Skip to main content
British Journal of Pharmacology logoLink to British Journal of Pharmacology
. 2003 Aug 11;140(2):315–322. doi: 10.1038/sj.bjp.0705447

Effects of GABA agonists on body temperature regulation in GABAB(1)−/− mice

Christophe Quéva 1,5, Marianne Bremner-Danielsen 2, Anders Edlund 3,6, A Jonas Ekstrand 3,7, Susanne Elg 4, Sven Erickson 1, Thore Johansson 3,8, Anders Lehmann 2,*, Jan P Mattsson 4
PMCID: PMC1574040  PMID: 12970075

Abstract

  1. Activation of GABAB receptors evokes hypothermia in wildtype (GABAB(1)+/+) but not in GABAB receptor knockout (GABAB(1)−/−) mice. The aim of the present study was to determine the hypothermic and behavioural effects of the putative GABAB receptor agonist γ-hydroxybutyrate (GHB), and of the GABAA receptor agonist muscimol. In addition, basal body temperature was determined in GABAB(1)+/+, GABAB(1)+/− and GABAB(1)−/− mice.

  2. GABAB(1)−/− mice were generated by homologous recombination in embryonic stem cells. Correct gene targeting was assessed by Southern blotting, PCR and Western blotting. GABAB receptor-binding sites were quantified with radioligand binding. Measurement of body temperature was done using subcutaneous temperature-sensitive chips, and behavioural changes after drug administration were scored according to a semiquantitative scale.

  3. GABAB(1)−/− mice had a short lifespan, probably caused by generalised seizure activity. No histopathological or blood chemistry changes were seen, but the expression of GABAB(2) receptor protein was below the detection limit in brains from GABAB(1)−/− mice, in the absence of changes in mRNA levels.

  4. GABAB receptor-binding sites were absent in brain membranes from GABAB(1)−/− mice.

  5. GABAB(1)−/− mice were hypothermic by approximately 1°C compared to GABAB(1)+/+ and GABAB(1)+/− mice.

  6. Injection of baclofen (9.6 mg kg−1) produced a large reduction in body temperature and behavioural effects in GABAB(1)+/+ and in GABAB(1)+/− mice, but GABAB(1)−/− mice were unaffected. The same pattern was seen after administration of GHB (400 mg kg−1). The GABAA receptor agonist muscimol (2 mg kg−1), on the other hand, produced a more pronounced hypothermia in GABAB(1)−/−mice. In GABAB(1)+/+ and GABAB(1)+/− mice, muscimol induced sedation and reduced locomotor activity. However, when given to GABAB(1)−/− mice, muscimol triggered periods of intense jumping and wild running.

  7. It is concluded that hypothermia should be added to the characteristics of the GABAB(1)−/−phenotype. Using this model, GHB was shown to be a selective GABAB receptor agonist. In addition, GABAB(1)−/− mice are hypersensitive to GABAA receptor stimulation, indicating that GABAB tone normally balances GABAA-mediated effects.

Keywords: Hypothermia, body temperature, GABA, gammahydroxybutyrate, baclofen, gene deletion, muscimol

Introduction

γ-Aminobutyric acid (GABA) is the major inhibitory transmitter in the mammalian brain. In the adult CNS, it mediates both fast and slow neuronal inhibition by activating the ligand-gated ionotropic GABAA and the G-protein-coupled metabotropic GABAB receptors, respectively. GABAB receptors inhibit adenylate cyclase through Gαi subunits, and activate the inwardly rectifying K+ channels, to mediate hyperpolarisation of postsynaptic membranes. Stimulation of presynaptic GABAB receptors suppresses neurotransmitter release by inhibition of voltage-sensitive P, N, and L types of Ca2+ channels (for review, see Couve et al., 2000).

Two GABAB receptor genes have been identified, GABAB(1) and GABAB(2) (Marshall et al., 1999; Bowery & Enna, 2000). In addition, GABAB(1) encodes several splice variants and the more abundant ones, GABAB(1a) and GABAB(1b), are detected in the mouse, rat and human, and differ in their amino-terminal ligand-binding domain, where the first 147 amino acids of GABAB(1a) are replaced by a sequence of 18 amino acids in GABAB(1b) (Kaupmann et al., 1997). When expressed alone in heterologous systems, GABAB(1a), GABAB(1b) and GABAB(2) display a low affinity for GABA, and a poor coupling to adenylate cyclase and inwardly rectifying K+ channels (see Marshall et al., 1999; Bowery & Enna, 2000). In addition, GABAB(1a,b) receptors are prevented from reaching the cell surface, due to the presence of an endoplasmic reticulum retention motif in the C-terminus (Couve et al., 1998). However, coexpression of GABAB(1a,b) and GABAB(2) allows trafficking of GABAB(1a,b) to the cell membrane and efficient GABAB coupling to ion channels (Marshall et al., 1999; Bowery & Enna, 2000). GABAB(1a,b) and GABAB(2) mRNA were also found to be coexpressed in the brain and coimmunoprecipitated in vitro and in vivo (Marshall et al., 1999; Bowery & Enna, 2000). The pharmacology and the affinity for agonist of the GABAB(1a,b)–GABAB(2) combination is comparable to the ones reported for the native receptors (Marshall et al., 1999; Bowery & Enna, 2000). Together, these results suggest that the heterodimer GABAB(1a,b)–GABAB(2) is the functional metabotropic GABAB receptor.

Among other effects in vivo, stimulation of the GABAB receptor is known to produce hypothermia which parallels behavioural changes (Gray et al., 1987). This effect is considered to be of central origin, since intracerebroventricular injection of the prototypic GABAB receptor agonist baclofen lowers body temperature (Gray et al., 1987). The specific site at which GABAB receptor agonists act is unknown, but it probably resides within the hypothalamic thermoregulation centre (Yakimova et al., 1996).

The aims of the current study were three-fold. Effects of baclofen on body temperature have been reported previously in GABAB(1)−/− mice (Schuler et al., 2001), but no alteration in basal temperature has been demonstrated. Here, we report on basal body temperature in GABAB(1)−/− mice. Also, gammahydroxybutyrate (GHB), a drug recently approved by the US Food and Drug Administration for treatment of narcolepsy with episodes of cataplexy but also a common drug of abuse, may produce CNS effects through stimulation of GABAB receptors (Carai et al., 2001), but the issue is controversial (Feigenbaum & Howard, 1996). It has been shown that GHB may be converted to GABA (Hechler et al., 1997), which acts on both GABAB and GABAA receptors. We used GABAB(1)−/− mice to resolve this issue as far as hypothermia and behavioural changes are concerned. Finally, similar to baclofen, the selective GABAA receptor agonist muscimol can induce hypothermia and sedation (Zarrindast & Oveissi, 1988). Given the close inter-relationship between GABAA and GABAB receptors, the effects of muscimol on body temperature and behaviour were studied in GABAB(1)−/− mice.

Methods

Generation of GABAB(1) mutant mice

The murine genomic GABAB(1) locus (17 kb) encompassing exons 1a1 to 5 (numbered according to Lamp et al., 2001) was isolated from a 129Sv/J genomic library (Stratagene, La Jolla, CA, U.S.A.; Figure 1, Ekstrand, 1999). The gene-targeting vector was prepared as follows: a 4126 bp EcoR1–Not1 fragment upstream of exon E1a1 was subcloned in the Sac2 Not1 sites of pSAβGeo(LoxP)2PGKDTA (obtained from Philippe Soriano, Fred Hutchinson Cancer Research Center, Seattle, WA, U.S.A.). The 5304 bp Hpa1–Sal1 fragment encompassing the 3′ end of exon E1a/b to exon 5 was subcloned into the Nhe1–Sal1 of the same vector. The expression of the positive selection cassette SAβGeo is under the control of the GABAB(1) promoter, and replaces a genomic region encompassing the exon E1a1 (start of the GABAB(1a) form) to the translation initiation of the GABAB(1b) form located in exon E1a/b (Figure 1a). A mock PCR-positive control for the targeted GABAB(1) locus was made by cloning a 4776 bp Not1 fragment corresponding to the 5′ part of the GABAB(1) locus into the Not1 site of pSAβGeo(LoxP)2PGKDTA.

Figure 1.

Figure 1

Strategy for deletion of the GABAB(1) gene. (a) Map of the murine GABAB(1) locus, showing the exon–intron structure of the gene. Exon nomenclature is according to Lamp et al. (2001). The targeting construct comprised the SAβGeo(LoxP)2, flanked by a 4126 bp EcoR1–Not1 5′ fragment and a 5304 bp Hpa1–Sal1 3′ fragment. Recombination between the GABAB(1) locus and the targeting construct resulted in the replacement of the genomic sequence from exon E1a1 to the 5′ part of exon E1a/b. Homologous recombination was selected by PCR, using the primer CQ113 and CQ114 (arrows), and later confirmed by Southern blotting. (b) Southern blot analysis of Sca1-digested genomic DNA confirming the correct junction. 5′ probe label indicates the 253 bp probe generated by PCR amplification using the primers 5′-ACTGAGCCTGGTCAAGGTCAG-3′ and 5′-CAACGCCACCGTGAAACCCT-3′. (c) Detection of GABAB(1) mRNA by RT–PCR gave rise to the expected 427 bp fragment, confirming the absence of GABAB(1) transcript in GABAB(1) mutant mice. GABAB(2) (970 bp RT–PCR product) transcript levels are normal in GABAB(1) mutant mice, despite the finding that the protein product could not be detected by Western blot. Amplification of the S16 sRNA was used as a control for efficient cDNA synthesis, and generated a 102 bp PCR product.

An Xho1 linearised targeting vector was electroporated into R1 embryonic stem (ES) cells and placed under G418 selection (300 mg l−1). Surviving clones were first selected by PCR on purified DNA, with the primers CQ113 5′-TATTGCCTTGGTTACTCTAAA-3′ and CQ 114 5′-CCCTGGACTACTGCGCCCTAC-3′. Positive clones were expanded and frozen. Correct targeting was ascertained by Southern blotting of Sca1-digested genomic DNA, using a 253 bp probe located outside the targeting construct, and generated by PCR amplification using the primers 5′-ACTGAGCCTGGTCAAGGTCAG-3′ and 5′-CAACGCCACCGTGAAACCCT-3′. The frequency of homologous recombination with the gene trap vector was 16%. Two ES cell clones were injected into C57Bl/6 blastocysts and then transplanted into pseudopregnant females according to standard procedures. They gave rise to chimeric founders that transmitted the targeted allele through the germ line. Heterozygote progeny was interbred to produce homozygous mice. Genotyping was performed by Southern blotting or by PCR from tail DNA with the primers CQ170 5′-TCGCATTGTCTGAGTAGGTGT-3′, CQ171 5′-GCCCTCTTGCCTCTCTAAATC-3′ and CQ172 5′-CCTCCTCAGAACGGCGTGCAG-3′. The wild-type and mutant PCR products are 552 and 692 bp, respectively (data not shown).

RNA preparation and analysis

Total RNA was prepared from cerebellum with Trizol reagent (Life Technologies, Carlsbad, CA, U.S.A.), according to the manufacturer's recommendation. The S16 control was amplified with the sense primer 5′-AGGAGCGATTTGCTGGTGTGGA-3′ and the antisense primers 5′-GCTACCAGGCCTTTGAGATGGA-3′ (20 cycles, primers generate a 102 bp fragment). GABAB(1) cDNA was amplified with the primer pair 5′-CTGCCCGGATGTGGAACCTTA-3′ and 5′-TCAGCATACCACCCGATGAGA-3′ and GABAB(2) cDNA with the primer pair 5′-ATCGAGCAGATCCGCAACGAG-3′ and 5′-ACACAACTTGACCCGTGACCC-3′, yielding respective fragments of 427 and 970 bp after 25 cycles.

Western blot analysis

Total protein extracts were prepared by lysis and sonication in a radioimmunoprecipitation buffer (10 mM Tris 7.4, 0.15 M NaCl, 1% NP40, 1% deoxycholate, 0.1% SDS, 0.5% aprotinin), in the presence of proteinase inhibitor cocktail (Boehringer Mannheim, Indianapolis, IN, U.S.A.). Brain protein extract (30 μg) was resolved on a 4–12% Bis(2-hydroxyethyl)-imino-tris(hydroxymethyl)methane (Bis-Tris) gel in 3-(N-morpholino)propanesulphonic acid sodium dodecylsulphate running buffer, and electroblotted to polyvinylidene fluoride membranes by semidry transfer. Filters were blocked in phosphate-buffered saline containing 0.1%Tween 20 (PBST) and 5% milk over night in +4°C. The primary antibody was diluted in PBST with 5% nonfat dry milk, and incubated for 1 h in room temperature. Filters were washed three times for 15 min in PBST in room temperature, and subsequently, a secondary horseradish peroxidase-linked antibody (Amersham, Aylesbury, U.K.) was added, diluted in PBST with 5% nonfat dry milk for 45 min in room temperature. Before the detection, reaction filters were washed as above. Antibody–antigen interaction was visualised by enhanced chemiluminescence (Amersham, Aylesbury, U.K.).

GABAB(1) receptor protein expression was detected with two different polyclonal antibodies, Grim raised in rabbits (epitope: CEDVNSRRDILPDYELKLIHH) at AstraZeneca (4 μg ml−1 dilution) and AB1531 (Chemicon International, Temecula, CA, U.S.A.) (1 : 5000 dilution). GABAB(2) receptor protein was analysed with three different antibodies: two polyclonal, Gunlög raised (epitope: TEPSRTCKDPIEDINSPEHI) at AstraZeneca (4 μg ml−1 dilution), AB5394 (Chemicon International, Temecula, CA, U.S.A.) (1 : 5000 dilution) and one monoclonal G12020 (Transduction Laboratories, Lexington, KY, U.S.A.) (1 : 2500 dilution).

Evaluation of phenotype

Routine histological, haematological and clinical chemistry analysis was made on the mice (Safety Assessment, AstraZeneca R&D Södertälje, Sweden). The histopathological evaluation included all major tissues which were fixed in 10% formalin, embedded after dehydration in Histowax, cut at 4–6 μm and stained with haematoxylin/eosin. Apart from careful observation of the behaviour, some mice were videofilmed for 24 h.

Preparation of synaptic membranes from mouse brain

Mouse brain synaptic membranes were prepared from the whole brain of GABAB(1)+/+ or GABAB(1)−/− mice, using the method described by Zukin et al. (1974) with some modifications. In brief, mouse whole brain was homogenised in 10 volumes of ice-cold buffer containing 0.32 M sucrose (Sigma-Aldrich, Steinheim, Germany), 10 mM Tris(hydroxymethyl)aminomethane (Tris; Sigma-Aldrich, Steinheim, Germany), 0.1 mM 4-(2-aminoethyl)benzenesulphonyl fluoride (AEBSF; Roche Diagnostics, Mannheim, Germany) and 20 μg ml−1 bacitracin, pH 7.4 (Sigma-Aldrich, Steinheim, Germany). The homogenate was centrifuged at 1000 × g for 10 min, and the supernatant was then centrifuged at 20,000 × g for 20 min. The pellet was resuspended (by vortex) in ice-cold distilled water containing 0.1 mM AEBSF and 20 μg ml−1 bacitracin (pH set to 7.0) and centrifuged at 8000 × g for 20 min. The supernatant and the upper layer of the pellet were centrifuged at 33,000 × g for 20 min. The pellet was resuspended in 50 mM Tris, pH 7.4, containing 0.1 mM AEBSF and 20 μg ml−1 bacitracin, and centrifuged at 33,000 × g for 20 min two times before it was snap frozen in methanol (Rathburn Chemicals, Walkerburn, Scotland)/dry ice and stored overnight at −70°C. Up to this point, the membranes were kept on ice at all times, and all centrifugation steps were performed at 4°C. The frozen pellet was thawed and washed six times in 50 mM Tris, pH 7.4, by centrifugation at 8000 × g for 10 min at 18°C. The resulting pellet was resuspended in TC buffer (50 mM Tris, 2.5 mM CaCl2, pH 7.4), snap frozen in methanol/dry ice and stored at −70°C. Protein concentration was determined according using Bio-Rad's protein assay kit (Bio-Rad, Hercules, CA, U.S.A.) with bovine gamma globulin as a standard.

[3H]CGP54626 saturation-binding analysis

GABAB receptor sites in mice brain membranes were studied by saturation-binding analysis using the GABAB receptor antagonist [3H]CGP54626A (Tocris, Bristol, U.K.). [3H]CGP54626A saturation binding was measured by incubation of [3H]CGP54626A (0.0051–100 nM, final concentrations) in 200 μl TC buffer containing 0.75% ethanol (Rathburn Chemicals, Walkerburn, Scotland) and 80 μg membrane protein. Nonspecific binding was determined in the presence of 1 mM GABA (Sigma-Aldrich, Steinheim, Germany). After incubation for 20 min at room temperature, incubations were terminated by rapid filtration through a glass fibre filter (Printed filtermat B filters, Wallac, Turku, Finland), which had been presoaked in 0.3% polyethyleneimine (Sigma-Aldrich, Steinheim, Germany), using a TOMTEC cell harvester (Orange, CT, U.S.A.). The filters were washed with 5 ml of 50 mM Tris (pH 7.4) at 4°C and dried for 1.5 min in a microwave followed by incubation at 55°C for 30 min. MeltiLex B/HS scintillator sheet was melted onto the filter, and radioactivity was determined in a Microbeta scintillation counter (Wallac, Turku, Finland). KD and Bmax values were calculated by fitting binding data to the equation B=Bmax/(1+KD/[L]), using Xlfit for Microsoft Excel.

Measurement of body temperature and behaviour

Age-matched, C57Bl6/129Sv F1 hybrid GABAB(1)+/+, GABAB(1)+/− and GABAB(1)−/− mice were used. They were kept in Perspex cages at an ambient temperature between 21 and 23°C and a relative humidity between 52 and 56%. The lights were on between 0700 and 1800, and there was a gradual dimming of the lights to complete darkness between 1800 and 1900, and a gradual increase in illumination back to normal daylight between 0600 and 0700. The distribution of sexes was relatively even across the groups (see the legend of Figure 4), and there was no tendency for any difference between males and females with respect to basal temperature or drug-induced changes. A thermosensitive chip (BioMedic Data Systems, Maywood, NJ, U.S.A.) was implanted in the interscapular region under brief isoflurane anaesthesia, and the animals were allowed to recover for at least 1 day. The animals had free access to food and water, except for during the experiment. On the experimental day, the mice were placed in individual cages between 0900 and 1000, and the ambient temperature was 20.5±1.0°C. After 30 min, three basal temperature recordings were made using a transponder (IPTT-100, BioMedic Data Systems, Maywood, NJ, U.S.A.) communicating with a PC for data acquisition. In preliminary experiments, the system was evaluated in mice by measuring the interscapular temperature and rectal temperature at the same time. There was a strong correlation between the two (R2=0.86, P<0.0001, n=23). The thermosensitive chips were calibrated by the producer in the range from 32 to 43°C, and they were calibrated against a thermistor in a water bath before implantation. The resolution of the chips was 0.1°C. Compounds (baclofen 9.6 mg kg−1, RBI, Natick, MA, U.S.A.; GHB 400 mg kg−1, Sigma-Aldrich, Steinheim, Germany, and muscimol 2 mg kg−1, Tocris Cookson, Bristol, U.K.) were injected subcutaneously at 5 ml kg−1 after the last measurement. The doses were chosen based on pilot experiments in which they were found to produce a significant hypothermia. Measurements were then made at regular intervals (see Figure 4). Behavioural scoring was made at each time point, and the behavioural data are presented as the maximal effect. The following definitions were used for behavioural effects:

Figure 4.

Figure 4

Effects of baclofen (a), GHB (b) and muscimol (c) on body temperature in GABAB+/+, GABAB+/− and GABAB−/− mice. Drugs were administered at time 0; n=11 (eight males), 12 (seven males) and 9 (five males) (GABAB+/+, GABAB+/− and GABAB−/− baclofen groups, respectively); n=9 (five males), 11 (three males) and 14 (eight males) (GABAB+/+, GABAB+/− and GABAB−/− GHB groups, respectively); n=8 (six males), 9 (five males) and 6 (four males) (GABAB+/+, GABAB+/− and GABAB−/− muscimol groups, respectively); ***P<0.001, Student's t-test.

0

no effect;

1

exophthalmus, slight motor impairment;

2

more pronounced motor impairment;

3

immobile with intact righting reflex;

4

no righting reflex, disturbed respiration, occasional seizures, detectable but very low muscle tonus;

5

paralysed, no muscle tonus, moribund (killed for ethical reasons).

The behaviour was scored by the same experienced observer (MBD) in all experiments. The doses used were obtained from pilot dose–response experiments. All animal experiments were approved by the Animal Ethics Committee of the Göteborg region.

The Student's t-test or Mann–Whitney U-test was used to compare the groups. A P-value below 0.05 was considered to be statistically significant. In the experiments on drug-induced changes in body temperature, only nadir temperatures were compared.

Results

Cloning and characterisation of the 5′-end of the mouse GABAB(1) gene

A mouse genomic library was screened and a 17 kb of the GABAB(1) gene was cloned and characterised. The cloned fragment encompassed exons E1a1 to 5. The exon/intron organisation was found to be similar to the human GABAB(1) gene (Ekstrand, 1999), and predicts the expression of the two receptor isoforms, 1a and 1b, as previously described for rats (Kaupmann et al., 1997) and for humans (Ekstrand, 1999). The ATG initiating the translation of the mouse GABAB(1a) isoform is located in exon E1a2, and the ATG for the GABAB(1b) isoform in exon E1a/b (Figure 1). The predicted partial mouse GABAB(1a) isoform amino-acid sequence encoded by exons E1a2 to 5 is 99.5% homologous to the human (Ekstrand, 1999) and 99.2% homologous to the rat sequence (Kaupmann et al., 1997), and the predicted partial mouse GABAB(1b) isoform amino-acid sequence encoded by exons E1a/b to 5 is 97.7% homologous to the human and 100% homologous to the rat sequence.

Phenotype of GABAB(1)−/− mice

Homologous recombination between the targeting vector and the GABAB(1) locus resulted in the deletion of the exons E1a1 to E1a/b and their replacement by the βGeo gene. Breeding of heterozygous mice resulted in the generation of GABAB(1)−/− mice at the expected Mendelian ratios. Southern blot analysis, RT–PCR and Western blot confirmed the absence of GABAB(1) receptor transcript and protein (Figures 1, 2).

Figure 2.

Figure 2

Western blot analysis showing the expression of GABAB receptor isoforms in whole brain and brain membrane extracts from GABAB(1)+/+, GABAB(1)+/− and GABAB(1)−/− mice. Equal amounts of proteins were separated by a 4–12% gradient Bis-Tris gel, and transferred on polyvinylidene difluoride membranes. Two bands with the apparent molecular weights of 130 and 105 kDa were detected when probing the Western blot with the polyclonal antibody Grim. These bands, corresponding to GABAB(1a) and GABAB(1b) receptor proteins, were identified in the brain extract and membrane extract of wild-type, heterozygote, but not mutant GABAB(1) mice, and in HEK293 cell lysate transfected with GABAB(1a) and GABAB(1b) expression vector. Similarly, GABAB(2), 110 kDa, was detected by the Gunlög polyclonal antibody in HEK293 cell transfected with a GABAB(2) expression vector, and in GABAB(1)+/+ and GABAB(1)+/− mice, but not in GABAB(1)−/−.

The lifespan of GABAB(1)−/− mice was greatly reduced so that only few GABAB(1)−/− lived longer than 8 weeks in the C57Bl6/129Sv F1 hybrid background used in this study. The exact cause of the premature death was not formally determined, but it is reasonable to assume that it was secondary to spontaneous seizures occurring intermittently. Apart from the epileptic behaviour, no overt behavioural changes were seen. In addition, no abnormalities were observed with respect to histological, clinical chemistry and haematological parameters. The latter two included the following: glucose, bilirubin (total), urea, creatinine, total protein, albumin, albumin/globulin ratio, cholesterol, alkaline phosphatase, aspartate aminotransferase, alanine aminotransferase, sodium, potassium, calcium, haematocrit, haemoglobin, erythrocytes, mean corpuscular volume, mean corpuscular haemoglobin, mean corpuscular haemoglobin concentration, leucocytes, platelets, neutrophils, eosinophils, basophils, lymphocytes and monocytes. There were no differences in body weight. For instance, male GABAB(1)(+/+) mice weighed 28.5±2.1 g (n=5) 60–69 days after birth. The corresponding figure for male GABAB(1)(−/−) mice was 26.1±1.0 g (n=8).

Expression of GABAB subunits

Successful targeting of the GABAB(1) gene was confirmed by complete absence of GABAB(1) mRNA and protein (Figures 1, 2). Furthermore, the expression of GABAB(2) receptor was abolished in whole-brain homogenates and in brain membranes (Figure 2).

Comparison of the receptor binding of CGP54626A in brain synaptic membranes from GABAB(1)(+/+) and GABA(−/−)B(1) mice

Ligand binding at GABAB receptor sites in whole-brain synaptic membranes from GABAB(1)(+/+) and GABAB(1)(−/−) mice was studied using [3H]CGP54626A. In membranes from GABAB(1)(+/+) mice, [3H]CGP54626 bound with an apparent KD of 4.3±0.65 nM (mean±s.e.m., n=3) and a Bmax of 4.6±3.0 pmol mg−1 (mean±s.e.m., n=3). In contrast, no specific [3H]CGP54626 binding could be detected in membranes from GABAB(1)(−/−) mice.

Hypothermia and behaviour

Analysis on basal body temperature was done on pooled data from all experiments before drug administration. Unexpectedly, the basal temperature of GABAB(1)−/− mice was significantly lower than that of the two other groups, the difference being about 1°C (P<0.001; Figure 3). There was no difference between males and females in this regard (data not shown).

Figure 3.

Figure 3

Basal temperature in GABAB+/+, GABAB+/− and GABAB−/− mice. Temperature was measured using temperature-sensitive subcutaneous chips, which allowed measurement without restraining the animals. ***P<0.001, Student's two-tailed, unpaired t-test, n=43 (28 males, GABAB(1)+/+), 49 (26 males, GABAB(1)+/−), 40 (23 males, GABAB(1)−/−). Predrug measurements from other studies were also included in the material.

Baclofen (9.6 mg kg−1) produced a marked hypothermia, which reached its nadir at 60–80 min after administration, and subsequently returned towards baseline levels (Figure 4). This effect was accompanied by behavioural alterations (Table 1 ). The hypothermic and behavioural effects of baclofen in GABAB(1)+/− mice were identical to those seen in GABAB(1)+/+ mice, but the GABAB(1)−/− mice did not respond to baclofen at all.

Table 1.

Effects of baclofen, GHB and muscimol on the behaviour in GABAB(1)+/+, GABAB(1)+/− and GABAB(1)−/− mice

Genotype Behavioural effect (mean±s.e.m., n)
  Baclofen GHB Muscimol
GABAB(1)+/+ 2.0±0.1(11) 4.5±0.0(9) 1.8±0.4(8)
GABAB(1)+/− 1.7±0.3(12) 4.3±0.2(11) 1.8±0.5(9)
GABAB(1)−/− 0.0±0.0(9)* 0.0±0.0(14)* Not assesseda

Values are means+s.e.m. (n).

*

P<0.01, GABAB(1)+/+ vs GABAB(1)−/− (Mann–Whitney U-test).

a

Could not be assessed due to a qualitative change in behaviour.

A high dose of GHB (400 mg kg−1) was required to elicit hypothermia and behavioural changes in wild-type and heterozygote mice (Figure 4, Table 1). As was the case after baclofen administration, GHB did not produce any hypothermic or behavioural effects in GABAB(1)−/− mice. The response to GHB was not sex-dependent in any of the groups.

Even if it can be assumed that the spectrum of behavioural changes would differ between muscimol and GABAB receptor agonists using more detailed evaluation, the behavioural scale used for mice given baclofen and GHB could also be applied to GABAB(1)+/+ and GABAB(1)+/− mice. These two groups did not differ in this regard (Table 1). However, the scale could not be used on GABAB(1)−/− mice, since they showed a completely different pattern of behaviour after muscimol treatment (2 mg kg−1) in that they were extremely sensitive to, for example, handling which induced wild running and jumping (‘popcorn') behaviour.

Discussion

The phenotype of the GABAB(1)−/− mice generated in our laboratory is similar to that described previously (Prosser et al., 2001; Schuler et al., 2001). In particular, the observations confirm the finding that the lifespan is short (a few weeks) in mice with a C57Bl6/129SvJ background (Prosser et al., 2001), and that this is most likely due to generalised epileptic seizures. In contrast, GABAB(1) null mutants with a Balb/c background also have seizures, but their lifespan seems to be normal (Schuler et al., 2001). We also confirm that there are no histopathological alterations nor any routine clinical chemistry changes associated with deletion of the GABAB(1) gene. Similar to the findings reported previously, GABAB(2) protein was undetectable in whole-brain extracts and in membrane preparations (Prosser et al., 2001; Schuler et al., 2001). Since the expression of GABAB(2) mRNA is normal in GABAB(1)−/− mice (present results and Schuler et al., 2001), this suggests that the absence of GABAB(2) protein is due to a translational or stability defect.

In agreement with the study of Schuler et al. (2001), baclofen neither affected temperature regulation nor behaviour in GABAB(1)−/− mice, which confirms the notion that the GABAB(1) subunit is necessary for these effects of baclofen. Since temperature was measured in a region where brown adipose tissue can be found, it may be argued that rectal measurement might provide a better reflection of core body temperature. However, baclofen has been shown in a number of studies to produce hypothermia, as measured with a rectal probe in awake rats and mice (e.g. Gray et al., 1987; Zarrindast & Oveissi, 1988; Schuler et al., 2001). Hyperthermia has been reported after systemic administration of baclofen to anaesthetised rats, but the temperature changes interscapularly and rectally parallel each other well (Addae et al., 1986). The hyperthermic response is only seen under anaesthesia, or after very high doses of baclofen given to conscious rats (Zarrindast & Oveissi, 1988).

An unexpected finding was that GABAB(1)−/− mice had a basal hypothermia which is particularly surprising in view of the hyperlocomotor activity observed in GABAB(1)−/− mice (data not shown; Schuler et al., 2001). Telemetric studies on mice have clearly demonstrated that locomotor activity and body temperature are closely correlated (Johansson et al., 1999). Although the hypothermic effects of GABAB receptor agonists have been known for a long time (e.g. Gray et al., 1987), antagonist experiments have to our knowledge not suggested that there would be any endogenous GABABergic tone in this system. In a previous study on GABAB(1)−/− mice, lack of effect of baclofen on body temperature was established, but no data on basal body temperature were presented (Schuler et al., 2001). While an earlier study revealed that GABAB(1)−/− mice weighed significantly less than wild-type mice by 3 weeks of age (Prosser et al., 2001), this could not be verified in our colony and therefore, differences in body weight cannot explain the disparities in body temperature. Whether the basal hypothermia reflects direct involvement of GABAB receptors in thermoregulatory circuits or if this depends on secondary alterations remains to be studied. At any rate, the effects of GABAB(1) deletion on body temperature warrant further studies on general energy metabolism in these mice. In addition, body temperature was measured at room temperature in the present experiments. Since this is well below the thermoneutral zone for the mouse, future studies should include body temperature measurements at thermoneutrality.

Several mechanisms for the actions of GHB in the CNS have been proposed. One hypothesis is that GHB directly stimulates GABAB receptors. However, even if very high doses of GHB are required to produce CNS effects, the affinity of GHB for GABAB receptors is very low (Lingenhoehl et al., 1999), or almost undetectable (unpublished own publications). There seem to exist specific GHB-binding sites distinct from GABAB receptor-binding sites, through which GHB can exert its effects (Snead, 2000). Further, GHB activates G-protein-coupled receptors separate from those stimulated by baclofen. However, it is also possible that administration of GHB elevates the cerebral levels of GABA acting on GABAB receptors (see Maitre (1997) for review). While the present findings do not rule out the existence of GHB receptors, they unambiguously show that the GABAB(1) receptor is necessary for the effects of GHB on body temperature and behaviour. Consistent with this is the report that selective GABAB receptor antagonists abolish the sedative/hypnotic effects of GHB, while a putative GHB antagonist has no inhibitory effect (Carai et al., 2001). Only one dose of GHB was used, and it caused profound effects on body temperature and behaviour. Therefore, although unlikely, it cannot be excluded that effects of lower doses of GHB may not be exclusively mediated by GABAB receptors.

Muscimol-induced hypothermia was significantly enhanced in GABAB(1)−/− mice. This finding indicates that there are tonically active GABAB(1) receptors which counteract the hypothermic effects of muscimol, and it reaffirms the notion that discovery of phenotypic changes in mice carrying gene deletions sometimes depends on the application of external challenges. The mechanism behind this effect is presently unknown. It may be due to global or region-specific changes in the expression of GABAA receptors. Alternatively, the absence of GABAB receptors may alter the characteristics of the circuitry controlling body temperature after GABAA receptor stimulation. Acute blockade of GABAB receptors with CGP62349 failed to affect muscimol-induced hypothermia in wild-type mice (unpublished own observations), which indicates that complete loss of the GABAB response is required to enhance the muscimol-induced hypothermia, or that absence of GABAB from the embryonic stage and onwards underlies the hypersensitivity to muscimol. The effect was not generalised for all hypothermia-inducing drugs in that preliminary experiments failed to reveal differences between GABAB(1)+/+ and GABAB(1)−/− mice in terms of hypothermia induced by the cholinomimetic oxotremorine. In addition, the markedly altered behaviour seen in GABAB(1)−/− mice after muscimol administration may reflect the inhibitory effects of the GABAB system on GABAA-ergic mechanisms. The ‘popcorn' behaviour observed in GABAB(1)−/− mice has been reported earlier after pharmacological treatment with cannabinoid receptor agonists (Patel & Hillard, 2001), and has been interpreted as hyper-reflexia. At the cellular level, inhibitory effects of GABAB receptor stimulation on GABAA-mediated currents (Obrietan & van den Pol, 1998) may underlie both enhanced hypothermic effects as well as changes in behavioural response recorded after muscimol administration. Additional studies are needed to further elucidate the effects of muscimol in GABAB(1)−/− mice. However, the present results clearly suggest that in contrast to some claims (Yamauchi et al., 2000), muscimol does not stimulate GABAB receptors in vivo.

In summary, the present study demonstrates that deletion of the GABAB(1) receptor gene is associated with basal hypothermia, abolition of responsiveness to GHB and hypersensitivity to GABAA receptor stimulation.

Acknowledgments

We gratefully acknowledge the histopathological and clinical chemistry investigations done by Alaa Saad, Ronny Fransson Steen and Britta Wenck at Safety Assessment, AstraZeneca R&D Södertälje, Sweden.

Abbreviations

AEBSF

4-(2-aminoethyl)benzenesulphonyl fluoride

ES

embryonic stem

GABA

γ-aminobutyric acid

GHB

gammahydroxybutyrate

PBST

phosphate-buffered saline/0.1%Tween 20

References

  1. ADDAE J.I., ROTHWELL N.J., STOCK M.J., STONE T.W. Activation of thermogenesis of brown fat in rats by baclofen. Neuropharmacology. 1986;25:627–631. doi: 10.1016/0028-3908(86)90215-7. [DOI] [PubMed] [Google Scholar]
  2. BOWERY N.G., ENNA S.J. Gamma-aminobutyric acidB receptors: first of the functional metabotropic heterodimers. J. Pharmacol. Exp. Ther. 2000;292:2–7. [PubMed] [Google Scholar]
  3. CARAI M.A., COLOMBO G., BRUNETTI G., MELIS S., SERRA S., VACCA G., MASTINU S., PISTUDDI A.M., SOLINAS C., CIGNARELLA G., MINARDI G., GESSA G.L. Role of GABAB receptors in the sedative/hypnotic effect of gamma-hydroxybutyric acid. Eur. J. Pharmacol. 2001;428:315–321. doi: 10.1016/s0014-2999(01)01334-6. [DOI] [PubMed] [Google Scholar]
  4. COUVE A., FILIPPOV A.K., CONNOLLY C.N., BETTLER B., BROWN D.A., MOSS S.J. Intracellular retention of recombinant GABAB receptors. J. Biol. Chem. 1998;273:26361–26367. doi: 10.1074/jbc.273.41.26361. [DOI] [PubMed] [Google Scholar]
  5. COUVE A., MOSS S.J., PANGALOS M.N. GABAB receptors: a new paradigm in G protein signaling. Mol. Cell. Neurosci. 2000;16:296–312. doi: 10.1006/mcne.2000.0908. [DOI] [PubMed] [Google Scholar]
  6. EKSTRAND J.New nucleotide sequences International Patent Application WO 1999. 99/21890
  7. FEIGENBAUM J.J., HOWARD S.G. Gamma hydroxybutyrate is not a GABA agonist. Prog. Neurobiol. 1996;50:1–7. doi: 10.1016/0301-0082(96)00029-9. [DOI] [PubMed] [Google Scholar]
  8. GRAY J.A., GOODWIN G.M., HEAL D.J., GREEN A.R. Hypothermia induced by baclofen, a possible index of GABAB receptor function in mice, is enhanced by antidepressant drugs and ECS. Br. J. Pharmacol. 1987;92:863–870. doi: 10.1111/j.1476-5381.1987.tb11392.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. HECHLER V., RATOMPONIRINA C., MAITRE M. Gamma-hydroxybutyrate conversion into GABA induces displacement of GABAB binding that is blocked by valproate and ethosuximide. J. Pharmacol. Exp. Ther. 1997;281:753–760. [PubMed] [Google Scholar]
  10. JOHANSSON C., GÖTHE S., FORREST D., VENNSTRÖM B., THORÉN P. Cardiovascular phenotype and temperature control in mice lacking thyroid hormone receptor-β or both α1 and β. Am. J. Physiol. 1999;276:H2006–H2012. doi: 10.1152/ajpheart.1999.276.6.H2006. [DOI] [PubMed] [Google Scholar]
  11. KAUPMANN K., HUGGEL K., HEID J., FLOR P.J., BISCHOFF S., MICKEL S.J., MCMASTER G., ANGST C., BITTIGER H., FROESTL W., BETTLER B. Expression cloning of GABAB receptors uncovers similarity to metabotropic glutamate receptors. Nature. 1997;386:239–246. doi: 10.1038/386239a0. [DOI] [PubMed] [Google Scholar]
  12. LAMP K., HUMENY A., NIKOLIC Z., IMAI K., ADAMSKI J., SCHIEBEL K., BECKER C.M. The murine GABAB receptor 1: cDNA cloning, tissue distribution, structure of the Gabbr1 gene, and mapping to chromosome 17. Cytogenet. Cell Genet. 2001;92:116–121. doi: 10.1159/000056880. [DOI] [PubMed] [Google Scholar]
  13. LINGENHOEHL K., BROM R., HEID J., BECK P., FROESTL W., KAUPMANN K., BETTLER B., MOSBACHER J. Gamma-hydroxybutyrate is a weak agonist at recombinant GABAB receptors. Neuropharmacology. 1999;38:1667–1673. doi: 10.1016/s0028-3908(99)00131-8. [DOI] [PubMed] [Google Scholar]
  14. MAITRE M. The gamma-hydroxybutyrate signalling system in brain: organization and functional implications. Prog. Neurobiol. 1997;51:337–361. doi: 10.1016/s0301-0082(96)00064-0. [DOI] [PubMed] [Google Scholar]
  15. MARSHALL F.H., JONES K.A., KAUPMANN K., BETTLER B. GABAB receptors – the first 7TM heterodimers. Trends Pharmacol. Sci. 1999;20:396–399. doi: 10.1016/s0165-6147(99)01383-8. [DOI] [PubMed] [Google Scholar]
  16. OBRIETAN K., VAN DEN POL A.N. GABAB receptor-mediated inhibition of GABAA receptor calcium elevations in developing hypothalamic neurons. J. Neurophysiol. 1998;79:1360–1370. doi: 10.1152/jn.1998.79.3.1360. [DOI] [PubMed] [Google Scholar]
  17. PATEL S., HILLARD C.J. Cannabinoid CB(1) receptor agonists produce cerebellar dysfunction in mice. J. Pharmacol. Exp. Ther. 2001;297:629–637. [PubMed] [Google Scholar]
  18. PROSSER H.M., GILL C.H., HIRST W.D., GRAU E., ROBBINS M., CALVER A., SOFFIN E.M., FARMER C.E., LANNEAU C., GRAY J., SCHENCK E., WARMERDAM B.S., CLAPHAM C., REAVILL C., ROGERS D.C., STEAN T., UPTON N., HUMPHREYS K., RANDALL A., GEPPERT M., DAVIES C.H, PANGALOS M.N. Epileptogenesis and enhanced prepulse inhibition in GABAB1-deficient mice. Mol. Cell. Neurosci. 2001;17:1059–1070. doi: 10.1006/mcne.2001.0995. [DOI] [PubMed] [Google Scholar]
  19. SCHULER V., LUSCHER C., BLANCHET C., KLIX N., SANSIG G., KLEBS K., SCHMUTZ M., HEID J., GENTRY C., URBAN L., FOX A., SPOOREN W., JATON A.L., VIGOURET J., POZZA M., KELLY P.H., MOSBACHER J., FROESTL W., KASLIN E., KORN R., BISCHOFF S., KAUPMANN K., VAN DER PUTTEN H., BETTLER B. Epilepsy, hyperalgesia, impaired memory, and loss of pre- and postsynaptic GABAB responses in mice lacking GABAB(1) Neuron. 2001;31:47–58. doi: 10.1016/s0896-6273(01)00345-2. [DOI] [PubMed] [Google Scholar]
  20. SNEAD O.C., III Evidence for a G protein-coupled gamma-hydroxybutyric acid receptor. J. Neurochem. 2000;75:1986–1996. doi: 10.1046/j.1471-4159.2000.0751986.x. [DOI] [PubMed] [Google Scholar]
  21. YAKIMOVA K., SANN H., SCHMID H.A., PIERAU F.K. Effects of GABA agonists and antagonists on temperature-sensitive neurones in the rat hypothalamus. J. Physiol. 1996;494:217–230. doi: 10.1113/jphysiol.1996.sp021486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. YAMAUCHI T., HORI T., TAKAHASHI T. Presynaptic inhibition by muscimol through GABAB receptors. Eur. J. Neurosci. 2000;12:3433–3436. doi: 10.1046/j.1460-9568.2000.00248.x. [DOI] [PubMed] [Google Scholar]
  23. ZARRINDAST M.R., OVEISSI Y. GABAA and GABAB receptor sites involvement in rat thermoregulation. Gen. Pharmacol. 1988;19:223–226. doi: 10.1016/0306-3623(88)90065-1. [DOI] [PubMed] [Google Scholar]
  24. ZUKIN S.R., YOUNG A.B., SNYDER S.H. Gamma-aminobutyric acid binding to receptor sites in the rat central nervous system. Proc. Natl. Acad. Sci. U.S.A. 1974;71:4802–4807. doi: 10.1073/pnas.71.12.4802. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from British Journal of Pharmacology are provided here courtesy of The British Pharmacological Society

RESOURCES