = 0.015) of Ca2+. These effects of isoflurane were nonsaturable up to the highest concentration tested (2.4 mM).

Figure 6
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 6

Concentration-dependent effects of isoflurane on basal release of l-[3H]glutamate and [14C]GABA in the absence (A) or presence (B) of 1.9 mM free Ca2+. Values for release evoked by 2 mM isoflurane are the mean ± S.E.M. (n = 6). Statistical comparisons between −Ca2+ and +Ca2+ (no significant differences) and betweenl-[3H]glutamate and [14C]GABA (†, P < 0.05; ††, P <0.01), determined by unpaired t tests, are shown. SeeResults for defined P values.

Discussion

The general anesthetics isoflurane and propofol significantly inhibited 4AP-evoked release of glutamate and GABA. Isoflurane preferentially inhibited 4AP-evoked glutamate release compared with GABA release over a range of clinical concentrations. This was due to greater efficacy rather than potency. The Na+channel blockers tetrodotoxin and riluzole also produced selective inhibition of 4AP-evoked release of glutamate versus GABA. This suggests that isoflurane, which inhibits neuronal voltage-gated Na+ channel currents (Rehberg et al., 1996;Rehberg and Duch, 1999; Lingameneni et al., 2000), selectively inhibits glutamate release via Na+ channel blockade (Tibbs et al., 1989). This conclusion is consistent with the recent finding that riluzole preferentially inhibits transmitter release from excitatory versus inhibitory neurons in hippocampal microcultures (Prakriya and Mennerick, 2000). The physiological difference(s) between glutamatergic and GABAergic nerve terminals that underlies this pharmacological selectivity remains to be determined.

In contrast to the secretagogue 4AP, neither isoflurane nor propofol affected glutamate or GABA release evoked by elevated K+. Previous studies support the conclusion that clinically relevant concentrations of volatile anesthetics do not affect K+-evoked excitatory amino acid release (Minchin, 1981; Kendall and Minchin, 1982; Lingamaneni et al., 2001) or GABA release (Minchin, 1981; Kendall and Minchin, 1982; Lecharney et al., 1995; Salford et al., 1997; Lingamaneni et al., 2001). Inhibition of K+-evoked release of glutamate (Miao et al., 1995; Larsen and Langmoen, 1998; Liachenko et al., 1999) or GABA (Larsen et al., 1998; Liachenko et al., 1999) by isoflurane has been reported. Interpretation of results obtained in brain slices (Liachenko et al., 1999), however, is confounded by indirect anesthetic effects mediated by postsynaptic GABAA receptors (Buggy et al., 2000). Failure to compare anesthetic effects on release evoked by additional secretagogues (Lingamaneni et al., 2001) makes it difficult to interpret results obtained only with elevated KCl. Intravenous anesthetics, including propofol, do not affect K+-evoked glutamate (Lingamaneni et al., 2001) or GABA (Mantz et al., 1995; Lingamaneni et al., 2001) release from synaptosomes at clinical concentrations, a finding reproduced in this study. Greater sensitivity of 4AP-evoked versus K+-evoked glutamate release to volatile anesthetics and propofol has been reported in isolated nerve terminals prepared from several brain regions and species using a fluorometric assay of endogenous glutamate (Schlame and Hemmings, 1995; Lingamaneni et al., 2001). This differential sensitivity between secretagogues was not due to differences in stimulus intensity. When [K+]o was reduced to 15 mM, which evoked glutamate and GABA release comparable to that evoked by 1 mM 4AP, isoflurane remained an ineffective inhibitor of K+-induced glutamate or GABA release.

Selective inhibition by isoflurane and propofol of 4AP-evoked over K+-evoked release implicates preferential blockade of voltage-gated Na+ channels over Ca2+ channels in their presynaptic actions. Repetitive depolarization due to blockade of A-type K+ channels by 4AP is amplified by activation of voltage-gated Na+ channels leading to the opening of voltage-gated Ca2+ channels coupled to transmitter release. Elevated K+ leads to synchronous activation of voltage-gated Ca2+channels followed by a plateau of residual Ca2+channel conductance (Tibbs et al., 1989). Blockade of Na+ channels is predicted to inhibit 4AP-evoked, but not K+-stimulated, transmitter release, as demonstrated in this study. In the presence of Ca2+, the Na+ channel blocker tetrodotoxin, like isoflurane, completely inhibited 4AP-evoked, but not K+-evoked, release. A role for Na+ channel blockade in the inhibition of glutamate release has been suggested previously for volatile anesthetics (Ratnakumari and Hemmings, 1998; Lingamaneni et al., 2001;Asai et al., 2002) and propofol (Ratnakumari and Hemmings, 1997).

Blockade of presynaptic voltage-gated Ca2+channels effectively inhibits evoked transmitter release by limiting Ca2+ influx (Wu and Saggau, 1997), which is essential for vesicular release (Südhof, 1995). Multiple Ca2+ channel subtypes can coexist in presynaptic terminals to regulate neurotransmitter release (Turner et al., 1993). Electrophysiological evidence indicates that certain voltage-gated Ca2+ channel subtypes are blocked by general anesthetics (Topf et al., 2003). In this study, anesthetics did not affect K+-evoked amino acid release, a process that involves Ca2+ influx primarily via voltage-gated Ca2+ channel opening. This may be explained by a high safety factor for inhibition of transmitter release by Ca2+ channel blockade or by reduced anesthetic sensitivity of the presynaptic Ca2+ channel subtypes coupled to transmitter release, possibly due to nerve terminal-specific modulation (Turner et al., 1993).

The Na+ channel blocker tetrodotoxin inhibited basal and Ca2+-independent K+-evoked glutamate and GABA release. Changes in the [Na+]o/[Na+]iratio alter Ca2+-independent reverse transport of transmitters that are coupled to Na+ cotransport (Lester et al., 1994). The disparity between inhibition of Ca2+-independent release by tetrodotoxin and by anesthetics may stem from the different potencies or efficacies by which they inhibit Na+ entry (Ratnakumari and Hemmings, 1997; Ratnakumari and Hemmings, 1998) and/or from anesthetic actions at other presynaptic targets (MacIver, 1997). Recent evidence suggests, however, that isoflurane or propofol do not directly affect presynaptic neuronal transporters of glutamate or GABA (Westphalen and Hemmings, 2003).

Isoflurane, like tetrodotoxin, inhibited basal glutamate release, apparently through Na+ channel blockade. In contrast, isoflurane enhanced basal GABA release, particularly at higher concentrations (>1.5 mM), whereas tetrodotoxin was inhibitory. This stimulatory effect of isoflurane occurred at higher than clinical concentrations (Taheri et al., 1991), was not saturable over the concentration range studied, and was Ca2+-independent. General anesthetics also evoke the release of norepinephrine (Pashkov and Hemmings, 2002), whereas halothane produces an increase in mini-excitatory postsynaptic potential frequency in a rat hippocampal slice preparation (Nishikawa and MacIver, 2000). These observations suggest that volatile anesthetics may stimulate a low level of spontaneous Ca2+-independent vesicular release from certain terminals by a mechanism resistant to inhibition by Na+ channel blockade. The mechanisms underlying Ca2+-independent vesicular transmitter release and biochemical differences between glutamatergic and GABAergic terminals remain unclear; whether these mechanisms are directly affected by anesthetics warrants further investigation.

Significantly more GABA was released compared with glutamate for a given concentration of 4AP (Tapia and Sitges, 1982). When the level of 4AP-evoked GABA release was equalized with that of glutamate by reducing the concentration of 4AP, selective inhibition of glutamate release by isoflurane was maintained, thus eliminating a difference in secretagogue intensity as a cause. Preferential stimulation of basal GABA release by isoflurane alone could also contribute to the lower maximal inhibition of evoked GABA release compared with that of glutamate. Selective inhibition of 4AP-evoked glutamate versus GABA release, however, was also produced by other Na+channel blockers that are not known to stimulate basal release.

Selective inhibition of amino acid release by general anesthetics and Na+ channel blockers and differential sensitivity to the secretagogue 4AP suggest fundamental physiological differences between glutamatergic and GABAergic nerve terminals. Mounting evidence supports the notion of distinct patterns of presynaptic ion channel distributions within and between neurons by demonstrating selective densities and/or function of K+ channels (Veh et al., 1995; Southan and Robertson, 1998) and Na+channels (Stuart et al., 1997; Martina et al., 2000). Differential involvement of a presynaptic phorbol ester receptor (Munc13-1) between glutamate and GABA terminals has also been reported (Augustin et al., 1999). Our findings, and those of Prakriya and Mennerick (2000), demonstrate the pharmacological implications of such transmitter-specific presynaptic specialization.

Inhibition of excitatory amino acid transmitter release, which may be enhanced by possible blockade of postsynaptic glutamate receptors (Perouansky and Antognini, 2003), appears to be an important mechanism of neuronal depression by clinical concentrations of volatile anesthetics. Concurrently, partial inhibition of inhibitory amino acid transmitter release may be balanced by potentiation of postsynaptic GABAA receptors. The presynaptic effects of the intravenous anesthetic propofol are less potent in relation to its clinically relevant concentrations; its marked effects on GABAA receptors (Hales and Lambert, 1991) may play a more important role in producing anesthesia. For isoflurane, selective depression of glutamate release, stimulation of spontaneous GABA release, and potentiation of postsynaptic GABAA receptors provide complementary actions to depress excitatory and enhance inhibitory central nervous system transmission.

Footnotes

  • This work was supported by a grant from the National Institutes of Health (GM 58055) and by the Department of Anesthesiology.

  • DOI: 10.1124/jpet.102.044685

  • Abbreviations:
    4AP
    4-aminopyridine
    FR
    fractional release
    • Received September 27, 2002.
    • Accepted December 4, 2002.
  • The American Society for Pharmacology and Experimental Therapeutics

References

  1. ↵
    1. Asai T,
    2. Kusudo K,
    3. Takenoshita M,
    4. Murase K
    (2002) Effect of halothane on neuronal excitation in the superficial dorsal horn of rat spinal cord slices: evidence for a presynaptic action. Eur J Neurosci 15:1278–1290.
    OpenUrlCrossRefPubMed
  2. ↵
    1. Augustin I,
    2. Rosenmund C,
    3. Südhof TC,
    4. Brose N
    (1999) Munc13–1 is essential for fusion competence of glutamatergic synaptic vesicles. Nature (Lond) 400:457–461.
    OpenUrlCrossRefPubMed
  3. ↵
    1. Buggy DJ,
    2. Nicol B,
    3. Rowbotham DJ,
    4. Lambert DG
    (2000) Effects of intravenous anesthetic agents on glutamate release. Anesthesiology 92:1067–1073.
    OpenUrlCrossRefPubMed
  4. ↵
    1. Dodd PR,
    2. Hardy JA,
    3. Oakley AE,
    4. Edwardson JA,
    5. Perry EK,
    6. Delaunoy J-P
    (1981) A rapid method for preparing synaptosomes: comparison, with alternative procedures. Brain Res 226:107–118.
    OpenUrlCrossRefPubMed
  5. ↵
    1. Franks NP,
    2. Lieb WR
    (1994) Molecular and cellular mechanisms of general anaesthesia. Nature (Lond) 367:607–613.
    OpenUrlCrossRefPubMed
  6. ↵
    1. Garcia-Sanz A,
    2. Badia A,
    3. Clos MV
    (2001) Superfusion of synaptosomes to study presynaptic mechanisms involved in neurotransmitter release from rat brain. Brain Res Protoc 7:94–102.
    OpenUrlCrossRefPubMed
  7. ↵
    1. Hales TG,
    2. Lambert JJ
    (1991) The actions of propofol on inhibitory amino acid receptors of bovine adrenomedullary chromaffin cells and rodent central neurones. Br J Pharmacol 104:619–628.
    OpenUrlPubMed
  8. ↵
    1. Kendall TJ,
    2. Minchin MC
    (1982) The effects of anesthetics on the uptake and release of amino acid neurotransmitters in thalamic slices. Br J Pharmacol 75:219–227.
    OpenUrlPubMed
  9. ↵
    1. Larsen M,
    2. Haugstad TS,
    3. Berg-Johnsen J,
    4. Langmoen IA
    (1998) Effect of isoflurane on release and uptake of γ-aminobutyric acid from cortical synaptosomes. Br J Anaesth 80:634–638.
    OpenUrlAbstract/FREE Full Text
  10. ↵
    1. Larsen M,
    2. Langmoen IA
    (1998) Effect of volatile anesthetics on synaptic release and uptake of glutamate. Toxicol Lett 100–101:59–64.
  11. ↵
    1. Lecharney J-P,
    2. Salford F,
    3. Henzel D,
    4. Desmones J-M,
    5. Mantz J
    (1995) Effects of thiopental, halothane and isoflurane on the calcium-dependent and -independent release of GABA from striatal synaptosomes in the rat. Brain Res 670:308–312.
    OpenUrlCrossRefPubMed
  12. ↵
    1. Lester HA,
    2. Mager S,
    3. Quick MW,
    4. Corey JL
    (1994) Permeation properties of neurotransmitter transporters. Annu Rev Pharmacol Toxicol 34:219–249.
    OpenUrlCrossRefPubMed
  13. ↵
    1. Liachenko S,
    2. Tang P,
    3. Somogyi G,
    4. Xu Y
    (1999) Concentration-dependent isoflurane effects on depolarization-evoked glutamate and GABA outflows from mouse brain slices. Br J Pharmacol 127:131–138.
    OpenUrlCrossRefPubMed
  14. ↵
    1. Lingamaneni R,
    2. Birch ML,
    3. Hemmings HC, Jr
    (2001) Widespread inhibition of sodium channel-dependent glutamate release from isolated nerve terminals by isoflurane and propofol. Anesthesiology 95:1460–1466.
    OpenUrlCrossRefPubMed
  15. ↵
    1. Lingameneni R,
    2. Vysotskaya TN,
    3. Duch DS,
    4. Hemmings HC, Jr
    (2000) Inhibition of voltage-dependent sodium channels by Ro 31–8220, a “specific” protein kinase inhibitor. FEBS Lett 473:265–268.
    OpenUrlCrossRefPubMed
  16. ↵
    1. Biebuyck JF,
    2. Lynch CI,
    3. Maze M,
    4. Saidman LJ,
    5. Yaksh TL,
    6. Zapol WM
    1. MacIver MB
    (1997) General anesthetic actions on transmission at glutamate and GABA synapses. in Anesthesia: Biological Foundations, eds Biebuyck JF, Lynch CI, Maze M, Saidman LJ, Yaksh TL, Zapol WM (Lippincott-Raven, New York), pp 277–286.
  17. ↵
    1. MacIver MB,
    2. Mikulec AA,
    3. Amagasu SM,
    4. Monroe FA
    (1996) Volatile anesthetics depress glutamate transmission via presynaptic actions. Anesthesiology 85:823–834.
    OpenUrlCrossRefPubMed
  18. ↵
    1. Mantz J,
    2. Lecharny JB,
    3. Laudenbach V,
    4. Henel D,
    5. Peytavin G,
    6. Desmonts JM
    (1995) Anesthetics affect the uptake but not the depolarization-evoked release of GABA in rat striatal synaptosomes. Anesthesiology 82:502–511.
    OpenUrlCrossRefPubMed
  19. ↵
    1. Martina M,
    2. Vida I,
    3. Jonas P
    (2000) Distal initiation and active propagation of action potentials in interneuron dendrites. Science (Wash DC) 287:295–300.
    OpenUrlAbstract/FREE Full Text
  20. ↵
    1. Miao N,
    2. Frazer MJ,
    3. Lynch CI
    (1995) Volatile anethetics depress Ca2+ transients and glutamate release in isolated cerebral synaptosomes. Anesthesiology 83:593–603.
    OpenUrlPubMed
  21. ↵
    1. Minchin MC
    (1981) The effect of anesthetics on the uptake and release of γ-aminobutyrate and d-aspartate in rat brain slices. Br J Pharmacol 73:681–689.
    OpenUrlPubMed
  22. ↵
    1. Nishikawa K,
    2. MacIver MB
    (2000) Membrane and synaptic actions of halothane on rat hippocampal pyramidal neurons and inhibitory interneurons. J Neurosci 20:5915–5923.
    OpenUrlAbstract/FREE Full Text
  23. ↵
    1. Ouanounou A,
    2. Carlen PL,
    3. El-Beheiry H
    (1998) Enhanced isoflurane suppression of excitatory synaptic transmission in the aged rat hippocampus. Br J Pharmacol 123:1075–1082.
    OpenUrlCrossRef
  24. ↵
    1. Pashkov VN,
    2. Hemmings HC, Jr
    (2002) The effects of general anesthetics on norepinephrine release from isolated rat cortical nerve terminals. Anesth Analg 95:1274–1281.
    OpenUrlCrossRefPubMed
  25. ↵
    1. Antognini JF,
    2. Carstens EE,
    3. Raines DE
    1. Perouansky M,
    2. Antognini JF
    (2003) Glutamate receptors: Physiology and anesthetic pharmacology. in Neural Mechanisms of Anesthesia, eds Antognini JF, Carstens EE, Raines DE (Humana Press, Inc. Totowa, NJ), pp 319–332.
  26. ↵
    1. Perouansky M,
    2. Baranov D,
    3. Salman M,
    4. Yaari Y
    (1995) Effects of halothane on glutamate receptor-mediated excitatory postsynaptic currents. Anesthesiology 83:109–119.
    OpenUrlCrossRefPubMed
  27. ↵
    1. Antognini JF,
    2. Carstens EE,
    3. Raines DE
    1. Perouansky M,
    2. Hemmings HC, Jr
    (2003) Presynaptic actions of general anesthetics. in Neural Mechanisms of Anesthesia, eds Antognini JF, Carstens EE, Raines DE (Humana Press, Inc. Totowa, NJ), pp 345–369.
  28. ↵
    1. Prakriya M,
    2. Mennerick S
    (2000) Selective depression of low-release probability excitatory synapses by sodium channel blockers. Neuron 26:671–682.
    OpenUrlCrossRefPubMed
  29. ↵
    1. Ratnakumari L,
    2. Hemmings HC, Jr
    (1997) Effects of propofol on sodium channel-dependent sodium influx and glutamate release in rat cerebrocortical synaptosomes. Anesthesiology 86:428–439.
    OpenUrlCrossRefPubMed
  30. ↵
    1. Ratnakumari L,
    2. Hemmings HC, Jr
    (1998) Inhibition of presynaptic sodium channels by halothane. Anesthesiology 88:1043–1054.
    OpenUrlCrossRefPubMed
  31. ↵
    1. Ratnakumari L,
    2. Hemmings HC, Jr
    (2000) Differential effects of anesthetic and nonanesthetic cyclobutanes on neuronal voltage-gated sodium channels. Anesthesiology 92:529–541.
    OpenUrlPubMed
  32. ↵
    1. Rehberg B,
    2. Duch DS
    (1999) Suppression of central nervous system sodium channels by propofol. Anesthesiology 91:512–520.
    OpenUrlCrossRefPubMed
  33. ↵
    1. Rehberg B,
    2. Xiao YH,
    3. Duch DS
    (1996) Central nervous system sodium channels are significantly suppressed at clinical concentrations of volatile anesthetics. Anesthesiology 84:1223–1233.
    OpenUrlCrossRefPubMed
  34. ↵
    1. Salford F,
    2. Keita H,
    3. Lecharney J-P,
    4. Henzel D,
    5. Desmonts JM,
    6. Mantz J
    (1997) Halothane and isoflurane differentially affect the regulation of dopamine and γ-aminobutyric acid release mediated by presynaptic acetylcholine receptors in the rat striatum. Anesthesiology 86:632–641.
    OpenUrlCrossRefPubMed
  35. ↵
    1. Scheller RH
    (1995) Membrane trafficking in the presynaptic nerve terminal. Neuron 14:893–897.
    OpenUrlCrossRefPubMed
  36. ↵
    1. Schlame M,
    2. Hemmings HC, Jr
    (1995) Inhibition by volatile anesthetics of endogenous glutamate release from synaptosomes by a presynaptic mechanism. Anesthesiology 82:1406–1416.
    OpenUrlCrossRefPubMed
  37. ↵
    1. Southan AP,
    2. Robertson B
    (1998) Patch-clamp recordings from cerebellar basket cell bodies and their presynaptic terminals reveal an asymmetric distribution of voltage-gated potassium channels. J Neurosci 18:948–955.
    OpenUrlAbstract/FREE Full Text
  38. ↵
    1. Stuart G,
    2. Spruston N,
    3. Sakmann B,
    4. Häusser M
    (1997) Action potential initiation and back propagation in neurons of the mammalian CNS. Trends Neurosci 20:125–131.
    OpenUrlCrossRefPubMed
  39. ↵
    1. Südhof TC
    (1995) The synaptic vesicle cycle: a cascade of protein-protein interactions. Nature (Lond) 375:645–653.
    OpenUrlCrossRefPubMed
  40. ↵
    1. Taheri S,
    2. Halsey MJ,
    3. Liu J,
    4. Eger EI, 2nd,
    5. Koblin DD,
    6. Laster MJ
    (1991) What solvent best represents the site of action on inhaled anesthetics in humans, rats and dogs. Anesth Analg 72:627–634.
    OpenUrlPubMed
  41. ↵
    1. Tapia R,
    2. Sitges M
    (1982) Effect of 4-aminopyridine on transmitter release in synaptosomes. Brain Res 250:291–299.
    OpenUrlCrossRefPubMed
  42. ↵
    1. Tibbs GR,
    2. Barrie AP,
    3. Van Mieghem FJE,
    4. McMahon HT,
    5. Nicholls DG
    (1989) Repetitive action potentials in isolated nerve terminals in the presence of 4-aminopyridine: effects on cytosolic free Ca2+ and glutamate release. J Neurochem 53:1693–1699.
    OpenUrlCrossRefPubMed
  43. ↵
    1. Tonner PH,
    2. Poppers DM,
    3. Miller KW
    (1992) The general anesthetic potency of propofol and its dependence on hydrostatic pressure. Anesthesiology 77:926–931.
    OpenUrlCrossRefPubMed
  44. ↵
    1. Antognini JF,
    2. Carstens EE,
    3. Raines DE
    1. Topf N,
    2. Recio-Pinto E,
    3. Blanck TJJ,
    4. Hemmings HC, Jr
    (2003) Actions of general anesthetics on voltage-gated ion channels. in Neural Mechanisms of Anesthesia, eds Antognini JF, Carstens EE, Raines DE (Humana Press, Inc. Totowa, NJ), pp 299–318.
  45. ↵
    1. Turner TJ,
    2. Adams ME,
    3. Dunlap K
    (1993) Multiple Ca2+ channel types coexist to regulate synaptosomal neurtransmitter release. Proc Nat Acad Sci 90:9518–9522.
    OpenUrlAbstract/FREE Full Text
  46. ↵
    1. Veh RW,
    2. Lichtinghagen R,
    3. Sewing S,
    4. Wunder F,
    5. Grumbach IM,
    6. Pongs O
    (1995) Immunochemical localization of five members of the KV1 channel subunits: contrasting subcellular locations and neuron-specific co-localizations in rat brain. Eur J Neurosci 7:2189–2205.
    OpenUrlCrossRefPubMed
  47. ↵
    1. Wakasugi M,
    2. Hitota K,
    3. Roth SH,
    4. Ito Y
    (1999) The effects of general anesthetics on excitatory and inhibitory synaptic transmission in area CA1 of the rat hippocampus in vitro. Anesth Analg 88:676–680.
    OpenUrlCrossRefPubMed
  48. ↵
    1. Westphalen RI,
    2. Hemmings HC, Jr
    (2003) The effects of isoflurane and propofol on glutamate and GABA transporters in isolated cortical nerve terminals. Anesthesiology 98:364–372.
    OpenUrlCrossRefPubMed
  49. ↵
    1. Wu L-G,
    2. Saggau P
    (1997) Presynaptic inhibition of elicited neurotransmitter release. Trends Neurosci 20:204–212.
    OpenUrlCrossRefPubMed
PreviousNext
Back to top

In this issue

Journal of Pharmacology and Experimental Therapeutics: 304 (3)
Journal of Pharmacology and Experimental Therapeutics
Vol. 304, Issue 3
1 Mar 2003
  • Table of Contents
  • About the Cover
  • Index by author
Download PDF
Article Alerts
Sign In to Email Alerts with your Email Address
Email Article

Thank you for sharing this Journal of Pharmacology and Experimental Therapeutics article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Selective Depression by General Anesthetics of Glutamate Versus GABA Release from Isolated Cortical Nerve Terminals
(Your Name) has forwarded a page to you from Journal of Pharmacology and Experimental Therapeutics
(Your Name) thought you would be interested in this article in Journal of Pharmacology and Experimental Therapeutics.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Research ArticleNEUROPHARMACOLOGY

Selective Depression by General Anesthetics of Glutamate Versus GABA Release from Isolated Cortical Nerve Terminals

Robert I. Westphalen and Hugh C. Hemmings
Journal of Pharmacology and Experimental Therapeutics March 1, 2003, 304 (3) 1188-1196; DOI: https://doi.org/10.1124/jpet.102.044685

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero

Share
Research ArticleNEUROPHARMACOLOGY

Selective Depression by General Anesthetics of Glutamate Versus GABA Release from Isolated Cortical Nerve Terminals

Robert I. Westphalen and Hugh C. Hemmings
Journal of Pharmacology and Experimental Therapeutics March 1, 2003, 304 (3) 1188-1196; DOI: https://doi.org/10.1124/jpet.102.044685
Reddit logo Twitter logo Facebook logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Materials and Methods
    • Results
    • Discussion
    • Footnotes
    • References
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Related Articles

Cited By...

More in this TOC Section

  • Substituted Tryptamine Activity at 5-HT Receptors and SERT
  • KRM-II-81 Analogs
  • VTA muscarinic M5 receptors and effort-choice behavior
Show more Neuropharmacology

Similar Articles

Advertisement
  • Home
  • Alerts
Facebook   Twitter   LinkedIn   RSS

Navigate

  • Current Issue
  • Fast Forward by date
  • Fast Forward by section
  • Latest Articles
  • Archive
  • Search for Articles
  • Feedback
  • ASPET

More Information

  • About JPET
  • Editorial Board
  • Instructions to Authors
  • Submit a Manuscript
  • Customized Alerts
  • RSS Feeds
  • Subscriptions
  • Permissions
  • Terms & Conditions of Use

ASPET's Other Journals

  • Drug Metabolism and Disposition
  • Molecular Pharmacology
  • Pharmacological Reviews
  • Pharmacology Research & Perspectives
ISSN 1521-0103 (Online)

Copyright © 2023 by the American Society for Pharmacology and Experimental Therapeutics