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

Volatile Anesthetic Effects on Glutamate versus GABA Release from Isolated Rat Cortical Nerve Terminals: 4-Aminopyridine-Evoked Release

Robert I. Westphalen and Hugh C. Hemmings Jr.
Journal of Pharmacology and Experimental Therapeutics January 2006, 316 (1) 216-223; DOI: https://doi.org/10.1124/jpet.105.090662
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

Inhibition of glutamatergic excitatory neurotransmission and potentiation of GABA-mediated inhibitory transmission are possible mechanisms involved in general anesthesia. We compared the effects of three volatile anesthetics (isoflurane, enflurane, or halothane) on 4-aminopyridine (4AP)-evoked 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 evoked by superfusion with pulses of 1 mM 4AP in the absence or presence of 1.9 mM free Ca2+. All three volatile anesthetics inhibited Ca2+-dependent glutamate and GABA release; IC50 values for glutamate were comparable to clinical concentrations (1–1.6× MAC), whereas IC50 values for GABA release exceeded clinical concentrations (>2.2× MAC). All three volatile anesthetics inhibited both Ca2+-independent and Ca2+-dependent 4AP-evoked glutamate release equipotently, whereas inhibition of Ca2+-dependent 4AP-evoked GABA release was less potent than inhibition of Ca2+-independent GABA release. Inhibition of Ca2+-independent 4AP-evoked glutamate release was more potent than that of GABA release for isoflurane and enflurane but equipotent for halothane. Tetrodotoxin inhibited both Ca2+-independent and Ca2+-dependent 4AP-evoked glutamate and GABA release equipotently, consistent with Na+ channel involvement. In contrast to tetrodotoxin, volatile anesthetics exhibited selective effects on 4AP-evoked glutamate versus GABA release, consistent with distinct mechanisms of action. Preferential inhibition of Ca2+-dependent 4AP-evoked glutamate release versus GABA release supports the hypothesis that reduced excitatory neurotransmission relative to inhibitory neurotransmission contributes to volatile anesthetic actions.

Depression of neuronal activity by general anesthetics appears to involve inhibition of excitatory glutamatergic neurotransmission and enhancement of inhibitory GABAergic neurotransmission (Hemmings et al., 2005a). Positive modulation of postsynaptic GABAA receptor function at GABAergic synapses is probably an important component of the depressant effects of volatile anesthetics (Jones et al., 1992; Franks and Lieb, 1994; MacIver, 1997). However, spinal GABAA receptor potentiation does not explain the immobilizing action characteristic of volatile anesthetics (Zhang et al., 2004). Accumulating evidence indicates that volatile anesthetics affect multiple synaptic targets, both presynaptically and postsynaptically, that interact to produce their clinical effects (MacIver, 1997; Stucke et al., 2002; Hemmings et al., 2005a).

Action potential-evoked neurotransmitter release involves nerve-terminal depolarization that triggers Ca2+ entry through voltage-gated Ca2+ channels coupled to synaptic vesicle exocytosis (Südhof, 2004). Basal neurotransmitter release includes spontaneous vesicular fusion and nonvesicular release via reverse transporter operation independent of Ca2+ entry. These pathways of release are mechanistically distinct and are therefore differentially sensitive to presynaptic neuromodulators (Engelman and MacDermott, 2004). In the accompanying report (Westphalen and Hemmings, 2005), we show that volatile anesthetics produce significant transmitter-selective and agent-specific effects on basal glutamate and GABA release that results in a consistent net increase in inhibitory relative to excitatory basal transmitter release.

Neurotransmitter release from isolated nerve terminals can be evoked pharmacologically by 4AP, a K+ channel blocker that induces release by mimicking action potential-evoked depolarizations (Tibbs et al., 1989). Inhibition of 4AP-evoked glutamate release by volatile anesthetics seems to involve suppression of presynaptic voltage-gated Na+ channels coupled to transmitter release (Schlame and Hemmings, 1995; Ratnakumari et al., 2000; Westphalen and Hemmings, 2003a). However, Na+ channel block does not affect spontaneous transmitter release (Engelman and MacDermott, 2004), and anesthetics modulate basal release independent of their Na+ channel antagonism. Here we compare the effects of three widely used volatile anesthetics, the ethers isoflurane and enflurane and the alkane halothane, and the selective voltage-gated Na+ channel blocker tetrodotoxin on 4AP-evoked release of glutamate and GABA while compensating for their agent-specific effects on basal release (Westphalen and Hemmings, 2005).

Materials and Methods

Materials. Sources of all compounds are listed in the accompanying article (Westphalen and Hemmings, 2005).

Evoked Glutamate and GABA Release ± Anesthetics. 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. Synaptosomes were prepared from rat cerebral cortex and prelabeled with l-[3H]glutamate and [14C]GABA, whereupon simultaneous l-[3H]glutamate and [14C]GABA release was evoked by pulses of 1 mM 4AP as described in the accompanying article (Westphalen and Hemmings, 2005). Using [3H]H2O as a tracer, a 2-min pulse resulted in a peak of 70 ± 4% of the applied concentration to the perfusion chamber, suggesting a maximum concentration of 0.7 mM 4AP exposure to synaptosomes (data not shown). 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, 2003b).

Anesthetics at 0.1 to 8× MAC (minimum alveolar concentration) were applied for 6 min (3 min before, during, and 1 min after stimulation with a 2-min pulse of 4AP) using a closed superfusion system (Westphalen and Hemming, 2005). This procedure insures the exposure to anesthetic throughout the 4AP-evoked release pulse (Westphalen and Hemmings, 2003a). After each experiment, anesthetic solutions were sampled and quantified by gas chromatography as described previously (Westphalen and Hemmings, 2005). In separate experiments, anesthetics were continuously applied 12 min prior to stimulation by 4AP and remained present throughout the perfusion, with comparable effects on release.

Data Analysis. Release in each fraction was expressed as a fraction of synaptosomal content of labeled transmitter prior to each fraction collected (fractional release [FR]). The magnitude of the release pulse was determined by subtracting baseline release (average of basal release before and after pulse) from cumulative fractional release values for the release pulse (sum ΔFR; Garcia-Sanz et al., 2001). Sum ΔFR data for each experiment were normalized by the ratio of the individual assay control to the mean of all assay controls for release in the absence or presence of Ca2+ for each anesthetic used prior to curve fitting.

Concentration-effect data for anesthetic effects on basal l-[3H]glutamate and [14C]GABA release were taken from Westphalen and Hemmings (2005). The basal effect on sum ΔFR for each anesthetic concentration was determined from the fitted equation and was subtracted from the sum ΔFR values determined for 4AP-evoked release. Likewise, the derived Ca2+-independent release values were subtracted from values obtained in the presence of Ca2+ at equivalent anesthetic concentrations to determine Ca2+-dependent release. This subtraction method is standard procedure when baseline measurements of x-axis data cannot be accurately duplicated (Prism, version 4.0; GraphPad Software Inc., San Diego, CA).

Data for anesthetic (0.1–8× MAC) and tetrodotoxin (0.3 nM-10 μM) inhibition of 4AP-evoked release (with basal effect subtracted) were fitted to concentration-effect curves by least-squares analysis to estimate Imax and IC50 with means ± S.E. (Prism, version 4.0). Each curve fit was concurrently tested for the bottom of the curve (Imax) differing significantly from 0. If the null hypothesis (Imax = 0) was not rejected, curve fit analyses were performed with Imax constrained to 0. Sum ΔFR values for all single concentration experimental groups followed Gaussian distributions, with some significantly differing in variance, as determined by a modification of the Kolmogorov and Smirnov test and Bartlett's test, respectively. Significant differences between mean sum ΔFR values and between concentration-effect curve parameters were determined by unpaired Student's t test with Welch correction for variances that were not assumed to be equal (Prism, version 4.0). Sum ΔFR values obtained for 4AP-evoked release in the presence of continuously applied anesthetics were compared with the corresponding values read from inhibition curves with the basal effect subtracted by paired Student's t test.

Results

4-Aminopyridine-Evoked Release. Stimulation of rat cerebrocortical synaptosomes with 1 mM 4AP in the presence of 1.9 mM free external Ca2+ evoked both l-[3H]glutamate and [14C]GABA release (Figs. 1, 2, 3, 4). Total release (sum ΔFR) was greater for [14C]GABA release compared with l-[3H]glutamate, as reported previously (Westphalen and Hemmings, 2003a).

Isoflurane. Isoflurane inhibited glutamate and GABA release in a concentration-dependent manner (Fig. 1). Concentration-effect curves showed that isoflurane inhibited 4AP-evoked release of l-[3H]glutamate (Fig. 1A) and [14C]GABA (Fig. 1B), both in the absence and presence of Ca2+. At concentrations up to 1 mM isoflurane, maximal inhibition of l-[3H]glutamate release was greater than maximal inhibition of [14C]GABA release. However, at concentrations >1 mM isoflurane, sum ΔFR values increased from maximal inhibition. This augmentation of release paralleled the effects on basal release previously observed at high isoflurane concentrations (Westphalen and Hemmings, 2005), as evident in Fig. 1, A and B. Subtraction of the effects of isoflurane on basal release from 4AP-evoked release yielded data that fitted sigmoidal concentration-effect curves with 100% inhibition for both l-[3H]glutamate release (Imax = 0, P = 0.84; Fig. 1C) and [14C]GABA release (Imax = 0, P = 0.37; Fig. 1D) and were comparable to release evoked during continuously perfused anesthetic (see legend to Fig. 1). Isoflurane inhibited evoked l-[3H]glutamate release with greater potency than [14C]GABA release in the presence of Ca2+ (P = 0.0001; Table 1). This selective inhibition by isoflurane is not due to the lower fractional release of glutamate versus GABA evoked by 1 mM 4AP, as shown previously by matching fractional release by lowering 4AP concentration (Westphalen and Hemmings, 2003a). In the absence of Ca2+, isoflurane inhibited 4AP-evoked release of l-[3H]glutamate with lower potency than [14C]GABA release (P < 0.0001; Table 1), with efficacies of 100% for both (Imax = 0, P = 0.77 and P = 0.78, respectively). Isoflurane inhibited Ca2+-dependent and Ca2+-independent 4AP-evoked l-[3H]glutamate release at clinically relevant concentrations (IC50 = 0.58 mM; 1.7× MAC) with the same potency (P = 0.98) and both with greater potency (P < 0.0001) than Ca2+-dependent 4AP-evoked [14C]GABA release (IC50 = 1.06 mM; 3.0× MAC; Table 1). Isoflurane inhibited Ca2+-independent 4AP-evoked [14C]GABA release with greater potency than Ca2+-dependent release (P < 0.0001; Table 1).

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

Inhibition potencies of 4-aminopyridine-evoked l-[3H]glutamate and [14C]GABA release

Values were derived by fitting data to sigmoidal concentration effect-curves (see Materials and Methods). Statistical significance between glutamate and GABA (*, P < 0.05; **, P < 0.01; ***, P < 0.001) and between –Ca2+ and Ca2+-dependent release (††, P < 0.01; †††, P < 0.001) were calculated using logIC50 values by unpaired Student's t test.

Enflurane. Enflurane inhibited 4AP-evoked release of l-[3H]glutamate and [14C]GABA in a concentration-dependent manner (Fig. 2, A and B). Maximum inhibition of l-[3H]glutamate release was indeterminable from these data, which trended negative at concentrations >0.8 mM, apparently as a consequence of the inhibitory effect of enflurane on basal glutamate release (Westphalen and Hemmings, 2005). With basal release subtracted, enflurane inhibited both l-[3H]glutamate and [14C]GABA release by 100% (Imax = 0, P = 0.98 and P = 0.25, respectively; Fig. 2, C and D). The potency of enflurane for inhibition of 4AP-evoked l-[3H]glutamate release was greater than that for [14C]GABA release (P < 0.0001; Table 1). In the absence of Ca2+, enflurane completely inhibited 4AP-evoked release of l-[3H]glutamate and [14C]GABA (Imax = 0, P = 0.57 and P = 0.12, respectively; Table 1; Fig. 2, C and D). Enflurane inhibited Ca2+-independent [14C]GABA release with greater potency than Ca2+-dependent release (P < 0.0001; Table 1). Enflurane inhibited Ca2+-dependent and Ca2+-independent 4AP-evoked l-[3H]glutamate release at clinically relevant concentrations (IC50 = 0.72 mM; 1.0× MAC) with the same potency (P = 0.66) but with greater potency (P = 0.0002) than Ca2+-dependent 4AP-evoked [14C]GABA release (IC50 = 2.34 mM; 3.1× MAC; Table 1). Enflurane inhibited Ca2+-independent 4AP-evoked [14C]GABA release with greater potency than Ca2+-dependent release (P < 0.0001; Table 1).

  Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Isoflurane effects on evoked release. Isoflurane effects on l-[3H]glutamate (A) and [14C]GABA release (B) evoked by 4AP from rat cortical synaptosomes were fitted to sigmoidal concentration-effect curves. Control data in the presence of Ca2+ (glutamate, n = 12; GABA, n = 13) or absence of Ca2+ (glutamate, n = 9; GABA, n = 11) are presented as mean ± S.D. Basal effects (taken from Westphalen and Hemmings, 2005) were subtracted from evoked release data and fitted to sigmoidal concentration-effect curves (C and D). Shaded area highlights the clinical concentration range (0.5–2× MAC). Symbols in gray present sum ΔFR data (mean ± S.D.) for 2-min pulses of 1 mM 4AP in the presence of Ca2+ plus ∼0.5× MAC (0.19 ± 0.003 mM, n = 3), ∼1× MAC (0.34 ± 0.03 mM, n = 4), or ∼2× MAC (0.76 ± 0.04 mM, n = 7) concentrations of isoflurane applied continuously, which do not differ from data obtained using the standard pulse application with basal effect subtracted (glutamate: P = 0.35, GABA: P = 0.17). Release in the absence of anesthetic presents mean ± S.D. of all assay controls.

Halothane. Halothane inhibited 4AP-evoked release of l-[3H]glutamate and [14C]GABA in a concentration-dependent manner (Fig. 3, A and B). With basal release subtracted, halothane inhibited both l-[3H]glutamate and [14C]GABA release by 100% (Imax = 0, P = 0.84 and P = 0.78, respectively; Fig. 3, C and D). Halothane inhibited 4AP-evoked l-[3H]glutamate release with greater potency than [14C]GABA release (P < 0.001; Table 1). In the absence of Ca2+, halothane inhibited 4AP-evoked release of l-[3H]glutamate and [14C]GABA with similar potencies (P = 0.93) and 100% efficacy (Imax = 0, P = 0.88 and P = 0.08, respectively; Table 1; Fig. 3, C and D). Halothane inhibited Ca2+-dependent l-[3H]glutamate release at clinically relevant concentrations (IC50 = 0.56 mM; 1.6× MAC) with the same potency as Ca2+-independent 4AP-evoked l-[3H]glutamate release (P = 0.085; Table 1). Halothane inhibited Ca2+-dependent [14C]GABA release (IC50 = 0.76 mM; 2.2× MAC) with less potency than Ca2+-independent release (P = 0.008; Table 1).

Tetrodotoxin. Tetrodotoxin inhibited 4AP-evoked release of l-[3H]glutamate and [14C]GABA in a concentration-dependent manner (Fig. 4, A and B). With basal effects subtracted, tetrodotoxin showed incomplete efficacy for inhibition of both 4AP-evoked l-[3H]glutamate release (66% inhibition; Imax = 0, P = 0.0024) and [14C]GABA release (46% inhibition; Imax = 0, P = 0.0022) in the presence of Ca2+ (Fig. 4, C and D). Tetrodotoxin inhibited 4AP-evoked l-[3H]glutamate and [14C]GABA release with the same potency (P = 0.78; Table 1). The effects of tetrodotoxin on Ca2+-independent 4AP-evoked l-[3H]glutamate and [14C]GABA release were not statistically significant with basal effects subtracted (Fig. 4, C and D). Thus, curve parameters for Ca2+-dependent inhibition of evoked release of glutamate and GABA were the same (Table 1). However, inhibition of Ca2+-dependent glutamate release was complete (Imax = 0, P = 0.86) but was incomplete for inhibition of Ca2+-dependent GABA release (Imax = 0, P = 0.026). This difference in maximal inhibition of Ca2+-dependent release by tetrodotoxin between glutamate and GABA was also statistically significant by a comparative curve fit (P = 0.044).

  Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Enflurane effects on evoked release. Enflurane effects on l-[3H]glutamate (A) and [14C]GABA (B) evoked by 4AP from rat cortical synaptosomes were fitted to sigmoidal concentration-effect curves. Control data in the presence of Ca2+ (glutamate, n = 20; GABA, n = 18) or absence of Ca2+ (glutamate, n = 8; GABA, n = 8) are presented as mean ± S.D. Basal effects (taken from Westphalen and Hemmings, 2005) were subtracted from evoked release data, and the difference was fitted to sigmoidal concentration-effect curves (C and D). Shaded area highlights the clinical concentration range (0.5–2× MAC). Symbols in gray present sum ΔFR data (mean ± S.D.) for 2-min pulses of 1 mM 4AP in the presence of Ca2+ plus ∼0.5× MAC (0.39 mM ± 0.007, n = 3), ∼1× MAC (0.74 ± 0.03 mM, n = 5), or ∼2× MAC (1.58 ± 0.02 mM, n = 5) concentrations of enflurane applied continuously, which do not differ from data obtained using the standard pulse application with basal effect subtracted (glutamate, P = 0.09; GABA, P = 0.26). Release in the absence of anesthetic presents mean ± S.D. of all assay controls.

Tetrodotoxin-Insensitive 4-Aminopyridine-Evoked Release. In the presence of Ca2+ and 1 μM tetrodotoxin (a concentration that maximally inhibited 4AP-evoked l-[3H]glutamate and [14C]GABA release; Fig. 4, C and D), the tetrodotoxin-insensitive 4AP-evoked [14C]GABA release was significantly inhibited by 0.7 mM isoflurane (P < 0.001), 1.8 mM enflurane (P < 0.001), or 0.7 mM halothane (P < 0.001) (Fig. 5). Inhibition of the residual tetrodotoxin-insensitive 4AP-evoked l-[3H]glutamate release was not significant for any anesthetic.

Discussion

Isoflurane, enflurane, and halothane all selectively inhibited Ca2+-dependent 4AP-evoked glutamate release compared with GABA release. Taking into consideration anesthetic-specific effects on basal release (Westphalen and Hemmings, 2005), the volatile anesthetics inhibited glutamate release with greater potencies within a clinically relevant concentration range and with full efficacies. This refined analysis clarifies the finding of different efficacies when basal effects are not considered (Westphalen and Hemmings, 2003a). Because depolarization-evoked vesicular transmitter release is Ca2+-dependent, these results support preferential inhibition of evoked vesicular glutamate release over GABA release by a variety of clinically used volatile anesthetics. The effects of volatile anesthetics on both basal and evoked release of the major excitatory and inhibitory transmitters may combine to depress overall central nervous system neuronal activity. The significant actions of volatile anesthetics on basal release highlight the need to consider these effects when analyzing drug effects on evoked release data.

  Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Halothane effects on evoked release. Halothane effects on l-[3H]glutamate (A) and [14C]GABA (B) evoked by 4AP from rat cortical synaptosomes were fitted to sigmoidal concentration-effect curves. Control data in the presence of Ca2+ (glutamate, n = 16; GABA, n = 16) or absence of Ca2+ (glutamate, n = 6; GABA, n = 6) are presented as mean ± S.D. Basal effects (taken from Westphalen and Hemmings, 2005) were subtracted from evoked release data and the difference fitted to sigmoidal concentration-effect curves (C and D). Shaded area highlights the clinical concentration range (0.5–2× MAC). Symbols in gray present sum ΔFR data (mean ± S.D.) for 2-min pulses of 1 mM 4AP in the presence of Ca2+ plus ∼0.5× MAC (0.14 ± 0.005 mM, n = 3), ∼1× MAC (0.33 ± 0.02 mM, n = 5), or ∼2× MAC (0.70 ± 0.01 mM, n = 4) concentrations of halothane applied continuously, which do not differ from data obtained using the standard pulse application with basal effect subtracted (glutamate, P = 0.06; GABA, P = 0.27). Release in the absence of anesthetic presents mean ± S.D. of all assay controls.

Different volatile anesthetic potencies for inhibition of glutamate versus GABA release by volatile anesthetics suggest underlying physiological difference(s) between glutamatergic and GABAergic nerve terminals resulting in distinct transmitter-selective pharmacologies. Voltage-gated Na+ channels are important in the generation of neuronal action potentials that trigger exocytosis of synaptic vesicles (Tibbs et al., 1989; Südhof, 2004). Mammalian Na+ channels show selective densities within and between neurons (Goldin, 2001; Novakovic et al., 2001), can modulate transmitter release (Engelman and MacDermott, 2004), and are inhibited by volatile anesthetics (see below). Thus, inhibition by volatile anesthetics of evoked neurotransmitter release might occur via Na+ channel block. This hypothesis is supported by the present demonstration that 4AP-evoked release is also inhibited by the selective Na+ channel blocker tetrodotoxin. Moreover, volatile anesthetics inhibit 4AP-evoked (action potential-evoked) transmitter release with greater potency than release evoked by elevated K+ (Schlame and Hemmings, 1995; Ratnakumari et al., 2000; Westphalen and Hemmings, 2003a; Hemmings et al., 2005b), which implicates an anesthetic target upstream of Ca2+ entry. Additional support for volatile anesthetic inhibition of mammalian neuronal Na+ channels includes anesthetic inhibition of synaptosomal influx of 22Na+ and [3H]batrachotoxinin-A binding to nerve-terminal Na+ channels (Ratnakumari and Hemmings, 1998), inhibition of heterogeneously expressed brain Na+ channels (Rehberg et al., 1996; Shiraishi and Harris, 2004), and inhibition of Na+ currents in isolated rat dorsal root ganglion neurons (Ratnakumari et al., 2000) and isolated neurohypophysial nerve terminals (Ouyang et al., 2003; Ouyang and Hemmings, 2005).

Isoflurane, enflurane, and halothane completely inhibited Ca2+-dependent 4AP-evoked glutamate and GABA release with potencies comparable to inhibition of mammalian voltage-gated Na+ channels (Rehberg et al., 1996; Ouyang et al., 2003). In contrast, the selective Na+ channel blocker tetrodotoxin completely inhibited Ca2+-dependent 4AP-evoked glutamate release but only partially inhibited GABA release. This is consistent with a previous study that showed tetrodotoxin-resistant Ca2+-dependent 4AP-evoked inhibitory postsynaptic currents in rat hippocampus (Akaike, 2002). Preferential inhibition of autaptic excitatory postsynaptic currents versus inhibitory postsynaptic currents by Na+ channel blockers in hippocampal neuron microcultures (Prakriya and Mennerick, 2000) suggests fundamental differences in the distribution and/or function of Na+ channels between glutamate and GABA nerve terminals. The potency of tetrodotoxin in inhibiting both glutamate and GABA release corresponds to the tetrodotoxin-sensitive Na+ channel subtypes NaV1.1, 1.2, 1.3, 1.6, and/or 1.7 (Novakovic et al., 2001; Lai et al., 2004). Perhaps the tetrodotoxin-insensitive fraction of 4AP-evoked GABA release involves expression of tetrodotoxin-resistant Na+ channel subtypes (e.g., NaV1.8 and/or 1.9; Novakovic et al., 2001; Lai et al., 2004) in a subset of GABAergic terminals. Variability in depolarization threshold and/or release probability of vesicular transmitter exocytosis between nerve terminals containing the same (Rosenmund et al., 1993; Murthy et al., 1997) and/or different transmitters (this study; Prakriya and Mennerick, 2000; Akaike, 2002) may provide the basis for preferential inhibition of evoked glutamate versus GABA release by volatile anesthetics. Because inhibition of 4AP-evoked glutamate and GABA release by tetrodotoxin was neither totally effective nor transmitter-selective in potency, volatile anesthetic effects on transmitter release that were fully efficacious and transmitter-selective can only partially be attributable to Na+ channel inhibition or require the distinct state-dependent Na+ channel-blocking effects of volatile anesthetics (Ouyang et al., 2003). The ability of clinical concentrations of volatile anesthetics to inhibit the tetrodotoxin-insensitive component of 4AP-evoked GABA release further supports the involvement of additional mechanism(s). GABAergic synaptic diversity in release probability, number of release sites, presynaptic GABAA receptor subtypes, modulation by endogenous and exogenous factors (Cherubini and Conti, 2001), and tetrodotoxin-insensitive nicotinic-mediated increases in GABA release (Zhu and Chiappinelli, 2002) suggest possible mechanisms for transmitter-selective anesthetic effects.

  Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Tetrodotoxin effects on evoked release. Tetrodotoxin effects on l-[3H]glutamate (A) and [14C]GABA (B) evoked by 4AP from rat cortical synaptosomes were fitted to sigmoidal concentration-effect curves. Control data in the presence of Ca2+ (glutamate, n = 7; GABA, n = 7) or absence of Ca2+ (glutamate, n = 6; GABA, n = 6) are presented as mean ± S.D. Basal effects (taken from Westphalen and Hemmings, 2005) were subtracted from evoked release data and the difference fitted to sigmoidal concentration-effect curves (C and D). Tetrodotoxin did not completely inhibit l-[3H]glutamate or GABA release in the presence of Ca2+, or significantly inhibit Ca2+-independent l-[3H]glutamate or GABA release. Symbols in gray present sum ΔFR data (mean ± S.D.) for 2-min pulses of 1 mM 4AP in the presence of Ca2+ and 10 nM (n = 4), 100 nM (n = 4), or 1 μM (n = 6) tetrodotoxin applied continuously, which do not differ from data obtained using the standard pulse application with basal effect subtracted (glutamate, P = 0.16; GABA, P = 0.10). Release in the absence of anesthetic presents mean ± of all assay controls.

  Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

Additivity experiments for the effects of volatile anesthetics plus tetrodotoxin on release. Effects of ∼2× MAC isoflurane (0.73 ± 0.02 mM, n = 4), enflurane (1.79 ± 0.05 mM, n = 4), or halothane (0.69 ± 0.12 mM, n = 4) on tetrodotoxin (TTX)-insensitive 4AP-evoked release of l-[3H]glutamate and [14C]GABA in the presence of Ca2+ are shown as mean ± S.E.M. TTX (1 μM, n = 8) significantly inhibited 4AP-evoked l-[3H]glutamate and [14C]GABA release versus control (†††, P < 0.001, n = 8). In the presence of 1 μM TTX, anesthetics significantly inhibited [14C]GABA release (**, P < 0.01), with no significant additive effect on l-[3H]glutamate release.

Other potential presynaptic targets of volatile anesthetics include K+ and Ca2+ channels (Topf et al., 2003; Franks and Honoré, 2004), ligand-gated ion channels (Franks and Lieb, 1994), cell signaling machinery, such as protein kinase C (Hemmings and Adamo, 1998) and G-protein coupled receptors (Yamakura et al., 2001), and synaptic vesicle fusion machinery (Hawasli et al., 2004), all of which can modulate transmitter release (Miller, 1998; Engelman and MacDermott, 2004). Genetic inactivation of components of the presynaptic release machinery differentially affect basal and evoked release (SNAP-25; Washbourne et al., 2001) or evoked glutamate and GABA release (Munc 13-1; Augustin et al., 1999), suggesting these proteins as potential targets for the pathway and transmitter-selective effects of volatile anesthetics. Like Na+ channels, these putative anesthetic targets may have distinct patterns of expression within and between neurons that may contribute to transmitter-selective effects.

Preferential inhibition of Ca2+-independent 4AP-evoked GABA versus glutamate release by isoflurane and enflurane, but not halothane, was in contrast to preferential inhibition of Ca2+-dependent glutamate versus GABA release by all three anesthetics tested. This suggests more than one presynaptic site of action for these effects. A potential target for anesthetic effects on Ca2+-independent transmitter release is inhibition of reverse transmitter transporter function (Attwell et al., 1993). Inhibition by isoflurane of glutamate uptake into rat cerebrocortical synaptosomes occurs with similar potency (IC50 = 0.7 mM) to inhibition of Ca2+-independent release, but inhibition of GABA uptake (IC50 = 0.8 mM) is less potent than inhibition of Ca2+-independent release (Westphalen and Hemmings, 2003b). The maximal efficacy of isoflurane inhibition of glutamate and GABA uptake also varies (Westphalen and Hemmings, 2003b), consistent with a greater role of transporter inhibition to glutamate release than GABA release. In view of the insensitivity of GABA uptake to volatile anesthetics, the greater sensitivity of Ca2+-independent 4AP-evoked GABA release to isoflurane further suggests that nerve-terminal GABA transporters are not sensitive to volatile anesthetics.

In summary, three volatile anesthetics consistently inhibited Ca2+-dependent 4AP-evoked glutamate release with greater potency than GABA release, which supports distinct presynaptic targets between glutamatergic and GABAergic nerve terminals. Although glutamate is released from synaptic vesicles at saturating concentrations for its receptors, even moderate reductions in release probability could affect the kinetics of synaptic glutamate concentration (Cavelier et al., 2005) and the release probability of low probability nerve terminals (Prakriya and Mennerick, 2000) to depress excitatory transmission. The less potent inhibition of GABA release by volatile anesthetics is balanced by postsynaptic potentiation of GABAA receptors, such that net inhibitory GABAergic transmission is enhanced (Stucke et al., 2002). The correlation between anesthetic potency in vivo (MAC) and potency of evoked glutamate release inhibition supports this action as a relevant volatile anesthetic target. The differing potencies for inhibition of Ca2+-dependent and Ca2+-independent 4AP-evoked GABA and glutamate release provide evidence for multiple transmitter and pathway-specific presynaptic targets. The biological difference(s) between glutamatergic and GABAergic nerve terminals that underlie this selectivity in volatile anesthetic action remains to be determined. Comparable transmitter-selective inhibition was not produced by tetrodotoxin, indicating that a mechanism(s) in addition to Na+ channel block must be involved. Selective depression of evoked glutamate versus GABA release depressed basal glutamate relative to GABA release, and potentiation of postsynaptic GABAA receptors provide synergistic mechanisms by which volatile anesthetics may depress excitatory and enhance net inhibitory central nervous system transmission.

Footnotes

  • This work was supported by National Institutes of Health Grant GM 58055 and by Department of Anesthesiology, Weill Medical College, Cornell University.

  • doi:10.1124/jpet.105.090662.

  • ABBREVIATIONS: 4AP, 4-aminopyridine; MAC, minimum alveolar concentration; FR, fractional release.

    • Received June 7, 2005.
    • Accepted September 15, 2005.
  • The American Society for Pharmacology and Experimental Therapeutics

References

  1. ↵
    Akaike N (2002) Focal stimulation of single GABA-ergic presynaptic boutons on the rat hippocampal neuron. Neurosci Res 42: 187–195.
    OpenUrlCrossRefPubMed
  2. ↵
    Attwell D, Barbour B, and Szatkowski M (1993) Nonvesicular release of neurotransmitter. Neuron 11: 401–407.
    OpenUrlCrossRefPubMed
  3. ↵
    Augustin I, Rosenmund C, Südhof TC, and Brose N (1999) Munc13-1 is essential for fusion competence of glutamatergic synaptic vesicles. Nature (Lond) 400: 457–461.
    OpenUrlCrossRefPubMed
  4. ↵
    Cavelier P, Hamann M, Rossi D, Mobbs P, and Attwell D (2005) Tonic excitation and inhibition of neurons: ambient transmitter sources and computational consequences. Prog Biophys Mol Biol 87: 3–16.
    OpenUrlCrossRefPubMed
  5. ↵
    Cherubini E and Conti F (2001) Generating diversity at GABAergic synapses. Trends Neurosci 24: 155–162.
    OpenUrlCrossRefPubMed
  6. ↵
    Engelman HS and MacDermott AB (2004) Presynaptic ionotropic receptors and control of transmitter release. Nat Rev Neurosci 5: 135–145.
    OpenUrlCrossRefPubMed
  7. ↵
    Franks NP and Honoré E (2004) The TREK K2P channels and their role in general anaesthesia and neuroprotection. Trends Pharmacol Sci 25: 601–608.
    OpenUrlCrossRefPubMed
  8. ↵
    Franks NP and Lieb WR (1994) Molecular and cellular mechanisms of general anaesthesia. Nature (Lond) 367: 607–613.
    OpenUrlCrossRefPubMed
  9. ↵
    Garcia-Sanz A, Badia A, and Clos MV (2001) Superfusion of synaptosomes to study presynaptic mechanisms involved in neurotransmitter release from rat brain. Brain Res Protoc 7: 94–102.
    OpenUrlCrossRefPubMed
  10. ↵
    Goldin AL (2001) Resurgence of sodium channel research. Ann Rev Physiol 63: 871–894.
    OpenUrlCrossRefPubMed
  11. ↵
    Hawasli AH, Saifee O, Liu C, Nonet ML, and Crowder CM (2004) Resistance to volatile anesthetics by mutations enhancing excitatory neurotransmitter release in Caenorhabditis elegans. Genetics 168: 831–843.
    OpenUrlAbstract/FREE Full Text
  12. ↵
    Hemmings HC Jr and Adamo AIB (1998) General anesthetic effects on protein kinase C. Toxicol Lett 100–101: 89–95.
    OpenUrl
  13. ↵
    Hemmings HC Jr, Akabas MH, Goldstein PA, Trudell JR, Orser BA, and Harrison NL (2005a) Emerging molecular mechanisms of general anesthetic action. Trends Pharmacol Sci 26: 503–510.
    OpenUrlCrossRefPubMed
  14. ↵
    Hemmings HC Jr, Yan W, Westphalen RI, and Ryan TA (2005b) The general anesthetic isoflurane depresses synaptic vesicle exocytosis. Mol Pharmacol 67: 1–9.
    OpenUrlFREE Full Text
  15. ↵
    Jones MV, Brooks PA, and Harrison NL (1992) Enhancement of γ-aminobutyric acid-activated Cl– currents in cultured rat hippocampal neurones by three volatile anaesthetics. J Physiol (Lond) 449: 279–293.
    OpenUrlPubMed
  16. ↵
    Lai J, Porreca F, Hunter JC, and Gold MS (2004) Voltage-gated sodium channels and hyperalgesia. Annu Rev Pharmacol Toxicol 44: 371–397.
    OpenUrlCrossRefPubMed
  17. ↵
    MacIver MB (1997) General anesthetic actions on transmission at glutamate and GABA synapses, in Anesthesia: Biological Foundations (Biebuyck JF, Lynch CI, Maze M, Saidman LJ, Yaksh TL and Zapol WM eds) pp 277–286, Lippincott-Raven, New York.
  18. ↵
    Miller RJ (1998) Presynaptic receptors. Annu Rev Pharmacol Toxicol 38: 201–227.
    OpenUrlCrossRefPubMed
  19. ↵
    Murthy VN, Sejnowski TJ, and Stevens CF (1997) Heterogeneous release properties of visualized individual hippocampal synapses. Neuron 18: 599–612.
    OpenUrlCrossRefPubMed
  20. ↵
    Novakovic SD, Eglen RM, and Hunter JC (2001) Regulation of Na+ channel distribution in the nervous system. Trends Neurosci 24: 473–478.
    OpenUrlCrossRefPubMed
  21. ↵
    Ouyang W and Hemmings HC Jr (2005) Depression by isoflurane of the action potential and underlying voltage-gated ion currents in isolated rat neurohypophysial nerve terminals. J Pharmacol Exp Ther 312: 801–808.
    OpenUrlAbstract/FREE Full Text
  22. ↵
    Ouyang W, Wang G, and Hemmings HC Jr (2003) Isoflurane and propofol inhibit voltage-gated sodium channels in isolated rat neurohypophysial nerve terminals. Mol Pharmacol 64: 373–381.
    OpenUrlAbstract/FREE Full Text
  23. ↵
    Prakriya M and Mennerick S (2000) Selective depression of low-release probability excitatory synapses by sodium channel blockers. Neuron 26: 671–682.
    OpenUrlCrossRefPubMed
  24. ↵
    Ratnakumari L and Hemmings HC Jr (1998) Inhibition of presynaptic sodium channels by halothane. Anesthesiology 88: 1043–1054.
    OpenUrlCrossRefPubMed
  25. ↵
    Ratnakumari L, Vysotskaya TN, Duch DS, and Hemmings HC Jr (2000) Differential effects of anesthetic and nonanesthetic cyclobutanes on neuronal voltage-gated sodium channels. Anesthesiology 92: 529–541.
    OpenUrlPubMed
  26. ↵
    Rehberg B, Xiao YH, and Duch DS (1996) Central nervous system sodium channels are significantly suppressed at clinical concentrations of volatile anesthetics. Anesthesiology 84: 1223–1233.
    OpenUrlCrossRefPubMed
  27. ↵
    Rosenmund C, Clements JD, and Westbrook GL (1993) Non-uniform probability of glutamate release at a hippocampal synapse. Science (Wash DC) 262: 754–757.
    OpenUrlAbstract/FREE Full Text
  28. ↵
    Schlame M and Hemmings HC Jr (1995) Inhibition by volatile anesthetics of endogenous glutamate release from synaptosomes by a presynaptic mechanism. Anesthesiology 82: 1406–1416.
    OpenUrlCrossRefPubMed
  29. ↵
    Shiraishi M and Harris RA (2004) Effects of alcohols and anesthetics on recombinant voltage-gated Na+ channels. J Pharmacol Exp Ther 309: 987–994.
    OpenUrlAbstract/FREE Full Text
  30. ↵
    Stucke AG, Stuth EAE, Tonkovic-Capin V, Tonkovic-Capin M, Hopp FA, Kampine JP, and Zuperku EJ (2002) Effects of halothane and sevoflurane on inhibitory neurotransmission to medullary expiratory neurons in a decerebrated dog model. Anesthesiology 96: 955–962.
    OpenUrlCrossRefPubMed
  31. ↵
    Südhof TC (2004) The synaptic vesicle cycle. Ann Rev Neurosci 27: 509–547.
    OpenUrlCrossRefPubMed
  32. ↵
    Tibbs GR, Barrie AP, Van Mieghem FJE, McMahon HT, and 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
  33. ↵
    Topf N, Recio-Pinto E, Blanck TJJ, and Hemmings HC Jr (2003) Actions of general anesthetics on voltage-gated ion channels, in Neural Mechanisms of Anesthesia (Antognini JF, Carstens EE, and Raines DE eds) pp 299–318, Humana Press Inc., Totowa, NJ.
  34. ↵
    Washbourne P, Thompson PM, Carta M, Costa ET, Mathews JR, Lopez-Bendito G, Molnár Z, Becher MW, Valenzuela CF, Partridge LD, et al. (2001) Genetic ablation of the t-SNARE SNAP-25 distinguishes mechanisms of neuroexocytosis. Nature Neurosci 5: 19–26.
    OpenUrl
  35. ↵
    Westphalen RI and Hemmings HC Jr (2003a) Selective depression by general anesthetics of glutamate versus GABA release from isolated cortical nerve terminals. J Pharmacol Exp Ther 304: 1188–1196.
    OpenUrlAbstract/FREE Full Text
  36. ↵
    Westphalen RI and Hemmings HC Jr (2003b) The effects of isoflurane and propofol on glutamate and GABA transporters in isolated cortical nerve terminals. Anesthesiology 98: 364–372.
    OpenUrlCrossRefPubMed
  37. ↵
    Westphalen RI and Hemmings HC Jr (2005) Volatile anesthetic effects on glutamate versus GABA release from isolated rat cortical nerve terminals: basal release. J Pharmacol Exp Ther 316: 208–215.
    OpenUrlCrossRefPubMed
  38. ↵
    Yamakura T, Lewohl JM, and Harris RA (2001) Differential effects of general anesthetics on G protein-coupled inwardly rectifying and other potassium channels. Anesthesiology 95: 144–153.
    OpenUrlCrossRefPubMed
  39. ↵
    Zhang Y, Sonner JM, Eger EII, Stabernack CR, Laster MJ, Raines DE, and Harris RA (2004) Gamma-aminobutyric acidA receptors do not mediate the immobility produced by isoflurane. Anesth Analg 99: 85–90.
    OpenUrlCrossRefPubMed
  40. ↵
    Zhu PJ and Chiappinelli VA (2002) Nicotinic receptors mediate increase GABA release in brain through a tetrodotoxin-insensitive mechanism during prolonged exposure to nicotine. Neuroscience 115: 137–144.
    OpenUrlCrossRefPubMed
PreviousNext
Back to top

In this issue

Journal of Pharmacology and Experimental Therapeutics: 316 (1)
Journal of Pharmacology and Experimental Therapeutics
Vol. 316, Issue 1
1 Jan 2006
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover
  • Index by author
  • Back Matter (PDF)
  • Editorial Board (PDF)
  • Front Matter (PDF)
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.
Volatile Anesthetic Effects on Glutamate versus GABA Release from Isolated Rat Cortical Nerve Terminals: 4-Aminopyridine-Evoked Release
(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

Volatile Anesthetic Effects on Glutamate versus GABA Release from Isolated Rat Cortical Nerve Terminals: 4-Aminopyridine-Evoked Release

Robert I. Westphalen and Hugh C. Hemmings
Journal of Pharmacology and Experimental Therapeutics January 1, 2006, 316 (1) 216-223; DOI: https://doi.org/10.1124/jpet.105.090662

Citation Manager Formats

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

Share
Research ArticleNEUROPHARMACOLOGY

Volatile Anesthetic Effects on Glutamate versus GABA Release from Isolated Rat Cortical Nerve Terminals: 4-Aminopyridine-Evoked Release

Robert I. Westphalen and Hugh C. Hemmings
Journal of Pharmacology and Experimental Therapeutics January 1, 2006, 316 (1) 216-223; DOI: https://doi.org/10.1124/jpet.105.090662
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