ArticlePDF Available

GABA Transporter Deficiency Causes Tremor, Ataxia, Nervousness, and Increased GABA-Induced Tonic Conductance in Cerebellum

Authors:

Abstract and Figures

GABA transporter subtype 1 (GAT1) knock-out (KO) mice display normal reproduction and life span but have reduced body weight (female, -10%; male, -20%) and higher body temperature fluctuations in the 0.2-1.5/h frequency range. Mouse GAT1 (mGAT1) KO mice exhibit motor disorders, including gait abnormality, constant 25-32 Hz tremor, which is aggravated by flunitrazepam, reduced rotarod performance, and reduced locomotor activity in the home cage. Open-field tests show delayed exploratory activity, reduced rearing, and reduced visits to the central area, with no change in the total distance traveled. The mGAT1 KO mice display no difference in acoustic startle response but exhibit a deficiency in prepulse inhibition. These open-field and prepulse inhibition results suggest that the mGAT1 KO mice display mild anxiety or nervousness. The compromised GABA uptake in mGAT1 KO mice results in an increased GABA(A) receptor-mediated tonic conductance in both cerebellar granule and Purkinje cells. The reduced rate of GABA clearance from the synaptic cleft is probably responsible for the slower decay of spontaneous IPSCs in cerebellar granule cells. There is little or no compensatory change in other proteins or structures related to GABA transmission in the mGAT1 KO mice, including GAT1-independent GABA uptake, number of GABAergic interneurons, and GABA(A)-, vesicular GABA transporter-, GAD65-, and GAT3-immunoreactive structures in cerebellum or hippocampus. Therefore, the excessive extracellular GABA present in mGAT1 KO mice results in behaviors that partially phenocopy the clinical side effects of tiagabine, suggesting that these side effects are inherent to a therapeutic strategy that targets the widely expressed GAT1 transporter system.
Content may be subject to copyright.
Neurobiology of Disease
GABA Transporter Deficiency Causes Tremor, Ataxia,
Nervousness, and Increased GABA-Induced Tonic
Conductance in Cerebellum
Chi-Sung Chiu,
1,6
Stephen Brickley,
2
Kimmo Jensen,
3,4
Amber Southwell,
1
Sheri Mckinney,
1
Stuart Cull-Candy,
5
Istvan Mody,
3
and Henry A. Lester
1
1
Division of Biology, California Institute of Technology, Pasadena, California 91125,
2
Biophysics Section, Imperial College London, London SW7 2AZ,
United Kingdom,
3
Departments of Neurology and Physiology, School of Medicine, University of California Los Angeles, Los Angeles, California 90095-1769,
4
Department of Physiology, University of Aarhus, DK-8000 Aarhus C, Denmark,
5
Department of Pharmacology, University College London, London WC1E
6BT, United Kingdom, and
6
Department of Neurobiology, Merck Research Laboratories, West Point, Pennsylvania 19486
GABA transporter subtype 1 (GAT1) knock-out (KO) mice display normal reproduction and life span but have reduced body weight
(female, 10%; male, 20%) and higher body temperature fluctuations in the 0.2–1.5/h frequency range. Mouse GAT1 (mGAT1) KO
mice exhibit motor disorders, including gait abnormality, constant 25–32 Hz tremor, which is aggravated by flunitrazepam, reduced
rotarod performance, and reduced locomotor activity in the home cage. Open-field tests show delayed exploratory activity, reduced
rearing, and reduced visits to the central area, with no change in the total distance traveled. The mGAT1 KO mice display no difference in
acoustic startle response but exhibit a deficiency in prepulse inhibition. These open-field and prepulse inhibition results suggest that the
mGAT1 KO mice display mild anxiety or nervousness. The compromised GABA uptake in mGAT1 KO mice results in an increased GABA
A
receptor-mediated tonic conductance in both cerebellar granule and Purkinje cells. The reduced rate of GABA clearance from the synaptic
cleft is probably responsible for the slower decay of spontaneous IPSCs in cerebellar granule cells. There is little or no compensatory
change in other proteins or structures related to GABA transmission in the mGAT1 KO mice, including GAT1-independent GABA uptake,
number of GABAergic interneurons, and GABA
A
-, vesicular GABA transporter-, GAD65-, and GAT3-immunoreactive structures in
cerebellum or hippocampus. Therefore, the excessive extracellular GABA present in mGAT1 KO mice results in behaviors that partially
phenocopy the clinical side effects of tiagabine, suggesting that these side effects are inherent to a therapeutic strategy that targets the
widely expressed GAT1 transporter system.
Key words: tiagabine; epilepsy; flunitrazepam; cerebellum; inhibition; tremor
Introduction
GABA is the principal inhibitory neurotransmitter in the mam-
malian brain, where it activates GABA
A
, GABA
B
, and GABA
C
receptors. GABA released from presynaptic terminals is removed
from the vicinity of the synaptic cleft by GABA transporters, and
this action is believed to be a key event in terminating synaptic
currents. GABA transporters are also involved in maintaining a
low extracellular GABA concentration throughout the brain, pre-
venting excessive tonic activation of synaptic and extrasynaptic
receptors. GABA transporters may also play a role in replenishing
the supply of presynaptic transmitter. Furthermore, GABA trans-
porters may reverse, under both normal and pathological cir-
cumstances, to release GABA (Richerson and Wu, 2003, 2004).
Of the three GABA transporters identified in the CNS, GABA
transporter subtype 1 (GAT1) is highly expressed in the olfactory
bulb, neocortex, cerebellum, superior colliculus, and substantia
nigra, where it is predominantly found in axons, presynaptic ter-
minals, and glial cells. GAT2 is weakly expressed throughout the
brain, primarily in arachnoid and ependymal cells. GAT3 expres-
sion is densest in the olfactory bulb, midbrain regions, and deep
cerebellar nuclei, where it is found predominantly on glial cells
(Radian et al., 1990; Ikegaki et al., 1994; Itouji et al., 1996; Yan et
al., 1997; Engel and Wu, 1998; Barakat and Bordey, 2002; Chiu et
al., 2002).
The GAT1 inhibitor tiagabine is a clinically useful antiepilep-
tic drug with few cognitive side effects (Aldenkamp et al., 2003),
but it also causes tremor (its major side effect), ataxia, dizziness,
asthenia, somnolence (sedation), and nonspecific nervousness
(Adkins and Noble, 1998; Pellock, 2001; Schachter, 2001). It is
important to know whether these side effects arise directly from
increased extracellular concentration of GABA in the CNS or,
instead, from actions on unintended targets. For instance, GAT1
Received Aug. 16, 2004; revised Jan. 25, 2005; accepted Jan. 25, 2005.
This research was supported by National Institutes of Health Grants DA-01921, NS-11756, MH-49176, NS-
030549, and DA-010509, National Science Foundation Grant 0119493, the Wellcome Trust, a Royal Society-Wolfson
Award (S.C.-C.), and a Della Martin Fellowship (C.-S.C.). We are indebted to members of Caltech and University of
California Los Angeles groups for advice, Limin Shi and Paul Patterson for use and help with the startle system, and
J. Crawley for comments on this manuscript.
Correspondence should be addressed to Henry A. Lester, Division of Biology, California Institute of Technology,
156-29, 1201 East California Boulevard, Pasadena, CA 91125. E-mail: lester@caltech.edu.
DOI:10.1523/JNEUROSCI.3364-04.2005
Copyright © 2005 Society for Neuroscience 0270-6474/05/253234-12$15.00/0
3234 The Journal of Neuroscience, March 23, 2005 25(12):3234 –3245
inhibitors may also inhibit GABA
A
receptors (Overstreet et al.,
2000; Jensen et al., 2003). If the latter mechanism holds, then a
more selective GAT1 inhibitor could be a more effective
antiepileptic.
To address this question, we examined the phenotype of the
ultimate GAT1-specific inhibitor: genetic interruption of GAT1
function. The homozygous and heterozygous mouse GAT1
(mGAT1) knock-out (KO) strain is viable and fertile, with a nor-
mal life span. Its hippocampal electrophysiology has been studied
previously (Jensen et al., 2003), but this is the first report of
several other phenotypes, including motor behavior, general
mood, cerebellar electrophysiology, and thermoregulation. We
emphasize measurements on the cerebellum, where GAT1 is
heavily expressed and has been quantified (Chiu et al., 2002).
GABA influences circadian rhythm (Liu and Reppert, 2000).
Because tiagabine-treated patients show dizziness, asthenia, and
somnolence, we determined whether the GAT1 KO mice display
altered activity in their habituated home cage. We also monitored
body temperature rhythm, which is synchronized with daily ac-
tivity (Weinert and Waterhouse, 1999). We found that the
mGAT1 KO mouse does phenocopy some effects of tiagabine,
which, in turn, suggests that the various clinical side effects of this
drug result, directly or indirectly, from its blockade of GAT1. We
measure altered synaptic physiology, deriving from increased
and prolonged extracellular [GABA], which provides a plausible
physiological basis for these effects.
Materials and Methods
GAT1 knock-out strain. The mGAT1 KO strain, previously termed
“intron-14-neo-mGAT1,” carries an intact neomycin selection marker
in intron 14. The details of the targeting construct, homologous recom-
bination, and genotyping were described previously (Chiu et al., 2002).
Synaptosomal GABA uptake assay. Details of synaptosomal prepara-
tion and GABA uptake assay were described previously (Chiu et al.,
2002). Briefly, mice were anesthetized with halothane (2-bromo-2-
chloro-1,1,1-trifluorothane), and brains were dissected and collected on
ice. The cerebellum (50 mg) was homogenized in 20(w/v) medium I
(0.32 Msucrose, 0.1 mMEDTA, and 5 mMHEPES, pH 7.5; 1 ml) (Nagy
and Delgado-Escueta, 1984). The P2 fraction (synaptosome fraction) was
suspended with 1 ml of medium I. Protein concentrations were analyzed
by using the Coomassie Plus kit (Pierce, Rockford, IL).
GABA uptake assays were performed by mixing 20
l of the suspen-
sion with 280
l of uptake buffer (in mM: 128 NaCl, 2.4 KCl, 3.2 CaCl
2
,
1.2 MgSO
4
, 1.2 KH
2
PO
4
, 10 glucose, 25 HEPES, pH 7.5) and then incu-
bated at 37°C for 10 min (Lu et al., 1998). GABA and [
3
H]GABA
in various concentrations (100
l) were added to the synaptosome sus-
pension and incubated for 10 min (final radioactive concentrations were
2.2– 8.8
Ci/ml). Uptake was terminated by placing the samples
in an ice-cold bath, followed by two washes with uptake buffer contain-
ing the same concentration of cold GABA at 10,000 g. The GABA
uptake inhibitor 1-[2-[[(diphenylmethylene)imino]oxy]ethyl]-1,2,5,6-
tetrahydro-3-pyridinecarboxylic acid hydrochloride (NO711) (final
concentration, 30
M) was included to measure the non-GAT1 uptake
activity; the NO711-sensitive fraction accounted for 75– 85% of wild-
type (WT) activity.
Tremor measurements. The mouse was placed ina2Lpolyethylene
freezer container. A piezoelectric transducer (LDT0 028K; Measure-
ment Specialties, Fairfield, NJ) was taped to the bottom of a 7.5 10 cm
plastic board (8 g), and this board was loosely attached to the bottom of
the container with a loop of paper tape. The mice were placed directly on
the board. The signal from the sensor was low-pass filtered at 200 Hz,
amplified by 100 (model 902; Frequency Devices, Haverhill, MA), and
led to the analog-to-digital inputs on an Axon DigiData 1200 interface
(Axon Instruments, Union City, CA). The signals were collected using
Clampex Gap-Free recording, and power spectra were computed in
ClampFit. We verified that the resonant frequency of this instrument was
far from the tremor frequency by replacing the mouse with 20 g of mass,
and the response of the instrument to constant-frequency mechanical
stimulation varied, with frequency, by 40% between 20 and 32 Hz.
Benzodiazepine modulation of the tremor. Mice were tested for baseline
tremor as described previously. They were then injected intraperitoneally
with either flunitrazepam in 20% FreAmine HBC (B Braun Medical,
Bethlehem, PA) or vehicle. After 15 min, the tremor was measured.
Footprint. Hindpaws were painted with black India ink, and mice were
placed in a cardboard box (90 12 12 cm) with a 75-cm-long white
paper floor. Paw angle is the hindpaw central axis relative to its walking
direction.
Rotarod. Mice were tested on a motorized rotarod (Ugo Basile, Com-
erio, Italy) consisting of a grooved metal roller (3 cm in diameter) and
separated 11-cm-wide compartments elevated 16 cm. The acceleration
rate was set at 0.15 rpm/s. Mice were placed on the roller, and the time
they remained on it during rotation was measured. The rotarod has an
increment of 4 rpm/step. Tests were performed for fixed speed at either
12 or 20 rpm and for accelerating speed. A maximum of 120 s was allowed
per animal for fixed speed tests.
Exploratory locomotor activity. An individual mouse was placed in a
novel environment of a square open field (50 50 cm), the floor of
which was divided into 25 smaller squares (5 5) by painted lines.
Within 10 cm of the chamber walls is termed the periphery (16 squares),
and the central region indicates the central nine squares. The animal
behavior in the open field was recorded by videotaping for 10 min and
analyzed subsequently. The measurements include delayed exploratory
activity (measuring the time required for mice to walk the first 50 cm),
frequency of visits to the central area, dwell time in the inner field, num-
ber of rearing events, total distance traveled, and walking speed. Mice
usually made short walks interrupted by brief stops. To make meaningful
walking-speed measurements, we chose uninterrupted walking for 25
cm and averaged 3–12 such walking-speed measurements for each ani-
mal. All animals were tested in a particular behavioral assay on the same
day during the light part of the light/dark cycle.
Elevated plus maze. Mice were allowed to habituate to the testing room
for 2 h. The maze consisted of two opposing open arms (40 10 cm) and
two opposing closed arms (40 10 cm, with 40 cm walls) on a platform
50 cm above the ground. Mice were placed in the center square (10 10
cm) facing an open arm and videotaped during a 5 min exploration. Arm
entries and duration were scored when all four paws entered the arm.
Partial arm entries were scored when one to three paws entered the
arm. Head dipping was scored when the head was dipped over the edge of
the maze. All animals were tested on the same day during the light part of
the light/dark cycle.
Home-cage activity. Mice were housed individually in cages with bed-
ding, food, and water. To assess activity, beam breaks were collected for
42 h with a photo beam system (San Diego Instruments, San Diego, CA).
Plots show the number of beam breaks for each 5 min interval.
Benzodiazepine hyperlocomotor activity. Mice were allowed to habitu-
ate to activity cages for 2 h. They were then injected intraperitoneally with
either flunitrazepam in 20% FreAmine HBC (5 or 15 mg/kg) or vehicle.
Activity (single beam breaks) and ambulation (successive beam breaks)
data were then collected for 1 h and plotted for each 5 min interval.
Acoustic startle and prepulse inhibition. Animals were tested in a Startle
Response system (SR-LAB; San Diego Instruments) consisting ofa5cm
Plexiglas cylinder mounted on a Plexiglas platform in a ventilated,
lighted, sound-attenuated chamber. Acoustic stimuli were presented by a
high-frequency loudspeaker mounted 28 cm above the cylinder. A piezo-
electric accelerometer attached to the Plexiglas base was used to detect
movement of the animals within the cylinder. Animal movement was
scored in arbitrary numbers between 0 and 1000. Ambient background
noise of 68 dB was maintained throughout each testing session. Each
session was initiated with a 5 min acclimation period followed by six 120
dB trials and concluded with another six 120 dB trials. These first and last
sets of six 120 dB pulses were not included in the analysis. For acoustic
startle-response (ASR) testing, seven different levels of acoustic startle
pulse (73, 78, 83, 85, 100, 110, and 120 dB) were presented along with a
trial containing only the background noise for 40 ms each in random
Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice J. Neurosci., March 23, 2005 25(12):3234 –3245 • 3235
order with variable intertrial intervals of 10 –20
s. At the onset of stimulus, 65 startle-amplitude
readings were taken for 1 ms each. Ten trials of
each decibel level were performed, and the av-
erage startle amplitude was determined. The
session used for prepulse inhibition (PPI) test-
ing consisted of five different trials presented 10
times each in random order. These include 120
dB startle pulse alone, 120 dB startle pulse pre-
ceded by a prepulse of 73, 78, or 83 dB (5, 10,
and 15 dB above background), and a trial con-
taining only the background noise. The per-
centage of prepulse inhibition was calculated as
follows: 100 [(average 120 dB startle pulse
average prepulse 120 dB startle pulse)/aver-
age 120 dB startle pulse].
Temperature measurements. Mini Mitter
(Sunriver, OR) ER-4000 telemetric tempera-
ture probes were used in 3- to 6-month-old
male mGAT1 KO mice. For implantation, mice
were anesthetized with halothane, anda1cm
incision was made at the back of the neck.
Probes were inserted subcutaneously into the
back. The incision was sealed with surgical glue.
The mice were housed with ad libitum water
and food at 24 2.5°C. Lights were on between
6:00 A.M. and 6:00 P.M. for 7–10 d after im-
plantations and then off for the period of data
collection. Temperature and activity data were
acquired using Vital View software (Mini-
Mitter) and analyzed (including fast Fourier
transforms) in Origin.
Seizure tests. Pentylenetetrazole (PTZ) was
solubilized in 0.9% NaCl saline solution, and
bicuculline was dissolved in 0.1N HCl, pH ad-
justed to 5.5 with 0.1N NaOH (Pericic and Bu-
jas, 1997). Animals were injected intraperitone-
ally with either PTZ or bicuculline. For PTZ,
animals were injected with either subthreshold
(40 mg/kg) or suprathreshold (70 mg/kg)
doses. For bicuculline, animals were injected
with 3, 4, or 5 mg/kg.
Brain slice electrophysiology. Cerebellar slices
were prepared using standard procedures
(Brickley et al., 1996). The brain was rapidly
dissected and submerged in cold slicing solu-
tion (4°C), which contained the following (in
mM): 125 NaCl, 2.5 KCl, 1 CaCl
2
, 4 MgCl
2
,25
NaHCO
3
, 1.25 NaH
2
PO
4
, and 25 glucose. All
extracellular solutions were bubbled with 95%
O
2
and 5% CO
2
, pH 7.4. After cutting on a
moving-blade microtome, slices were main-
tained at 32°C for 60 min before transfer to a
recording chamber. For granule cell recordings,
slices were constantly perfused (1.5 ml/min)
with recording solution containing the follow-
ing (in mM): 125 NaCl, 2.5 KCl, 2 CaCl
2
,1
MgCl
2
, 26 NaHCO
3
, 1.25 NaH
2
PO
4
, and 25 glucose. For Purkinje cell
recordings, slices were perfused with the following (in mM): 126 NaCl, 2.5
KCl, 2 CaCl
2
, 2 MgCl
2
, 26 NaHCO
3
, 1.25 NaH
2
PO
4
, 10 glucose, 0.2
L-ascorbic acid, 1 pyruvic acid, and 3 kynurenic acid. All experiments
were performed at room temperature, and whole-cell voltage-clamp re-
cordings were made using Axopatch 1D or 200B amplifiers (Axon Instru-
ments). The pipette solution contained the following (in mM): 140 CsCl,
4 NaCl, 0.5 CaCl
2
, 10 HEPES, 5 EGTA, 2 Mg-ATP, adjusted to pH 7.3
with CsOH.
Currents were filtered at 2–3 kHz and digitized at 10 kHz. The tonic
GABA
A
receptor-mediated conductance (G
GABA
) was measured from the
reduction in holding current recorded in the presence of the GABA
A
recep-
torantagonist 2-(3-carboxypropyl)-3-amino-6-(4-methoxyphenyl)-pyrida-
zinium bromide (SR95531) (100
M). All-points histograms were con-
structed from sections of data not containing synaptic currents and mean
values calculated from a Gaussian fit to the histogram. Spontaneous IPSCs
(sIPSCs) were detected with amplitude- and kinetics-based criteria (events
were accepted when they exceeded a threshold of 6 8 pA for 0.5 ms) using
custom-written LabView-5.1-based software (National Instruments, Aus-
tin, TX). All IPSCs were also inspected visually, and sweeps were rejected or
accepted manually. Individual spontaneously occurring IPSCs were then
aligned on their initial rising phase, and average IPSC waveforms were con-
structed from those events that exhibited a clear monotonic rise and re-
turned to baseline before the occurrence of later sIPSCs. The decay of average
Figure 1. mGAT1 KO cerebellar images, synaptosomal GABA uptake, and body weight. A, Fluorescent image of an mGAT1-GFP
knock-in mouse cerebellar cortex, showing typical GAT1 expression pattern. B, Fluorescent image of GAT1 KO showing no obvious
GAT1 expression pattern. C, WT mouse shows no obvious fluorescence. Band Cwere exposed to 20-fold greater photo power
thanA.GL,Granulecelllayer;ML,molecularlayer; P, Purkinje cell. Scale bar, 50
m.D,QuantificationofGAD65,vGAT,andGABA
A
receptor-containing boutons in WT and KO mice based on the immunocytochemical staining (see supplemental figure for the
actual images, available at www.jneurosci.org as supplemental material). E, The NO711-sensitive synaptosomal [
3
H]GABA up-
take activities among the three genotypes (mean SEM; triplicate assays from each of two experiments with all three geno-
types).E,DecreasedbodyweightoftheGAT1KOmouse.Data measured from 11 litters of het/het matings between the ages of 50
and 66 d are shown. Compared with WT littermates, male homozygotes weigh 20% less, whereas female homozygotes weigh
10% less. M, Male; F, female. WT, Het, KO: n8, 17, 15 for males; n11, 14, 8 for females. Differences from WT at *p0.05
and **p0.01.
3236 J. Neurosci., March 23, 2005 25(12):3234 –3245 Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice
sIPSC waveforms was quantified as a weighted
value calculated from the
charge transfer of normalized averages (
integral
).
Immunocytochemistry. Detailed procedures for immunocytochemis-
try were described previously (Chiu et al., 2002; Jensen et al., 2003). Mice
were anesthetized with halothane and perfused with 4% paraformalde-
hyde in PBS, pH adjusted to 7.6 with Na
2
HPO
4
. Brains were dissected
and kept in 4% paraformaldehyde for1hin4°Candthen incubated in
30% sucrose in PBS for 20 h. The brains were embedded in OCT
medium (Tissue-Tek; Miles, Elkhart, IN) for either horizontal or sagittal
sections and sliced by cryostat at 35
m. Brain slices were stored in a
solution containing the following (in mM): 11 NaH
2
PO
4
,20Na
2
HPO
4
,
30% ethylene glycol, and 30% glycerol, pH 7.5, at 20°C.
Sections were incubated for2hatroom temperature in a blocking
solution (10% normal goat serum and 0.3% Triton X-100 in PBS, pH
7.6), followed by incubation with the primary antibody for2dat4°C
with rotational mixing. Primary antibodies and their dilutions were rab-
bit anti-GAT3 (1:200 dilution; Chemicon, Temecula, CA), rabbit anti-
GABA
A
receptor
1 (1:100; Upstate Biotechnology, Lake Placid, NY),
rabbit anti-glutamate decarboxylase 65 (GAD65) (1:1000; Chemicon),
and rabbit anti-vesicular GABA transporter (vGAT) (1:100; Synaptic
Systems, Goettingen, Germany). The brain slices were first washed with
PBS containing 0.5% Triton X-100 followed by two additional washes
with PBS. The slices were then incubated in solutions containing the
appropriate rhodamine red-x-conjugated secondary antibodies. These
secondary antibodies include goat anti-rabbit, goat anti-guinea pig, or
donkey anti-goat secondary antibodies (1:200; Jackson ImmunoRe-
search, West Grove, PA). After three washes with PBS, slices were rinsed
with PBS, mounted with Vectashield (Vector Laboratories, Burlingame,
CA), and subjected to confocal microscope imaging.
Results
Evidence for functional knock-out of GAT1
The knock-in mouse strain studied here, previously termed
intron-14-neo-intact-mGAT1, harbors a neomycin resistance
cassette (neo) in intron 14 as well as a green
fluorescent protein (GFP) moiety fused to
the C terminus of the mGAT1 coding re-
gion in exon 14 (Jensen et al., 2003). This
strain was originally constructed as a ge-
netic intermediate in the eventual con-
struction of a neo-deleted mGAT1-GFP
knock-in strain that has also been de-
scribed previously (Chiu et al., 2002).
However, we found that the present strain
appears to have essentially no functional
mGAT1 (Jensen et al., 2003), presumably
because the neo sequences interfere with
mRNA or protein. Figure 1 shows addi-
tional evidence on this point in the cere-
bellum, for which we later provide electro-
physiological data. First, the GFP moiety at
the C terminus of the GAT1 construct pro-
vides a fluorescent label for the level of
GAT1 expression (Chiu et al., 2002). The
mGAT1 KO strain shows 2% as much
fluorescence as the mGAT1-GFP strain
(Chiu et al., 2002) and no more fluores-
cence than WT mice (Fig. 1A–C). Second,
to measure mGAT1 function, we per-
formed GABA uptake assays on cerebellar
synaptosomes. The NO711-sensitive
GABA uptake activity from mutant mice
synaptosomes was 2% of that of WT lit-
termates, whereas heterozygotes displayed
intermediate GABA uptake activity (Fig.
1E), indicating that mutant mice have little
or no functional presynaptic GAT1 activity. mGAT1-deficient mice
also display reduced body weight, 20 and 10% less than WT for
males and females, respectively (Fig. 1F).
Cerebellar immunocytochemistry
To test whether the mGAT1 KO mouse has abnormalities in the
GABAergic system, we performed immunocytochemistry on sev-
eral proteins related to GABA function. Immunocytochemistry
using antibodies against GAD65, the GABA
A
1 subunit and the
vGAT indicated that mGAT1 KO mice do not change GABAergic
synapse densities and related receptor expression in the molecu-
lar layer of the cerebellum (summarized in Fig. 1D) (based on
images in the supplemental figure, available at www.jneurosci.
org as supplemental material). We also found no qualitative dif-
ferences in GABA
A
1 subunit staining in the granule cell layer
(data not shown). These data agree with previous data on the
hippocampus (Jensen et al., 2003). Also, the expression pattern
for GAT3 is not changed, suggesting that no compensatory
changes occurred because of the GAT1 deficit (see supplemental
figure, available at www.jneurosci.org as supplemental material).
Immunocytochemistry using antibodies against GABAergic,
interneuron-specific, calcium-binding proteins showed no
changes in GABAergic interneuron density in the hippocampus
and in the cerebellum.
Behavioral characterizations of GAT1 KO mice
Tremor
The mGAT1 KO mice display readily observable, nearly contin-
uous tremor in the limbs and tail. Measured by a simple instru-
ment (Fig. 2A) (see Materials and Methods), the tremor fre-
Figure 2. Characterization of mGAT1 KO tremor. A, Recordings from the vibration transducer. Arrows (higher amplitude)
indicate activities when forepaws were raised. B, Power spectrum of the transducer signal for all genotypes. All genotypes shared
a minor peak at 80 Hz; however, only the KO showed a significant tremor at 25–32 Hz. C, Modulation of tremor frequency and
amplitude by flunitrazepam. Error bars represent SEM.
Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice J. Neurosci., March 23, 2005 25(12):3234 –3245 • 3237
quency is 25–32 Hz (Fig. 2B). In addition,
KO and WT share an additional lower am-
plitude tremor at 80 Hz (Fig. 2B)(n
6). Vibrations in both frequency ranges are
highest during rearing episodes (Fig. 2A,
arrows). Acute high-dose NO711 treat-
ment caused complete sedation in WT
mice, vitiating any observations on tremor
in NO711-treated WT mice.
Flunitrazepam treatment decreased the
frequency and increased the amplitude of
the tremor in mGAT1 KO mice (Fig. 2C) but
had very little effect on the power spectrum
of WT mice (data not shown). These effects
in KO mice were both significant for 15
mg/kg flunitrazepam; both effects were in-
termediate for 10 mg/kg flunitrazepam, but
only the frequency change was significant for
10 mg/kg flunitrazepam.
Ataxia
Ataxia is associated with cerebellar defects
in many strains of mice (Mullen et al.,
1976; Watanabe et al., 1998; Rico et al.,
2002). The mGAT1 KO mice walk with an
abnormally large paw angle relative to the
direction of walking: 23 0.7 versus
12.5 1.4° for KO and WT, respectively
(Fig. 3E,F). In another indication of
ataxia, mGAT1 KO mice display flattened
stance and lowered hip on the rotarod
(Fig. 3B). The mGAT1 KO mice show re-
duced time on the rotarod in both fixed
speed (Fig. 3C) and accelerating speed
(Fig. 3D) tests, indicating ataxia. Both WT
and mutant mice improved performance
on the rotarod after training; however, the
difference of latency to fall remained sig-
nificant between WT and KO mice (data
not shown). The mGAT1 KO and WT
mice displayed equal muscle strength in
hanging-wire activity tests (data not
shown).
Mild anxiety: flexor contraction,
exploratory activity, elevated plus maze,
and startle
When suspended by the tail, the mGAT1-
deficient mice display trembling and
flexor contraction (front paws held to-
gether and rear paws flexed) (Fig. 3A).
This gesture resembles typical mouse
models for anxiety. WT littermates dis-
played normal extension without trem-
bling (Fig. 3A). The flexor contraction was
also observed in WT mice treated with a high dose of NO711
(10 40 mg/kg; data not shown).
The open field was used as an additional test of anxiety-like
behavior (Fig. 4) (Prut and Belzung, 2003). Because an open field
is a novel environment, rodents tend to prefer the periphery of
the apparatus, later exploring the central parts of the open field.
We observed several aspects of behavior in this apparatus. The
mGAT1 KO mice tend to remain longer in the corner of the open
field (Fig. 4A) and then tend to walk slowly along the wall; thus,
there was markedly reduced frequency of visits to the central area
(Fig. 4B), reduced dwell time in the central area (Fig. 4C), and
reduced rearing activity (Fig. 4D). These results may signify anx-
iety of the mGAT1 KO mice. There was modestly reduced walk-
ing speed (Fig. 4E). Although WT and heterozygotes walk faster
than KO mice, they spend more time in rearing; as a consequence,
all three genotypes traveled about the same distance (Fig. 4F).
Several of the observations in the open-field test suggest that the
Figure 3. mGAT1 KO displays abnormal motor behavior. A, WT (left) and mGAT1 KO (right) mice showed different gestures
whenhungbytheirtails.WT mice showed a typical extensor gesture, whereas KO mice showed flexor contraction. B,StanceofWT
and KO mice on the rotarod. The KO mice show flattened and lowered hips, and their paws move more slowly than WT mice. C,
Mice were tested at fixed speed (either 12 or 20 rpm) on the rotarod. n6, 5, and 8 (WT, Het, and KO, respectively). D, Mice were
tested at accelerating speeds. KO mice fell significantly sooner than WT mice. E, Abnormal gait. Hindpaw footprint pattern of WT,
heterozygotes, and homozygotes is shown. The hindpaws of mGAT1 KO mice show a wider angle with respect to the direction of
walking. The KO mouse seems to waddle. F, Comparison of the average paw angles among WT (n8), Het (n9), and KO (n
17) mice. The paw angle of the KO mouse is approximately twice as large as that of WT and heterozygotes (23 1 vs 12.5 1°).
Differences from WT at *p0.05 and **p0.01.
3238 J. Neurosci., March 23, 2005 25(12):3234 –3245 Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice
heterozygote is the least anxious phenotype; we did not explore
this observation systemically.
We observed the mGAT1 KO mice in the elevated plus maze,
another test of anxiety (Fig. 5A,B). The mGAT1 KO mice dis-
played increased partial arm entries (Fig. 5A) and time spent in
the central square (Fig. 5B) compared with WT mice. Homozy-
gous mutant mice showed reduced open-arm entries and re-
duced total time spent in the open arms (data not shown). There
was a trend toward reduced closed-arm entries; however, this
difference was not statistically significant compared with WT.
Mutant mice spent the majority of the testing time in the central
square engaging in partial-arm entries, indicating no reduction in
locomotor activity. No difference was seen in head dipping. Thus,
the elevated plus maze provided some additional evidence for
anxiety.
Startle is a fast twitch of facial and body muscles evoked by
sudden and intense tactile, visual, or acoustic stimulations. Many
anxious mouse strains display both enhanced ASR and reduced
PPI. The mGAT1-deficient mice display normal ASR (Fig. 5C)
but reduced PPI (Fig. 5D), compared with their WT littermates.
The baseline movement of the mGAT1 mutant in the absence of
acoustic stimulation was elevated above the WT. This most likely
reflects the constant tremor of these mice.
Ambulation activity
The mGAT1 KO mice show reduced ambulation in their cages; as
a consequence, the 24 h activity cycle becomes less obvious (Fig.
6A). The total ambulation activities were 2425 395 versus
965 146 times during 42 h for WT and mGAT1 KO mice,
respectively (Fig. 6B) (mean SEM). Flunitrazepam treatment
caused hyperlocomotor activity in KO an-
imals and sedation in WT animals (data
not shown).
Autonomic regulation: body
temperature fluctuations
The mGAT1 KO mice display a striking
pattern of abnormal temperature regula-
tion (Fig. 7A,B). There is a normal circa-
dian temperature rhythm, but in addition,
there are many fluctuations, primarily hy-
perthermic episodes on a time scale of sev-
eral minutes to 2 h. To quantify these
fluctuations, we computed and averaged
the power spectral density of temperature
fluctuations in WT or KO mice (Fig. 7C).
The data have been normalized to the peak
at 0.0416 h
1
(corresponding to the circa-
dian rhythm). It is clear that mGAT1 KO
mice display increased relative noise
power in the frequency range from 0.2 to
1.5 h
1
. The mGAT1 hyperthermic epi-
sodes are larger, especially during high ac-
tivity (i.e., higher body temperature), but
no more frequent than in WT mice (Fig.
7B). Two additional animals in each group
provided similar data, but these animals
were not included in the averaged power
spectra because of differences in sample
rate.
Sensitivity to convulsants
The mGAT1 KO mouse is slightly more
sensitive than the WT mouse to PTZ-induced seizures, but there
is no obvious change in bicuculline-induced seizure susceptibil-
ity. Bicuculline (i.p.) at 5 mg/kg kills WT and mGAT1 KO mice,
whereas at 3 and 4 mg/kg, both WT and KO mice survived with
moderate seizure (n2 each). PTZ at a subthreshold dose (40
mg/kg, i.p.) decreased observable activity in WT and heterozy-
gotes while causing preconvulsive states and mild seizures in
mGAT1 KO mice (n3 each). At a suprathreshold dose (70
mg/kg), all WT and heterozygotes survived with severe seizures,
whereas mGAT1 KO mice showed severe seizures, and one of
three died (n3 each).
Cerebellar slice electrophysiology
GABA
A
receptor-mediated currents, recorded from wild-type
mice, are similar to those reported previously (Brickley et al.,
2001) (Fig. 8). Granule cells dialyzed with high-internal Cl
and
voltage clamped at 70 mV (see Materials and Methods) exhibit
sIPSCs, with a frequency of 0.8 0.6 Hz (n4). In addition, a
tonic GABA
A
receptor-mediated conductance (G
GABA
) is clearly
present in all recordings. The phasic and tonic conductances are
both blocked by the GABA
A
receptor antagonist SR95531 (100
M) (Fig. 8A). The magnitude of G
GABA
(84.2 50.4 pS/pF) is
similar to previous reports for animals of this age, as are the peak
amplitude (388.5 143.3 pS/pF) and kinetics (
integral
17.6
3.3 ms) of average sIPSCs (Brickley et al., 2001).
Recordings from mGAT1 KO cerebellar granule cells reveal
marked differences in both the tonic and phasic conductances,
consistent with the removal of a GABA transporter. In all seven
recordings from mGAT1 KO mice, G
GABA
is significantly in-
creased (Fig. 8B) to an average value of 318.9 65.6 pS/pF ( p
Figure 4. Characterization of mGAT1 KO exploratory activity in the open field. A, The time required for mice to walk the first 50
cm in open field. Most WT and Het mice take 10 s, whereas KO mice spend 13–240 s. Points show mean SEM. B,C, The KO
mouse shows reduced frequency (B) and reduced duration (C) visiting the central area in the open-field test. Total visits to the
central area were 15 2, 25 3, and 7 3, and total time to stay in the central area were 67 11, 106 23, and 21 7s
for WT, Het, and KO, respectively. D, The GAT1 KO mouse showed reduced frequencies of rearing (73 2, 87 9, and 29 10
for WT, Het, and KO, respectively). E, The average walking speeds for WT, Het, and KO mice were 16.2 0.5, 19.4 0.6, and
12.3 0.6 cm/s, respectively. F, mGAT1 KO mice showed no obvious difference in total walking distance within 10 min (2000
210, 2840 320, and 2190 300 for WT, Het, and KO, respectively). E,n7, 12, and 10 (WT, Het, and KO, respectively). For all
other panels, n6, 8, and 8 (WT, Het, and KO, respectively). *p0.05; **p0.01.
Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice J. Neurosci., March 23, 2005 25(12):3234 –3245 • 3239
0.05) (Fig. 8C). Conventional sIPSCs (Fig. 8D,E) are still detect-
able within the current record, albeit at an apparently lower av-
erage frequency (0.4 0.2 Hz). The increased current variance
associated with G
GABA
(Fig. 8E, histogram) made resolution of
small sIPSCs more difficult. Nevertheless, it appears that the av-
erage peak amplitude is not significantly different in the mGAT1
KO recordings (270.5 31.5 pS/pF). However, as shown in Fig-
ure 8F, the decay of sIPSCs is slower in the mGAT1 KO cells
(
integral
36.9 5.7 ms compared with 17.6 3.3 ms in wild-
type granule cells). Therefore, in mature cerebellar granule cells,
we observed an 300% increase in the magnitude of G
GABA
and
a 100% increase in the decay time of sIPSCs after the removal of
GAT1.
In Purkinje cell recordings, a standing inward GABA
A
receptor-mediated conductance, defined by sensitivity to the
GABA
A
receptor antagonist SR95531 (100
M), was observed
4
Figure 5. Additional anxiety-related behaviors: elevated plus maze and acoustic startle.
mGAT1KOmicedisplay increased partial arm entries(A)andtime spent in the central square(B)
compared with WT mice. *p0.01; n4 mice in each group. C, Acoustic startle response of
mutant (open circle; n4) and WT (filled square; n4) measured in arbitrary units. D,
Prepulse inhibition of mutant (open column; n7) and WT (filled column; n7). The 5, 10,
and 15 under the x-axis refer to prepulses at 5, 10, and 15 dB, respectively, above the back-
ground level of 68 dB. The difference between KO and WT is significant at *p0.05 and **p
0.01. Error bars represent SEM.
Figure 6. GAT1 KO mice showed reduced ambulation in home cages. A, Profiles of ambula-
tion activity of WT (top) and KO (bottom) mice over a 42 h recording period. WT displays a 24 h
rhythm, whereas KO shows lower activity. B, Total ambulation activity of KO and WT mice
(2425 395 and 965 146 counts). Error bars represent SEM.
3240 J. Neurosci., March 23, 2005 25(12):3234 –3245 Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice
in mGAT1 KO mice, which is much larger than in WT mice (75
19 pS/pF, n10 vs 13 5 pS/pF, n9, respectively) (Fig. 9).
However, the high frequency of sIPSCs consistently observed in
Purkinje cells (10 Hz) indicates that a comparison of sIPSC
kinetics between WT and mGAT KO mice is not possible in this
cell type because of the considerable superimposition of events.
Moreover, it is not feasible to selectively analyze the tonic and phasic
components of the GABA
A
-mediated conductance in a similar man-
ner to the granule cell recordings. Nonetheless, as shown in Figure 9,
it is clear that in Purkinje cell recordings, the magnitude of a standing
inward GABA
A
receptor-mediated conductance is significantly in-
creased in mGAT1 KO mice.
Discussion
The mGAT1 KO mouse as a model for tiagabine side effects
The distinct phenotype of mGAT1 KO mice includes ataxia,
tremor, sedation, nervousness (mild anxiety), increased fre-
quency and amplitude of body temperature fluctuations, and
reduced body weight. Similar behavioral patterns were also ob-
served in WT mice treated with either tiagabine or NO711, both
GAT1 inhibitors (Nielsen et al., 1991; Suzdak et al., 1992; Suzdak,
1994). Epileptic patients treated with Tiagabine display similar
side effects, including dizziness, asthenia, somnolence (sedation),
nonspecific nervousness, tremor, and ataxia (Adkins and Noble,
1998). The fact that the mGAT1 KO mice phenocopy many ef-
fects of both mice and humans treated with GAT1 inhibitors
suggests that the clinical side effects might be expected from any
systemically administered drug that tar-
gets GAT1, no matter how selective.
Synaptic basis of the tremor
GAT1 inhibition causes elevated extracel-
lular [GABA] and therefore generates an
increased tonic GABA
A
-mediated con-
ductance, perhaps primarily by acting at
areas that typically express high-affinity,
nondesensitizing GABA
A
receptors
(Brickley et al., 1996; Wall and Usowicz,
1997; Hamann et al., 2002; Jensen et al.,
2003). Our data for cerebellar granule (Fig.
8) and Purkinje (Fig. 9) cells support these
ideas. Previous studies also report a pro-
longation of the evoked GABA
A
receptor-
mediated synaptic decay after block of
GABA transporters (Dingledine and Korn,
1985; Roepstorff and Lambert, 1992, 1994;
Thompson and Gahwiler, 1992; Draguhn
and Heinemann, 1996; Rossi and Ha-
mann, 1998; Overstreet et al., 2000). This
phenomenon is not observed after action
potential-independent release (Thomp-
son and Gahwiler, 1992; Isaacson et al.,
1993), suggesting that GAT1 transporters
are likely to be more important in limiting
the GABA profile after multivesicular re-
lease. However, the use of GABA transport
blockers in previous assays may be compli-
cated by the fact that GAT1 inhibitors are
also competitive antagonists of GABA
A
re-
ceptors (Overstreet et al., 2000; Jensen et
al., 2003).
The cerebellar glomerulus, like the bas-
ket cell–Purkinje cell pinceau synapse and
the chandelier cell–pyramidal cell car-
tridge of cortex, is a highly organized synaptic structure that con-
tains many synaptic contacts produced by just a few presynaptic
inhibitory axons (Jakab and Hamori, 1988) and features a dense
level of GAT1 expression (Chiu et al., 2002). The dramatically
prolonged granule cell IPSC waveforms in mGAT1 KO mice are
certainly consistent with the idea that GAT1 plays a more impor-
tant role in clearing GABA after multivesicular release in struc-
tures such as the glomerulus, where diffusion is limited (Nielsen
et al., 2004). This may explain the greater prolongation of sIPSCs
we observe in mGAT1 KO granule cells (Fig. 8) than previously
observed in hippocampus (Jensen et al., 2003). The unchanged
level of GAD65 (Fig. 1 D), vGAT (Fig. 1D), the GABA receptor
1
subunit (Fig. 1D), GABA
B
receptors (Jensen et al., 2003), and
GAT3 (see supplemental figure, available at www.jneurosci.org
as supplemental material) in the mGAT1 KO mice argues against
some classes of compensatory changes in response to the chron-
ically elevated [GABA]. Furthermore, the NO711-insensitive cer-
ebellar synaptosomal GABA uptake was only 15–25% of the total
activity in WT, and the absolute value of NO711-insensitive
GABA uptake activity showed no difference between WT and
mGAT1 KO. However, we cannot rule out other changes such as
altered subunit composition of GABA
A
receptors or an altered
waveform of synaptically released [GABA]. Whatever the under-
lying synaptic mechanisms, the distorted inhibitory waveform
observed in granule cells suggests that inhibition in one or more
motor control nuclei provides a reasonable, although not quan-
Figure 7. mGAT1-deficient mice display more body temperature fluctuations in the 0.2–1.5/h frequency range than WT mice.
A, Raw traces of body temperature fluctuation from one WT (top) and one mutant (bottom) mouse. Mutant mice display multiple
hyperthermic episodes, especially during periods of higher activity (i.e., higher body temperature). B, Expanded traces from A.C,
Powerspectrumanalysis.They-axisrepresentstheaveragepower (n4 KO, 3 WT) normalized to the peak at 24 h cycle as 100%.
The x-axis represents the frequency (inverse hours).
Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice J. Neurosci., March 23, 2005 25(12):3234 –3245 • 3241
titative, explanation for the tremor that we
observed in the mGAT1 KO mouse. An
oscillation between excitation and inhibi-
tion underlies many neuronal pacemak-
ers, and in mGAT KO mice, this oscillation
is apparently timed in part by the accentu-
ated inhibitory phase that results from in-
creased and prolonged [GABA]. Fluni-
trazepam, an allosteric activator of GABA
A
receptors, increased the period and in-
creased the amplitude of the tremor (Fig.
2C), consistent with the idea that one
phase of the oscillation is governed by the
waveform of GABA
A
-mediated inhibition.
Which inhibitory synapse(s) domi-
nates the tremor? We do not imply that the
oscillation is solely determined by the tim-
ing of a cerebellar inhibitory synapse such
as the Golgi cell– granule cell contact.
The removal of GAT1 presumably alters
characteristics of GABA-mediated trans-
mission in many nuclei. GABA
A
receptor
1 subunit knock-out mice tremble at
18 Hz (Kralic et al., 2002), suggesting
that a tremor can arise from either too
little or too much GABAergic transmis-
sion throughout the brain. However, the
tremor in mGAT1 KO mice is inconsis-
tent with the low-frequency tremors
generally associated with basal ganglia
and midbrain pathology. The tremor
also has a higher frequency (25–32 Hz)
(Fig. 2) than most previously reported
mouse tremors but equal to that of mice
expressing the hypofunctional glycine
receptor (GlyR) oscillator
1 subunit
(Simon, 1997) or a human hyper-
ekplexia-related GlyR mutant (Becker et
al., 2002). Glycine transporter 2 knock-
out mice also display 15–25 Hz tremor
(Gomeza et al., 2003). These observa-
tions on the glycinergic system suggest
that the tremor is primarily spinal in
origin.
Ataxia
The ataxia exhibited by mGAT1-deficient mice (i.e., rotarod def-
icits) (Fig. 3C,D; broader paw angle in E,F) is more likely to
originate from a specific cerebellar defect, because ablation of
GABAergic neurons in the cerebellum also causes ataxia in sev-
eral classic mouse mutants. Overall, these results illustrate that
normal motor control depends on maintaining appropriate lev-
els of both phasic and tonic GABA
A
receptor-mediated inhibition
in the cerebellum.
Nervousness versus anxiety
Nervousness describes the clinical side effects of tiagabine (Do-
drill et al., 1997, 1998, 2000; Adkins and Noble, 1998). In the
absence of an accepted test for nervousness in rodents, we as-
sumed that it can be assessed as a mild form of anxiety. The GAT1
KO mice show such a phenotype. In the open-field test, mGAT1-
deficient mice display delayed exploratory activity and decreased
frequency of visits to the central area (Fig. 4A–C). However, the
reduced rearing (Fig. 4D) could be caused by simply the de-
creased motor ability that leads to the lowered stance (Fig. 3B);
there was only moderately reduced walking speed (Fig. 4E) and
no reduction in total distance traveled (Fig. 4F). Furthermore,
mutant mice show no difference in acoustic startle response com-
pared with WT (Fig. 5C), but they display a dramatic decrease in
prepulse inhibition of the acoustic startle response (Fig. 5D). The
mGAT1 KO displays reduced home-cage activity, but the modest
decrement in open-field walking speed suggests that mutant mice
remain active when encountering novel environments, whereas
they display reduced activity in a habituated environment (Fig.
5A,B). In contrast, 5-HT transporter null mice exhibit a classical
pattern of increased anxiety-like behavior in the elevated plus
maze, in light– dark exploration and emergence tests, and in
open-field tests (Holmes et al., 2003).
It is also true that many classical anxiolytic drugs operate by
increasing the activity of GABA
A
receptors. Likewise, reduced
GABA also causes anxiety; for example, GAD65 knock-out mice
exhibit increased anxiety-like behavior in both the open-field and
elevated-zero maze assays (Kash et al., 1999).
Figure 8. mGAT1 KO cerebellar granule cells are characterized by an increased tonic GABA
A
-mediated conductance and pro-
longedIPSCs.A,B,Continuous current records from typical wild-type (A) andmGAT1KO(B) internal granule cells voltage clamped
at 70 mV. The horizontal line indicates the 0 current level in each recording. There is an increased inward current in mGAT1 KO
cells and a substantial increase in the current variance associated with this conductance. This increased tonic conductance is
completely blocked by the GABA
A
receptor antagonist SR95531 (gabazine). C, The bar graph illustrates that, on average, G
GABA
in
GAT1 KO granule cells was 319 65 pS/pF (n7) compared with 84 50 pS/pF (n4) in control littermates. This resembles
the 98 20 pS/pF G
GABA
recorded previously in the C57BL/6 strain (Brickley et al., 2001). Therefore, the tonic conductance tripled
after the removal of GAT1, indicating a raised concentration of ambient GABA in the slice preparation. D,E, Two average sIPSC
waveforms recorded from a wild-type (D) and an mGAT1 KO (E) granule cell are shown on the same scale. The waveforms have
similar peak amplitudes but very different decays. The histograms also illustrate the peak amplitude distribution of all sIPSCs
recorded in these cells. The open histograms were constructed from periods of baseline noise. As shown by the increase in the
width of the baseline histogram for mGAT1 KO, the increased current variance associated with mGAT1 KO recordings does
complicateinterpretationofpeakamplitudemeasurements. It is possible that we are missing a significant fraction of small events
in the mGAT1 KO, because they would be unresolved in the noisy mGAT1 KO recordings. However, this possible artifact does not
affect the decay estimates, because the decay of sIPSCs is not correlated with peak amplitude in granule cells (data not shown). F,
The significant increase in the decay of sIPSC recorded from mGAT1 KO granule cells. The decay was defined as
integral
(see
Materials and Methods). The
integral
of control littermates was 13 5ms(n4) compared with 37 6ms(n5) in the
mGAT1 KO animals. In contrast, there was no significant difference between the average peak amplitudes recorded in the two
strains. Error bars represent SEM.
3242 J. Neurosci., March 23, 2005 25(12):3234 –3245 Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice
Reduced body weight
The reduced body weight of mGAT1 KO mice (Fig. 1 F) contrasts
with obesity of transgenic mice overexpressing mGAT1 under
nonspecific or pan-neuronal promoters (Ma et al., 2000). GABA-
related regulatory mechanism of feeding behavior in the ventro-
medial hypothalamus may be responsible for impaired responses
to glucoprivation in genetically obese rats (Tsujii and Bray, 1991).
Benzodiazepine-treated rats lose body weight, presumably via
activation of GABA
A
receptors (Blasi, 2000). Excess GABA in the
anterior piriform cortex region reduces feeding (Truong et al.,
2002). We believe that the reduced body weight and tremor is not
related to delayed (or retarded) development, because mGAT1
KO mice are reproductive at the same time as WT and display
muscle strength and balance similar to WT. Additional detailed
studies are required.
Thermoregulation and circadian rhythm
Thermoregulation is controlled by several brain regions, includ-
ing the horizontal limb of the diagonal band of Broca (HDB), the
basal forebrain, the preoptic area (POA), and the rostral part of
the raphe pallidus nucleus (rRPa). Many neurons in these areas
are GABAergic. In the HDB, muscimol reduces thermosensitivity
(Hays et al., 1999) and, in the rRPa, muscimol to rRPa blocks
fever and thermogenesis in brown adipose tissue induced by
intra-POA as well as by intracerebroventricular prostaglandin E2
applications (Nakamura et al., 2002).
We know of no clinical studies on temperature effects of tiaga-
bine. However, the higher amplitude of hyperthermic episodes in
the mGAT1 KO mouse (Fig. 7) clearly does not phenocopy the
acute hypothermic effects of tiagabine in rodents (Inglefield et al.,
1995). Interestingly, GABA
B
activation leads to hypothermia
(Schuler et al., 2001), but we found previously that the presynap-
tic GABA
B
response is diminished or lost in mGAT1 KO mice
(Jensen et al., 2003), which may explain the discrepancy.
Although GABA has been related to circadian rhythm in many
publications (Liu and Reppert, 2000; Wagner et al., 2001), the
mGAT1-deficient mice did not display obvious changes in circa-
dian rhythm during5doftesting either in constant dark or in a
12 h light/dark cycle environment. These results suggest that ex-
cess GABA does not affect circadian rhythm.
An additional use for knock-out mice strains
To the other useful information obtained from knock-out mouse
strains, we may add the decision regarding whether the clinical
side effects of a drug (in this case, tiagabine) arise from either
widespread expression of its target or nonselective actions on
other targets. Such information is particularly valuable when the
pleiotropic effects cannot readily be predicted from, but are cer-
tainly consistent with, the widespread and varied roles of the
target molecule. Of course, such a study is rather straightforward
when it is believed that the effects are mostly acute and subject to
straightforward neurological tests (as in the present case), rather
than delayed and primarily psychiatric (as for serotonin and per-
haps dopaminergic and noradrenergic transporters).
References
Adkins JC, Noble S (1998) Tiagabine. A review of its pharmacodynamic and
pharmacokinetic properties and therapeutic potential in the management
of epilepsy. Drugs 55:437–460.
Aldenkamp AP, De Krom M, Reijs R (2003) Newer antiepileptic drugs and
cognitive issues. Epilepsia 44 [Suppl 4]:21–29.
Barakat L, Bordey A (2002) GAT-1 and reversible GABA transport in Berg-
mann glia in slices. J Neurophysiol 88:1407–1419.
Becker L, von Wegerer J, Schenkel J, Zeilhofer HU, Swandulla D, Weiher H
(2002) Disease-specific human glycine receptor
1 subunit causes hy-
perekplexia phenotype and impaired glycine- and GABA
A
-receptor trans-
mission in transgenic mice. J Neurosci 22:2505–2512.
Blasi C (2000) Influence of benzodiazepines on body weight and food intake
Figure 9. mGAT1 KO mice display higher tonic currents in cerebellar Purkinje cells. In both
WT and mGAT1 KO slices, sIPSCs were recorded in Purkinje cells by holding at 70 mV. Zero
current levels are shown in the light trace. Injection of SR95531 into the bath (100
M; heavy
trace) blocked tonic current and sIPSCs. Average tonic currents in mGAT1 KO cells are approxi-
mately six times larger than in WT cells [75 19 and 13 5 pS/pF for KO (n10) and WT
(n9), respectively]. Error bars represent SEM.
Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice J. Neurosci., March 23, 2005 25(12):3234 –3245 • 3243
in obese and lean Zucker rats. Prog Neuropsychopharmacol Biol Psychi-
atry 24:561–577.
Brickley SG, Cull-Candy SG, Farrant M (1996) Development of a tonic
form of synaptic inhibition in rat cerebellar granule cells resulting from
persistent activation of GABA
A
receptors. J Physiol (Lond) 497:753–759.
Brickley SG, Revilla V, Cull-Candy SG, Wisden W, Farrant M (2001) Adap-
tive regulation of neuronal excitability by a voltage-independent potas-
sium conductance. Nature 409:88–92.
Chiu CS, Jensen K, Sokolova I, Wang D, Li M, Deshpande P, Davidson N,
Mody I, Quick MW, Quake SR, Lester HA (2002) Number, density, and
surface/cytoplasmic distribution of GABA transporters at presynaptic
structures of knock-in mice carrying GABA transporter subtype 1-green
fluorescent protein fusions. J Neurosci 22:10251–10266.
Dingledine R, Korn SJ (1985)
-Aminobutyric acid uptake and the termi-
nation of inhibitory synaptic potentials in the rat hippocampal slice.
J Physiol (Lond) 366:387–409.
Dodrill CB, Arnett JL, Sommerville KW, Shu V (1997) Cognitive and qual-
ity of life effects of differing dosages of tiagabine in epilepsy. Neurology
48:1025–1031.
Dodrill CB, Arnett JL, Shu V, Pixton GC, Lenz GT, Sommerville KW (1998)
Effects of tiagabine monotherapy on abilities, adjustment, and mood.
Epilepsia 39:33–42.
Dodrill CB, Arnett JL, Deaton R, Lenz GT, Sommerville KW (2000) Tiaga-
bine versus phenytoin and carbamazepine as add-on therapies: effects on
abilities, adjustment, and mood. Epilepsy Res 42:123–132.
Draguhn A, Heinemann U (1996) Different mechanisms regulate IPSC ki-
netics in early postnatal and juvenile hippocampal granule cells. J Neuro-
physiol 76:3983–3993.
Engel JE, Wu CF (1998) Genetic dissection of functional contributions of
specific potassium channel subunits in habituation of an escape circuit in
Drosophila. J Neurosci 18:2254–2267.
Gomeza J, Ohno K, Hulsmann S, Armsen W, Eulenburg V, Richter DW,
Laube B, Betz H (2003) Deletion of the mouse glycine transporter 2
results in a hyperekplexia phenotype and postnatal lethality. Neuron
40:797–806.
Hamann M, Rossi DJ, Attwell D (2002) Tonic and spillover inhibition of
granule cells control information flow through cerebellar cortex. Neuron
33:625–633.
Hays TC, Szymusiak R, McGinty D (1999) GABA
A
receptor modulation of
temperature sensitive neurons in the diagonal band of Broca in vitro.
Brain Res 845:215–223.
Holmes A, Yang RJ, Lesch KP, Crawley JN, Murphy DL (2003) Mice lacking
the serotonin transporter exhibit 5-HT
1A
receptor-mediated abnormali-
ties in tests for anxiety-like behavior. Neuropsychopharmacology
28:2077–2088.
Ikegaki N, Saito N, Hashima M, Tanaka C (1994) Production of specific
antibodies against GABA transporter subtypes (GAT1, GAT2, GAT3) and
their application to immunocytochemistry. Brain Res Mol Brain Res
26:47–54.
Inglefield JR, Perry JM, Schwartz RD (1995) Postischemic inhibition of
GABA reuptake by tiagabine slows neuronal death in the gerbil hip-
pocampus. Hippocampus 5:460468.
Isaacson JS, Solis JM, Nicoll RA (1993) Local and diffuse synaptic actions of
GABA in the hippocampus. Neuron 10:165–175.
Itouji A, Sakai N, Tanaka C, Saito N (1996) Neuronal and glial localization
of two GABA transporters (GAT1 and GAT3) in the rat cerebellum. Brain
Res Mol Brain Res 37:309–316.
Jakab RL, Hamori J (1988) Quantitative morphology and synaptology of
cerebellar glomeruli in the rat. Anat Embryol (Berl) 179:81–88.
Jensen K, Chiu CS, Sokolova I, Lester HA, Mody I (2003) GABA
transporter-1 (GAT1)-deficient mice: differential tonic activation of
GABA
A
versus GABA
B
receptors in the hippocampus. J Neurophysiol
90:2690–2701.
Kash SF, Tecott LH, Hodge C, Baekkeskov S (1999) Increased anxiety and
altered responses to anxiolytics in mice deficient in the 65-kDa isoform of
glutamic acid decarboxylase. Proc Natl Acad Sci USA 96:1698–1703.
Kralic JE, O’Buckley TK, Khisti RT, Hodge CW, Homanics GE, Morrow AL
(2002) Deletion of GABA
A
receptor (GABA
A
-R)
1 subunit alters ben-
zodiazepine (BZD) site pharmacology, function and related behavior. Soc
Neurosci Abstr 28:39.12.
Liu C, Reppert SM (2000) GABA synchronizes clock cells within the supra-
chiasmatic circadian clock. Neuron 25:123–128.
Lu Y, Grady S, Marks MJ, Picciotto M, Changeux JP, Collins AC (1998)
Pharmacological characterization of nicotinic receptor-stimulated GABA
release from mouse brain synaptosomes. J Pharmacol Exp Ther
287:648657.
Ma YH, Hu JH, Zhou XG, Zeng RW, Mei ZT, Fei J, Guo LH (2000) Trans-
genic mice overexpressing gamma-aminobutyric acid transporter sub-
type I develop obesity. Cell Res 10:303–310.
Mullen RJ, Eicher EM, Sidman RL (1976) Purkinje cell degeneration, a new
neurological mutation in the mouse. Proc Natl Acad Sci USA 73:208 –212.
Nagy A, Delgado-Escueta AV (1984) Rapid preparation of synaptosomes
from mammalian brain using nontoxic isoosmotic gradient material
(Percoll). J Neurochem 43:1114–1123.
Nakamura K, Matsumura K, Kaneko T, Kobayashi S, Katoh H, Negishi M
(2002) The rostral raphe pallidus nucleus mediates pyrogenic transmis-
sion from the preoptic area. J Neurosci 22:46004610.
Nielsen EB, Suzdak PD, Andersen KE, Knutsen LJS, Sonnewald U, Braestrup
C (1991) Characterization of Tiagabine (NO-328), a new potent and
selective GABA uptake inhibitor. Eur J Pharmacol 196:257–266.
Nielsen TA, DiGregorio DA, Silver RA (2004) Modulation of glutamate
mobility reveals the mechanism underlying slow-rising AMPAR EPSCs
and the diffusion coefficient in the synaptic cleft. Neuron 42:757–771.
Overstreet LS, Jones MV, Westbrook GL (2000) Slow desensitization regu-
lates the availability of synaptic GABA
A
receptors. J Neurosci
20:7914–7921.
Pellock JM (2001) Tiagabine (gabitril) experience in children. Epilepsia 42
[Suppl 3]:49–51.
Pericic D, Bujas M (1997) Sex differences in the response to GABA antago-
nists depend on the route of drug administration. Exp Brain Res
115:187–190.
Prut L, Belzung C (2003) The open field as a paradigm to measure the effects
of drugs on anxiety-like behaviors: a review. Eur J Pharmacol 463:3–33.
Radian R, Ottersen OP, Strorm-Mathisen J, Castel M, Kanner BI (1990)
Immunocytochemical localization of the GABA transporter in rat brain.
J Neurosci 10:1319–1330.
Richerson GB, Wu Y (2003) Dynamic equilibrium of neurotransmitter
transporters: not just for reuptake anymore. J Neurophysiol
90:1363–1374.
Richerson GB, Wu Y (2004) Role of the GABA transporter in epilepsy. Adv
Exp Med Biol 548:76–91.
Rico B, Xu B, Reichardt LF (2002) TrkB receptor signaling is required for
establishment of GABAergic synapses in the cerebellum. Nat Neurosci
5:225–233.
Roepstorff A, Lambert JD (1992) Comparison of the effect of the GABA
uptake blockers, tiagabine and nipecotic acid, on inhibitory synaptic effi-
cacy in hippocampal CA1 neurones. Neurosci Lett 146:131–134.
Roepstorff A, Lambert JDC (1994) Factors contributing to the decay of the
stimulus-evoked IPSC in rat hippocampal CA1 neurons. J Neurophysiol
72:2911–2926.
Rossi DJ, Hamann M (1998) Spillover-mediated transmission at inhibitory
synapses promoted by high affinity
6 subunit GABA
A
receptors and
glomerular geometry. Neuron 20:783–795.
Schachter SC (2001) Pharmacology and clinical experience with tiagabine.
Expert Opin Pharmacother 2:179–187.
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 AL, Vigouret J, Pozza M,
Kelly PH, Mosbacher J, Froestl W, Kaslin E, Korn R, et al. (2001) Epi-
lepsy, hyperalgesia, impaired memory, and loss of pre- and postsynaptic
GABA
B
responses in mice lacking GABA
B(1)
. Neuron 31:47–58.
Simon ES (1997) Phenotypic heterogeneity and disease course in three mu-
rine strains with mutations in genes encoding for alpha 1 and beta glycine
receptor subunits. Mov Disord 12:221–228.
Suzdak PD (1994) Lack of tolerance to the anticonvulsant effects of tiagab-
ine following chronic (21 day) treatment. Epilepsy Res 19:205–213.
Suzdak PD, Frederiksen K, Andersen KE, Sorensen PO, Knutsen LJ, Nielsen
EB (1992) NNC-711, a novel potent and selective gamma-aminobutyric
acid uptake inhibitor: pharmacological characterization. Eur J Pharmacol
224:189–198.
Thompson SM, Gahwiler BH (1992) Effects of the GABA uptake inhibitor
tiagabine on inhibitory synaptic potentials in rat hippocampal slice cul-
tures. J Neurophysiol 67:1698–1701.
3244 J. Neurosci., March 23, 2005 25(12):3234 –3245 Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice
Truong BG, Magrum LJ, Gietzen DW (2002) GABA(A) and GABA(B) re-
ceptors in the anterior piriform cortex modulate feeding in rats. Brain Res
924:1–9.
Tsujii S, Bray GA (1991) GABA-related feeding control in genetically obese
rats. Brain Res 540:48–54.
Wagner S, Sagiv N, Yarom Y (2001) GABA-induced current and circadian
regulation of chloride in neurones of the rat suprachiasmatic nucleus.
J Physiol (Lond) 537:853–869.
Wall MJ, Usowicz MM (1997) Development of action potential-dependent
and independent spontaneous GABA
A
receptor-mediated currents in
granule cells of postnatal rat cerebellum. Eur J Neurosci 9:533–548.
Watanabe D, Inokawa H, Hashimoto K, Suzuki N, Kano M, Shigemoto R,
Hirano T, Toyama K, Kaneko S, Yokoi M, Moriyoshi K, Suzuki M, Koba-
yashi K, Nagatsu T, Kreitman RJ, Pastan I, Nakanishi S (1998) Ablation
of cerebellar Golgi cells disrupts synaptic integration involving GABA
inhibition and NMDA receptor activation in motor coordination. Cell
95:17–27.
Weinert D, Waterhouse J (1999) Daily activity and body temperature
rhythms do not change simultaneously with age in laboratory mice.
Physiol Behav 66:605–612.
Yan XX, Cariaga WA, Ribak CE (1997) Immunoreactivity for GABA plasma
membrane transporter, GAT-1, in the developing rat cerebral cortex:
transient presence in the somata of neocortical and hippocampal neu-
rons. Brain Res Dev Brain Res 99:1–19.
Chiu et al. Behavior and Electrophysiology of GAT1 KO Mice J. Neurosci., March 23, 2005 25(12):3234 –3245 • 3245
... So far, the GAT1-selective inhibitor tiagabine is the only FDA (US Food and Drug Administration) approved drug targeting GAT, indicated for the treatment of partial epileptic seizures ( Fig. 1) [10]. Recently, in drug discovery, selective inhibition of extrasynaptic GAT3 and BGT1 have been pursued to avoid the numerous sideeffects linked to GAT1 inhibition, such as asthenia, dizziness, nervousness, and depression, providing some proof-of-concept for inhibiting extrasynaptic GATs as a way to enhance tonic inhibition [12,13]. The focus of this study has been GAT3, which is a key target in regulating GABA metabolism via uptake of GABA into astrocytic processes [14] and relevant for both pathologies characterized by increased tonic GABAergic inhibition such as the chronic phase after an ischemic stroke [15] and certain types of absence epilepsy [16]. ...
Article
Full-text available
The GABA transporter 3 (GAT3) is a member of the GABA transporter (GAT) family proposed to have a role in regulating tonic inhibition. The GAT3-preferring substrate (S)-isoserine has shown beneficial effects in a mouse model of stroke accompanied by an increased GAT3 expression, indicating a molecular mechanism mediated by GAT3. However, (S)-isoserine is not ideally suited for in vivo studies due to a lack of selectivity and brain permeability. To elucidate the structural determinants of (S)-isoserine for GAT3 inhibition, and to optimize and inform further ligand development, we here present the design, synthesis and pharmacological evaluation of a series of conformationally constrained isoserine analogs with defined stereochemistry. Using [³H]GABA uptake assays at recombinant human GAT3, we identified the azetidine and pyrrolidine analogs (2S,2´S)-6 and (2S,2´S)-7 as the most potent inhibitors. To further elaborate on the selectivity profile both compounds were tested at all GATs, the taurine transporter (TauT) and GABAA receptors. Although (2S,2´S)-6 and (2S,2´S)-7 are comparable to (S)-isoserine with respect to potency, the selectivity vs. the taurine transporter was significantly improved (at least 6 and 53 times more activity at hGAT3, respectively). A subsequent comprehensive structure-activity study showed that different connectivity approaches, stereochemical variations, simple or larger α- and N-substituents, and even minor size enlargement of the heterocyclic ring all abrogated GAT3 inhibition, indicating very strict stereochemical and size requirements. The observed structure activity relationships may guide future ligand optimization and the novel ligands ((2S,2´S)-6 and (2S,2´S)-7) can serve as valuable tools to validate the proposed GAT3-mediated effect of (S)-isoserine such as in functional recovery after stroke and thus help corroborate the relevance of targeting GAT3 and tonic inhibition in relevant brain pathologies.
... Interestingly, in the cortex, such inhibition of either GAT1 or GAT3 did not increase tonic inhibition, whereas simultaneous inhibition of both did 84 , indicating synergistic cortical GABA uptake by GAT1 and GAT3. In addition to the hippocampus and cortex, active GABA uptake by GATs decreased tonic inhibition in various other brain regions such as the thalamus 42 , cerebellum 85 , striatum 86 , globus pallidus 79 and suprachiasmatic nucleus 87 , emphasizing the essential role of GATs in regulating GABA tone. It is worth mentioning that GAT expression is not exclusive to neurons or astrocytes. ...
Article
γ-Aminobutyric acid (GABA) is the major inhibitory neurotransmitter released at GABAergic synapses, mediating fast-acting phasic inhibition. Emerging lines of evidence unequivocally indicate that a small amount of extracellular GABA — GABA tone — exists in the brain and induces a tonic GABA current that controls neuronal activity on a slow timescale relative to that of phasic inhibition. Surprisingly, studies indicate that glial cells that synthesize GABA, such as astrocytes, release GABA through non-vesicular mechanisms, such as channel-mediated release, and thereby act as the source of GABA tone in the brain. In this Review, we first provide an overview of major advances in our understanding of the cell-specific molecular and cellular mechanisms of GABA synthesis, release and clearance that regulate GABA tone in various brain regions. We next examine the diverse ways in which the tonic GABA current regulates synaptic transmission and synaptic plasticity through extrasynaptic GABAA-receptor-mediated mechanisms. Last, we discuss the physiological mechanisms through which tonic inhibition modulates cognitive function on a slow timescale. In this Review, we emphasize that the cognitive functions of tonic GABA current extend beyond mere inhibition, laying a foundation for future research on the physiological and pathophysiological roles of GABA tone regulation in normal and abnormal psychiatric conditions.
... So far, the GAT1-selective inhibitor tiagabine is the only FDA (US Food and Drug Administration) approved drug targeting GAT, indicated for the treatment of partial epileptic seizures (Fig. 1) [10]. Recently, in drug discovery, selective inhibition of extrasynaptic GAT3 and BGT1 have been pursued to avoid the numerous side-effects linked to GAT1 inhibition, such as asthenia, dizziness, nervousness, and depression, providing some proof-of-concept for inhibiting extrasynaptic GATs as a way to enhance tonic inhibition [12, 13]. On the other hand, in pathologies characterized by increased tonic GABAergic inhibition such as the chronic phase after an ischemic stroke [14] and certain types of absence epilepsy [15], another strategy is needed as GAT inhibitors would only exacerbate GABA signalling further. ...
Preprint
Full-text available
The GABA transporter 3 (GAT3) is a member of the GABA transporter (GAT) family proposed to have a role in regulating tonic inhibition. The GAT3-preferring substrate ( S )-isoserine has shown beneficial effects in a mouse model of stroke accompanied by an increased GAT3 expression, indicating a molecular mechanism mediated by GAT3. However, ( S )-isoserine is not ideally suited for in vivo studies due to a lack of selectivity and brain permeability. To elucidate the structural determinants of ( S )-isoserine for GAT3 inhibition, and to optimize and inform further ligand development, we here present the design, synthesis and pharmacological evaluation of a series of conformationally constrained isoserine analogues with defined stereochemistry. Using [ ³ H]GABA uptake assays at recombinant human GAT3, we identified the azetidine and pyrrolidine analogs (( S,S )- 6a and ( S,S )- 7a ) as the most potent inhibitors. To further elaborate on the selectivity profile both compounds were tested at all GATs, the taurine transporter (TauT) and GABA A receptors. Although ( S,S )- 6a and ( S,S )- 7a are comparable to ( S )-isoserine with respect to potency, the selectivity versus the taurine transporter was significantly improved (at least 6 and 53 times more activity at hGAT3, respectively). A subsequent comprehensive structure-activity study showed that different connectivity approaches, stereochemical variations, simple or larger α- and N - substituents, and even minor size enlargement of the alicyclic ring all abrogated GAT3 inhibition, indicating very strict stereochemical and size requirements. The observed structure activity relationships may guide future ligand optimization and the novel ligands (( S,S )- 6a and ( S,S )- 7a ) can serve as valuable tools to validate the proposed GAT3-mediated effect of ( S )-isoserine such as in functional recovery after stroke and thus help corroborate the relevance of targeting GAT3 and tonic inhibition in relevant brain pathologies.
Preprint
Full-text available
The beam walk is widely used to study coordination and balance in rodents. While the task has ethological validity, the main endpoints of foot slip counts and time to cross are prone to human-rater variability and offer limited sensitivity and specificity. We asked if machine learning-based methods could reveal previously hidden, but biologically relevant, insights from the task. Marker-less pose estimation, using DeepLabCut, was deployed to label 13 anatomical points on mice traversing the beam. Next, we automated classical endpoint detection, including foot slips, with high recall (>90%) and precision (>80%). A total of 395 features were engineered and a random-forest classifier deployed that, together with skeletal visualizations, could test for group differences and identify determinant features. This workflow, named Forestwalk, uncovered pharmacological treatment effects in C57BL/6J mice, revealed phenotypes in transgenic mice used to study Angelman syndrome and SLC6A1-related neurodevelopmental disorder, and will facilitate a deeper understanding of how the brain controls balance in health and disease.
Article
Full-text available
Enhanced GABAergic neurotransmission contributes to impairment of motor coordination and gait and of cognitive function in different pathologies, including hyperammonemia and hepatic encephalopathy. Neuroinflammation is a main contributor to enhancement of GABAergic neurotransmission through increased activation of different pathways. For example, enhanced activation of the TNFα–TNFR1-NF-κB-glutaminase-GAT3 pathway and the TNFα-TNFR1-S1PR2-CCL2-BDNF-TrkB pathway in cerebellum of hyperammonemic rats enhances GABAergic neurotransmission. This is mediated by mechanisms affecting GABA synthesizing enzymes GAD67 and GAD65, total and extracellular GABA levels, membrane expression of GABA A receptor subunits, of GABA transporters GAT1 and GAT three and of chloride co-transporters. Reducing neuroinflammation reverses these changes, normalizes GABAergic neurotransmission and restores motor coordination. There is an interplay between GABAergic neurotransmission and neuroinflammation, which modulate each other and altogether modulate motor coordination and cognitive function. In this way, neuroinflammation may be also reduced by reducing GABAergic neurotransmission, which may also improve cognitive and motor function in pathologies associated to neuroinflammation and enhanced GABAergic neurotransmission such as hyperammonemia, hepatic encephalopathy or Parkinson’s disease. This provides therapeutic targets that may be modulated to improve cognitive and motor function and other alterations such as fatigue in a wide range of pathologies. As a proof of concept it has been shown that antagonists of GABA A receptors such as bicuculline reduces neuroinflammation and improves cognitive and motor function impairment in rat models of hyperammonemia and hepatic encephalopathy. Antagonists of GABA A receptors are not ideal therapeutic tools because they can induce secondary effects. As a more effective treatment to reduce GABAergic neurotransmission new compounds modulating it by other mechanisms are being developed. Golexanolone reduces GABAergic neurotransmission by reducing the potentiation of GABA A receptor activation by neurosteroids such as allopregnanolone. Golexanolone reduces neuroinflammation and GABAergic neurotransmission in animal models of hyperammonemia, hepatic encephalopathy and cholestasis and this is associated with improvement of fatigue, cognitive impairment and motor incoordination. This type of compounds may be useful therapeutic tools to improve cognitive and motor function in different pathologies associated with neuroinflammation and increased GABAergic neurotransmission.
Article
Depression and anxiety are two mental disorders prevailing among adolescents. However, issues regarding the trajectory of depression and anxiety are still controversial on both the disease and symptom dimensions. The novel method of network analysis was used to provide insight into the symptom dimension. 20,544 adolescents (female = 10,743, 52.3%) aged between 14 and 24 years (age mean ± sd = 16.9 ± 2.94) were divided into three subgroups according to age so that the course of depression and anxiety could be traced. Network analysis and the Bayesian network model were used in the current study. The results indicated that uncontrollable worry - excessive worry was the most significant edge for all adolescents, whereas concentration - motor had the highest edge weights for early adolescents, and anhedonia - energy was the most critical pairwise symptom for middle and late adolescents. Irritability can bridge anxiety and depression in the early and middle stages of adolescence, while suicide plays a bridging role in the early and late stages of adolescence. Restlessness and guilt can bridge anxiety and depression in middle- and late-stage adolescents, and feeling afraid plays a unique role in middle-stage adolescents. Except for sad mood, which can trigger middle adolescents' anxiety and depression, the other three subgroups were mainly triggered by nervousness. In addition, all results in our current study were shown to be stable and accurate. In treatment, targeting central and triggering symptoms at different stages of adolescence may be critical to alleviating the comorbidity of anxiety and depression.
Article
Pyrethroids (PYRs) are a group of synthetic organic chemicals that mimic natural pyrethrins. Due to their low toxicity and persistence in mammals, they are widely used today. PYRs exhibit higher lipophilicity than other insecticides, which allows them to easily penetrate the blood-brain barrier and directly induce toxic effects on the central nervous system. Several studies have shown that the cerebellum appears to be one of the regions with the largest changes in biomarkers. The cerebellum, which is extremely responsive to PYRs, functions as a crucial region for storing motor learning memories. Exposure to low doses of various types of PYRs during rat development resulted in diverse long-term effects on motor activity and coordination functions. Reduced motor activity may result from developmental exposure to PYRs in rats, as indicated by delayed cerebellar morphogenesis and maturation. PYRs also caused adverse histopathological and biochemical changes in the cerebellum of mothers and their offspring. By some studies, PYRs may affect granule cells and Purkinje cells, causing damage to cerebellar structures. Destruction of cerebellar structures and morphological defects in Purkinje cells are known to be directly related to functional impairment of motor coordination. Although numerous data support that PYRs cause damage to cerebellar structures, function and development, the mechanisms are not completely understood and require further in-depth studies. This paper reviews the available evidence on the relationship between the use of PYRs and cerebellar damage and discusses the mechanisms of PYRs.
Article
Full-text available
Hereditary hyperekplexia is caused by disinhibition of motoneurons resulting from mutations in the ionotropic receptor for the inhibitory neurotransmitter glycine (GlyR). To study the pathomechanisms involved in vivo, we generated and analyzed transgenic mice expressing the hyperekplexia-specific dominant mutant human GlyR alpha1 subunit 271Q. Tg271Q transgenic mice, in contrast to transgenic animals expressing a wild-type human alpha1 subunit (tg271R), display a dramatic phenotype similar to spontaneous and engineered mouse mutations expressing reduced levels of GlyR. Electrophysiological analysis in the ventral horn of the spinal cord of tg271Q mice revealed a diminished GlyR transmission. Intriguingly, an even larger reduction was found for GABA(A)-receptor-mediated inhibitory transmission, indicating that the expression of this disease gene not only affects the glycinergic system but also leads to a drastic downregulation of the entire postsynaptic inhibition. Therefore, the transgenic mice generated here provide a new animal model of systemic receptor interaction to study inherited and acquired neuromotor deficiencies at different functional levels and to develop novel therapeutic concepts for these diseases.
Article
Although glial GABA uptake and release have been studied in vitro, GABA transporters (GATs) have not been characterized in glia in slices. Whole cell patch-clamp recordings were obtained from Bergmann glia in rat cerebellar slices to characterize carrier-mediated GABA influx and efflux. GABA induced inward currents at −70 mV that could be pharmacologically separated into GABA A receptor and GAT currents. In the presence of GABA A/B/C receptor blockers, mean GABA-induced currents measured −48 pA at −70 mV, were inwardly rectifying between −70 and +50 mV, were inhibited by external Na ⁺ removal, and were diminished by reduction of external Cl ⁻ . Nontransportable blockers of GAT-1 (SKF89976-A and NNC-711) and a transportable blocker of all the GAT subtypes (nipecotic acid) reversibly reduced GABA-induced transport currents by 68 and 100%, respectively. A blocker of BGT-1 (betaine) had no effect. SKF89976-A and NNC-711 also suppressed baseline inward currents that likely result from tonic GAT activation by background GABA. The substrate agonists, nipecotic acid and β-alanine but not betaine, induced voltage- and Na ⁺ -dependent currents. With Na ⁺ and GABA inside the patch pipette or intracellular GABA perfusion during the recording, SKF89976-A blocked baseline outward currents that activated at −60 mV and increased with more depolarized potentials. This carrier-mediated GABA efflux induced a local accumulation of extracellular GABA detected by GABA A receptor activation on the recorded cell. Overall, these results indicate that Bergmann glia express GAT-1 that are activated by ambient GABA. In addition, GAT-1 in glia can work in reverse and release sufficient GABA to activate nearby GABA receptors.
Article
Synopsis Tiagabine is a γ-aminobutyric acid (GABA) uptake inhibitor which is structurally related to nipecotic acid but has an improved ability to cross the blood-brain barrier. Clinical trials have shown that tiagabine is effective as add-on therapy in the management of patients with refractory partial epilepsy. In short term studies of this indication, tiagabine ≤ 64 mg/day for 7 to 12 weeks reduced the complex partial and simple partial seizure frequency by ≥ 50% in 8 to 31 and 28.2 to 37% of patients, respectively. Tiagabine appeared to produce a sustained reduction in seizure frequency in studies of up to 12 months’ duration. Data from preliminary studies are currently insufficient to confirm the usefulness of tiagabine when used as monotherapy or in the treatment of children with epilepsy. Further studies are, therefore, necessary to more fully elucidate the efficacy of the drug in these settings. Adverse events associated with tiagabine are primarily CNS-related and include dizziness, asthenia, nonspecific nervousness and tremor. Skin rash or psychosis occurred with similar frequencies among tiagabine- and placebo-treated patients. With long term administration (≥ 1 year for many patients), the profile and incidence of adverse events was similar to that for short term therapy. Tiagabine does not appear to affect the hepatic metabolism of other drugs such as carbamazepine and phenytoin. Possible disadvantages of tiagabine include its short plasma elimination half-life, necessitating 2 to 4 times daily administration, and its inducible hepatic metabolism. Thus, tiagabine is a new antiepileptic agent with a novel mechanism of action, which has demonstrated efficacy in the adjunctive treatment of patients with refractory partial epilepsy. Further investigation of the efficacy of tiagabine is expected to provide a clearer definition of its place in the treatment of epilepsy and its relative merits in relation to other antiepileptic drugs. Pharmacodynamic Properties Tiagabine increases synaptosomal concentrations of the inhibitory neurotransmitter γ-aminobutyric acid (GABA) via inhibition of the GABA transporter GAT-1. The increase in synaptic concentrations of GABA leads to potentiation of GABA-mediated inhibitory neurotransmission. Tiagabine lacks appreciable affinity for other receptor or uptake sites including benzodiazepine, histamine H1, serotonin 5-HT1B or dopamine D1 or D2 receptors or β1− or β2−adrenoceptors. It is not a substrate for the GABA uptake carrier and is therefore unlikely to act as a false transmitter. Tiagabine is active in a number of animal seizure models, protecting against seizures induced by chemical [e.g. methyl-6,7-dimethoxy-4-ethyl-β-carboline-3-carboxylate (DMCM) and pentylenetetrazol (PTZ)] and nonchemical stimuli (e.g. audiogenic and kindling). It is a more potent anticonvulsant than the conventional antiepileptics phenytoin, phenobarbital, carbamazepine and valproic acid against audiogenic and DMCM- and PTZ-induced tonic or clonic seizures in mice and rats. Tiagabine is also more potent than lamotrigine, gabapentin and vigabatrin in protecting against audiogenic and DMCM- and PTZ-induced tonic or clonic seizures in mice and was the only drug able to block PTZ-induced clonic seizures in mice. Tiagabine may be proconvulsant in animal models of non-convulsive epilepsy. Results from short and long term studies in patients with epilepsy have revealed no clinically significant deterioration in cognitive performance or electroencephalographic changes during tiagabine therapy and suggest that the drug may even have a slight beneficial effect on cognition under certain circumstances. Pharmacokinetic Properties Tiagabine is well absorbed after oral administration and has an absolute oral bioavailability of 90%. Peak plasma concentrations occurred approximately 1 hour after administration and measured 43 to 552 μg/L in healthy volunteers after single-dose administration of tiagabine 2 to 24mg. Multiple-dose administration of tiagabine does not result in significant drug accumulation. The rate, but not the extent, of absorption of tiagabine is decreased by concomitant food intake. Tiagabine is widely distributed throughout the body (volume of distribution is approximately 1 L/kg) and approximately 96% of the drug in human plasma is bound to plasma proteins. Tiagabine is extensively metabolised by the hepatic cytochrome P450 enzyme CYP3A. 63% of a radiolabelled orally administered dose was excreted in the faeces and 25% in the urine. The plasma clearance of tiagabine is 21.4 L/h in patients receiving concomitant enzyme-inducing antiepileptic drugs such as carbamazepine, phenytoin and primidone and 12.8 L/h in patients with epilepsy not receiving concomitant treatment with these agents. The elimination half-life of tiagabine ranges from 3.8 to 9 hours in patients with epilepsy; patients receiving concomitant treatment with enzyme-inducing antiepileptic drug therapy exhibit values at the lower end of this range. Conversely, the clearance of tiagabine is reduced in patients with hepatic impairment, which may necessitate dosage reduction. Renal impairment or old age do not appear to significantly reduce the clearance of tiagabine; however, clearance of the drug appears to be slightly increased in children. Although there is some evidence to suggest a relationship between tiagabine plasma concentration and therapeutic effect, there are currently insufficient data to recommend routine monitoring of plasma tiagabine concentrations in patients receiving the drug. Concomitant administration of tiagabine has been shown not to influence the pharmacokinetics of conventional antiepileptic drugs such as carbamazepine and phenytoin or other drugs including oral contraceptives, theophylline, warfarin and digoxin. A significant decrease (approximately 10 to 12%) in the peak plasma concentration and area under the plasma concentration-time curve for valproic acid has been reported in patients receiving concomitant tiagabine therapy; however, because valproic acid has a wide therapeutic range (50 to 100 mg/L), thisdecrease was considered to be of limited clinical significance. Therapeutic Efficacy The results of 5 double-blind placebo-controlled studies (2 crossover and 3 parallel-group studies) have shown tiagabine ≤64 mg/day to be effective as add-on therapy in patients with refractory partial epilepsy. According to a pooled analysis of these studies, 23% of patients treated with tiagabine, compared with 9% of placebo-treated patients, experienced a reduction in total seizure frequency of at least 50% during 7 to 12 weeks of treatment. The reduction in 4-week overall seizure frequency (25 vs 0.1%) and increase in the number of seizure-free days(6 vs 0%) were significantly greater for tiagabine than for placebo recipients. Individual results from 3 of these studies showed a ≥50% decrease in seizure rate relative to baseline in 8 to 31% and 28.2 to 37% of patients with complex partial and simple partial seizures, respectively. A corresponding response rate of 63% was reported for patients who experienced secondarily generalised tonic-clonic seizures in 1 study. The clinical benefits of adjunctive tiagabine therapy appear to be maintained during long term treatment. Approximately 30 to 40% of patients treated with tiagabine ≤80 mg/day for up to 12 months in nonblind extension studies continued to experience a ≥50% reduction in overall seizure frequency. Initial studies have reported a beneficial effect with tiagabine monotherapy in some patients with epilepsy refractory to monotherapy with other antiepileptic drugs. Among 31 patients recruited to a dose-ranging study, 19 were converted to tiagabine monotherapy for ≥2 weeks and 12 of these patients completed the study at a mean dosage of 38.4 mg/day. In a second study, efficacy rates (≥50% reduction in complex partial seizure frequency from baseline) were 31 and 18%, respectively, after treatment with tiagabine 36 and 6 mg/day (intent-to-treat analysis). Notably, both studies reported a high rate of study withdrawal (>50%); this was not an unexpected finding, as the refractory nature of the disease in patients recruited to these studies made successful conversion to tiagabine monotherapy less likely. The few data available from the only comparative study of tiagabine monotherapy (vs carbamazepine monotherapy) in patients with newly diagnosed partial epilepsy do not allow an accurate assessment of efficacy. Studies are therefore required to establish the efficacy of tiagabine in relation to conventional and newer antiepileptic agents. In small noncomparative studies, tiagabine (0.25 to 1.5 mg/kg/day or 32 to 80 mg/day) also demonstrated efficacy as add-on therapy in children and adolescents (aged 2 to 16 years) with refractory partial epilepsy, producing a >50% reduction in partial seizure frequency in approximately 20 to 50% of patients. Tolerability The most frequent adverse events in 675 patients with refractory epilepsy treated with tiagabine as add-on therapy (typically ≤64 mg/day for up to 12 weeks) were dizziness (30%), asthenia (24%), nonspecific nervousness (12%) and tremor (9%) [pooled data from placebo-controlled studies]. This compared with a significantly lower incidence of 13, 12, 3 and 3%, respectively, among placebo recipients (n = 363). Diarrhoea, depressed mood and emotional lability were also significantly more frequent with tiagabine (4 to 7% vs 1 to 2%). Other adverse events, including psychosis and skin rash which are frequently associated with other antiepileptic drugs, occurred with a similar frequency among tiagabine- and placebo-treated patients. These and other symptoms were generally mild or moderate and occurred early in the course of therapy. Treatment discontinuation due to adverse effects has been reported in 15% of patients treated with adjunctive tiagabine therapy and 33% of patients treated with tiagabine monotherapy. Longer term tolerability data (including results from 814 patients treated with tiagabine for ≥1 year), suggest that the profile and incidence of adverse effects associated with long term tiagabine therapy are similar to those reported during short term therapy. Clinically significant changes in haematological or liver function tests have not been observed during tiagabine therapy. Furthermore, the risk of status epilepticus appears to be minimal and similar to that reported for patients with epilepsy not treated with tiagabine. Dosage and Administration Dosage recommendations for the use of tiagabine as an adjunct to enzyme-inducing antiepileptic drug therapy in adult and adolescent patients with partial epilepsy differ between the US and Europe. In the US in adults, tiagabine should be initiated at a dosage of 4 mg once daily, which may be increased by 4 to 8 mg/day at weekly intervals until clinical response is achieved or a maximum dosage of 32 to 56 mg/day (administered in 2 to 4 divided doses) is reached. In adolescents (aged 12 to 18 years), the maximum recommended dosage is ≤32 mg/day. In contrast, European guidelines recommend an initial dosage of 7.5 to 15 mg/day in adults and adolescents (aged >12 years), followed by weekly incremental increases of 5 to 15 mg/day; the usual maintenance dosage is 30 to 50 mg/day administered in 3 divided doses. If cessation of tiagabine therapy is necessary, the dosage should be reduced gradually over a period of 2 to 3 weeks. Tiagabine should be administered with food, and dosage reduction may be necessary in patients with mild or moderate hepatic impairment. In Europe tiagabine is contraindicated in patients with severe hepatic impairment. There are currently insufficient data to recommend the use of tiagabine in pregnant or lactating women.
Article
GABA(B) (gamma-aminobutyric acid type B) receptors are important for keeping neuronal excitability under control. Cloned GABA(B) receptors do not show the expected pharmacological diversity of native receptors and it is unknown whether they contribute to pre- as well as postsynaptic functions. Here, we demonstrate that Balb/c mice lacking the GABA(B(1)) subunit are viable, exhibit spontaneous seizures, hyperalgesia, hyperlocomotor activity, and memory impairment. Upon GABA(B) agonist application, null mutant mice show neither the typical muscle relaxation, hypothermia, or delta EEG waves. These behavioral findings are paralleled by a loss of all biochemical and electrophysiological GABA(B) responses in null mutant mice. This demonstrates that GABA(B(1)) is an essential component of pre- and postsynaptic GABA(B) receptors and casts doubt on the existence of proposed receptor subtypes.
Article
A new procedure is described for the isolation of synaptosomes from various parts of mammalian brain. This method utilizes an isoosmotic Percoll/sucrose discontinuous gradient and has some advantages over the traditionally used synaptosomal isolation techniques: (1) it is possible to prepare suitable gradients while retaining isoosmolarity; (2) the time of the preparation is remarkably short (approximately 1 h); (3) if necessary, the gradient material can be easily removed from the samples. Intact synaptosomes were recovered from the 10%/16% (vol/vol) Percoll interphase. The fractions were identified and characterized by electron microscopy and by several biochemical markers for synaptosomes and other subcellular organelles. The homogeneity of the preparations is comparable to or better than that of synaptosomes prepared by the conventional methods. This procedure has been successfully used for the isolation of synaptosomes from very small tissue samples of various experimental animals and human brain.
Article
The neuroprotective effects of enhancing neuronal inhibition with a -aminobutyric acid (GABA) uptake inhibitor were studied in gerbil hippocampus following transient ischemia. We used in vivo microdialysis to determine a suitable dosing regimen for tiagabine (NNC 328) to elevate extracellular levels of GABA within the hippocampus. In anesthetized (normothermic) gerbils, tiagabine (45 mg/kg, i. p.) selectively elevated extracellular GABA levels 450% in area CA1 of the hippocampus. In gerbils subjected to cerebral ischemia via 5-min bilateral carotid occlusion, extracellular GABA levels increased 13-fold in area CA1, returning to baseline within 30–45 min. When tiagabine was injected 10 min following onset of reperfusion, GABA levels remained elevated (200–470%) for 90 min. In addition, tiagabine significantly reduced the ischemic-induced elevation of glutamate levels in area CA1 during the postischemic period when GABA levels were elevated. There was no effect of postischemic tiagabine on aspartate or six other amino acids. Using the same dosing regimen, we evaluated the degree of neuroprotection in the hippocampus of gerbils 4 and 21 days after ischemia. Tiagabine decreased body temperature a maximum of 2.7°C beginning 30 min into reperfusion and lasting 90 min. In untreated gerbils sacrificed 4 and 21 days after ischemia, there was severe necrosis (99%) of the pyramidal cell layer in area CA1. Whereas tiagabine significantly protected the CA1 pyramidal cell layer in ischemic gerbils at 4 days (overt necrosis confined to about 17% of area CA1), the protection diminished significantly 21 days postischemia. When normothermia was maintained both during and after ischemia in a separate group of tiagabine-treated animals, approximately 77% of the CA1 pyramidal cell layer was necrotic at 4 days. Based on these findings, we suggest that (1) tiagabine slows the development of hippocampal degeneration following ischemia, and (2) that mild, postischemic hypothermia is responsible, in large part, for the neuroprotective actions of this drug. We conclude that the histological outcome after administration of cerebral neuroprotectants should be assessed following long-term survival. © 1995 Wiley-Liss, Inc.
Article
Purpose: We evaluated the dose-related impacts of tiagabine (TGB) on cognition and mood in a monotherapy study. Methods: Patients were 123 adults with uncontrolled partial seizures, each treated with a single currently available antiepileptic drug (AED) for management of clinical epilepsy. They completed a battery of neuropsychological tests during an 8 week prospective baseline period and once again at the end of the 12-week fixed-dose period (or earlier if they dropped out of the study). Sixty-six patients were randomized to 6 mg/day TGB and 57 were randomized to 36 mg/day TGB. Results: Few changes in either abilities or adjustment and mood were noted when all patients were considered as a single group. However, analysis of both dose and attainment of TGB monotherapy showed that patients receiving TGB monotherapy did best, improving particularly in the areas of adjustment and mood with low-dose TGB and in the area of abilities with high-dose TGB. Patients who did not attain monotherapy showed no change except that the high-dose group did not perform as well on measures of mood and adjustment. Baseline AED and changes in seizure control did not affect the results. Conclusions: Patients attainment of TGB monotherapy was associated with their achievement of positive changes of varying degree on psychological tests. Failure to attain TGB monotherapy was associated with no changes on the tests except in patients receiving high-dose TGB where it appeared that some alterations in mood might have been avoided if a slower titration schedule had been used.