You are on page 1of 14

J Neurophysiol 91: 1036 –1049, 2004.

First published September 10, 2003; 10.1152/jn.00364.2003.

Developmental-Dependent Action of Microtubule Depolymerization on the


Function and Structure of Synaptic Glycine Receptor Clusters in Spinal
Neurons

Brigitte van Zundert,1,2 Francisco J. Alvarez,2 Juan Carlos Tapia,1 Hermes H. Yeh,3 Emilio Diaz,1 and
Luis G. Aguayo1
1
Laboratory of Neurophysiology, Department of Physiology, University of Concepción, Concepción, Chile; 2Department of Anatomy,
Wright State University, Dayton, Ohio 45435; and 3Center for Aging and Developmental Biology and Department of Pharmacology and
Physiology, University of Rochester Medical Center, Rochester, New York 14627
Submitted 11 April 2003; accepted in final form 2 September 2003

van Zundert, Brigitte, Francisco J. Alvarez, Juan Carlos Tapia, Cotman 1978; Matus and Taff-Jones 1978), and several bridg-
Hermes H. Yeh, Emilio Diaz, and Luis G. Aguayo. Developmen- ing proteins have been proposed to link membrane macromol-
tal-dependent action of microtubule depolymerization on the function ecules to the underlying subsynaptic cytoskeleton within the
and structure of synaptic glycine receptor clusters in spinal neurons. J PSD (reviewed in Sheng and Pak 2000).
Neurophysiol 91: 1036 –1049, 2004. First published September 10,
2003; 10.1152/jn.00364.2003. Microtubules have been proposed to
Gephyrin, the first protein identified as a synaptic bridging
interact with gephyrin/glycine receptors (GlyRs) in synaptic aggre- component, is a key constituent of the PSD at inhibitory
gates. However, the consequence of microtubule disruption on the synapses (Triller et al. 1985, 1987). Gephyrin antisense and
structure of postsynaptic GlyR/gephyrin clusters is controversial and knock-out experiments indicated that gephyrin is associated
possible alterations in function are largely unknown. In this study, we with glycine receptors (GlyRs) and GABAA receptors
have examined the physiological and morphological properties of (GABAARs) in postsynaptic clusters (Essrich et al. 1999; Feng
GlyR/gephyrin clusters after colchicine treatment in cultured spinal et al. 1998; Kirsch et al. 1993; Kneussel et al. 1999). While it
neurons during development. In immature neurons (5–7 DIV), dis- is known that gephyrin strongly binds GlyRs through a peptide
ruption of microtubules resulted in a 33 ⫾ 4% decrease in the peak domain within the M3-M4 cytoplasmic loop of the GlyR ␤
amplitude and a 72 ⫾ 15% reduction in the frequency of spontaneous subunit (Meyer et al. 1995), the mechanism of interaction with
glycinergic miniature postsynaptic currents (mIPSCs) recorded in
whole cell mode. However, similar colchicine treatments resulted in
GABAARs is not understood. In addition, in vitro binding
smaller effects on 10 –12 DIV neurons and no effect on mature studies suggested possible interactions of gephyrin with the
neurons (15–17 DIV). The decrease in glycinergic mIPSC amplitude cytoskeleton. Gephyrin binds tubulin and microtubules directly
and frequency reflects postsynaptic actions of colchicine, since with high affinity and significant cooperativity (Kirsch et al.
postsynaptic stabilization of microtubules with GTP prevented both 1991) and can interact with microfilaments through intermedi-
actions and similar reductions in mIPSC frequency were obtained by ate actin-binding proteins such as profilin (Mammoto et al.
modifying the Cl⫺ driving force to obtain parallel reductions in 1998) or regulators of actin filament dynamics like collybistin
mIPSC amplitude. Confocal microscopy revealed that colchicine re- (Kins et al. 2000). Hence, a commonly proposed model for the
duced the average length and immunofluorescence intensity of syn- structure of glycinergic synapses hypothesizes that gephyrin
aptic gephyrin/GlyR clusters in immature (approximately 30%) and anchors GlyRs to subsynaptic cytoskeleton (Kirsch 1999; Kneus-
intermediate (approximately 15%) neurons, but not in mature clusters.
Thus the structural and functional changes of postsynaptic gephyrin/
sel and Betz 2000). Tubulin depolymerization would then be
GlyR clusters after colchicine treatment were tightly correlated. Fi- expected to disrupt synaptic GlyR clusters.
nally, RT-PCR, kinetic analysis and picrotoxin blockade of glyciner- Morphological studies performed in spinal cord cultures
gic mIPSCs indicated a reorganization of the postsynaptic region from showed that microtubule depolymerization affected the num-
containing both ␣2␤ and ␣1␤ GlyRs in immature neurons to only ␣1␤ ber, structure, and intensity of gephyrin/GlyR postsynaptic
GlyRs in mature neurons. Microtubule disruption preferentially clusters (Kirsch and Betz 1995). However, microtubule disrup-
affected postsynaptic sites containing ␣2␤-containing synaptic tion in hippocampal cultures failed to alter gephyrin/GABAA
receptors. clusters (Allison et al. 2000) and led to the conclusion that
neither microtubules nor cytoplasmic tubulin-dimers play a
significant role in the organization of gephyrin-containing in-
INTRODUCTION
hibitory synapses. The reasons for these conflicting results are
Postsynaptic densities (PSD) harbor neurotransmitter recep- unknown.
tors that are anchored in the membrane at precise locations Using the whole cell patch-clamp technique, we previously
opposite to presynaptic active zones releasing neurotransmit- reported that the function of synaptic GlyRs, but not extrasyn-
ters. Cytoskeletal components are believed to stabilize these aptic receptors, was changed in cultured spinal cord neurons
molecular complexes. As such, both tubulin and actin are dialyzed with the microtubule depolymerization agent colchi-
major components of the PSD (Cotman et al. 1974; Kelly and cine (van Zundert et al. 2002). In contrast, the function of

Address for reprint requests and other correspondence: L. G. Aguayo, Dept. The costs of publication of this article were defrayed in part by the payment
of Physiology, University of Concepción, P.O. Box 160-C, Concepción, Chile of page charges. The article must therefore be hereby marked ‘‘advertisement’’
(E-mail: laguayo@udec.cl). in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

1036 0022-3077/04 $5.00 Copyright © 2004 The American Physiological Society www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
COLCHICINE AFFECTS SYNAPTIC GLYRS IN IMMATURE NEURONS 1037

synaptic ␣-amino-3-hydroxy-5-methyl-4-isoxazole propionic Immunocytochemistry


acid receptors (AMPARs) and gamma amino butyric acid type
Neurons were fixed for 5 (GlyR) or 15 min (GABAAR␥2) with 4%
A receptors (GABAARs) was unchanged. In this study, we paraformaldehyde in 0.1 M phosphate buffer (pH 7.4). Fixing time
have further analyzed the developmental profile of colchicine was increased to 30 min for other antibodies. Cultures were perme-
action on the function and structure of synaptic GlyR clusters. abilized with a pretreatment of 0.25% Triton X-100 (15 min) to allow
We found a strong inhibitory effect of colchicine on immature better access to intracellular epitopes and then blocked with 10%
glycinergic synapses, which contain the neonatal ␣2␤ receptor. normal horse serum (NHS) for 30 min to reduce nonspecific staining.
After synaptic maturation, however, mainly the adult ␣1␤ Cells were incubated for 1 h with pair combinations of the following
GlyR remained in the postsynaptic membrane and colchicine primary antibodies: mouse monoclonal antibody (MmAb) against
was not effective. GABAARs postsynaptic clusters were unal- gephyrin (mAb7a; 1:100; Boehringer-Mannheim, Indianapolis, IN),
tered at any developmental stage. Preliminary results were rabbit polyclonal antibody (RpAb) against the ␣1 and ␣2 subunit of
presented in abstract form (van Zundert et al. 2001). the GlyR (1:10; Calbiochem, La Jolla, CA), guinea pig polyclonal
antibody (GpAb) against the ␥2 subunit of the GABAAR (1:1,000;
kindly provided by Dr. J. M. Fritschy, University of Zurich), MmAb
METHODS against synaptophysin (1:100; Oncogene, San Diego, CA), RpAb
against synaptophysin (1:500; Zymed Laboratories Inc., San Fran-
Cell culture cisco, CA), MmAb against SV2 (1:200; kindly provided by K. M.
Mouse (C57BL/J6) spinal cord neurons obtained from five to six Buckley, Harvard Medical School, Boston, MA), and RpAb against
embryos (13–14 days) were plated at 300,000 cells/ml into 35-mm synapsin I (1:1,000; Calbiochem). We tested the following combina-
tissue culture dishes coated with poly-L-lysine (MW ⬎ 350 kDa, tions of antibodies: gephyrin/synapsin I, synaptophysin/SV2, synap-
Sigma Chemical, St. Louis, MO). The neuronal feeding medium tophysin/synapsin I, GlyR/gephyrin, GABAAR␥2/gephyrin, GlyR/
consisted of 90% minimal essential medium (GIBCO, Grand Island, SV2, GABAAR␥2/SV2, and GABAAR␥2/gephyrin/synapsin I. Im-
NY), 5% heat-inactivated horse serum (GIBCO), 5% fetal bovine munoreactive (IR) sites were visualized after incubation for 1 h with
serum (GIBCO), and a mixture of nutrient supplements (Aguayo and appropriate secondary antibodies raised in donkey and conjugated to
Pancetti 1994). Fresh media was replaced every 3 days. Experiments FITC or Cy3 (1:50; Jackson ImmunoResearch Laboratories, West
were performed on 5–17 DIV neurons. The microtubule interacting Grove, PA). With every combination, we used a MmAb mixed with
molecules colchicine, ␥-lumicolchicine, nocodazole, and taxol (20 a RpAb or GpAb. Presynaptic markers were usually labeled with
␮M) were added to the culture media for 3 h or dialyzed into the FITC-conjugated secondary antibodies and postsynaptic gephyrin and
neurons via the recording patch pipette as previously reported (Kirsch GABAAR␥2 antibodies with Cy3-conjugated antibodies. For the tri-
and Betz 1995; Rosenmund and Westbrook 1993; van Zundert et al. ple immunolocalization, together with the normal GABAAR␥2/
2002). gephyrin double immunolabeling, synapsin I was labeled with a
biotin-SI-conjugated RpAb (1:50) and stained with Cy5-conjugated
streptavidin (1:250; Jackson ImmunoResearch Laboratories). The
Electrophysiology cells were coverslipped using Vectashield (Vector Laboratories, Bur-
For voltage- and current-clamp recordings in the whole cell con- lingame, CA).
figuration, patch electrodes were filled with (in mM) 140 KCl, 10
BAPTA, 10 HEPES (pH 7.4), 4 MgCl2, and 2 ATP-Na2. The external Image visualization and sampling
solution contained (in mM) 150 NaCl, 10 KCl, 2.0 CaCl2, 1.0 MgCl2,
10 HEPES (pH 7.4), and 10 glucose. After the whole cell configura- For quantitative analysis of immunocytochemical data, samples of
tion was established, the capacitance of the cell and the series resis- spinal neurons (10 –15 cells from 2 separate experiments of paired
tance were compensated using the amplifier (⬎80%). Spontaneous control and colchicine-treated cells) were chosen randomly for imag-
glycinergic miniature postsynaptic currents (mIPSCs) were isolated ing using confocal microscopy (Olympus Fluoview; 100⫻ oil immer-
by the addition of CNQX (2 ␮M, RBI), bicuculline (2 ␮M, RBI), TTX sion objective, N.A. 1.35, digitally zoomed 3 times, pixel size ⫽ 0.07
(0.1 ␮M, Sigma), and MgCl2 (2 mM) to the external solution (van ␮m). Stacks of optical sections separated by 0.5 ␮m in the z axis were
Zundert et al. 2002). They were confirmed as glycinergic mIPSCs by acquired throughout whole cells. Dual color immunofluorescent im-
a complete blockade with 750 nM strychnine. At 1 and 25 min after ages were captured in simultaneous two-channel mode. Antibody
establishing the whole cell configuration, the spontaneous events dilutions were chosen for adequate immunofluorescent intensity to
occurring during a 3-min interval were analyzed. The properties of the minimize cross-talk between the channels. If even minimal cross-talk
mean mIPSC amplitude and frequency at 1 min were used as an between FITC and Cy3 fluorescence was observed (FITC usually
internal control (van Zundert et al. 2002). By presenting the data as labeled the presynaptic marker and resulted in brighter fluorescence),
normalized (25 min/1 min), the results obtained with different drug this was abolished by collecting Cy3 fluorescence above 610 nm. The
treatments were pooled and analyzed in terms of their statistical number of synaptic gephyrin and GlyR clusters was determined as the
differences. The noise was not significantly changed when neurons number of clusters apposed/co-localized to punctate synapsin I or SV2
were dialyzed with any of the cytoskeletal disrupters. Synaptic cur- immunoreactivity. Co-localization was studied by superimposing both
rents were recorded using Axotape 7.0 software and analyzed off-line color channels. For triple immunolocalizations, neurons were imaged
(Axon Instruments, Union City, CA). Every identified synaptic event using a Leica TCS confocal microscope using multitracking imaging
collected during the test intervals displaying an amplitude above the of each channel independently and optical and confocal conditions
background noise (12–15 pA) was analyzed using MiniAnalysis 5.0 similar to those previously used and described above.
software (Synaptosoft). Data were excluded if the uncompensated
series resistance (⬍8 M⍀) increased by ⬎15% during the recordings. Analysis of cluster size
Unless otherwise noted, all reagents were purchased from Sigma.
Analysis of mIPSC properties included peak amplitude, inter-event In contrast to previous confocal quantitative studies on gephyrin
interval, rise-time, and decay-time constant. The decay-phase of cluster sizes done in tissue sections (Geiman et al. 2000; Lim et al.
mIPSCs was best fitted with a single exponential curve. To avoid 1999; Oleskevich et al. 1999; Triller et al. 1990), the majority of
analysis of data points with excess noise at the peak and baseline, the synapses in our cultures were positioned on the lateral sides of the
rise and decay-phase were fitted between 10 and 90% of its amplitude. cells, and relatively few crossed over the top of neurons. Therefore

J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org


Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
1038 VAN ZUNDERT ET AL.

there were only few examples of gephyrin clusters viewed “en face” anti-sense: 5⬘-ccatccagatgtcaattgcctt-3⬘); ␤ (sense: 5⬘-ggaattcgggatg-
where surface areas could be resolved accurately. Hence, we mea- gagacgtcc-3⬘; anti-sense: 5⬘-gctctcaagttgcattttgc-3⬘). The sequence of
sured cluster lengths. Frequently, individual clusters were imaged in all the primers used as the amplified fragments (sequence, restriction
more than one serial optical section. The optical section that contained maps) has been previously described (Kirchhoff et al. 1996). Negative
the largest and brightest immunofluorescence for each individual (omitting RNA, template sense primers) and positive (adult and em-
cluster was selected for measurement. For cluster size measurements, bryonic spinal cord cDNAs) controls were routinely included. At least
the stacks of confocal optical sections were analyzed using ImagePro three independent experiments per condition were considered.
software (ver. 4.1; Media Cybernetics, Silver Spring, MD). Individual
clusters were detected in single optical sections by thresholding the Data analysis
image from the maximum arbitrary gray level (AGL) pixel intensity
(4095) to 25% of this value. These image segmentation parameters Data are expressed as means ⫾ SE. Statistical comparisons were
allowed us to detect even the fainter clusters while outlining intensely performed using Student’s t-test or ANOVA. Cumulative probability
fluorescent clusters inside their diffraction halo. Using this method, distributions of mIPSC amplitudes were compared with Kolmogorov-
individual clusters were automatically segmented and outlined, and Smirnov test. P ⬍ 0.05 was considered significant.
their maximum length was measured. For each experiment, imaging
conditions were kept constant, and no postcapture modifications were
RESULTS
done in images used for quantitative analysis. The data were compiled
in Microsoft Excel, analyzed in Statview, and plotted using Origin. Colchicine decreases spontaneous glycinergic mIPSCs in a
Importantly, our measurements of gephyrin cluster size gave similar developmentally regulated manner
results to previous studies in spinal cord cultures (Meier et al. 2000;
Rosenberg et al. 2001) and spinal cord sections visualized with either In agreement with a previous study (van Zundert et al.
confocal microscopy (Oleskevich et al. 1999), conventional fluores- 2002), intracellular dialysis with colchicine decreased glycin-
cence microscopy, or three-dimensional reconstruction with electron ergic transmission in young (⬍12 DIV) cultured spinal neu-
microscopy (Alvarez et al. 1997). For illustration, the neuron was rons. The amplitude of control glycinergic mIPSCs, isolated in
reconstructed from the stack of optical sections, and further figure
the presence of CNQX, bicuculline, and TTX, was rather stable
composition and labeling was done in CorelDraw 3.0, CorelDraw 8.0,
or SigmaPlot 4.0. over a recording period of 25 min (Fig. 1A). In contrast,
neurons dialyzed with 20 ␮M colchicine in the patch pipette
displayed a large change in the amplitude histogram distribu-
RT-PCR: Total RNA extraction and cDNA synthesis tion after 25 min of recording (Fig. 1B, left). This decrease in
To extract total RNA from the spinal cord cultures, cells were lysed glycinergic activity resulted in a shift toward smaller current
with 1 ml Trizol (5 min, 20°C, GIBCO), and 0.2 ml of chloroform was amplitudes as shown in the cumulative probability amplitude
added to each sample. After vigorous shaking, the mixture was distributions (P ⬍ 0.001, Fig. 1B, right). Colchicine also
centrifuged (12,000g, 15 min, 4°C), and RNA was precipitated by significantly reduced mean glycinergic mIPSC amplitude and
mixing the aqueous phase with isopropyl alcohol (0.5 ml, 10 min, frequency by 29 ⫾ 4% (P ⬍ 0.001) and 69 ⫾ 8% (P ⬍ 0.001),
20°C) and centrifuged at 12,000g (15 min). Finally, the RNA was respectively.
washed two times with ethyl alcohol (75%) and spun down at 7,500g,
dried under the hood (5 min, 20°C), and dissolved in sterile RNAse
We investigated whether the colchicine-induced alterations
free H2O. The integrity and purity of each RNA sample was evaluated of glycinergic mIPSCs was dependent on developmental stage.
by agarose gel electrophoresis (1–2%). The RNA concentration was To facilitate analysis, the spinal neurons were grouped accord-
determined by: [RNA] ⫽ 40 ␮g/ml ⫻ A260 ⫻ Fd, where A260 is the ing to their stage of development as immature (5–7 DIV),
absorbance (260 nm) of the sample and Fd represents the dilution intermediate (10 –12 DIV), and mature (15–17 DIV) neurons.
factor. cDNA synthesis employed 12 ␮l of total RNA, 1 ␮l of random It was found that the mean amplitude of glycinergic mIPSCs
hexamer primers (50 ng, Promega), 1 ␮l of 10 mM dNTPs, 4 ␮l of 5⫻ increased from 33 ⫾ 4 pA in immature neurons to 56 ⫾ 8 pA
reaction buffer, 2 ␮l of DTT (0.1 M), 1 ␮l of RNAsin (40 U/␮l), and in mature neurons (Fig. 1C1). At the same time, the frequency
1 ␮l of RT-Superscript II (200 U). The RT reactions (42°C) were increased from 0.84 ⫾ 0.22 Hz in immature neurons to 1.64 ⫾
stopped after 1.5 h by cooling to ⫺20°C. cDNA was determined in 0.56 Hz in mature neurons (Fig. 1D1). Interestingly, we found
each case: [cDNA] ⫽ 33 ␮g /ml ⫻ A260 ⫻ Fd.
that colchicine dialysis reduced both the amplitude (37 ⫾ 5%
reduction, P ⬍ 0.001) and frequency (72 ⫾ 15% reduction,
cDNA amplification and temporal subunit expression P ⬍ 0.01) in immature neurons (n ⫽ 5), while having no effect
PCR mixtures (25 ␮l final) contained 2.5 ␮l 10⫻ Taq polymerase in mature neurons (n ⫽ 8; Fig. 1, C2 and D2). Smaller
buffer, 0.5 ␮l dNTPs mix (10 mM), 0.75 ␮l MgCl2 (50 mM), 0.75 ␮l colchicine effects on both the amplitude (18 ⫾ 3% reduction;
sense (10 ␮M) and 0.75 ␮l antisense primers (10 ␮M), 3 ␮l cDNA P ⬍ 0.01) and the frequency (61 ⫾ 7% reduction; P ⬍ 0.01)
template (50 ng), 16.5 ␮l H2O, and 0.25 ␮l Taq polymerase (5 U/␮l). were found in neurons of intermediate stage of development
A Biorad thermocycler was used with PCR conditions set to 94°C (3 (n ⫽ 7).
min) for primary denaturation, 30 cycles of amplification (94°C de-
naturation, 30 s; 45–51°C annealing, 30 s; 72°C elongation, 30 s), and
10 min at 72°C for final elongation. The predicted sizes of ␣1, ␣2, and Colchicine-induced reductions of glycinergic mIPSC activity
␣3 amplicons were 529, 384, and 470 bp, respectively, and the ␤ are due to postsynaptic microtubule depolymerization
amplicon was 183 bp. The PCR amplified products were resolved by
agarose gel electrophoresis (1–2%), photographed (Polaroid films),
A summary of the effects produced by different agents
and digitized (Cannon device) using Adobe Photoshop (4.0). The acting on microtubules on the normalized mean mIPSC am-
following primers were used to amplify mouse GlyR subunit cDNA: plitude is shown in Fig. 2A. The amplitude of glycinergic
␣1 (sense: 5⬘-aggcccaacttcaaaggtcc-3⬘; anti-sense: 5⬘-ctgtgttgtagtgct- mIPSCs remained highly stable (1.07 ⫾ 0.07 presented as the
tggtgc-3⬘); ␣2 (sense: 5⬘-ggtacaccatgaatgacctg-3⬘; anti-sense: 5⬘- amplitude ratio 25 min/1 min, from 34.7 ⫾ 7 pA at 1 min)
ccatccagatgtcaattgcctt-3⬘); ␣3 (sense: 5⬘-gctgacattaacactctcttgtcc-3⬘; when the neurons were dialyzed with normal internal solution
J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
COLCHICINE AFFECTS SYNAPTIC GLYRS IN IMMATURE NEURONS 1039

FIG. 1. Glycinergic miniature postsynaptic currents (mIPSCs)


become insensitive to colchicine with maturation of spinal
neurons. A: no differences were found in synaptic glycinergic
mIPSCs after 25 min of dialysis with normal (control) intracel-
lular solution. Amplitude histograms, obtained from 5–12 DIV
neurons (n ⫽ 7), at 1 and 25 min, dialyzed with normal
intracellular solution, are shown at left (bin size, 1 pA). Bath
solution contained CNQX (2 ␮M), bicuculline (2 ␮M), MgCl2
(2 mM), and TTX (100 nM). Examples of current traces at these
2 time points are shown above the corresponding histograms.
Cumulative amplitude distributions (n ⫽ 7 neurons) are shown
at right. B: current traces and amplitude histograms generated
during intracellular application of 20 ␮M colchicine (n ⫽ 10)
showed significant differences. After 25 min colchicine dialy-
sis, glycinergic mIPSC activity is strongly reduced, and the
cumulative probability distribution (right) is shifted toward
smaller amplitudes (P ⬍ 0.001). C and D: mean peak-amplitude
(C1) and frequency (D1) of glycinergic mIPSCs increased
during maturation. Inhibitory action of colchicine on both am-
plitude (C2) and frequency (D2) decreased with development.
Each symbol represents means ⫾ SE obtained from 5–10
neurons. **P ⬍ 0.01, ***P ⬍ 0.001. Asterisks indicate a
statistically significant action of colchicine compared with con-
trol.

for 25 min (n ⫽ 7). On the other hand, dialysis of sister neurons units with IC50s of approximately 64 and 324 ␮M, respectively
with the microtubule disrupter colchicine (20 ␮M, n ⫽ 10) or (Machu 1998). We obtained four lines of experimental evi-
nocodazole (20 ␮M, n ⫽ 5) reduced the normalized mIPSC dence indicating that our results with 20 ␮M intracellular
amplitude to 0.71 ⫾ 0.04 (P ⬍ 0.001) and 0.88 ⫾ 0.01 (P ⬍ colchicine cannot be explained by this competitive inhibition.
0.05), respectively. When dialyzed with the inactive analog of First, analysis with extracellular colchicine (1–1,000 ␮M)
colchicine, ␥-lumicolchicine (20 ␮M, n ⫽ 5), the amplitude showed that the whole cell glycine-activated current (30 ␮M)
was not significantly changed (0.93 ⫾ 0.09, P ⬎ 0.05). Taxol was inhibited with very similar affinities in immature and
(20 ␮M, n ⫽ 4), a cytotoxin that increases the stabilization of mature neurons (IC50 of 143 ⫾ 28 and 187 ⫾ 13 ␮M, respec-
tubulin dimers (Nogales 2000), was unable to change the tively; P ⬎ 0.05, Fig. 2C). Second, internal nocodazole, which
amplitude of the glycinergic transmission (0.98 ⫾ 0.04, P ⬎ was reported to be devoid of antagonistic properties on the
0.05). Dialysis with taxol for even 60 min (data not shown) was GlyR when applied extracellularly (Machu 1998), was still
unable to affect the glycinergic transmission, and extracellular able to reduce the mIPSC amplitude (Fig. 2A). Third, the
application of this alkaloid (20 ␮M; 3 h to the media) did not application of 20 ␮M external colchicine for 20 min did not
alter the number and morphology of gephyrin clusters (data not affect the average mIPSC amplitude (0.93 ⫾ 0.03, n ⫽ 8, P ⬎
shown; Kirsch and Betz 1995), suggesting that stabilization of 0.05, Fig. 2A) or the cumulative probability amplitude distri-
actin filaments and actin dynamics did not affect the integrity butions (P ⬎ 0.05, Fig. 2B). Finally, intracellular GTP (500
of postsynaptic gephyrin/GlyRs clusters. ␮M), a guanine nucleotide with low membrane permeability
It was previously shown in Xenopus oocytes that extracel- and known for its capacity to stabilize microtubules (Nogales
lular colchicine inhibited overexpressed ␣2 and ␣1 GlyR sub- 2000), blocked the colchicine-induced reduction of glycinergic
J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
1040 VAN ZUNDERT ET AL.

FIG. 2. Colchicine inhibits mIPSC by affecting


postsynaptic microtubules. A: effects of different agents
that are able to interact with microtubules on the normal-
ized mean glycinergic mIPSC amplitude. Note that only
intracellular colchicine and nocodazole reduced the
mIPSC amplitude, while external colchicine (also 20 ␮M
for 20 min) had no effect. To avoid variability due to
different levels of spontaneous activity in different neu-
rons that could mask changes induced by the drugs, all
data at 20 –25 min after intracellular dialysis of drugs or
control recording solution were normalized with respect to
activity measured in the same cell at 1 min (control ⫽
34.7 ⫾ 7 pA). Symbols represent means ⫾ SE, *P ⬍ 0.05,
***P ⬍ 0.001. B: cumulative probability distribution is
similar after 25 min of 20 ␮M external colchicine appli-
cation. C: effect of increasing concentrations of colchicine
in immature (E) and mature (F) neurons on 30 ␮M glycine.
IC50 values were 143 ⫾ 28 and 187 ⫾ 13 ␮M in immature
and mature neurons, respectively. D and E: absence of
autapses in cultured neurons. D: current response with a
2-ms test pulse (from ⫺80 to 0 mV). This pulse activated
a large “unclamped” Na⫹ current, which was followed by
a 2nd downward reflection corresponding to a combina-
tion of capacitative and Na⫹ tail currents. Lack of effect
with strychnine (100 nM) indicates the absence of a gly-
cinergic autaptic current (dashed line). Repetitive stimu-
lation (5 Hz) was unable to increase the IPSC frequency
(inset). E: voltage responses from the same neuron stim-
ulated with a 0.6-nA depolarizing current pulse (5 ms).
Dashed line shows that strychnine did not change the
shape of the voltage response. Inset: the neuron exhibited
spontaneous glycinergic synaptic activity.

mIPSCs amplitude (1.0 ⫾ 0.08, n ⫽ 4, Fig. 2A). Control frequency can be interpreted as a reduction in the probability of
experiments showed that GTP had no effect when added alone neurotransmitter release (Walmsley et al. 1998). Intracellular
to the internal solution (0.98 ⫾ 0.04, n ⫽ 4, Fig. 2A). dialysis of membrane permeable colchicine could therefore
have altered presynaptic cytoskeleton structures associated
Effect of colchicine cannot be explained by a presynaptic with neurotransmitter release after diffusion into the presynap-
site of action tic terminal. However, our results with extracellular colchicine
and GTP do not support this idea.
We also analyzed the number of mIPSC events and found Alternatively, functional autaptic synapses could form under
that the frequency was decreased when the neurons were our culture conditions (Bekkers and Stevens 1991), and col-
dialyzed with either colchicine (0.31 ⫾ 0.08 presented as the chicine action could be partly due to alterations in autaptic
frequency ratio 25 min/1 min, P ⬍ 0.001) or nocodazole transmission. To investigate this possibility, we examined
(0.49 ⫾ 0.11, P ⬍ 0.05). On the other hand, normal internal whether the neurons used in our studies showed evidence of
solution (0.84 ⫾ 0.15, P ⬎ 0.05) or addition of ␥-lumicolchi- glycinergic autaptic transmission. Under the low Cl⫺ gradient
cine (0.85 ⫾ 0.17, P ⬎ 0.05) to the patch pipette was unable used, GlyR activation should be associated with a negative sign
to significantly alter mIPSC frequency. A decrease in mIPSC current. Figure 2D illustrates the effect of applying an 80-mV
J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
COLCHICINE AFFECTS SYNAPTIC GLYRS IN IMMATURE NEURONS 1041

FIG. 3. Colchicine altered the properties of


gephyrin clusters in immature neurons. A1–D1:
confocal images of synapsin I-IR terminals (green)
in the soma and processes of spinal interneurons.
A2–D2: gephyrin clusters (red) in the same cells
from A1–D1. A3–D3: superimposed images of
gephyrin and synapsin I immunofluorescence. Syn-
aptic gephyrin clusters are found apposed to or
co-localized (yellow) with synapsin I-IR. A: im-
ages obtained from an immature control neuron.
Note that only approximately 60% of the gephyrin
clusters are apposed/co-localized with synapsin I,
suggesting that many nonsynaptic gephyrin clus-
ters are present in immature neurons. B: application
of colchicine (20 ␮M for 3 h) caused a reduction in
both size and intensity of synaptic gephyrin clus-
ters in immature neurons (B3). C: mature neurons
show more synaptic terminals and gephyrin clus-
ters compared with immature neurons. Size of the
gephyrin clusters is also increased, and the major-
ity (⬎95%) are localized at synapses. D: colchicine
does not affect morphology of gephyrin clusters in
mature neurons. Images were reconstructed from a
stack of optical sections. Insets: magnified views of
boxed areas showing examples of synaptic and
nonsynaptic gephyrin clusters. Scale bar: 10 ␮m.

depolarizing voltage step into a neuron held at ⫺80 mV. This Fig. 2A). As found with colchicine, the decrease in mIPSC
depolarizing pulse was able to activate a large somatic “un- quantal size was accompanied by a large (54 ⫾ 5%) reduction
clamped” Na⫹ inward current (approximately 2 nA) but no in frequency, supporting the idea that the reduction in fre-
autaptic glycinergic postsynaptic current was detected (Fig. quency after 25 min of colchicine treatment is best explained
2D). In addition, the frequency of spontaneous glycinergic by an alteration in current amplitude.
mIPSCs in the same neuron was unchanged even after a
sustained depolarization elicited by a high-frequency (5 Hz) Colchicine treatment decreased the size and intensity of
train of depolarizing pulses (Fig. 2D, inset). Additional exper- gephyrin/GlyR clusters in immature spinal neurons but not
iments using current-clamp recordings showed that soma-elic- in mature neurons
ited action potentials were not able to induce strychnine-sen-
sitive autaptic synaptic potentials (Fig. 2E). From the 11 neu- We next investigated whether the developmentally regulated
rons examined, only 1 displayed an autaptic response, action of colchicine might be correlated with morphological
suggesting that it is unlikely that the colchicine effect is pro- changes in postsynaptic gephyrin clusters on microtubule dis-
duced by alterations in autaptic transmission. ruption (Kirsch and Betz 1995). Immature, intermediate, and
The above experiments ruled out the possibility that a pre- mature neurons were treated for 3 h with 20 ␮M colchicine
synaptic mechanism was associated to the action of colchicine (Kirsch and Betz 1995; van Zundert et al. 2002). After fixation,
on the glycinergic mIPSC frequency. Therefore we can argue gephyrin clusters were immunolabeled with monoclonal anti-
that colchicine decreased the frequency of mIPSCs by reducing body 7a, and the size (maximal length), density (luminosity),
the current amplitude in such a way that a number of synaptic and numbers of synaptic and extrasynaptic gephyrin clusters
events fall under the threshold of detection (approximately 12 were analyzed. Synaptic gephyrin clusters were detected by
pA). To determine whether a change in mIPSC peak amplitude combining gephyrin-IR with a presynaptic marker (see METH-
might account for the decrease in frequency, the mIPSC am- ODS). We considered gephyrin clusters synaptic when they
plitude was reduced 27 ⫾ 2% (37 ⫾ 4 to 27 ⫾ 2 pA, n ⫽ 5) were opposite to immunolabeled boutons and extrasynaptic if
by lowering the Cl⫺ driving force. This closely simulates the they were not. We compared synapsin I to other presynaptic
colchicine-induced reduction in mIPSC amplitude (shown in markers like synaptophysin or SV2 and found that these three
J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
1042 VAN ZUNDERT ET AL.

presynaptic markers labeled a similar population of boutons. number of extrasynaptic clusters was reduced to 4.4 ⫾ 1.1 in
Synaptophysin and synapsin I showed 78 ⫾ 9% co-localization immature neurons (P ⬍ 0.05).
(n ⫽ 4 neurons), whereas synaptophysin and SV2 were co- Interestingly, it was found that the size and immunofluores-
localized in ⬎90% of the boutons (n ⫽ 4 neurons). We found cent intensity of synaptic gephyrin clusters were altered after
that synapsin I was the brighter marker and labeled a few more treating immature neurons with colchicine (Figs. 3B and 4, A
varicosities and thus was chosen for further studies. Analysis and B). Thus colchicine treatment resulted in a significant
of synapsin I-IR showed that colchicine treatment was unable decrease (77 ⫾ 5% of control) in the average length of gephy-
to significantly alter the bouton size in both immature (0.93 ⫾ rin clusters in immature neurons from 0.39 ⫾ 0.02 ␮m in
0.04 ␮m in control vs. 0.89 ⫾ 0.05 ␮m in treated) and mature control to 0.30 ⫾ 0.02 ␮m after treatment (P ⬍ 0.01, Fig. 4A).
neurons (0.80 ⫾ 0.03 ␮m in control vs. 0.77 ⫾ 0.04 ␮m in In addition, analysis of cumulative probability cluster size
treated). distribution revealed an apparent shift toward the left (P ⬍
Immature control neurons showed gephyrin-IR clusters at 0.001; data not shown), indicating that large clusters were more
the periphery of the neuronal soma and along proximal neurites sensitive to the colchicine treatment than smaller clusters. On
(Fig. 3A). Not all gephyrin clusters were apposed or co-local- colchicine application, the immunofluorescent intensity was
ized (yellow) with synapsin I (Fig. 3A3), indicating the exis- also reduced from 916 ⫾ 37 to 746 ⫾ 21 arbitrary gray level
tence of extrasynaptic gephyrin clusters at this developmental units (Fig. 4B, P ⬍ 0.001). At intermediate developmental
stage. Synapsin I immunoreactivity was very robust in all stages, a smaller but significant decrease in gephyrin cluster
cultures; however, we cannot completely rule out the possibil- size was detected after colchicine treatment (85 ⫾ 5% of
ity that some newly formed synaptic varicosities might contain control; P ⬍ 0.05; from 0.39 ⫾ 0.01 ␮m in control to 0.33 ⫾
low synapsin I antigenicity, below our detection threshold. 0.02 ␮m after treatment), but the immunofluorescent intensity
Nevertheless, previous ultrastructural analysis has demon- was unchanged (P ⬎ 0.05).
strated the existence of extrasynaptic gephyrin clusters at early During in vitro development, more synaptic interactions
stages of development in spinal cord cultures (Colin et al. among the neurons are formed as indicated by an increased
1996). In addition, the proportion of extrasynaptic to synaptic number of synapsin I contacts in mature neurons (Fig. 3C1).
gephyrin clusters are well matched to another study of cultured Correspondingly, mature neurons expressed approximately
spinal cord neurons that used a different presynaptic marker three times more gephyrin clusters, and these were always
(Dumoulin et al. 2000). During development, the number of juxtaposed to synapsin I (Fig. 3C3). Synaptic gephyrin clusters
synaptic gephyrin clusters per cell increased from 14.8 ⫾ 1.6 grew bigger in size to a mean length of 0.48 ⫾ 0.01 ␮m (Fig.
in immature neurons (n ⫽ 14) to 29.3 ⫾ 3.5 in intermediate 3C2). Extrasynaptic clusters were of similar size (0.33 ⫾ 0.01
(n ⫽ 14) and 52.1 ⫾ 4.7 in mature neurons (n ⫽ 14). The ␮m) compared with the previous developmental stages (0.31 ⫾
number of extrasynaptic gephyrin clusters per cell decreased 0.02 ␮m in both immature and intermediate neurons). Consis-
from 9.6 ⫾ 2.2 in immature neurons to 4.9 ⫾ 0.9 and 1.9 ⫾ 0.4 tent with our electrophysiological data, colchicine did not alter
in intermediate and mature neurons, respectively. Colchicine synaptic or extrasynaptic gephyrin cluster number, size, or
treatment did not change the number of synaptic gephyrin immunofluorescent intensity in mature neurons (Figs. 3D and
clusters per cell at any developmental stage; however, the 4, A and B).

FIG. 4. Size and luminosity of gephyrin clusters were af-


fected in a maturation-dependent manner by colchicine. Imma-
ture, intermediate, and mature spinal neurons were treated with
colchicine and stained with antibodies against gephyrin and
synapsin I to identify synaptic gephyrin clusters. A: colchicine
decreased the mean synaptic gephyrin cluster size of immature
and intermediate neurons. Size of gephyrin clusters in mature
neurons was not significantly affected by colchicine. Note the
increase in cluster size with time of development in vitro. B:
luminosity of synaptic gephyrin clusters expressed in arbitrary
gray levels (AGL) is decreased following colchicine treatment
in immature neurons but not in intermediate or mature neurons.
Each symbol represents means ⫾ SE obtained from 14 neurons.
Asterisks indicate a statistically significant action of colchicine
compared with control. *P ⬍ 0.05, **P ⬍ 0.01, ***P ⬍ 0.001.
C and D: mean gephyrin cluster size plotted as a function of
mean glycinergic mIPSC amplitude at the 3 developmental
stages in control (open symbols) and colchicine-treated (closed
symbols) neurons. High correlations were found with intracel-
lular (C) and extracellular colchicine treatments (D).

J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org


Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
COLCHICINE AFFECTS SYNAPTIC GLYRS IN IMMATURE NEURONS 1043

These results, as well as previous studies (Lim et al. 1999; postsynaptic cluster size and structure due to its favorable
Oleskevich et al. 1999), suggest a positive correlation between fixation tolerance and epitope display characteristics (Alvarez
gephyrin/GlyR cluster size and glycinergic mIPSC amplitude. et al. 1997; Kirsch and Betz 1995; Triller et al. 1985, 1990).
Next, we performed a correlative analysis between the average Additional experiments were performed with an antibody that
mean peak amplitude of glycinergic mIPSCs and gephyrin targets GlyR ␣1/␣2 subunits, which co-localized with 75 ⫾ 2%
cluster size in control and after colchicine application either by of gephyrin clusters (Fig. 5A). Similar to the action on gephyrin
dialysis into the neuron via the patch pipette or to the media for clusters, colchicine (3 h to the media) decreased the size of
3 h, similar to the protocol used to obtain the structural data synaptic GlyR clusters in immature neurons (71 ⫾ 6% of
(see Fig. 3). As shown in Fig. 4, C and D, both colchicine control, P ⬍ 0.01; Fig. 5, C and D), and the effect became less
treatments gave a high positive correlation which ranged from evident in intermediate neurons (90 ⫾ 7% of control, P ⬎
0.92 to 0.97. Similar results were found when the quantal 0.05). Mature neurons were not analyzed because colchicine
amplitude was correlated with the gephyrin cluster intensity had no action on either glycinergic mIPSCs or gephyrin cluster
(r ⫽ 0.88). The results support the conclusion that colchicine size/intensity.
action on glycinergic activity is produced by a modification in Previous studies have indicated that gephyrin is also asso-
postsynaptic receptor clusters. ciated to GABAARs (Essrich et al. 1999; Kneussel et al. 1999),
The data in Fig. 3 was obtained using an antibody against and we found that 48 ⫾ 4% of ␥2-containing GABAARs
gephyrin. Gephyrin immunolabeling constitutes a good method clusters co-localized with gephyrin in immature spinal neurons
for performing high-resolution morphological analysis of (Fig. 5B). Therefore it is possible that colchicine treatment

FIG. 5. Colchicine affects the morphology of


synaptic gephyrin/glycine receptor (GlyR) clusters,
but not synaptic GABAAR␥2 clusters. A: images
show that GlyR (green) and gephyrin (red) clusters
are highly (80%) co-localized (yellow, arrows).
Few isolated gephyrin and GlyR clusters (double
arrowheads) are found in these neurons. B:
GABAAR␥2 (red) and gephyrin clusters (green)
are found to co-localize at approximately 50% (yel-
low, arrows). Isolated GABAAR␥2 (double arrow-
heads) and gephyrin clusters (arrow heads) are
frequently found in these neurons. C: images of
GlyR clusters alone (green, C1) and superimposed
with SV2 (red, C2) indicate the existence of syn-
aptic GlyR clusters (yellow, arrows). Note that not
all GlyR clusters are apposed/co-localized with
SV2 (arrowheads). D: application of colchicine (20
␮M for 3 h) reduced the size of synaptic GlyR
clusters. E: similar colchicine treatment did not
alter the morphology of GABAAR␥2 clusters. F:
triple immunolocalization shows that GABAAR␥2
(red) gephyrin (green) clusters are apposed (blue)
and co-localized with synapsin I (white), suggest-
ing the synaptic localization of gephyrin/
GABAAR␥2 cluster complexes. Images are recon-
structed from a stack of optical sections. All con-
focal images were obtained from immature spinal
interneurons. Insets: magnified views of boxed ar-
eas demonstrating examples of synaptic and non-
synaptic GlyR clusters. Scale bar: 10 ␮m.

J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org


Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
1044 VAN ZUNDERT ET AL.

might also affect the properties of synaptic GABAAR clusters. Previous reports have proposed that acceleration of mIPSC
Contrary to this idea, our results showed that colchicine did not kinetics with synaptic development results from a switch be-
alter the size of synaptic GABAAR␥2 clusters in immature tween neonate ␣2␤ GlyRs (slow) to adult ␣1␤ GlyRs (fast)
(108 ⫾ 3% of control) and intermediate (106 ⫾ 3% of control) (Ali et al. 2000; Krupp et al. 1994; Singer et al. 1998; Taka-
neurons (Fig. 5E). This lack of effect is in agreement with a hashi et al. 1992). Two independent experiments support the
previous study from our laboratory that showed that colchicine idea that these changes are recapitulated in cultured mouse
was unable to alter GABAergic synaptic currents in these spinal cord neurons. First, using RT-PCR, we found that ␣1
spinal neurons (van Zundert et al. 2002). and ␤-subunit mRNA levels were high at all three develop-
mental stages, while the ␣2-subunit expression was strongly
Immature cultured spinal cord neurons express ␣2␤ and down-regulated in mature neurons (Fig. 7B). No mRNA for the
␣1␤ GlyRs in their synapses ␣3-subunit was detected during any of the three developmental
stages. Similar expression patterns were previously described
During the course of our study, we noted that colchicine during the development of cultured rat spinal neurons
treatment not only caused a reduction in the amplitude and (Bechade et al. 1996) and brain stem motoneurons (Singer et
frequency of glycinergic mIPSCs, but it also accelerated its al. 1998). Second, we found that 100 ␮M picrotoxin, a toxin
kinetics in immature neurons (Fig. 6A). Interestingly, the de- known to differentially affect ␣2␤ (IC50 ⫽ 0.3 mM) and ␣1␤
cay-phase of mIPSCs became approximately three times faster (IC50 ⫽ 1 mM) GlyRs (Pribilla et al. 1992), significantly
with neuronal maturation, and colchicine was no longer able to decreased the amplitude and decay-time of glycinergic mIPSCs
accelerate these rapid currents (Fig. 6, B and C). For instance, in immature neurons (Fig. 7C), but not in mature neurons (Fig.
while glycinergic mIPSCs in control immature neurons dis- 7D). Supporting the idea that the ␤-subunit is required to
played a mean decay-time constant of 13.1 ⫾ 1.8 ms (n ⫽ 6 cluster the GlyR in the postsynaptic membrane (Meyer et al.
cells), this value was 8.3 ⫾ 1.1 ms in colchicine-treated neu- 1995), we found that continuous application of 10 ␮M picro-
rons (n ⫽ 5 cells, P ⬍ 0.05). In intermediate (7.9 ⫾ 1.4 ms, toxin, a concentration known to selectively inhibit ␣ homo-
n ⫽ 5) and mature neurons (4.3 ⫾ 0.4 ms, n ⫽ 5 cells), the meric (IC50 ⫽ 7 ␮M; Pribilla et al. 1992), was unable to affect
glycinergic currents displayed a faster decay-time constant and glycinergic mIPSCs (data not shown). Taken together, these
colchicine did not significantly modify these values (Fig. 6, B data suggest the presence of both ␣2␤ and ␣1␤ GlyRs in the
and C). The mean rise-time of mIPSCs, on the other hand, postsynaptic membrane of immature neurons and predomi-
slightly decreased during maturation (15% from 2.6 ⫾ 0.3, nantly ␣1␤ receptor in mature synapses.
P ⬎ 0.05) and was unaffected by colchicine. The properties of
developing mIPSCs in spinal neurons are in agreement with
those reported in developing brain stem motoneurons (Singer Colchicine and picrotoxin target the same population of
et al. 1998). postsynaptic GlyRs
The previous results suggest that colchicine and 100 ␮M
picrotoxin could affect the same population of slow immature
␣2␤ GlyRs. Therefore, in the next series of experiments, we
analyzed whether glycinergic activity of colchicine treated
neurons (3 h in the media; 20 ␮M) displayed a different
sensitivity to picrotoxin. We reasoned that if colchicine pro-
voked the reduction of ␣2␤ GlyRs, this would reduce picro-
toxin sensitivity of the synaptic currents. The open circles in
the graph of Fig. 8A summarize data from five immature
control neurons. In the absence of picrotoxin, glycinergic
mIPSC amplitudes ranged from 15 to 150 pA (mean, 42 ⫾ 7
pA), and the decay-time constant varied from 2 to 35 ms
(mean, 12.5 ⫾ 2.4 ms). Picrotoxin caused a strong reduction in
glycinergic events with a large amplitude and slow decay-
phase (Fig. 8A, F). After treatment of sister cultures with
colchicine, we found that the average amplitude and decay-
time constant of mIPSCs was reduced to 32 ⫾ 4 pA and 7.7 ⫾
FIG. 6. Acceleration of mIPSCs decay-phase kinetics with colchicine and
during spinal maturation. A: traces represent glycinergic currents recorded in 1.4 ms, respectively (Fig. 8B, E). Interestingly, most glycin-
an immature neuron at 1 min and after 25 min of dialysis with colchicine. Note ergic mIPSCs in these colchicine-treated neurons (n ⫽ 4) were
that colchicine affects not only the amplitude, but also accelerates the decay- resistant to picrotoxin application (Fig. 8B, F).
phase. B: mIPSCs recorded from a mature neuron show fast decay-kinetics that A correlation (r ⫽ 0.65) between the decay-time constant and
are unaffected by colchicine dialysis. C: mean decay-time constant of mIPSCs the amplitude of mIPSCs was found in control and treated neurons
at different times of development in control conditions (䡺) and in the presence
of colchicine (f). Top traces are typical averaged control responses obtained in (Fig. 8, A and B, lines). This correlation is unlikely to be due to
immature (␶1) and mature (␶3) neurons. Bottom traces are averaged responses dendritic cable properties because no correlation was detected
obtained from immature (␶2) and mature (␶4) colchicine-treated neurons. Thin between decay-time constant and rise-time (r ⫽ 0.35; Fig. 8, C
dotted line in the top control traces corresponds to the superimposed bottom and D, lines). The mean rise-time under control conditions was
traces obtained in the presence of colchicine and shows acceleration of decay
induced by treatment. Subscript for ␶ represents the number indicated in the 2.9 ⫾ 0.4 ms and remained similar after treatment with colchicine
graph. Each symbol represents means ⫾ SE obtained from 5– 6 cells. *P ⬍ and picrotoxin (P ⬎ 0.05, data not shown).
0.05. In summary, these results suggest that glycinergic mIPSCs
J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
COLCHICINE AFFECTS SYNAPTIC GLYRS IN IMMATURE NEURONS 1045

FIG. 7. Immature neurons express ␣2, ␣1,


and ␤-GlyR subunits, while mature neurons
down-regulated ␣2.A: bars represent GlyR tran-
scripts derived from murine ␣ and ␤ gene struc-
ture (closed regions are the signal peptides).
Selected primers for amplification of GlyR ␣-
and ␤-subunit isoforms are indicated by hori-
zontal arrows. cDNA fragment size for each
subunit is shown. B: profiles of GlyR ␣1 (529
bp), ␣2 (385 bp), and ␤ (183 bp) subunits
obtained from cultures with different stages of
development are visualized by agarose gel elec-
trophoresis. Note that immature neurons, but
not mature cells, displayed high levels of ␣2
GlyR cDNA. C and D: picrotoxin (PTX)
changed the properties of mIPSCs in immature
neurons but not in mature neurons. C: top
traces show glycinergic mIPSCs recorded from
an immature spinal neuron before and during
the application of 100 ␮M PTX. Bottom cur-
rent traces are typical averaged responses. Thin
dotted line in the left trace corresponds to su-
perimposed right trace obtained in the presence
of PTX and shows an acceleration of decay
phase. Bars illustrate the inhibitory effect of
PTX on the amplitude and decay time constant
in immature neurons. D: effect of PTX on
events recorded from mature neurons. Super-
imposed trace obtained in the presence of PTX
shows a small effect of the toxin on decay.
Each bar shows means ⫾ SE obtained from
5– 6 neurons. Normalized data in the presence
of PTX was obtained using the response at 1
min as 100%. *P ⬍ 0.05, **P ⬍ 0.01.

become less sensitive to picrotoxin after colchicine treatment, and mitter release would affect glycinergic and leave GABAAergic
support the hypothesis that colchicine predominantly affects neo- transmission unaffected. In addition, application of extracellu-
natal ␣2␤ GlyRs, leaving the adult ␣1␤ GlyRs unaffected. lar colchicine (20 min) did not reduce the frequency or ampli-
tude of glycinergic mIPSCs. Further evidence supporting a
DISCUSSION postsynaptic action of colchicine was obtained during this
study. First, intracellular application of the microtubule stabi-
Disruption of postsynaptic clusters affects glycinergic lizer GTP completely blocked the colchicine-induced reduction
transmission in glycinergic activity. Second, the rare occurrence of autapses
in these cultured spinal neurons indicate that the effect of
The data presented here and in a previous report (van
colchicine cannot be explained by depolymerization of vesicle-
Zundert et al. 2002) strongly suggest that colchicine actions on
associated microtubules in presynaptic autaptic terminals. Fi-
glycinergic currents are due to structural alterations of postsyn-
aptic receptor clusters consequent to microtubule depolymer- nally, we found that the reduction in event frequency with
ization. We found no evidence supporting the idea that intra- colchicine could be well reproduced by changing the Cl⫺
cellular dialysis of colchicine in the postsynaptic neurons driving force to obtain mIPSCs with smaller amplitudes.
might affect neurotransmitter release from presynaptic termi- Colchicine has been shown to competitively inhibit ␣2 and
nals. For instance, although glycinergic transmission was re- ␣1 GlyR subunits in Xenopus oocytes with IC50s of approxi-
duced by intracellular dialysis of colchicine, the synaptic trans- mately 64 and 324 ␮M, respectively (Machu 1998). Analysis
missions mediated by GABAARs and AMPARs remained the whole cell glycine-activated current during in vitro devel-
largely unmodified (van Zundert et al. 2002). In cultured spinal opment of spinal cord cultures demonstrated that immature and
neurons, GlyRs and GABAARs frequently co-localize in mature neurons display a similar sensitivity to colchicine with
postsynaptic gephyrin-containing clusters opposite to presyn- IC50s of 143 ⫾ 28 and 187 ⫾ 13 ␮M, respectively. In addition,
aptic boutons containing inhibitory amino acids (Dumoulin et extracellular application of 20 ␮M colchicine had no effect on
al. 2000). Hence, it is not likely that alterations in neurotrans- glycinergic mIPSCs. Taken together, the effects observed with
J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
1046 VAN ZUNDERT ET AL.

decay-phase of glycinergic mIPSCs during development de-


pends on a transition between slow ␣2 and fast ␣1 subunits
(Krupp et al. 1994; Legendre 2001; Singer et al. 1998; Taka-
hashi et al. 1992). We also found that the time-course of the
glycinergic mIPSCs markedly accelerated with maturation of
spinal neurons in culture. Analysis of mIPSCs properties
showed that immature slow events were more sensitive to
picrotoxin than mature faster events. However, our data sug-
gest that immature glycinergic currents are primarily associ-
ated to ␣2␤ subunits since these slow decay mIPSCs were
affected by higher picrotoxin concentrations than those used to
inhibit (IC50 ⫽ 7 ␮M) homomeric ␣ receptors (Legendre 1997;
Pribilla et al. 1992; Tapia and Aguayo 1998; Ye 2000). This
conclusion is in agreement with other previous studies in
immature spinal neurons showing the presence of a 48 pS
channel, typical of heteromeric ␣2␤ receptors, and not a ⬎85
pS associated to homomeric receptors (Bormann et al. 1987,
1993; Takahashi and Momiyama 1991; Twyman and Mac-
donald 1991). Similar to results obtained in newborn brain
motoneurons (Singer et al. 1998), our results indicate that
FIG. 8. Colchicine inhibits the effect of picrotoxin on mIPSCs. A: correla-
immature neurons express GlyRs that are formed by ␣2 and ␤
tion between decay-time constant and mIPSC amplitude before (E) and during
(F) application of 100 ␮M PTX in immature neurons. B: lack of PTX effect (F) subunits. During maturation, the mRNA level for the ␣2-
on colchicine-treated neurons (20 ␮M for 3 h). Note that amplitudes and subunit decreased, shifting the balance of mRNA expression
decay-time constants are already reduced after colchicine treatment (E). r ⬍ toward ␣1 and ␤-subunits. In addition, analysis of the decay-
0.7 in all treatments. Plots of decay-time constant vs. rise-time before (䡺) and time constant and the sensitivity to picrotoxin suggests that
during (f) application of 100 ␮M PTX in untreated (C) and colchicine-treated
neurons (D). No significant correlation was found (r ⫽ 0.35). Data points are mature neurons express primarily ␣1␤ GlyRs. This finding is
obtained from 4 –5 neurons. in agreement with previous studies (Krupp et al. 1994; Singer
et al. 1998; Takahashi et al. 1992).
internal dialysis of 20 ␮M colchicine in immature and mature These data suggest that mature and immature GlyRs are
neurons are best explained by a postsynaptic mechanism con- composed of ␤-subunits and that they are linked to underlying
sequent to microtubule disruption (see van Zundert et al. 2002). postsynaptic microtubules. However, disruption of microtu-
In addition, our results provide good evidence supporting the bules in immature neurons selectively affected a population of
idea that microtubules are important to maintain gephyrin/ glycinergic currents displaying slow decay-times, thus leaving
GlyRs clustered in the postsynaptic membrane of immature faster events unaffected. These results indicate that microtu-
neurons, since we found that colchicine efficiently reduced the bules regulate the immature ␣2␤ GlyR, but not the adult ␣1␤
size and luminosity of postsynaptic gephyrin/GlyR clusters. receptor. Two lines of evidence support this idea. First, the
Previous studies (Lim et al. 1999; Oleskevich et al. 1999), in picrotoxin-sensitive population of mIPSCs was no longer evi-
addition to our present data, suggest that the amplitude of dent after disruption of microtubules with colchicine, suggest-
glycine-mediated mIPSCs and the size of gephyrin clusters are ing they represent the same population of receptors. Second,
positively correlated. Colchicine reduced both parameters in mature neurons with fast ␣1␤ GlyRs were insensitive to col-
immature, but not in mature neurons. In conclusion, the reduc- chicine.
tion in the amplitude of the glycinergic mIPSCs after micro- Based on the current understanding regarding functional and
tubule disruption can be best explained by a structural alter- structural properties of synaptic GlyRs, we postulate a model
ation of the postsynaptic density that results in decreased in which immature neurons express both ␣2␤ and ␣1␤ GlyRs
numbers of functional GlyRs in the synapse. at the postsynaptic membrane, while mature neurons contain
mainly ␣1␤ GlyRs at the synapse (Fig. 9). Microtubules, via
Microtubule disruption preferentially affects immature gephyrin, anchor both types of GlyRs in the postsynaptic
␣2␤-GlyRs membrane. On microtubule depolymerization, the molecular
complex made up by gephyrin and ␣2␤ GlyRs is liberated and
While the amplitude of the mIPSC appears to reflect the diffuses to the extrasynaptic membrane, reducing glycinergic
number of postsynaptic receptors, its decay is dictated by the mIPSCs and gephyrin/GlyR cluster size and density. In con-
kinetic properties of single channels. For example, in overex- trast, colchicine treatment does not affect ␣1␤ GlyR/gephyrin
pression studies, ␣1 and ␣2 subunits assemble homomeric clusters in immature and mature neurons (Fig. 9). This can be
GlyRs with mean open times of 2 and 174 ms, respectively explained by the linkage of GlyR/gephyrin complexes to mi-
(Takahashi et al. 1992). In parallel with the known subunits crotubules that are stabilized by posttranslational modifica-
switching from ␣2 to ␣1 subunits during postnatal develop- tions, such as acetylation and/or the binding of molecules such
ment (Becker et al. 1988), patch-clamp analysis in spinal cord as MAPs (Nogales 2000). In addition, the interaction of GlyRs
slices demonstrated that the open time of native GlyR channels with subsynaptic microtubules could depend on the type of ␣
decreased from 40 ms at E20 to 6 ms at P22 (Takahashi et al. subunit present in the receptor rather than to differences in
1992). Therefore it is believed that the differences in the cytoskeleton properties. Interestingly, more than 10 potential
J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org
Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
COLCHICINE AFFECTS SYNAPTIC GLYRS IN IMMATURE NEURONS 1047

FIG. 9. Proposed mechanisms for anchoring ␣2␤ and ␣1␤


GlyRs in glycinergic synapses. Microtubules, via gephyrin,
anchor ␣2␤- and ␣1␤-containing GlyRs in the postsynaptic
membrane. While immature synapses are enriched with ␣2
GlyRs, these are down-regulated at mature synapses that are
composed almost exclusively of ␣1-containing GlyRs. Depo-
lymerization of microtubules by colchicine causes the release of
␣2␤ GlyR/gephyrin complexes and diffusion to the extrasyn-
aptic membrane. In contrast, gephyrin-coupled ␣1␤ GlyRs are
linked to colchicine insensitive microtubules, which could be
stabilized by MAPs or acetylation (Ac). Alternatively, stabili-
zation of ␣1␤ GlyR/gephyrin clusters can be achieved by
assembling gephyrin into hexagonal lattices. Therefore the in-
sensitivity of mature receptors to cytoskeleton disruption can be
also associated to the presence of ␣1␤ GlyR/gephyrin com-
plexes rather than to microtubule-dependent mechanisms. Ver-
tical lines indicate the location of synaptic region.

gephyrin isoforms can be generated by alternative splicing, In conclusion, this study shows that disruption of microtu-
allowing ␣1␤- and ␣2␤-GlyRs to associate with gephyrin bules by colchicine reduced both postsynaptic glycinergic cur-
variants that have distinct affinities for cytoskeletal elements. rents and the size and density of synaptic gephyrin clusters in
Thus the developmental dependent change of the GlyR molec- immature spinal neurons. Thus this study expands those on
ular composition could affect the manner in which the receptor regulation of n-methyl-D-aspartic acid receptors (NMDARs),
interacts with the gephyrin scaffold and subsynaptic microtu- AMPARs, and nicotinic acetylcholine receptors (nAChRs) by
bules. the state of the cytoskeleton (Allison et al. 1998, 2000; Sattler
Another possibility is that, after enough ␣1␤ GlyR/gephyrin et al. 2000; Shoop et al. 2000). Moreover, our data suggest that
complexes have accumulated in the synapse during maturation, microtubule disruption affects immature synapses rather than
gephyrin forms stable hexagonal scaffolds in the PSD, which mature synapses. This property could explain previous contro-
maintain ␣1␤ GlyRs anchored at the postsynaptic membrane versial findings in which disruption of microtubules altered the
and independent of microtubules (Fig. 9; see also Kneussel and number and morphology of gephyrin clusters in 10 –12 DIV
Betz 2000; Liu et al. 2000). Thus as proposed for the role of spinal neurons (Kirsch and Betz 1995), while gephyrin distri-
actin filaments in excitatory hippocampal synapses (Allison et bution in ⬎21 DIV hippocampal neurons was not affected by
al. 2000; Zhang and Benson 2001), the state of microtubules microtubule depolymerization (Allison et al. 2000).
could play a role in the development and maintenance of
predominantly immature glycinergic synapses, which need to ACKNOWLEDGMENTS
be highly plastic to shape efficient synaptic transmission. In We thank Drs. P. Legendre and B. Walmsley for reading and suggestions on
agreement with this idea, electron microscopy studies have not this manuscript. The authors thank Dr. J. M. Fritschy for making the
consistently found microtubules within the 20-to 30-nm region GABAAR␥2 antibody available and L. J. Aguayo for expert technical assis-
tance.
beneath adult symmetric (inhibitory) or gephyrin-containing
synapses (Alvarez et al. 1997; Peters et al. 1991; Triller et al.
GRANTS
1985, 1987). It is possible that postsynaptic microtubules are
not well preserved in routine or immunocytochemical electron Research support was provided by Fondo Nacional de Ciencia y Tecnologia
microscopy preparations as previously shown for presynaptic (FONDECYT) 2000135 and Programa de Mejoramiento de la Calidad y
Equidad de la Educación Superior del Ministerio de Educaión UCH9903 to B.
microtubules (Gray 1975). Accordingly, some microtubules van Zundert, FONDECYT 1980106, 1020475, and GIA-DIUC to L. G.
were shown in close relationship to immature and mature Aquayo, FONDECYT 2990063 to J. C. Tapia, the National Science Founda-
postsynaptic densities (not necessarily inhibitory) in tissue tion 9984441 to F. J. Alvarez, and National Institute of Neurological Disorders
sections or cultures pretreated with taxol, cryosubstitution and Stroke Grants NS-24830 and NS-41489 to H. H. Yeh.
techniques, or other procedures aimed to stabilize cytoskeletal
components (Bird 1989; Ichimura and Hashimoto 1988; Le- REFERENCES
Beux and Willemot 1975; Westrum and Gray 1977). However, Aguayo LG and Pancetti FC. Ethanol modulation of the gamma-aminobu-
the most convincing ultrastructural evidence available on mi- tyric acidA- and glycine-activated Cl⫺ current in cultured mouse neurons.
crotubule presence in the PSD was obtained in neonatal syn- J Pharmacol Exp Ther 270: 61– 69, 1994.
apses (Westrum and Gray 1977). To our knowledge, no sys- Ali DW, Drapeau P, and Legendre P. Development of spontaneous glycin-
ergic currents in the Mauthner neuron of the zebrafish embryo. J Neuro-
tematic ultrastructural analysis of the spatial relationships be- physiol 84: 1726 –1736, 2000.
tween microtubules and the postsynaptic density of inhibitory Allison DW, Chervin AS, Gelfand VI, and Craig AM. Postsynaptic scaf-
synapses is available. folds of excitatory and inhibitory synapses in hippocampal neurons: main-

J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org


Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
1048 VAN ZUNDERT ET AL.

tenance of core components independent of actin filaments and microtu- Krupp J, Larmet Y, and Feltz P. Postnatal change of glycinergic IPSC decay
bules. J Neurosci 20: 4545– 4554, 2000. in sympathetic preganglionic neurons. Neuroreport 5: 2437–2440, 1994.
Allison DW, Gelfand VI, Spector I, and Craig AM. Role of actin in LeBeux YJ and Willemot J. An ultrastructural study of the microfilaments in
anchoring postsynaptic receptors in cultured hippocampal neurons: differ- rat barin by means of E-PTA staining and heavy meromyosin labeling. II.
ential attachment of NMDA versus AMPA receptors. J Neurosci 18: 2423– The synapses. Cell Tissue Res 160: 37– 68, 1975.
2436, 1998. Legendre P. Pharmacological evidence for two types of postsynaptic glycin-
Alvarez FJ, Dewey DE, Harrington DA, and Fyffe RE. Cell-type specific ergic receptors on the Mauthner cell of 52-h-old zebrafish larvae. J Neuro-
organization of glycine receptor clusters in the mammalian spinal cord. physiol 77: 2400 –2415, 1997.
J Comp Neurol 379: 150 –170, 1997. Legendre P. The glycinergic inhibitory synapse. Cell Mol Life Sci 58: 760 –
Bechade C, Colin I, Kirsch J, Betz H, and Triller A. Expression of glycine 793, 2001.
receptor alpha subunits and gephyrin in cultured spinal neurons. Eur J Neu- Lim R, Alvarez FJ, and Walmsley B. Quantal size is correlated with receptor
rosci 8: 429 – 435, 1996. cluster area at glycinergic synapses in the rat brainstem. J Physiol 516:
Becker D-M, Hich W, and Betz H. Glycine receptor heterogeneity in rat 505–512, 1999.
spinal cord during postnatal development. EMBO J 7: 3717–3726, 1988. Liu MT, Wuebbens MM, Rajagopalan KV, and Schindelin H. Crystal
Bekkers JM and Stevens CF. Excitatory and inhibitory autaptic currents in structure of the gephyrin-related molybdenum cofactor biosynthesis protein
isolated hippocampal neurons maintained in cell culture. Proc Natl Acad Sci MogA from Escherichia coli. J Biol Chem 275: 1814 –1822, 2000.
USA 88: 7834 –7838, 1991. Machu T. Colchicine competitively antagonizes glycine receptors expressed
Bird MM. Microtubules and their relationships with other cytoskeletal com- in Xenopus oocytes. Neuropharmacology 37: 391–396, 1998.
ponents at cholinergic tectal synapses in culture. J Anat 166: 1– 6, 1989. Mammoto A, Sasaki T, Asakura T, Hotta I, Imamura H, Takahashi K,
Bormann J, Hamill OP, and Sakmann B. Mechanism of anion permeation Matsuura Y, Shirao T, and Takai Y. Interactions of drebrin and gephyrin
through channels gated by glycine and gamma-aminobutyric acid in mouse with profilin. Biochem Biophys Res Commun 243: 86 – 89, 1998.
cultured spinal neurones. J Physiol 385: 243–286, 1987. Matus AI and Taff-Jones DH. Morphology and molecular composition of
Bormann J, Rundstrom N, Betz H, and Langosch D. Residues within isolated postsynaptic junctional structures. Proc R Soc Lond B Biol Sci 203:
transmembrane segment M2 determine chloride conductance of glycine 135–151, 1978.
receptor homo- and hetero-oligomers. EMBO J 12: 3729 –3737, 1993. Meier J, Meunier-Durmort C, Forest C, Triller A, and Vannier C. For-
Colin I, Rostaing P, and Triller A. Gephyrin accumulates at specific plas- mation of glycine receptor clusters and their accumulation at synapses.
malemma loci during neuronal maturation in vitro. J Comp Neurol 374: J Cell Sci 113: 2783–2795, 2000.
467– 479, 1996. Meyer G, Kirsch J, Betz H, and Langosch D. Identification of a gephyrin
Cotman CW, Banker G, Churchill L, and Taylor D. Isolation of postsyn- binding motif on the glycine receptor beta subunit. Neuron 15: 563–572,
aptic densities form rat brain. J Cell Biol 63: 441– 455, 1974. 1995.
Dumoulin A, Levi S, Riveau B, Gasnier B, and Triller A. Formation of Nogales E. A structural view of microtubule dynamics. Cell Mol Life Sci 56:
mixed glycine and GABAergic synapses in cultured spinal cord neurons. 133–142, 2000.
Eur J Neurosci 12: 3883–3892, 2000. Oleskevich S, Alvarez FJ, and Walmsley B. Glycinergic miniature synaptic
Essrich C, Lorez M, Benson JA, Fritschy JM, and Luscher B. Postsynaptic currents and receptor cluster sizes differ between spinal cord interneurons.
clustering of major GABAA receptor subtypes requires the gamma 2 subunit J Neurophysiol 82: 312–319, 1999.
and gephyrin. Nat Neurosci 1: 563–571, 1999. Peters A, Palay SL, and Webster H. The Fine Structure of the Nervous
Feng G, Tintrup H, Kirsch J, Nichol MC, Kuhse J, Betz H, and Sanes JR. System: The Neurons and Supporting Cells. Philadelphia, PA: W. B. Saun-
Dual requirement for gephyrin in glycine receptor clustering and molyb- ders, 1991.
doenzyme activity. Science 282: 1321–1324, 1998. Pribilla I, Takagi T, Langosch D, Bormann J, and Betz H. The atypical M2
Geiman EJ, Knox MC, and Alvarez FJ. Postnatal maturation of gephyrin/ segment of the beta subunit confers picrotoxinin resistance to inhibitory
glycine receptor clusters on developing Renshaw cells. J Comp Neurol 426: glycine receptor channels. EMBO J 11: 4305– 4311, 1992.
130 –142, 2000. Rosenberg M, Meier J, Triller A, and Vannier C. Dynamics of glycine
Gray EG. Presynaptic microtubules and their association with synaptic vesi- receptor insertion in the neuronal plasma membrane. J Neurosci 21: 5036 –
cles. Proc R Soc Lond Biol Sci 190: 367–372, 1975. 5044, 2001.
Ichimura T and Hashimoto PH. Structural components in the synaptic clefts Rosenmund C and Westbrook GL. Calcium-induced actin depolymerization
captured by freeze substitution and deep etching of directly frozen cerebellar reduces NMDA channel activity. Neuron 10: 805– 814, 1993.
cortex. J Neurocytol 17: 3–12, 1988. Sattler R, Xiong Z, Lu WY, MacDonald JF, and Tymianski M. Distinct
Kelly PT and Cotman CW. Synaptic proteins. Characterization of tubulin and roles of synaptic and extrasynaptic NMDA receptors in excitotoxicity.
actin and identification of a distinct postsynaptic density polypeptide. J Cell J Neurosci 20: 22–33, 2000.
Biol 79: 173–183, 1978. Sheng M and Pak DT. Ligand-gated ion channel interactions with cytoskel-
Kins S, Betz H, and Kirsch J. Collybistin, a newly identified brain-specific etal and signaling proteins. Annu Rev Physiol 62: 755–778, 2000.
GEF, induces submembrane clustering of gephyrin. Nat Neurosci 3: 22–29, Shoop RD, Yamada N, and Berg DK. Cytoskeletal links of neuronal ace-
2000. tylcholine receptors containing alpha 7 subunits. J Neurosci 20: 4021– 4029,
Kirchhoff F, Mulhardt C, Pastor A, Becker CM, and Kettenmann H. 2000.
Expression of glycine receptor subunits in glial cells of the rat spinal cord. Singer JH, Talley EM, Bayliss DA, and Berger AJ. Development of gly-
J Neurochem 66: 1383–1390, 1996. cinergic synaptic transmission to rat brain stem motoneurons. J Neuro-
Kirsch J. Assembly of signaling machinery at the postsynaptic membrane. physiol 80: 2608 –2620, 1998.
Curr Opin Neurobiol 9: 329 –335, 1999. Takahashi T and Momiyama A. Single-channel currents underlying glycin-
Kirsch J and Betz H. The postsynaptic localization of the glycine receptor- ergic inhibitory postsynaptic responses in spinal neurons. Neuron 7: 965–
associated protein gephyrin is regulated by the cytoskeleton. J Neurosci 15: 969, 1991.
4148 – 4156, 1995. Takahashi T, Momiyama A, Hirai K, Hishinuma F, and Akagi H. Func-
Kirsch J, Langosch D, Prior P, Littauer UZ, Schmitt B, and Betz H. The tional correlation of fetal and adult forms of glycine receptors with devel-
93-kDa glycine receptor-associated protein binds to tubulin. J Biol Chem opmental changes in inhibitory synaptic receptor channels. Neuron 9: 1155–
266: 22242–22245, 1991. 1161, 1992.
Kirsch J, Wolters I, Triller A, and Betz H. Gephyrin antisense oligonucle- Tapia JC and Aguayo LG. Changes in the properties of developing glycine
otides prevent glycine receptor clustering in spinal neurons. Nature 366: receptors in cultured mouse spinal neurons. Synapse 28: 185–194, 1998.
745–748, 1993. Triller A, Cluzeaud F, and Korn H. Gamma-aminobutyric acid-containing
Kneussel M and Betz H. Clustering of inhibitory neurotransmitter receptors terminals can be apposed to glycine receptors at central synapses. J Cell Biol
at developing postsynaptic sites: the membrane activation model. Trends 104: 947–956, 1987.
Neurosci 23: 429 – 435, 2000. Triller A, Cluzeaud F, Pfeiffer F, Betz H, and Korn H. Distribution of
Kneussel M, Brandstatter JH, Laube B, Stahl S, Muller U, and Betz H. glycine receptors at central synapses: an immunoelectron microscopy study.
Loss of postsynaptic GABA(A) receptor clustering in gephyrin-deficient J Cell Biol 101: 683– 688, 1985.
mice. J Neurosci 19: 9289 –9297, 1999. Triller A, Seitanidou T, Franksson O, and Korn H. Size and shape of

J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org


Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.
COLCHICINE AFFECTS SYNAPTIC GLYRS IN IMMATURE NEURONS 1049

glycine receptor clusters in a central neuron exhibit a somato-dendritic tively regulated by the cytoskeleton in mouse spinal neurons. J Neurophysiol
gradient. New Biol 2: 637– 641, 1990. 87: 640 – 644, 2002.
Twyman RE and Macdonald RL. Kinetic properties of the glycine receptor Walmsley B, Alvarez FJ, and Fyffe RE. Diversity of structure and function
main- and sub-conductance states of mouse spinal cord neurones in culture. at mammalian central synapses. Trends Neurosci 21: 81– 88, 1998.
J Physiol 435: 303–331, 1991. Westrum LE and Gray EG. Microtubules associated with postsynaptic
van Zundert B, Alvarez FJ, and Aguayo LG. Colchicine affects the ampli- thickenings. J Neurocytol 6: 505–518, 1977.
tude of glycinergic synaptic currents and the size of glycine receptor clusters Ye J. Physiology and pharmacology of native glycine receptors in developing
in the postsynaptic membrane. Soc Neurosci Abstr 27: 728, 2001. rat ventral tegmental area neurons. Brain Res 862: 74 – 82, 2000.
van Zundert B, Alvarez FJ, Yevenes GE, Cárcamo JG, Vera JC, and Zhang W and Benson DL. Stages of synapse development defined by
Aguayo LG. Glycine receptors involved in synaptic transmission are selec- dependence on F-actin. J Neurosci 21: 5169 –5181, 2001.

J Neurophysiol • VOL 91 • FEBRUARY 2004 • www.jn.org


Downloaded from www.physiology.org/journal/jn (131.000.054.235) on April 12, 2019.

You might also like