Background

With the exception of gamma-aminobutyric acidA (GABAA) receptors, the major molecular targets underlying the anesthetizing actions of thiopental have yet to be established. Neuronal nicotinic acetylcholine receptors (nAChRs) are closely related to GABAA receptors and hence might also be major targets. If so, they might be expected to be substantially inhibited by surgical concentrations (EC50 = 25 micrometer) of thiopental and to display the same stereoselectivity as does general anesthesia.

Methods

Neuronal alpha4beta2, neuronal alpha7 and muscle alphabetagammadelta nAChRs were expressed in Xenopus oocytes. Peak acetylcholine-activated currents were measured at -70 mV using the two-electrode voltage clamp technique. Racemic thiopental and its two optical isomers were applied with and without preincubation and at high and low concentrations of acetylcholine.

Results

Inhibition of all three nAChRs was enhanced by preincubation with thiopental, a protocol that mimics the pharmacologic situation in vivo. Using this protocol, inhibition was further enhanced by high concentrations of acetylcholine, with IC50 = 18 +/- 2, 34 +/- 4, and 20 +/- 2 micrometer (mean +/- SEM) thiopental for the neuronal alpha4beta2, neuronal alpha7 and muscle alphabetagammadelta nAChRs, respectively, with Hill coefficients near unity. Neither the neuronal alpha7 nor the muscle alphabetagammadelta nAChR differentiated between the optical isomers of thiopental. However, R(+)-thiopental was significantly more effective than the S(-) isomer at inhibiting the neuronal alpha4beta2 nAChR; interestingly, this is diametrically opposite to their stereoselectivity for general anesthesia.

Conclusions

Both central neuronal and peripheral muscle nAChRs can be substantially inhibited by thiopental at surgical EC50 concentrations but with either no stereoselectivity or one opposite to that for general anesthesia. Thus, nAChRs are probably not crucial targets for producing thiopental anesthesia, although nAChRs may play a part in the side effects produced by this agent.

HISTORICALLY, thiopental has been one of the most popular and important of the intravenous general anesthetics in routine clinical use. Nonetheless, with the exception of γ-aminobutyric acidA(GABAA) receptors, 1–4the molecular targets that underlie its anesthetizing actions remain unknown. We consider here, as possible and plausible targets, neuronal nicotinic acetylcholine receptors (nAChRs) because they belong to the same superfamily of fast neurotransmitter-gated receptors as do GABAAreceptors. 5–7We have expressed in Xenopus  oocytes what are possibly the major vertebrate central nervous system (CNS) species 8of neuronal heteromeric (α4β2) and neuronal homomeric (α7) nAChRs, together with (for comparison) the muscle heteromeric (αβγδ) nAChR. Recent studies 9,10with this expression system have found the neuronal α4β2 (but not the neuronal α7 nor the muscle αβγδ) nAChR to be extremely sensitive to inhibition by volatile anesthetics.

Thiopental is invariably administered surgically as an equimolar (racemic) mixture of its two optical isomers (enantiomers), R  (+) and S  (−), although the S  (−) enantiomer appears to be approximately twice as potent as the R  (+) enantiomer at abolishing the righting reflex in mice. 11–13The scarcity of experimental data on the effects of the pure thiopental enantiomers is largely a result of their being unavailable commercially and their chemical syntheses being difficult. Fortunately, it is now feasible to prepare pure R  (+)- and S  (−)-thiopental by separation of racemic thiopental on a chiral column. 2Indeed, these enantiomers have been prepared and used in this laboratory to study the stereoselective interactions of thiopental with GABAAreceptors. Using pure enantiomers of not only thiopental, but also of pentobarbital and hexobarbital, it was found 2that the rank orders and degrees of chirality observed for potentiating the activity of GABAAreceptors matched those for barbiturate general anesthesia. (A similar study using GABAAreceptors and chemically synthesized thiopental enantiomers was subsequently published. 14)

Many criteria can be used to judge the relevance of a putative target to the anesthetizing actions of a particular agent. 1,15Here we use two particular criteria: anesthetic sensitivity and stereoselectivity. For a given neuronal nAChR and thiopental, these criteria lead to the following tests for relevance: (1) The activity of the nAChR should be substantially inhibited by a clinical EC50concentration of thiopental (approximately 25 μm 1). (2) This inhibition should not only be stereoselective, but S  (−)-thiopental should be substantially more effectual than R  (+)-thiopental. Here we describe the design and results of experiments that test these predictions for different classes of nAChR channels.

This study conforms to the United Kingdom Animals (Scientific Procedures) Act of 1986.

Preparation and Injection of Xenopus  Oocytes

Adult female Xenopus laevis  frogs (Blades Biological, Cowden, Kent, United Kingdom) were maintained in fresh-water holding tanks with a 12-h light–dark cycle in a temperature-controlled room at 20–22°C. Animals were anesthetized by immersion in a 0.2% (wt/vol) solution of tricaine (3-aminobenzoic acid ethyl ester methanesulfonate). A surgical incision into the abdomen was made, and portions of the ovaries containing oocytes were removed and teased apart with forceps. The ovarian fragments were then briefly washed in calcium-free oocyte incubation buffer (Ca2+-free OIB; composition: 88 mm NaCl, 1 mm KCl, 2.4 mm NaHCO3, 0.8 mm MgSO4, 15 mm HEPES, titrated to pH 7.5 with NaOH). Individual oocytes were obtained by digestion of these fragments in Ca2+-free OIB containing 2 mg/ml of type 1A collagenase (Sigma Chemical Co., Poole, Dorset, United Kingdom), for 3 h at room temperature. The oocytes were then carefully washed in Ca2+-free OIB to remove all traces of collagenase, before being transferred into normal OIB (composition 88 mm NaCl, 1 mm KCl, 2.4 mm NaHCO3, 0.8 mm MgSO4, 0.4 mm CaCl2, 0.3 mm Ca(NO3)2, 15 mm HEPES, titrated to pH 7.5 with NaOH). Oocytes were then scrutinized under a dissection microscope, and healthy eggs at stages 5–6 of development were set aside for injection with either DNA (chick α7 subunit; rat α4and β2 subunits) or RNA (mouse α, β, γ, and δ subunits). Injections of cDNA were made with 10 nl of diethyl pyrocarbonate–treated water containing 0.1– 1.0 pg of the appropriate cDNA directly into the nucleus of the oocyte, using the “blind” method. In the case of the α4β2 combination, a premixed solution with an equimolar ratio of the α4 and β2 cDNAs was used. Injections of cRNA were made into the cytosol of the eggs with 50 nl of diethyl pyrocarbonate–treated water containing a total of 10–40 ng of the cRNAs premixed at an equimolar ratio of the α, β, γ, and δ subunits. All injections were conducted with use of calibrated micropipettes (10–16-μm tip diameter) in combination with a Picospritzer II valve (General Valve Corp., Fairfield, NJ), which provided short pressure pulses of nitrogen gas to expel known volumes of the nucleic acid solutions from the micropipettes. Injected oocytes were maintained in a cooled incubator at 19–20°C in normal OIB supplemented with antibiotics (penicillin 100 IU/ml, streptomycin 100 μg/ml; Life Technologies, Paisley, United Kingdom) for 2–7 days before use. All chemicals, unless otherwise stated, were obtained from Sigma Chemical Company (Poole, Dorset, United Kingdom).

The nAChR cDNAs used here were kindly supplied as follows. Rat neuronal nAChR cDNAs for the α4 and β2 subunits (in the pSM expression vector) were from Jim Patrick (Baylor College of Medicine, Houston, TX). Mouse muscle nAChR cDNAs encoding the α, β, γ, and δ subunits were from Jim Boulter (Salk Institute, San Diego, CA), contained in either the pSP64 (γ) or pSP65 (α, β, and δ) vectors. The chick α7 cDNA was supplied by Marc Ballivet (Université de Genève, Switzerland) in the “flip” expression vector. 16The pSM- and “flip”-derived plasmids contain SV40 promoter sequences, which allows direct use of these cDNAs by the nuclear injection method. However, in the case of the mouse muscle subunits, individual cRNA transcripts suitable for injection were prepared from the cDNAs supplied with use of a Riboprobe System cRNA production kit (Promega, Southampton, United Kingdom).

Recording Saline and Drugs Used

The extracellular saline used for electrophysiological recordings had the following composition: 110 mm NaCl, 2 mm KCl, 1 mm MgCl2, 2 mm BaCl2, and 10 mm HEPES (titrated to pH 7.5 with NaOH). Additionally, atropine (100 nm) was added to all solutions to suppress the activity of endogenous muscarinic receptors. At this concentration atropine did not affect the acetylcholine-evoked nicotinic currents. The R  (+) and S  (−) enantiomers of thiopental were prepared from racemic thiopental stock (94.3% sodium salt–sodium carbonate mixture) to a purity level of 99.0% or more, with use of high-performance liquid chromatography with a permethylated β-cyclodextrin column, as described elsewhere. 2Acetylcholine chloride and thiopental solutions were prepared freshly from stock salts on the day of the experiment.

Recording Technique for Xenopus  Oocytes

Inward currents evoked by bath application of acetylcholine chloride (0.3 μm to 10.0 mm) were recorded at −70 mV from oocytes expressing the desired subunit combinations and situated in a bath (volume ≈ 50 μl), where they were continuously perfused at a rate of 2 ml/min with appropriate solutions. Currents were recorded with use of an Axoclamp 2A amplifier (Axon Instruments, Foster City, CA) in the two-electrode voltage clamp mode. Recording electrodes were made from thin-walled filamented borosilicate glass capillary tubes (GC150TF, Clark Electromedical Instruments, Reading, Berkshire, United Kingdom); the passive electrode was filled with 3 m KCl and the current-passing electrode with 3 m KCl supplemented with BAPTA [1,2-bis(2-aminophenoxy)ethane-N,N,N’,N’-tetraacetic acid, tetrapotassium salt, 100 mm]. These electrodes typically had resistances of 0.4–0.8 MΩ. Currents were filtered at 10 Hz (−3 dB, 8-pole Bessel filter; Model 902, Frequency Devices, Haverhill, MA) before being digitized (sampling rate 20 Hz) and stored on a computer for offline analysis. All experiments were performed at room temperature (21–23°C).

Current responses were evoked by continuous application of acetylcholine-containing solutions until a well-defined peak in the response was observed (usually at 20–30 s); this peak was then used as the measure of receptor activity. During data collection a standard acetylcholine concentration was chosen and applied alternately throughout the experiment, to allow the data to be normalized and to correct for any fluctuation in the control response. Two protocols were used concurrently to assess the effects of thiopental. In the first instance, a protocol was used where appropriate concentrations of acetylcholine were simply coapplied with the various concentrations of thiopental. In the second, or preapplied protocol, the oocyte was bathed in the appropriate concentration of thiopental for 1 min before its coapplication with acetylcholine. (The effect of preincubation was complete after approximately 30 s.) In both cases, application of acetylcholine was maintained until an unambiguous peak in the response was observed. Such applications were repeated to ensure consistency of the response. Thiopental by itself had no significant effect on the resting current. Inhibition by thiopental of the response to acetylcholine was shown to be fully reversible at the thiopental concentrations used here. Inhibition of the acetylcholine-evoked current by thiopental was calculated by averaging controls before and after application of anesthetic and comparing these mean control values with the current responses in the presence of thiopental.

Data Analysis

Control acetylcholine concentration–response data were fitted to a standard activation Hill equation of the form where E is the peak acetylcholine-induced current expressed as a percentage of the maximal current, A is the concentration of acetylcholine, n H  is the Hill coefficient, and EC 50  is the acetylcholine concentration that gives a half-maximal effect.

Similarly, data describing the inhibition of the acetylcholine-evoked response by thiopental were fitted to an inhibition Hill equation of the form where y is the percentage of the control acetylcholine current remaining in the presence of thiopental at a concentration I, n H  is the Hill coefficient, and IC 50  is the concentration of thiopental required to inhibit the control response by 50%.

Statistical Analysis

Numerical fits to these Hill equations were obtained using the method of unweighted least-squares (Marquardt-Levenberg algorithm, KaleidaGraph version 3.0.2, Abelbeck Software, Reading, PA) applied to individual (not mean) data points. Values throughout are given as mean ± SEM, and statistical significance was assessed with use of the Student t  test.

Control Acetylcholine Concentration–Response Characteristics

Acetylcholine concentration–response relationships were determined for Xenopus  oocytes expressing rat neuronal heteromeric α4β2, chick neuronal homomeric α7, and mouse muscle heteromeric αβγδ nAChRs. For each nAChR, the data were fitted to the activation Hill equation (equation 1) with the results given in figure 1and table 1. The insets to figure 1show typical inward current responses to acetylcholine; note the differing degrees of desensitization, depending on receptor type and acetylcholine concentration. The EC 50  concentrations and Hill coefficients (n H ) for the relatively slowly desensitizing heteromeric α4β2 and αβγδ receptors are comparable to those reported previously from this laboratory. 9For the homomeric α7 receptor, our EC 50  value is in reasonable agreement with previously published values, 16–18although these are all, almost certainly, substantial overestimates (perhaps by a factor of 10). This is an unavoidable consequence of combining fast α7 receptor desensitization with the slow-mixing rates characteristic of the Xenopus  oocyte perfusion system, and it is because of the peak of the acetlycholine receptor response occurring before the steady state acetylcholine concentration is attained. 19Experiments (not shown) comparing the time courses of receptor activation and solution exchange suggested that this was a problem for the homomeric α7 but not for the heteromeric α4β2 and αβγδ nAChRs.

Fig. 1. Acetylcholine concentration–response data for oocytes expressing (A ) α4β2 neuronal receptors, (B ) α7 neuronal receptors, and (C ) αβγδ muscle receptors. Each data point gives the mean peak current from at least four oocytes, expressed as a percentage of the maximum current. The errors are SEM and where not shown are smaller than the size of the symbols. The lines are least squares fits (table 1) to the activation Hill equation (equation 1). The insets show typical current traces in response to acetylcholine delivered (solid bars) at the stated concentrations; the calibration bars refer to both left and right insets.

Fig. 1. Acetylcholine concentration–response data for oocytes expressing (A ) α4β2 neuronal receptors, (B ) α7 neuronal receptors, and (C ) αβγδ muscle receptors. Each data point gives the mean peak current from at least four oocytes, expressed as a percentage of the maximum current. The errors are SEM and where not shown are smaller than the size of the symbols. The lines are least squares fits (table 1) to the activation Hill equation (equation 1). The insets show typical current traces in response to acetylcholine delivered (solid bars) at the stated concentrations; the calibration bars refer to both left and right insets.

Close modal

Table 1. EC50Concentrations and Hill Coefficients for Activation of Nicotinic Acetylcholine Receptor Channels by Acetylcholine

Activation Hill equation parameters are given as mean ± SEM for the specified degrees of freedom and were obtained from fits of the data to equation 1.

Table 1. EC50Concentrations and Hill Coefficients for Activation of Nicotinic Acetylcholine Receptor Channels by Acetylcholine
Table 1. EC50Concentrations and Hill Coefficients for Activation of Nicotinic Acetylcholine Receptor Channels by Acetylcholine

Racemic Thiopental Concentration–Response Relationships for Inhibition of the Nicotinic Acetylcholine Receptors

For each nAChR, thiopental concentration-dependent inhibition of acetylcholine-activated inward currents was determined during four conditions: at a low and a high concentration of acetylcholine, both with and without preincubation with thiopental. The low and high ACh concentrations for each nAChR (table 2) differed by at least an order of magnitude, with the low concentration being selected for experimental convenience as one that activated conveniently large but essentially nondesensitizing currents (for the heteromeric α4β2 and αβγδ receptors) or minimally desensitizing but still measurable currents (for the homomeric α7 receptor). Taking into account a possible 10-fold overestimate of our EC 50  for the α7 receptor (see above), the low acetylcholine concentrations for all three nAChRs turned out to be near to their actual EC 50  concentrations.

Table 2. IC50Concentrations and Hill Coefficients for Inhibition of ACh-activated Nicotinic ACh Receptor Channels by Racemic Thiopental

Inhibitory Hill equation parameters are given as mean ± SEM for the specified degrees of freedom and were obtained from fits of the data to equation 2.

ACh = acetylcholine; NS = no significant fit.

Table 2. IC50Concentrations and Hill Coefficients for Inhibition of ACh-activated Nicotinic ACh Receptor Channels by Racemic Thiopental
Table 2. IC50Concentrations and Hill Coefficients for Inhibition of ACh-activated Nicotinic ACh Receptor Channels by Racemic Thiopental

Preincubation with racemic thiopental of oocytes expressing nAChRs increased their sensitivity to inhibition by this agent (fig. 2and table 2). This was true for all three nAChRs at both low and high concentrations of acetylcholine, with the effects of preincubation being greater at the high acetylcholine concentrations. These enhancements of inhibition could be surprisingly large. For example, consider the neuronal α4β2 nAChR. The IC 50  for thiopental at the high (1,000 μm) concentration of acetylcholine decreased from 204 ± 38 to 18 ± 2 μm thiopental because of preincubation, whereas the corresponding decrease at the low (100 μm) concentration of acetylcholine was from 81 ± 5 to 26 ± 2 μm thiopental. (For comparison, EC 50 = 25 μm thiopental for clinical general anesthesia, using as the anesthetic end point a lack of response of patients to a painful stimulus. 1)

Fig. 2. Concentration–response data showing inhibition by thiopental of (A ) α4β2 neuronal receptors, (B ) α7 neuronal receptors, and (C ) αβγδ muscle receptors. Each data point gives the mean peak current from at least four oocytes, normalized to 100 for the control response in the absence of thiopental. The errors are SEM and where not shown are smaller than the size of the symbols. The open symbols are for the protocol of simply coapplying thiopental with acetylcholine, whereas the filled symbols refer to the pharmacologically more relevant protocol of preapplying thiopental for 1 min before coapplying it with acetylcholine; notice that preapplying thiopental almost invariably resulted in more inhibition. The circles and triangles refer, respectively, to low and high concentrations of acetylcholine: (A ) 100 and 1,000 μm, (B ) 30 and 400 μm, (C ) 10 and 300 μm. The lines are least squares fits to the inhibition Hill equation (equation 2and table 2). Each inset shows four typical traces (identified by arrows to the current peaks) for currents from an oocyte activated by the low concentration of acetylcholine, in the absence (initial control and control on wash) and presence (coapplied and preapplied) of 25 μm thiopental (the human EC50for general anesthesia).

Fig. 2. Concentration–response data showing inhibition by thiopental of (A ) α4β2 neuronal receptors, (B ) α7 neuronal receptors, and (C ) αβγδ muscle receptors. Each data point gives the mean peak current from at least four oocytes, normalized to 100 for the control response in the absence of thiopental. The errors are SEM and where not shown are smaller than the size of the symbols. The open symbols are for the protocol of simply coapplying thiopental with acetylcholine, whereas the filled symbols refer to the pharmacologically more relevant protocol of preapplying thiopental for 1 min before coapplying it with acetylcholine; notice that preapplying thiopental almost invariably resulted in more inhibition. The circles and triangles refer, respectively, to low and high concentrations of acetylcholine: (A ) 100 and 1,000 μm, (B ) 30 and 400 μm, (C ) 10 and 300 μm. The lines are least squares fits to the inhibition Hill equation (equation 2and table 2). Each inset shows four typical traces (identified by arrows to the current peaks) for currents from an oocyte activated by the low concentration of acetylcholine, in the absence (initial control and control on wash) and presence (coapplied and preapplied) of 25 μm thiopental (the human EC50for general anesthesia).

Close modal

Using the thiopental preapplication protocol (which is arguably more relevant pharmacologically; see Discussion), for all three nAChRs the racemic thiopental concentration–response curves were shifted to the left by the high concentrations of acetylcholine (fig. 2). This effect, which corresponds to an acetylcholine-induced increase in thiopental sensitivity, was very much more marked for the muscle (αβγδ) nAChR than for the neuronal (α4β2 and α7) nAChRs. All the nAChRs, however, could be sensitive to thiopental. For example, at the high concentrations of acetylcholine, the IC 50  concentrations for inhibition of the three receptors (α4β2, α7, and αβγδ) were 18 ± 2, 34 ± 4, and 20 ± 2 μm thiopental, respectively (table 2). Comparing these IC 50  values with the EC 50  value for general anesthesia (25 μm thiopental), a case can be made for the substantial inhibition of all three nAChRs during maintenance of anesthesia with racemic thiopental.

Effects of the Optical Isomers of Thiopental on Inhibition of the Nicotinic Acetylcholine Receptors

In addition to the aforementioned experiments with racemic thiopental, we separated the two optical isomers (enantiomers) of thiopental, S  (−) and R  (+), and determined their individual effects on each of the three nAChRs at a fixed concentration (25 μm, the EC 50  of the racemate for general anesthesia). For each nAChR we used eight conditions: a low and high concentration of ACh, with and without preapplication of each thiopental enantiomer.

The results of these experiments with the enantiomers are given in figure 3. Neither the neuronal α7 nor the muscle αβγδ nAChR differentiated between the R  (+) and S  (−) enantiomers of thiopental during any of the conditions used (P > 0.5). The α4β2 nAChR, conversely, was consistently more sensitive to the R  (+) enantiomer, and this difference was statistically significant (P < 0.01 or P < 0.05), except during the condition of high acetylcholine concentration and no preincubation with the anesthetic enantiomers. Overall, then, using the pharmacologically relevant preincubation protocol at both high and low concentrations of acetylcholine, thiopental was found to stereoselectively inhibit the neuronal α4β2 but not the neuronal α7 or the muscle αβγδ nAChR. Interestingly, the stereoselectively observed with the rat neuronal α4β2 nAChR is the opposite of that observed for general anesthesia in mice. 11–13 

Fig. 3. Summary of the effects of the thiopental enantiomers (25 μm) on (A ) α4β2 neuronal receptors, (B ) α7 neuronal receptors, and (C ) αβγδ muscle receptors. Each pair of histograms shows the mean ± SEM of currents activated by a low or high concentration of acetylcholine in the presence of (left ) S  (−)- and (right ) R  (+)-thiopental either preapplied or coapplied; the currents are expressed as percentages of the controls in the absence of thiopental. The number of oocytes is given by n, and significant differences between the enantiomers are denoted by *(P < 0.05) or **(P < 0.01).

Fig. 3. Summary of the effects of the thiopental enantiomers (25 μm) on (A ) α4β2 neuronal receptors, (B ) α7 neuronal receptors, and (C ) αβγδ muscle receptors. Each pair of histograms shows the mean ± SEM of currents activated by a low or high concentration of acetylcholine in the presence of (left ) S  (−)- and (right ) R  (+)-thiopental either preapplied or coapplied; the currents are expressed as percentages of the controls in the absence of thiopental. The number of oocytes is given by n, and significant differences between the enantiomers are denoted by *(P < 0.05) or **(P < 0.01).

Close modal

Actions of Intravenous General Anesthetics on Members of a Superfamily of Neurotransmitter-gated Receptor Channels

Nicotinic acetylcholine receptors are members of an anesthetic-sensitive superfamily of structurally and genetically related fast neurotransmitter-gated receptor channels. 5–7This superfamily also includes GABAA, glycine, and 5-hydroxytryptamine type 3 receptors (but not glutamate-gated receptors). Most members of this superfamily are very sensitive to minimum alveolar concentration levels of volatile general anesthetics but show varying susceptibilities to anesthetizing EC 50  concentrations of intravenous general anesthetics such as propofol, etomidate, and thiopental. Low concentrations of propofol and etomidate, for example, substantially potentiate GABAAreceptors, 20–22while having relatively modest effects on glycine receptors, 20,21,235-hydroxytryptamine type 3 receptors 24,25and neuronal nAChRs. 9,26In addition, thiopental and etomidate potentiate GABAAreceptors in a stereoselective manner (propofol is achiral and cannot act stereoselectively), with the same rank orders of potency found for general anesthesia. 2,22These results are consistent with major roles for GABAAreceptors in the production of general anesthesia with propofol, etomidate, or thiopental.

Thiopental, however, appears to be a more promiscuous intravenous agent than either propofol or etomidate, substantially affecting not only GABAAreceptors but some other superfamily members as well. For example, although 5-hydroxytryptamine type 3 receptors in N1E-115 neuroblastoma cells are insensitive to thiopental, 25,27recombinant glycine receptors in the Xenopus  expression system are sensitive. 28Furthermore, a steady current (but not the peak current) through nicotine-activated ganglionic-type neuronal nAChRs in rat PC12 cells can be more than half-inhibited by surgical EC 50  concentrations of thiopental. 29The situation with respect to thiopental inhibition of CNS-type neuronal nAChRs is unknown and is addressed in the current study.

General Properties of the Major Central Nervous System Neuronal Nicotinic Acetylcholine Receptors

Neuronal nAChRs in vertebrates have a wide range of compositions. 30–33Each functional receptor is a pentamer consisting of two to five subunits of the α type (α2, α3, α4, α5, α6, α7, α8, α9) and zero to three subunits of the β type (β2, β3, β4). (Muscle-type nAChRs are made up of other subunits: two α1, one β1, one δ, and either one γ or one ε.) The heteromeric α4β2 (two α4 and three β2 subunits) and homomeric α7 (five α7 subunits) nAChRs are representative of two important classes of neuronal CNS receptors that account for most of the high-affinity binding sites in the brain for nicotine and α-bungarotoxin, respectively. In addition, the α7 nAChR is especially permeable to the physiologically important cation Ca2+. Furthermore, the α4β2 receptor is supersensitive to inhibition by volatile anesthetics, 9whereas the α7 receptor is insensitive to these agents. 10For these reasons, and taking into account the availability or otherwise of specific subunit clones, we chose to study the rat neuronal heteromeric α4β2 nAChR, the chick neuronal homomeric α7 nAChR, and (for comparison) the embryonic mouse muscle heteromeric αβγδ nAChR. We expressed all three nAChRs in the same system (the Xenopus  oocyte) to make fair comparisons of their behaviors toward thiopental. However, the use of clones from different species, although necessary at the time this study began, may have introduced some variability into these comparisons.

The physiologic functions of neuronal nAChRs in the CNS are still largely unknown, 32and much of the interest in these nicotinic receptors stems from the agonist effects of tobacco-derived nicotine. There is little evidence for fast nicotinic cholinergic synaptic transmission in the brain, which is surprising in view of its important roles in peripheral ganglionic and neuromuscular transmission. 34This points to a presynaptic rather than a postsynaptic role for central nAChRs. It is partly for this reason that presynaptic neuronal nAChRs in the brain are currently receiving so much attention, 34,35the idea being that acetylcholine or nicotine binding to such receptors modulates the release of a number of different types of neurotransmitters (e.g. , dopamine, norepinephrine, glutamate, GABA, and acetylcholine) from presynaptic terminals. Although this is an attractive hypothesis that has received some experimental support, many questions remain to be answered. For example, although the source of nicotinic agonist poses no conceptual problems for cholinergic synapses or for experiments with exogenous nicotine, it is still not clear in most instances what are the physiologic sources of endogenous agonist (presumably acetylcholine) and what are the relevant concentrations at the presynaptic receptors. Whatever the answers, it seems likely that elucidation of the mechanisms and roles of neuronal nAChRs will produce more surprises.

Inhibition of Nicotinic Acetylcholine Receptors by Thiopental: Effects of the Different Experimental Protocols

Thiopental inhibited the peak current responses to acetylcholine in a protocol-dependent manner. We found that preincubation of oocytes with thiopental resulted in enhanced inhibition for all nAChRs (fig. 2and table 2). This effect, which could be very large, was presumably a result of preincubation favoring the attainment of equilibrium between receptor and anesthetic before the application of the acetylcholine agonist. (This is the simplest explanation, although we cannot rule out slow second messenger effects.) Pharmacologically, preapplication of anesthetic would appear to be the protocol most relevant to the steady state maintenance of general anesthesia. For induction, the situation is less clear but would still seem to favor preapplication, because the thiopental concentration is at a quasi–steady state for times on the order of seconds, whereas acetylcholine release, presumably from nerve terminals, occurs at times on the order of milliseconds. For interpreting mechanisms at the molecular level, anesthetic–acetylcholine coapplication without anesthetic preapplication is the easier option but should be resisted, we believe, because it can be misleading. In our experiments, for example, we would have erroneously concluded from the results using the coapplication protocol alone that both the neuronal α4β2 and α7 nAChRs are relatively insensitive to thiopental.

Our results using the pharmacologically more relevant thiopental preincubation protocol, conversely, show that both neuronal and muscle nAChRs can be very sensitive to inhibition by thiopental. Indeed, at high concentrations of acetylcholine agonist IC 50 = 18 ± 2, 34 ± 4, and 20 ± 2 μm thiopental for the neuronal α4β2, neuronal α7 and muscle αβγδ nAChRs, respectively, all with Hill coefficients near unity (table 2). These IC 50  concentrations for inhibiting the nAChRs can be compared with the EC 50  of 25 μm thiopental for surgical general anesthesia 1; at this anesthetizing concentration of thiopental, the aforementioned IC 50  concentrations translate into nAChR inhibitions of 58, 42, and 56%, respectively, for the neuronal α4β2, neuronal α7,and muscle αβγδ nAChRs. At the low concentrations of acetylcholine, the neuronal nAChRs are slightly less sensitive to thiopental than at the high concentrations of acetylcholine (this is significant at the 5% level for the α4β2 but not the α7 nAChR), whereas the muscle nAChR is profoundly less sensitive to thiopental at low concentrations of acetylcholine. This result is consistent with substantial open-channel block for the muscle-type receptor 36,37but not for the neuronal α4β2 receptor, 9as has been found for volatile anesthetics.

Testing for a Role of Neuronal Nicotinic Acetylcholine Receptors in the Anesthetizing Actions of Thiopental

Does thiopental inhibition of central neuronal nAChRs play an important role in producing general anesthesia? If it does, these receptor channels might be expected to be substantially inhibited at thiopental EC 50  concentrations for surgical general anesthesia. Using as an anesthetic end point a failure of patients to respond purposefully to a painful stimulus equivalent to a surgical incision, we previously calculated 1that this EC 50  is approximately 25 μm thiopental. It can be seen from figures 2A and 2Bthat, using the thiopental preincubation protocol with both high and low concentrations of acetylcholine, the α4β2 and α7 neuronal nAChRs were indeed substantially (35–60%) inhibited at this anesthetizing EC 50  concentration of thiopental (arrows in fig. 2). Although this suggests that the activities of neuronal nAChRs in the CNS are severely compromised during thiopental general anesthesia, it does not logically follow that thiopental disruption of nAChR function actually causes general anesthesia.

Clearly, another criterion (in addition to sensitivity) is needed to judge relevance. A convenient yet powerful test 1,7exploits the fact that general anesthesia is stereoselective for many chiral (i.e. , optically active) anesthetics, including thiopental. 13If a given receptor is one of at most a few important targets for the anesthetizing actions of a particular chiral agent, then the rank order and ratio of effectiveness of its optical isomers should be similar for producing general anesthesia and for modulating the activity of the receptor being tested. Using this test with optical isomers of not only thiopental but also of pentobarbital, hexobarbital, etomidate, and anesthetic neurosteroids, previous studies from this 2,22and other 38laboratories have identified potentiation of GABAAre-ceptor activity as a major factor underlying the production of general anesthesia by these intravenous agents. Conversely, using this test with ketamine enantiomers identifies the glutamate N -methyl-d-aspartate receptor as a major target for this agent. 1,39,40 

What happens when this test is applied to our results for the R  (+) and S  (−) optical isomers of thiopental acting on neuronal nAChRs? S  (−)-thiopental is more potent than R  (+)-thiopental at abolishing the mouse righting reflex, 11,12whereas it is less effective at inhibiting the neuronal heteromeric α4β2 nAChR and equally effective at inhibiting the neuronal homomeric α7 nAChR (fig. 3). (For a comparison, the muscle αβγδ nAChR, like the neuronal α7 nAChR, showed no discrimination between the thiopental enantiomers.) Thus both neuronal nAChRs fail to pass the stereoselectivity criterion for important roles in thiopental anesthesia.

A very recent study 41using depressant and convulsant enantiomers of the barbiturate MPPB (1-methyl-5-phenyl-5-propyl barbituric acid) examined their effects on ganglion-type nAChRs in PC12 cells. The two MPPB enantiomers had opposite (inhibitory and excitatory) effects on animals, whereas they were found to have uniformly inhibitory effects on the activities of the neuronal nAChRs. The investigators concluded that inhibition of these ganglion-type neuronal nAChRs does not contribute to the anesthetic action of barbiturates.

Implications for Molecular Mechanisms of Thiopental General Anesthesia

What are the implications of our results for understanding the molecular mechanisms of thiopental general anesthesia? First, we conclude from our studies with the pure optical isomers of thiopental that neither of the neuronal nAChRs tested plays a major role in the anesthetizing actions of thiopental. Instead, from the positive results of our previous GABAAreceptor study 2with these same optical isomers, it would appear that potentiation of the activities of inhibitory GABAAreceptors is the main factor. Second, although thiopental inhibition of nAChRs does not actually produce anesthesia, the fact that these receptors can be very sensitive to anesthetizing concentrations of this intravenous agent suggests that they may be involved in the generation of anesthetic side effects.

The authors thank the Medical Research Council for its support and Marc Ballivet, Jim Boulter, and Jim Patrick for their kind gifts of the individual nAChR subunit cDNAs.

1.
Franks NP, Lieb WR: Molecular and cellular mechanisms of general anaesthesia. Nature 1994; 367: 607–14
2.
Tomlin SL, Jenkins A, Lieb WR, Franks NP: Preparation of barbiturate optical isomers and their effects on GABAAreceptors. A nesthesiology 1999; 90: 1714–22
3.
Tanelian DL, Kosek P, Mody I, MacIver MB: The role of the GABAAreceptor/chloride channel complex in anesthesia. A nesthesiology 1993; 78: 757–76
4.
Olsen RW, Fischer JB, Dunwiddie TV: Barbiturate enhancement of γ-aminobutyric acid receptor binding and function as a mechanism of anesthesia, Molecular and Cellular Mechanisms of Anesthetics. Edited by Roth SH, Miller KW. New York, Plenum Medical Book Company, 1986, pp 165–77
5.
Unwin N: Neurotransmitter action: Opening of ligand-gated ion channels. Neuron 1993; 10 (suppl): 31–41
6.
Ortells MO, Lunt GG: Evolutionary history of the ligand-gated ion-channel superfamily of receptors. Trends Neurosci 1995; 18: 121–7
7.
Franks NP, Lieb WR: An anesthetic-sensitive superfamily of neurotransmitter-gated ion channels. J Clin Anesth 1996; 8: S3–7
8.
Role LW, Berg DK: Nicotinic receptors in the development and modulation of CNS synapses. Neuron 1996; 16: 1077–85
9.
Violet JM, Downie DL, Nakisa RC, Lieb WR, Franks NP: Differential sensitivities of mammalian neuronal and muscle nicotinic acetylcholine receptors to general anesthetics. A nesthesiology 1997; 86: 866–74
10.
Flood P, Ramirez-Latorre J, Role L:α4β2 neuronal nicotinic acetylcholine receptors in the central nervous system are inhibited by isoflurane and propofol, but α7-type nicotinic acetylcholine receptors are unaffected. A nesthesiology 1997; 86: 859–65
11.
Christensen HD, Lee IS: Anesthetic potency and acute toxicity of optically active disubstituted barbituric acids. Toxicol Appl Pharmacol 1973; 26: 495–503
12.
Haley TJ, Gidley JT: Pharmacological comparison of R(+), S(-) and racemic thiopentone in mice. Eur J Pharmacol 1976; 36: 211–4
13.
Andrews PR, Mark LC: Structural specificity of barbiturates and related drugs. A nesthesiology 1982; 57: 314–20
14.
Cordato DJ, Chebib M, Mather LE, Herkes GK, Johnston GAR: Stereoselective interaction of thiopentone enantiomers with the GABA(A) receptor. Br J Pharmacol 1999; 128: 77–82
15.
Krasowski MD, Harrison NL: General anaesthetic actions on ligand-gated ion channels. Cell Mol Life Sci 1999; 55: 1278–303
16.
Couturier S, Bertrand D, Matter J-M, Hernandez M-C, Bertrand S, Millar N, Valera S, Barkas T, Ballivet M: A neuronal nicotinic acetylcholine receptor subunit (α7) is developmentally regulated and forms a homo-oligomeric channel blocked by α-BTX. Neuron 1990; 5: 847–56
17.
Amar M, Thomas P, Johnson C, Lunt GG, Wonnacott S: Agonist pharmacology of the neuronal α7 nicotinic receptor expressed in Xenopus  oocytes. FEBS Lett 1993; 327: 284–8
18.
Gerzanich V, Anand R, Lindstrom J: Homomers of α8 and α7 subunits of nicotinic receptors exhibit similar channel but contrasting binding site properties. Mol Pharmacol 1994; 45: 212–20
19.
Papke RL, Thinschmidt JS: The correction of alpha7 nicotinic acetylcholine receptor concentration-response relationships in Xenopus  oocytes. Neurosci Lett 1998; 256: 163–6
20.
Pistis M, Belelli D, Peters JA, Lambert JJ: The interaction of general anaesthetics with recombinant GABAAand glycine receptors expressed in Xenopus laevis  oocytes: A comparative study. Br J Pharmacol 1997; 122: 1707–19
21.
Hara M, Kai Y, Ikemoto Y: Enhancement by propofol of the γ-aminobutyric acidAresponse in dissociated hippocampal pyramidal neurons of the rat. A nesthesiology 1994; 81: 988–94
22.
Tomlin SL, Jenkins A, Lieb WR, Franks NP: Stereoselective effects of etomidate optical isomers on gamma-aminobutyric acid type A receptors and animals. A nesthesiology 1998; 88: 708–17
23.
Mascia MP, Machu TK, Harris RA: Enhancement of homomeric glycine receptor function by long-chain alcohols and anaesthetics. Br J Pharmacol 1996; 119: 1331–6
24.
Machu TK, Harris RA: Alcohols and anesthetics enhance the function of 5-hydroxytryptamine3receptors expressed in Xenopus laevis  oocytes. J Pharmacol Exp Ther 1994; 271: 898–905
25.
Appadu BL, Lambert DG: Interaction of i.v. anaesthetic agents with 5-HT3receptors. Br J Anaesth 1996; 76: 271–3
26.
Charlesworth P, Richards CD: Anaesthetic modulation of nicotinic ion channel kinetics in bovine chromaffin cells. Br J Pharmacol 1995; 114: 909–17
27.
Jenkins A, Franks NP, Lieb WR: Actions of general anaesthetics on 5-HT3receptors in N1E-115 neuroblastoma cells. Br J Pharmacol 1996; 117: 1507–15
28.
Daniels S, Roberts RJ: Post-synaptic inhibitory mechanisms of anaesthesia: Glycine receptors. Toxicol Lett 1998; 100: 71–6
29.
Andoh T, Furuya R, Oka K, Hattori S, Watanabe I, Kamiya Y, Okumura F: Differential effects of thiopental on neuronal nicotinic acetylcholine receptors and P2Xpurinergic receptors in PC12 cells. A nesthesiology 1997; 87: 1199–209
30.
Deneris ES, Connolly J, Rogers SW, Duvoisin R: Pharmacological and functional diversity of neuronal nicotinic acetylcholine receptors. Trends Pharmacol Sci 1991; 12: 34–40
31.
McGehee DS, Role LW: Physiological diversity of nicotinic acetylcholine receptors expressed by vertebrate neurons. Annu Rev Physiol 1995; 57: 521–46
32.
Sivilotti L, Colquhoun D: Acetylcholine receptors: Too many channels, too few functions. Science 1995; 269: 1681–2
33.
Lukas RJ, Changeux J-P, Le Novère N, Albuquerque EX, Balfour DJK, Berg DK, Bertrand D, Chiappinelli VA, Clarke PBS, Collins AC, Dani JA, Grady SR, Kellar KJ, Lindstrom JM, Marks MJ, Quik M, Taylor PW, Wonnacott S: International Union of Pharmacology. XX. Current status of the nomenclature for nicotinic acetylcholine receptors and their subunits. Pharmacol Rev 1999; 51: 397–401
34.
Wonnacott S: Presynaptic nicotinic ACh receptors. Trends Neurosci 1997; 20: 92–8
35.
MacDermott AB, Role LW, Siegelbaum SA: Presynaptic ionotropic receptors and the control of transmitter release. Annu Rev Neurosci 1999; 22: 443–85
36.
Forman SA, Miller KW, Yellen G: A discrete site for general anesthetics on a postsynaptic receptor. Mol Pharmacol 1995; 48: 574–81
37.
Dilger JP, Brett RS, Mody HI: The effects of isoflurane on acetylcholine receptor channels. 2. Currents elicited by rapid perfusion of acetylcholine. Mol Pharmacol 1993; 44: 1056–63
38.
Wittmer LL, Hu Y, Kalkbrenner M, Evers AS, Zorumski CF, Covey DF: Enantioselectivity of steroid-induced γ-aminobutyric acidAreceptor modulation and anesthesia. Mol Pharmacol 1996; 50: 1581–6
39.
Lodge D, Anis NA, Burton NR: Effects of optical isomers of ketamine on excitation of cat and rat spinal neurones by amino acids and acetylcholine. Neurosci Lett 1982; 29: 281–6
40.
Zeilhofer HU, Swandulla D, Geisslinger G, Brune K: Differential effects of ketamine enantiomers on NMDA receptor currents in cultured neurons. Eur J Pharmacol 1992; 213: 155–8
41.
Watanabe I, Andou T, Furuya R, Sasaki T, Kamiya Y, Itoh H: Depressant and convulsant barbiturates both inhibit neuronal nicotinic acetylcholine receptors. Anesth Analg 1999; 88: 1406–11