Ikemoto 2002

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 278, No. 8, Issue of February 21, pp.

5929 –5940, 2003


© 2003 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

Glycolysis and Glutamate Accumulation into Synaptic Vesicles


ROLE OF GLYCERALDEHYDE PHOSPHATE DEHYDROGENASE AND 3-PHOSPHOGLYCERATE KINASE*

Received for publication, November 14, 2002


Published, JBC Papers in Press, December 17, 2002, DOI 10.1074/jbc.M211617200

Atsushi Ikemoto‡§, David G. Bole‡, and Tetsufumi Ueda‡¶储**


From the ‡Mental Health Research Institute, Departments of ¶Pharmacology and 储Psychiatry,
University of Michigan Medical School, Ann Arbor, Michigan 48109-0669

Glucose is the major source of brain energy and is synaptic transmission caused by hypoglycemia occurs in part if
essential for maintaining normal brain and neuronal not entirely by a presynaptic mechanism (7, 11, 12). Fleck et al.
function. Hypoglycemia causes impaired synaptic trans- (7) have shown that substantial reduction of extracellular glu-
mission. This occurs even before significant reduction cose results in a decrease in stimulus-evoked Glu release, with
in global cellular ATP concentration, and relationships no changes in ATP levels. These studies together suggest that
among glycolysis, ATP supply, and synaptic transmis- glycolysis or glycolytic intermediate(s) are necessary for nor-
sion are not well understood. We demonstrate that the mal synaptic transmission independent of global cellular ATP
glycolytic enzymes glyceraldehyde phosphate dehydro- levels.
genase (GAPDH) and 3-phosphoglycerate kinase In an attempt to reveal the underlying mechanism of hypo-
(3-PGK) are enriched in synaptic vesicles, forming a

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


glycemia-induced aberrant synaptic transmission, we previ-
functional complex, and that synaptic vesicles are capa-
ously explored the possibility that glycolytic intermediates
ble of accumulating the excitatory neurotransmitter
could modify proteins localized in the nerve ending (13, 14).
glutamate by harnessing ATP produced by vesicle-
3-Phosphoglycerate (3-PG)1 was demonstrated to stimulate
bound GAPDH/3-PGK at the expense of their substrates.
The GAPDH inhibitor iodoacetate suppressed GAPDH/ phosphorylation of 155- and 72-kDa proteins. The latter was
3-PGK-dependent, but not exogenous ATP-dependent, identified as glucose-1,6-bisphosphate synthetase, and 1,3-
[3H]glutamate uptake into isolated synaptic vesicles. It bisphosphoglycerate (1,3-BPG) was found to serve as the direct
also decreased vesicular [3H]glutamate content in the substrate for phosphorylation of this enzyme, by donating
nerve ending preparation synaptosome; this decrease 1-phosphate. Both of these phosphorylated proteins are en-
was reflected in reduction of depolarization-induced riched in the synaptosomal (nerve ending preparation) as well
[3H]glutamate release. In contrast, oligomycin, a mito- as cell body soluble fractions, but the significance of these
chondrial ATP synthase inhibitor, had minimal effect on modifications in synaptic transmission remains unclear.
any of these parameters. ADP at concentrations above In this paper, we show that the glycolytic intermediate 1,3-
0.1 mM inhibited vesicular glutamate and dissipated BPG forms an acyl-enzyme intermediate with vesicle-bound
membrane potential. This suggests that the coupled glyceraldehyde phosphate dehydrogenase (GAPDH), that ves-
GAPDH/3-PGK system, which converts ADP to ATP, en- icle-bound GAPDH exists in a complex with 3-phosphoglycer-
sures maximal glutamate accumulation into presynap- ate kinase (3-PGK), and that activation of vesicle-associated
tic vesicles. Together, these observations provide in- GAPDH and 3-PGK is sufficient to support vesicular uptake of
sight into the essential nature of glycolysis in sustaining Glu. Glutamate is now recognized as the major excitatory neu-
normal synaptic transmission. rotransmitter responsible for triggering neuronal firing, in the
vertebrate central nervous system (15–23). As such, proper Glu
synaptic transmission is not only essential for basic neuronal
Glycolysis plays a vital role in maintaining normal brain communication but also is involved in learning and memory
function. Glucose is known to serve as the major substrate for formation (19, 20). Glutamate accumulation into synaptic ves-
cerebral energy under normal conditions (1). Recent evidence icles in the nerve terminal is an initial crucial step in Glu
suggests a direct correlation between glucose utilization and transmission (18, 22–26). This process requires ATP to gener-
cognitive function (2). Reduction of glucose levels results in ate an electrochemical gradient, which is the driving force for
pathophysiological states and abnormal electrophysiological Glu uptake into synaptic vesicles (27–34).
activity; however, this occurs long before significant alteration We present evidence that glycolytically produced ATP, in
in tissue ATP levels is detected (3–7). Substitution of pyruvate particular that produced by GAPDH and 3-PGK, but not
for glucose does not support normal evoked neuronal activity, mitochondria-derived ATP, is harnessed for accumulation of
although tissue ATP level returns to normal (8 –10). Abnormal Glu into synaptic vesicles in synaptosomes; Glu transported
into synaptic vesicles in this manner is released upon depo-
* This work was supported in part by National Institutes of Health larization. These findings could provide an explanation for
Grants NS 24384, NS 36656, and NS 42200 and a grant from Taisho hypoglycemia-induced aberrant synaptic transmission and
Pharmaceutical Co., Ltd. (Tokyo, Japan). The costs of publication of this
article were defrayed in part by the payment of page charges. This
article must therefore be hereby marked “advertisement” in accordance
1
with 18 U.S.C. Section 1734 solely to indicate this fact. The abbreviations used are: 3-PG, 3-phosphoglycerate; t-ACPD,
§ On leave from the Dept. of Biological Chemistry, Faculty of Phar- trans-1-aminocyclopentane-1,3-dicarboxylic acid; 4-AP, 4-aminopyri-
maceutical Sciences, Nagoya City University, 467-8603 Nagoya, Japan. dine; 1,3-BPG, 1,3-bisphosphoglycerate; DHAP, dihydroxyacetone
** To whom correspondence should be addressed: Mental Health phosphate; FCCP, carbonyl cyanide p-(trifluoromethoxy)-phenylhydra-
Research Institute at MSRB II, C570D, The University of Michigan, zone; GAP, glyceraldehyde-3-phosphate; GAPDH, glyceraldehyde-3-
1150 W. Medical Center Dr., Ann Arbor, MI 48109-0669. Tel.: 734-763- phosphate dehydrogenase; 3-PGK, 3-phosphoglycerate kinase; PGM,
3790; Fax: 734-936-2690; E-mail: tueda@umich.edu. monophosphoglycerate mutase; ACSF, artificial cerebrospinal fluid.

This paper is available on line at http://www.jbc.org 5929


5930 GAPDH and Vesicular Glu Accumulation
insight into the essential nature of glycolysis in normal
synaptic transmission.

EXPERIMENTAL PROCEDURES
Materials—[␥-32P]ATP (6,000 Ci/mmol) was obtained from
PerkinElmer Life Sciences. L-[G-3H]Glutamic acid (42.0 Ci/mmol) was
purchased from Amersham Biosciences. The affinity-purified polyclonal
antibodies (rabbit IgG) specific to recombinant monophosphoglycerate
mutase (PGM) type B was kindly provided by Oriental Yeast Co., Ltd.
(Tokyo, Japan). Anti-GAPDH monoclonal antibody (mouse IgG) and
anti-3-PGK polyclonal antibody (rabbit IgG) were purchased from U.S.
Biological (Swampscott, MA) and Accurate Chemical & Science Co.
(Westbury, NY), respectively. Glycolytic enzymes and all other chemi-
cals were purchased from Sigma-Aldrich unless mentioned elsewhere.
Preparation of Subcellular Fractions—Synaptic vesicles were pre-
pared from bovine cerebrum through the discontinuous sucrose gradi-
ent procedure as described previously (35). The subcellular fractions of
bovine cerebrum were prepared as described previously (36). Synapto-
somes were prepared from cerebra of male Sprague-Dawley rats (150 –
200 g) and purified through the Percoll gradient centrifugation step, as
described by Dunkley et al. (37). Protein concentration was determined
by the method of Bradford (38) with a Coomassie protein assay reagent
kit (Pierce) with bovine serum albumin as standard protein.
Synthesis of [3-32P]1,3-BPG—[32P]Dihydroxyacetone phosphate
(DHAP) was prepared by phosphorylation of dihydroxyacetone by glyc-
erol kinase with [␥-32P]ATP in a mixture containing 5 mM Tris-HCl (pH

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


7.0), 40 ␮M MgSO4, 10 mM dihydroxyacetone, 20 ␮M [␥-32P]ATP (6,000
Ci/mmol), and 0.5 units of glycerol kinase (Bacillus stearothermophi-
lus). The reaction was performed in a volume of 40 ␮l at 37 °C for 20
min. [3-32P]1,3-BPG was prepared from [32P]DHAP by conversion to
[32P]GAP with triose-phosphate isomerase, followed by a GAPDH reac-
tion in the presence of the lactate dehydrogenase-coupled NAD-regen-
erating system. The reaction mixture (400 ␮l) contained 12.5 mM tri-
ethanolamine (pH 8.0), 0.2 mM EDTA, 2 mM NAD, 2 mM sodium
pyruvate, 2 mM KH2PO4, 50 ␮M DHAP, 3.2 units of GAPDH (rabbit
muscle), 20 units of triose-phosphate isomerase (rabbit muscle), 10
units of lactate dehydrogenase (rabbit muscle), and 40 ␮l of the [32P]D-
HAP reaction mixture. The entire mixture was incubated at 25 °C for 5
min, filtered to remove the enzymes by an Amicon Centricon-10 con-
centrator at 4 °C, and put onto DEAE-cellulose (Whatman DE32, 1.2 ⫻
2.2 cm) previously equilibrated with 10 mM glycylglycine, pH 7.4. Elu-
tion was carried out with stepwise increases in the NaCl concentration:
50, 75, 100, 125, 150, and 200 mM. DHAP, GAP, and inorganic phos-
phate (Pi) were eluted with 75 mM NaCl in the same buffer. 3-PG and
1,3-BPG were eluted with 125 and 150 mM NaCl, respectively, in the
same buffer. All the compounds thus prepared were stored at ⫺80 °C
until use. FIG. 1. GAPDH is labeled with [3-32P]1,3-BPG and enriched in
Radioactive compounds were analyzed by high pressure liquid chro- synaptic vesicles. Synaptic vesicles were incubated at 37 °C for 10 s
matography on a Whatman Partisil 10 SAX WCS column (4.6 ⫻ 250 with [3-32P]1,3-BPG in the absence or presence of 4 mM MgSO4 (Mg2⫹)
mm), comparing their retention times with those of nonradioactive or 50 ␮M 3-PG and subjected to SDS-PAGE under neutral pH condi-
tions, according to the method of Fairbanks (A) or under alkaline pH
authentic standards monitored at 214 nm. The column was equilibrated
conditions (standard SDS-PAGE) according to the method of Laemmli
with 0.4 M sodium phosphate buffer (pH 3.2), and glycolytic intermedi-
(B), followed by autoradiography. C, the 0.2 M NaCl extract of the
ates and nucleotides were eluted isocratically, as described previously synaptic vesicle fraction was incubated at 37 °C for 10 s with [3-32P]1,3-
(14). Retention times for glycolytic intermediates were 4.1 min for BPG in the presence (anti-GAPDH, mouse IgG) or absence (anti-PGM,
DHAP and GAP, 5.3 min for Pi, 5.9 min for 2-phosphoglycerate and rabbit IgG) of 4 mM Mg2⫹; immunoprecipitated with the anti-GAPDH
3-PG, 6.0 min for ADP, 7.4 min for phosphoenolpyruvate, 15.4 min for antibody, anti-PGM antibody, mouse IgG (control for anti-GAPDH), or
1,3-BPG, and 30.0 min for ATP. rabbit IgG (control for anti-PGM); and subjected to SDS-PAGE under
Protein Labeling with [3-32P]1,3-BPG—The synaptic vesicle fraction neutral pH conditions, followed by autoradiography. D, various subcel-
(30 ␮g of protein) was preincubated at 37 °C for 30 s in 27 ␮l of 5 mM lular fractions (30 ␮g of protein) as indicated were incubated at 37 °C
Tris-maleate (pH 7.4). The reaction was initiated by the addition of 3 ␮l for 10 s with 140 nM [3-32P]1,3-BPG and subjected to SDS-PAGE under
of [3-32P]1,3-BPG (240 Ci/mmol) to a final concentration of 140 nM and neutral pH conditions, followed by autoradiography. E, the subcellular
allowed to continue for 10 s. For electrophoretic protein separation fractions were separated by standard SDS-PAGE, blotted onto polyvi-
under neutral pH conditions, the reaction was terminated by the addi- nylidene difluoride membranes, and detected with anti-GAPDH. F, the
tion of 10 ␮l of buffer containing 4% SDS, 30% sucrose, 40 mM Tris-HCl synaptic vesicle fraction was treated with various concentrations of
(pH 7.0), 4 mM EDTA, 160 mM 2-mercaptoethanol, and 50 ␮g/ml brom- NaCl, and the salt extracts (supernatant (Sup)) and the salt-insoluble
phenol blue; aliquots (25 ␮l) were subjected to polyacrylamide gel (1% material (Pellet) were analyzed for GAPDH as described above. The
data shown are representative of results obtained from three separate
SDS and 5.6% acrylamide) without stacking gel according to the method
experiments. Syn. cytosol, synaptosomal cytosol; Per. cytosol, perikaryal
of Fairbanks et al. (39). For electrophoretic protein separation under
cytosol (cell body cytosol); Pl. membrane, plasma membrane.
alkaline pH conditions (standard SDS-PAGE), the reaction was termi-
nated by the addition of 10 ␮l of SDS sample buffer containing 4% SDS,
40% glycerol, 0.2 M Tris-HCl (pH 6.8), 20% 2-mercaptoethanol, and 50 bilized protein G (0.1 ml as 50% slurry) and chemically cross-linked,
␮g/ml bromphenol blue; aliquots (25 ␮l) were subjected to polyacryl- using the Seize X mammalian immunoprecipitation kit (Pierce). The
amide gel (0.1% SDS and 12% acrylamide) according to the method of synaptic vesicle fraction (50 ␮g of protein) was stirred in 0.2 ml of buffer
Laemmli (40), except for omission of sample boiling. Autoradiography containing 0.32 M sucrose, 4 mM Tris-maleate (pH 7.4), and 0.2 M NaCl
was carried out as described previously (41) and analyzed using an for 5 min at 4 °C and centrifuged at 200,000 gmax for 1 h at 4 °C. When
32
image analyzer (Bio-Rad Gel Doc 2000). P-labeled protein was detected, the synaptic vesicle fraction was in-
Immunoprecipitation—Antibodies (10 ␮g) were absorbed onto immo- cubated at 37 °C for 10 s with the same buffer containing 140 nM (240
GAPDH and Vesicular Glu Accumulation 5931

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


FIG. 2. Association of 3-PGK with synaptic vesicles mediated
by GAPDH. A, various subcellular fractions as indicated were ana-
lyzed for 3-PGK by SDS-PAGE/Western blot as described in the Fig. 1
legend, except for use of anti-3-PGK antibodies instead of anti-GAPDH
antibody. B, the synaptic vesicle fraction was treated with various
concentrations of NaCl as indicated, and the salt extracts (supernatant
(Sup)) and the salt-insoluble material (Pellet) were analyzed for 3-PGK
by SDS-PAGE/Western blotting probed with anti-3-PGK antibodies.
The 0.2 M NaCl extract of synaptic vesicles was subjected to immuno-
precipitation with rabbit IgG or anti-3-PGK antibodies (C) and with FIG. 3. Dependence of vesicular Glu uptake on the GAPDH/3-
mouse IgG or the anti-GAPDH antibody (D); the immunoprecipitates PGK substrate. A, time course of vesicular Glu uptake induced by
were analyzed for GAPDH (C) and 3-PGK (D), respectively. E, synaptic activation of vesicle-bound GAPDH and 3-PGK. Glu uptake into syn-
vesicles (SV) were washed twice with 0.8 M NaCl and incubated in the aptic vesicles was measured at 30 °C for the indicated periods in the
absence or presence of GAPDH, each with or without 3-PGK at 37 °C for absence (‚) or presence (●) of 2 mM GAP, 2 mM NAD, 2 mM Pi, and 0.1
10 min, followed by two additional washings with low salt buffer. The mM ADP. B, GAPDH and 3-PGK substrate requirement for vesicular
synaptic vesicle pellets were then analyzed for GAPDH and 3-PGK. The Glu uptake. Glu uptake into synaptic vesicles was measured at 30 °C
data shown are representative of results obtained from three separate after a 10-min incubation in the absence or presence of various agents
experiments. or exogenous 3-PGK (1 unit) as indicated. C, effect of various ADP
concentrations on Glu uptake into synaptic vesicles induced by activa-
tion of vesicle-bound GAPDH and 3-PGK. Glu uptake was determined
Ci/mmol) [3-32P]1,3-BPG. The supernatant (180 ␮l) was subjected to
at 30 °C after 10-min incubation in the presence of 2 mM GAP, 2 mM
immunoprecipitation with immobilized antibody, according to the man- NAD, and 2 mM Pi. Values are the mean ⫾ S.D. of four separate
ufacturer’s protocol. Aliquots (20 ␮l) were subjected to SDS-PAGE, experiments.
except for omission of sample boiling, according to Fairbanks et al. (39)
or Laemmli (40), followed by Western blot analysis or autoradiography,
as appropriate. precipitated with 15% trichloroacetic acid and dissolved in 30 ␮l of SDS
Western Blot Analysis—For analysis of subcellular fractions, 30 ␮g of sample buffer. Aliquots (25 ␮l) of the 200,000 ⫻ gmax pellet and super-
protein were subjected to standard SDS-PAGE (40). In NaCl solubili- natant fractions were subjected to standard SDS-PAGE.
zation experiments, the synaptic vesicle fraction (50 ␮g of protein) was For analysis of GAPDH and 3-PGK binding to synaptic vesicles,
stirred in 0.2 ml of buffer containing 0.32 M sucrose, 4 mM Tris-maleate synaptic vesicles (1 mg of protein) were washed twice by 20 ml of buffer
(pH 7.4), and various concentrations of NaCl for 5 min at 4 °C and then containing 0.32 M sucrose, 4 mM Tris-maleate (pH 7.4), and 0.8 M NaCl;
centrifuged at 200,000 ⫻ gmax for 1 h. The pellet was dissolved in 30 ␮l bound NaCl was then removed by washing twice with 20 ml of the same
of SDS sample buffer. The protein in the supernatant (180 ␮l) was buffer without NaCl. The washed synaptic vesicles (40 ␮g of protein)
5932 GAPDH and Vesicular Glu Accumulation

FIG. 4. Effect of various GAP and


exogenous ATP concentrations on
vesicular Glu uptake and ATP levels
after incubation. Glu uptake into syn-
aptic vesicles (A and C) and ATP levels in
the synaptic vesicle suspension (B and D),
after 10-min incubation at 30 °C, were
measured in the presence of 2 mM NAD, 2
mM Pi, 0.1 mM ADP, and the indicated
concentrations of GAP (A and B) or of

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


various concentrations of exogenous ATP
(C and D). E, Glu uptake into synaptic
vesicles measured in the presence of var-
ious concentrations of GAP, 2 mM NAD, 2
mM Pi, and 0.1 mM ADP (●) or various
concentrations of exogenous ATP (‚) was
plotted as a function of ATP level in the
synaptic vesicle suspension, each deter-
mined after a 10-min incubation at 30 °C.
Values are the mean ⫾ S.D. of four sepa-
rate experiments.

were incubated at 37 °C for 10 min in 0.1 ml of buffer in the absence or incubated at 30 °C for 10 min with 100 ␮M [3H]Glu (a specific activity of
presence of 2 ␮g of purified GAPDH or 3-PGK. The synaptic vesicles 7.4 GBq/mmol was obtained by the addition of unlabeled Glu to
were pelleted by centrifugation at 200,000 ⫻ gmax for 1 h at 4 °C and [3H]Glu) in 0.1 ml of an incubation medium (pH 7.4) containing 20 mM
washed twice with the buffer. The pellet was suspended in 80 ␮l of SDS Hepes-KOH, 0.25 M sucrose, 4 mM MgSO4, 4 mM KCl, and 2 mM
sample buffer, and an aliquot (25 ␮l) was subjected to standard L-aspartic acid in the absence or presence of 2 mM ATP (pH adjusted to
SDS-PAGE. 7.4 by the addition of Tris-base). Prior to incubation, preincubation
Proteins were electrotransferred onto the Immobilon-P polyvinyli- without ATP and [3H]Glu was carried out for 1 min at 30 °C. When
dene difluoride membrane (Millipore Corp.) using a semidry transfer inhibitor effect was tested, test agents were added at the start of the
apparatus (Bio-Rad Trans-Blot SD). The membrane was treated with preincubation period. In some experiments, ATP was replaced with a
5% nonfat dry milk in a solution containing 50 mM Tris-HCl (pH 7.4), mixture of 2 mM GAP, 2 mM Pi, 2 mM NAD, and 0.1 mM ADP (pH
0.5 M NaCl, and 0.1% Tween 20 (TBS-T) for 1 h and then incubated for adjusted to 7.4 by the addition of Tris-base) as a source of a Glu uptake
2 h at room temperature with anti-GAPDH monoclonal antibody at a activator.
1:50 dilution or anti-3-PGK polyclonal antibodies at 1:500 dilution, Glu Content in Synaptic Vesicles within Synaptosomes—[3H]Glu con-
followed by incubation with alkaline phosphatase-conjugated goat anti- tent in synaptic vesicles within the synaptosome was assayed as de-
mouse IgG or anti-rabbit IgG, respectively, at room temperature for scribed previously (42, 43) with minor modifications. Synaptosomes
1.5 h. Unbound antibodies were washed out with TBS-T. 5-Bromo-4- (100 ␮g of protein) were suspended in 0.1 ml of oxygenated (95% O2, 5%
chloro-3-indolyl phosphate and nitro blue tetrazolium (Bio-Rad) were CO2) Krebs-Ringer buffer containing 150 mM NaCl, 2.4 mM KCl, 1.2 mM
used as substrates for color development and analyzed using an image Na2HPO4, 1.2 mM CaCl2, 1.2 mM MgSO4, 5 mM Hepes-Tris (pH 7.4), and
analyzer (Bio-Rad Gel Doc 2000). 10 mM glucose and preincubated at 37 °C for 10 min in the absence or
Glu Uptake into Synaptic Vesicles—Vesicular glutamate uptake was presence of iodoacetate, oligomycin (Calbiochem), and pyruvate at in-
measured by the filtration-based assay using Whatman GF/C filters, as dicated concentrations. After 3 ␮Ci of [3H]Glu (42.0 Ci/mmol) were
described previously (27, 28), with minor modifications. In the standard added to the medium, synaptosomes were incubated for an additional
assay, aliquots (10 ␮g of protein) of bovine synaptic vesicles were 10 min. Aliquots (10 ␮l) were removed and filtered on Whatman GF/C
GAPDH and Vesicular Glu Accumulation 5933
filters to determine the total amount of [3H]Glu taken up by the syn-
aptosomes, and the rest were immediately frozen on dry ice. For vesic-
ular [3H]Glu content determination, the frozen synaptosomes (90 ␮l)
were thawed by adding 1.5 ml of ice-cold hypotonic solution containing
6 mM Tris-maleate (pH 8.1) and 2 mM aspartate and incubated for 20
min at 0 °C. Aliquots (1 ml) were filtered on Whatman GF/C filters, and
radioactivity retained on filters was determined in a Beckman LS 6500
scintillation spectrophotometer.
Glu Release from Synaptosomes—Glutamate release from synapto-
somes was assayed using the superfusion technique described previ-
ously (43), with minor modifications. Synaptosomes (200 ␮g of protein)
were suspended in 1.5 ml of oxygenated (95% O2/5% CO2) artificial
cerebrospinal fluid (ACSF) containing 124 mM NaCl, 5 mM KCl, 2 mM
MgSO4, 1.25 mM NaH2PO4, 22 mM NaHCO3, and 10 mM D-glucose and
preincubated with or without 300 ␮M iodoacetate or 2 ␮M oligomycin at
37 °C for 10 min. After 3.5 ␮Ci of [3H]Glu (42.0 Ci/mmol) were added to
the medium, synaptosomes were incubated for an additional 10 min. An
aliquot (1.0 ml) of [3H]Glu-loaded synaptosomal suspension was layered
onto a cellulose-acetate membrane filter (pore size 0.45 ␮m) placed in a
superfusion chamber. The synaptosomes were superfused (0.5 ml/min)
with ACSF for 60 min before application of 50 ␮M 4-aminopyridine
(4-AP) plus 2 mM CaCl2 to trigger depolarization of the synaptosomal
membrane. In some control experiments, synaptosomes were super-
fused with ACSF containing 300 ␮M iodoacetate or 2 ␮M oligomycin for
the first 20 min of the superfusion period. This period was the same as
the sum of the synaptosomal preincubation and [3H]Glu-loading peri-

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


ods. These procedures were all carried out at 37 °C. Fractions were
collected every 10 s for 3 min, from 30 s prior to 4-AP application. The
amount of [3H]Glu released into each fraction was expressed as a
percentage of the total [3H]Glu taken up into synaptosomes at the end
of 10 min of [3H]Glu loading. The total [3H]Glu loaded into synapto-
somes was calculated from the amount of [3H]Glu in 0.1 ml of loaded
synaptosomes, determined by the same filtration method used for the
vesicular Glu uptake assay.
Other Biochemical Analyses—GAPDH activity was measured using
20 ␮g of protein of synaptic vesicles in a reaction mixture (1 ml)
containing 0.1 M Tris-HCl (pH 8.5), 1.7 mM sodium arsenate, 20 mM
sodium fluoride, 1 mM NAD, 1 mM GAP, and 5 mM KH2PO4 (44).
Activity was calculated from the rate of increase in NADH formation by
monitoring absorbance at 340 nm at 25 °C.
ATP levels in the synaptic vesicle and synaptosome suspensions were
FIG. 5. Iodoacetate selectively inhibits GAPDH/3-PGK activa-
determined under the same conditions described for vesicular Glu up- tion-dependent over exogenous ATP-dependent vesicular Glu
take and vesicular Glu content assays, respectively, by the luciferin/ uptake. A, effect of various iodoacetate concentrations on GAPDH/3-
luciferase method, using an ATP bioluminescent assay kit (Sigma- PGK activation-dependent versus exogenous ATP-dependent vesicular
Aldrich), according to the manufacturer’s protocol. Glu uptake. The synaptic vesicle fraction (10 ␮g of protein) was prein-
Generation of the membrane potential across the synaptic vesicle cubated at 30 °C for 10 min in the presence of various concentrations of
membrane was monitored by ATP-induced fluorescence quenching of iodoacetate, followed by an additional 10-min incubation at 30 °C after
the membrane potential-sensitive dye oxonol V (Molecular Probes, Inc., the addition of a mixture of [3H]Glu, 2 mM GAP, 2 mM NAD, 2 mM Pi,
Eugene, OR) using a Fluorolog III fluorospectrophotometer (Horiba and 0.1 mM ADP (●) or of a mixture of [3H]Glu and 2 mM ATP (‚).
Jobin Yvon Co., Ltd., Tokyo, Japan) as described previously (32, 43). [3H]Glu accumulated into synaptic vesicles was determined as de-
Vesicular ATPase activity was assayed by determining free inorganic scribed under “Experimental Procedures.” B, effect of various iodoac-
phosphate liberated upon incubation of synaptic vesicles (30 ␮g of etate concentrations on vesicle-bound GAPDH activity and on ATP
protein) with ATP, according to the method of Lanzetta et al. (45), with produced by activation of vesicle-bound GAPDH/3-PGK. The synaptic
minor modifications as described previously (43). vesicle fraction (10 ␮g of protein) was incubated in the presence of
various concentrations of iodoacetate for a total period of 20 min as
RESULTS described for A, except for replacement of [3H]Glu with nonradioactive
Glu. After incubation, GAPDH activity (●) and the amount of ATP
A 37-kDa Protein Enriched in Synaptic Vesicles Is Labeled produced (䡺) were determined, as described under “Experimental Pro-
with [3-32P]1,3-BPG and Identified as GAPDH—Purified syn- cedures.” Values are the mean ⫾ S.D. of four separate experiments.
aptic vesicles were allowed to react with [3-32P]1,3-BPG for 10 s
and subjected to SDS-PAGE at either neutral (Fig. 1A) or proteins might be GAPDH and PGM, respectively. This was
alkaline (Fig. 1B) pH (39, 40). Electrophoresis at neutral pH confirmed by immunoprecipitation with antibodies directed
revealed incorporation of a radioactive moiety of [3-32P]1,3- against these proteins (Fig. 1C). Thus, labeling of the 37-kDa
BPG into vesicular proteins of Mr ⫽ 37,000 and 29,000. In GAPDH with [3-32P]1,3-BPG probably occurs by formation of a
contrast, electrophoresis at alkaline pH revealed only the 29- thioester bond between a cysteine residue and the 3-phospho-
kDa protein, which appears to retain more of the 32P label than glyceroyl moiety of 1,3-BPG; thioester bonds are known to be
when electrophoresis was conducted at neutral pH. These re- labile at alkaline pH. Labeling of the 29-kDa PGM probably
sults indicate that both the 29- and 37-kDa vesicular proteins represents phosphorylation of a histidine residue that is labile
can incorporate a 32P-containing moiety; however, the stability at either neutral or acidic pH.
of the labeled proteins differs depending on pH. Incorporation In order to determine the subcellular distribution of GAPDH
of the radioactive moiety into the 37-kDa protein was stimu- and PGM, we examined the amount of [3-32P]1,3-BPG radioac-
lated by Mg2⫹ but not affected by 3-PG. Labeling of the 29-kDa tivity incorporated into these two proteins present in synapto-
protein was inhibited by Mg2⫹ and completely blocked by 3-PG. somal cytosol, perikaryal cytosol, microsome, synaptic vesicle,
The linkage of the 32P-containing moiety to the 37-kDa protein and plasma membrane fractions (Fig. 1D). For each subcellular
was labile in the presence of 0.2 M hydroxylamine (data not fraction, an equivalent amount of protein was incubated with
shown). These observations suggested that the 37- and 29-kDa [3-32P]1,3-BPG. Labeled GAPDH was found in all subcellular
5934 GAPDH and Vesicular Glu Accumulation

FIG. 6. Comparison in responsive-


ness to chloride, aspartate, t-ACPD,
Rose Bengal, and FCCP between
GAPDH/3-PGK activation-dependent
and exogenous ATP-dependent vesic-
ular Glu uptake. The effect of various
chloride concentrations on Glu uptake
into synaptic vesicles was determined at
30 °C after a 10-min incubation in the
presence of 2 mM GAP, 2 mM NAD, 2 mM
Pi, and 0.1 mM ADP (A) or of 2 mM ATP
(B). Effect of 2 mM Asp, 2 mM t-ACPD, 0.5
␮M Rose Bengal (RB), and 25 ␮M FCCP on
Glu uptake into synaptic vesicles was de-
termined at 30 °C after 10-min incubation
in the presence of 2 mM GAP, 2 mM NAD,
2 mM Pi, and 0.1 mM ADP (C) or of 2 mM
ATP (D). Values are the mean ⫾ S.D. of
four separate experiments.

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


fractions but was enriched in the synaptic vesicle fraction. In noprecipitates, indicating that GAPDH and 3-PGK must exist
contrast, PGM was most enriched in the synaptosomal cytosol. in a tight complex. A second co-immunoprecipitation experi-
Subcellular fractions were also subjected to Western blotting ment was carried out, which showed that 3-PGK can be co-
with anti-GAPDH antibody. GAPDH was found in the greatest precipitated with anti-GAPDH antibody (Fig. 2D). In order to
concentration in the synaptic vesicle fraction (Fig. 1E), in ac- determine whether GAPDH and 3-PGK directly bind to the
cord with the results obtained with the labeling method. Anal- synaptic vesicles, synaptic vesicles were first washed with 0.8 M
ysis of subcellular fractions indicates that GAPDH not only NaCl and then incubated with purified GAPDH and/or 3-PGK.
occurs in the cytosol but also can bind to various types of The vesicles were pelleted by centrifugation, and the amount of
membranes. To determine the nature of the interaction of vesicle-bound GAPDH or 3-PGK was determined by immuno-
GAPDH with synaptic vesicles, synaptic vesicles were treated blotting (Fig. 2E). GAPDH was found capable of binding to the
with various concentrations of NaCl (Fig. 1F). GAPDH was synaptic vesicle with or without the addition of 3-PGK. In
dissociated from synaptic vesicles with increasing salt concen- contrast, binding of 3-PGK to synaptic vesicles was greatly
trations. At 0.8 M NaCl, only about 10% of GAPDH was found increased when GAPDH was included with 3-PGK. These re-
to remain associated with synaptic vesicle membranes. At sults suggest that GAPDH directly binds to synaptic vesicles
physiologic ionic strength, GAPDH is associated with synaptic and that 3-PGK binds to vesicle-bound GAPDH.
vesicle membranes, although it does also occur in the cytosol Synaptic Vesicles Are Capable of Accumulating Glu via Ac-
fractions. tivation of Endogenous GAPDH/3-PGK—Synaptic vesicles
3-PGK Is Associated with Synaptic Vesicles via Interaction bearing both GAPDH and 3-PGK should be able to synthesize
with GAPDH—In the glycolytic pathway, GAPDH is known to ATP when GAPDH substrates and ADP are supplied. GAPDH
exist in a complex with 3-PGK (46). The activities of GAPDH and 3-PGK activity could regenerate ATP for vesicular uptake
and 3-PGK are energetically coupled, utilizing GAP, NAD, Pi, of Glu. Synaptic vesicles were incubated for various periods
and ADP, to yield 3-PG, ATP, NADH, and H⫹ (46, 47). Because with [3H]Glu in the presence or absence of 2 mM GAP, NAD,
GAPDH is enriched in synaptic vesicles, we considered the and Pi and 0.1 mM ADP. Vesicular Glu uptake was dependent
possibility that 3-PGK may also preferentially localize to syn- on the presence of the GAPDH/3-PGK substrate mixture
aptic vesicle membranes. Western blots were conducted on throughout the incubation period(s) tested and was approxi-
subcellular fractions using anti-3-PGK antibody (Fig. 2A). Like mately linear up to 6 min under these conditions (Fig. 3A). When
GAPDH, 3-PGK was found enriched in the synaptic vesicle any one of the components was omitted from the incubation
fraction. Immunoreactive 3-PGK was also dissociated from syn- mixture, vesicular Glu uptake was not supported (Fig. 3B).
aptic vesicles by increasing concentrations of NaCl in a manner [3H]Glu uptake resulting from ATP generated by vesicular
similar to that observed for GAPDH (Fig. 2B). To determine GAPDH and 3-PGK was comparable with that observed with 2
whether GAPDH and 3-PGK exist in a complex on synaptic mM ATP and not increased by the addition of exogenous 3-PGK.
vesicles, we performed experiments to determine whether ADP supported uptake of [3H]Glu in a concentration-dependent
GAPDH and 3-PGK could be co-immunoprecipitated. Synaptic manner but was inhibitory at concentrations above 0.1 mM (Fig.
vesicle salt extracts were incubated with anti-3-PGK antibody, 3C). These results indicate that vesicular Glu uptake can occur
and the resulting immunoprecipitates were subjected to West- by harnessing the activity of vesicle-bound GAPDH and 3-PGK.
ern blotting with anti-GAPDH antibody (Fig. 2C). Immunore- ATP Synthesized by Vesicle-bound GAPDH/3-PGK Is More
active GAPDH was detected in the anti-3-PGK antibody immu- Effective than Exogenous ATP in Glu Uptake—Experiments
GAPDH and Vesicular Glu Accumulation 5935

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


FIG. 8. ATP-dependent vesicular Glu uptake is inhibited by
FIG. 7. Effect of GAPDH substrates on vesicular Glu uptake in ADP, which has no effect on vesicular ATPase activity. A, effect
the presence of ATP. Glu uptake into isolated synaptic vesicles at of various concentrations of ADP (●), AMP (䡺), or adenosine (‚) on Glu
30 °C after 10 min incubation in the presence of 30 ␮M ATP (A) or 2 mM uptake into synaptic vesicles at 30 °C after 10 min incubation was
ATP (B) was measured in the absence or presence of GAPDH substrates determined in the presence of 2 mM ATP. B, effect of various ADP
(2 mM GAP, 2 mM NAD, and 2 mM Pi) as described under “Experimental concentrations on vesicular ATPase activity was determined as de-
Procedures.” Values are the mean ⫾ S.D. of four separate experiments. scribed under “Experimental Procedures.” Values are the mean ⫾ S.D.
of four separate experiments.
shown in Fig. 3 suggested that GAP-mediated vesicular Glu
uptake probably occurs as a result of ATP synthesis. To directly Glu accumulated per ATP level after 10 min of incubation is
demonstrate this, synaptic vesicles were incubated with vari- greater when GAPDH/3-PGK substrates are utilized. These
ous concentrations of GAP in the presence of 2 mM NAD, Pi, and results suggest that the ATP generated by vesicle-bound
0.1 mM ADP, and the amount of ATP produced, as well as Glu GAPDH/3-PGK is more effective than exogenous ATP in sup-
uptake, was determined (Fig. 4, A and B). ATP was synthesized porting vesicular Glu uptake.
in a GAP-concentration-dependent manner in parallel to vesic- GAP-dependent Vesicular Glu Uptake Is Inhibited by Iodoac-
ular Glu uptake. Under these conditions, 12 ␮M ATP were etate—In an effort to provide further evidence that GAP-de-
sufficient to give rise to maximal vesicular Glu uptake. This pendent Glu uptake involves GAPDH, we have determined the
concentration is substantially lower than the apparent Km effect of the GAPDH-selective inhibitor iodoacetate on vesicu-
value for ATP (0.8 –1.2 mM), determined using exogenous ATP lar Glu uptake in the presence of GAP, NAD, Pi, and ADP. As
under standard assay conditions for vesicular Glu uptake (28) shown in Fig. 5A, iodoacetate inhibited vesicular GAP-depend-
(see Fig. 9B). These observations raised the possibility that ent Glu uptake in a concentration-dependent manner. When
ATP produced by vesicle-bound GAPDH/3-PGK might be more ATP was substituted for the GAPDH/3-PGK substrates, iodoac-
effective than exogenous ATP in supporting vesicular Glu up- etate exhibited little inhibition in the concentration range
take. To further explore this possibility, exogenous ATP-de- tested. Iodoacetate also inhibited GAPDH activity and im-
pendent vesicular Glu uptake was compared with GAP-sup- paired the ability of synaptic vesicles to regenerate ATP from
ported Glu uptake. Synaptic vesicles were incubated with ADP in a concentration-dependent manner. These results in-
various concentrations of exogenous ATP, and the amount of dicate that iodoacetate prevents, via inhibition of GAPDH, ATP
ATP, as well as Glu uptake, after 10 min of incubation was generation necessary for vesicular Glu uptake.
determined (Fig. 4, C and D). At least 80 –100 ␮M exogenous Characterization of the GAPDH/3-PGK Activation-dependent
ATP were needed to support the same maximal Glu uptake Vesicular Glu Uptake System—The ATP-dependent vesicular
observed for GAP-supported uptake. This ATP concentration is Glu uptake system is known to be markedly stimulated by
much higher than that observed with the endogenous ATP- low millimolar chloride (28, 34, 48 –50) and harnesses an
generating system. In Fig. 4E, the amount of Glu taken up into electrochemical proton gradient as the driving force (28, 29 –
synaptic vesicles was plotted as a function of the ATP concen- 34). In contrast to Na⫹-dependent plasma membrane Glu
tration determined after a 10-min incubation. The amount of uptake, vesicular Glu uptake is insensitive to aspartate (28,
5936 GAPDH and Vesicular Glu Accumulation

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


FIG. 9. Kinetics of ADP inhibition of vesicular Glu uptake. A,
the initial rate of vesicular Glu uptake at 30 °C in the presence of 2 mM
ATP was determined with various Glu concentrations, each in the
absence (E) or presence of 0.5 mM (Œ) or 2 mM ADP (f), by incubating
the synaptic vesicle fraction (10 ␮g of protein) for 1.5 min, as described
under “Experimental Procedures.” B, the initial rate of vesicular Glu
uptake with a fixed Glu concentration (100 ␮M) was determined with
various ATP concentrations in the absence (E) or presence of 0.5 mM (Œ)
or 2 mM ADP (f), as described above. Values are the mean ⫾ S.D. of four FIG. 10. Effect of ADP on membrane potential in synaptic ves-
separate experiments. icles. After oxonol V (1.3 ␮M) was allowed to equilibrate with synaptic
vesicles as described under “Experimental Procedures,” ATP was added
(2 mM as a final concentration). Various amounts of ADP were then
31, 34, 48, 50) but potently inhibited by the membrane- added, resulting in indicated final concentrations, followed by the ad-
permeant polyhalogenated fluorescein Rose Bengal (43). t- dition of FCCP (8.33 ␮M as a final concentration). A, time course tracing
ACPD also inhibits vesicular Glu uptake by serving as a of fluorescence quenching before and after the addition of 0.3, 1, and 2
substrate for the vesicular Glu transporter (51). mM ADP. The data shown are representative of three separate experi-
ments. B, membrane potential as a function of various concentrations
We have studied the characteristics of GAPDH/3-PGK acti- of ADP as indicated. Values are the mean ⫾ S.D. of three separate
vation-induced vesicular Glu uptake with respect to the prop- experiments.
erties of the ATP-dependent vesicular Glu uptake system.
When vesicle-bound GAPDH and 3-PGK were activated in the by vesicle-bound GAPDH/3-PGK appeared more effective than
presence of their substrates and various concentrations of chlo- exogenous ATP in supporting vesicular Glu uptake, it was of
ride, maximal Glu uptake occurred at 4 mM chloride; at higher interest to see whether GAPDH substrates are effective in
chloride concentration, Glu uptake was attenuated (Fig. 6A). augmenting vesicular Glu uptake in the presence of limiting
This dependence on chloride is indistinguishable from that concentrations of ATP. Fig. 7 shows the effect of GAPDH sub-
observed for exogenous ATP-induced vesicular Glu uptake (Fig. strates on vesicular Glu uptake in the presence of ATP. The
6B). Moreover, GAPDH/3-PGK activation-induced vesicular addition of a mixture of GAP, NAD, and Pi, which alone exhib-
Glu uptake was insensitive to Asp but affected by t-ACPD and ited no Glu uptake, elevated Glu uptake in the presence of a
Rose Bengal as well as by the proton ionophore FCCP, in a low concentration of exogenous ATP (Fig. 7A). This suggests
manner similar to that observed with exogenous ATP-depend- that ADP generated by ATPase is recycled to ATP by 3-PGK,
ent Glu uptake (Fig. 6, C and D). These results indicate that when vesicle-bound GAPDH is activated, and that the regen-
the Glu transporter responsible for GAPDH/3-PGK activation- erated ATP is utilized to give rise to additional Glu uptake. At
induced vesicular Glu uptake is the same as that responsible a much higher ATP concentration, this augmentation was
for ATP-dependent vesicular Glu uptake. much smaller, however (Fig. 7B). This could be partly attrib-
GAPDH Substrates Enhance Vesicular Glu Uptake in the uted to saturation of proton pump ATPase with ATP. These
Presence of a Low Concentration of ATP—Since ATP generated results are compatible with the notion that GAPDH/3-PGK-
GAPDH and Vesicular Glu Accumulation 5937

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


FIG. 12. Effect of pyruvate, glucose, oligomycin, and iodoac-
etate on vesicular [3H]Glu content and ATP levels in synapto-
somes. Synaptosomes were preincubated at 37 °C for 10 min in the
absence or presence of 10 mM glucose, 2 ␮M oligomycin, 300 ␮M iodoac-
etate, or 1 mM pyruvate; ATP level (A) and vesicular [3H]Glu content (B)
FIG. 11. Iodoacetate reduces vesicular [3H]Glu content in syn- were determined at 37 °C after an additional 10-min incubation with
aptosomes more effectively than oligomycin, whereas both [3H]Glu, as described in the legend to Fig. 11. Vesicular [3H]Glu content
agents have similar effects on synaptosomal ATP levels in the was expressed as percentage of total synaptosomal [3H]Glu uptake.
concentration ranges tested. Synaptosomes were preincubated at Values are the mean ⫾ S.D. of four separate experiments.
37 °C for 10 min in the presence of various concentrations of iodoacetate
(A, C, E, and G) or oligomycin (B, D, F, and H), followed by the addition in the presence of 0, 0.5, and 2.0 mM ADP were 1.6, 1.6, and 1.5
of [3H]Glu; ATP level (A and B), synaptosomal [3H]Glu uptake (C and
D), and vesicular [3H]Glu content (E and F) were determined at 37 °C
mM, respectively, and those for ATP in the presence of 0, 0.5,
after 10 min, as described under “Experimental Procedures.” Vesicular and 2.0 mM ADP were 0.82, 0.76, and 0.79 mM, respectively.
[3H]Glu content was expressed as percentage of total synaptosomal The noncompetitive inhibition with respect to ATP was rather
[3H]Glu uptake (G and H). Values are the mean ⫾ S.D. of four separate surprising but was consistent with ADP’s inability to inhibit
experiments.
ATPase activity (Fig. 8B).
The effect of ADP on membrane potential was examined in
mediated vesicular Glu uptake occurs through the ATP-de- an effort to gain further insight into the mechanism of ADP
pendent vesicular Glu uptake system. inhibition. Fig. 10A shows that ADP is capable of dissipating
ADP Inhibits ATP-dependent Vesicular Glu Uptake—In or- preformed membrane potential in a concentration-dependent
der to understand the ADP inhibition observed earlier (Fig. manner. Disruption of membrane potential by 10 mM ADP was
3C), we have studied the effect of ADP on vesicular Glu uptake 47% (Fig. 10B). ADP did not completely dissipate membrane
supported by exogenous ATP. It was found that exogenous potential at the concentrations tested. The ability of ADP to
ATP-dependent vesicular Glu uptake was also inhibited by reduce membrane potential would not entirely account for its
ADP, but not by AMP or adenosine, at concentrations higher inhibitory effect on Glu uptake.
than 0.1 mM (Fig. 8A). ADP had no effect on ATPase activity at ATP Produced by Glycolysis Is Predominantly Utilized for
the concentrations tested up to 10 mM (Fig. 8B). Kinetic exper- Vesicular Accumulation of Glu in the Nerve Ending—The data
iments indicate that this inhibition is noncompetitive with indicating that ATP produced by vesicular GAPDH and 3-PGK
respect to Glu or ATP (Fig. 9, A and B). Thus, neither Km for can support vesicular uptake of Glu prompted us to postulate
Glu nor Km for ATP was altered significantly; Km values for Glu that Glu accumulation into synaptic vesicles in the nerve ter-
5938 GAPDH and Vesicular Glu Accumulation

FIG. 14. Proposed role of GAPDH and 3-PGK in vesicular Glu


uptake. The glycolytic intermediate GAP is converted to 1,3-BPG by
synaptic vesicle-associated GAPDH in the presence of NAD and Pi.
1,3-BPG is then converted to 3-PG by GAPDH-associated 3-PGK, ac-
companied by phosphorylation of ADP to form ATP. ATP thus produced
FIG. 13. Treatment of synaptosomes with iodoacetate prior to is efficiently consumed by vesicular H⫹-ATPase, generating an electro-
and during the [3H]Glu-loading period reduces [3H]Glu release. chemical proton gradient in the presence of low concentrations of chlo-
Synaptosomes were preincubated at 37 °C for 10 min in the absence (E, ride; this provides the driving force for Glu uptake, which is carried out
●, 䡺, and ‚) or presence of 300 ␮M iodoacetate (IA; f) or 2 ␮M oligo- by the vesicular Glu transporter (VGLUT). ADP generated by vesicular
mycin (OM; Œ); [3H]Glu was loaded into the synaptosome at 37 °C for 10 H⫹-ATPase is efficiently converted to ATP by 3-PGK. When ADP is
min. The [3H]Glu-loaded synaptosomes were layered onto the mem- accumulated, Glu uptake into synaptic vesicles would be inhibited. ⌬⌿,
brane in the superfusion chamber and superfused at 37 °C for 20 min membrane potential; ⌬pH, pH gradient.

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


with ACSF in the absence (E, ●, f, and Œ) or presence of 300 ␮M
iodoacetate (䡺) or 2 ␮M oligomycin (‚). After further superfusion with mitochondria. However, pyruvate failed not only to enhance
ACSF at 37 °C for 40 min, synaptosomes were subjected to depolariza- vesicular Glu accumulation in the absence of exogenous glucose
tion with 50 ␮M 4-aminopyridine (4-AP; ●, f, Œ, 䡺, ‚) or not (E); but also to overcome the reduction of vesicular Glu content
[3H]Glu release was measured as described under “Experimental Pro-
cedures.” Values are the mean ⫾ S.D. of six separate experiments.
caused by iodoacetate (Fig. 12B). Oligomycin blocked pyruvate-
induced increases in ATP levels (Fig. 12A) but had no signifi-
cant effect on vesicular Glu content corrected for synaptosomal
minal may be supported by glycolytically produced ATP. In
uptake (Fig. 12B). In the presence of glucose, pyruvate failed to
order to test this hypothesis, we compared the effects of the
elevate ATP levels or vesicular Glu content, either in the ab-
GAPDH inhibitor iodoacetate and the mitochondrial ATP syn-
sence or presence of oligomycin. These results indicate that
thesis inhibitor oligomycin on vesicular accumulation of Glu in
ATP produced by glycolysis is predominantly utilized, whereas
purified synaptosomes. When GAPDH was inhibited by iodoac-
mitochondria-generated ATP is insufficient to support vesicu-
etate, or mitochondrial ATP synthesis was inhibited by oligo-
lar uptake of Glu in nerve endings. These observations support
mycin, ATP levels were reduced in a concentration-dependent
the notion that synaptic vesicle-bound GAPDH and 3-PGK play
manner (Fig. 11, A and B). The reduction in synaptosomal ATP
an important role by providing ATP necessary for Glu uptake
caused by 500 ␮M iodoacetate and 2 ␮M oligomycin was 67 and
into synaptic vesicles in the nerve terminal.
64%, respectively. Both reagents decreased synaptosomal
Glu Accumulated into Vesicles, at the Expense of Glycolyti-
[3H]Glu uptake to a similar extent (Fig. 11, C and D). Inhibition
cally Produced ATP, Is Released upon Depolarization—In order
of synaptosomal [3H]Glu uptake by 500 ␮M iodoacetate and 2
to determine whether the Glu accumulated into synaptic vesi-
␮M oligomycin was 33 and 37%, respectively. In contrast, io-
cles by glycolysis-generated ATP is released upon nerve ending
doacetate was more effective than oligomycin in reducing ve-
stimulation, we examined depolarization-induced release in
sicular [3H]Glu content (Fig. 11, E and F). When vesicular
synaptosomes. Synaptosomes treated with either iodoacetate
content was corrected for synaptosomal uptake (Fig. 11, G and
or oligomycin, prior to (and during) loading with [3H]Glu, were
H), the inhibition of vesicular uptake within the synaptosomes
depolarized by the potassium channel blocker 4-AP. As shown
by 500 ␮M iodoacetate was 78%, whereas the inhibition by 2 ␮M
in Fig. 13, iodoacetate treatment led to substantial reduction in
oligomycin was only 16%. Although the reduction of [3H]Glu
the amount of [3H]Glu released. In contrast, oligomycin pre-
content by iodoacetate is probably due to inhibition of GAPDH,
treatment had a much smaller effect. The reduction in depo-
the possibility is not entirely ruled out that iodoacetate could
larization-induced Glu release is similar to the reduction of
inhibit vesicular Glu uptake within the synaptosomes by a
vesicular Glu content observed with each inhibition (Fig. 11, G
mechanism not involving GAPDH inhibition. However, iodoac-
and H). When synaptosomes were treated with iodoacetate or
etate did not inhibit [3H]Glu uptake by purified synaptic vesi-
oligomycin after loading with [3H]Glu, release was affected
cles in the presence of ATP (Fig. 5A).
only to a small extent. The magnitude of this reduction was
Inhibition of GAPDH by iodoacetate would lead to reduction
similar to that observed when synaptosomes were treated with
of pyruvate, the precursor for the tricarboxylic acid cycle sub-
oligomycin before loading with [3H]Glu.
strate acetyl-CoA and thereby would decrease ATP production
The large differential of [3H]Glu release observed between
in mitochondria. To better understand the mechanism of vesic-
iodoacetate- and oligomycin-treated synaptosomes indicates
ular Glu accumulation, we further investigated the effect of
that [3H]Glu accumulated in synaptic vesicles utilizing ATP
pyruvate on vesicular Glu uptake in synaptosomes in the ab-
synthesized via glycolysis can be released by depolarization.
sence or presence of iodoacetate or oligomycin. Pyruvate led to
Thus, the effect of GAPDH inhibition by iodoacetate on presyn-
an increase in ATP levels when added to synaptosomes incu-
aptic vesicular Glu accumulation is reflected in a decrease of
bated without glucose or to iodoacetate-treated synaptosomes
Glu released by exocytosis.
incubated with glucose (Fig. 12A). Increased levels of ATP
resulting from incubation with pyruvate were not observed in DISCUSSION
the presence of oligomycin. These results are consistent with We have presented evidence that GAPDH is modified by the
the concept that pyruvate-induced ATP production occurs in high energy glycolytic intermediate 1,3-BPG via covalent at-
GAPDH and Vesicular Glu Accumulation 5939
tachment of the 3-phosphoglyceroyl moiety. The radioactive tion of refilling Glu would be more efficiently served if ATP
moiety was removed by treatment with neutral hydroxylamine. were generated “locally” by vesicle-bound GAPDH/3-PGK. In
This suggests that this modification involves thioester forma- isolated synaptic vesicles, GAPDH/3-PGK-generated ATP ap-
tion between the 3-phosphoglyceroyl group and a GAPDH cys- pears more effective than exogenous ATP in supporting vesic-
teine residue, most likely representing the acylated enzyme ular Glu uptake (Fig. 4). In neurons, mitochondria located in
intermediate (52, 53). the nerve ending may not be close enough to synaptic vesicles
GAPDH was found to be bound to purified synaptic vesicles. to provide sufficient ATP in a timely fashion for sustaining
This observation is in agreement with the findings by Schlaefer adequate Glu accumulation. Moreover, ATP synthesis in mito-
et al. (54) and Rogalski-Wilk and Cohen (55). Schlaefer et al. chondria occurs after ATP formation during glycolysis. Thus,
have postulated that vesicle-associated GAPDH might be func- the synaptic vesicle-bound GAPDH-3-PGK system would have
tionally coupled to ATP uptake into synaptic vesicles (54). We an advantage over mitochondrial ATP synthase in its prompt
have provided evidence that synaptic vesicle-bound GAPDH, response to energy demand for refilling vesicles with the trans-
coupled with 3-PGK, can produce ATP sufficient for Glu uptake mitter Glu.
by the vesicular Glu uptake system. In the presence of the In conclusion, we propose that the ATP produced by vesicle-
glycolytic intermediate GAP, GAPDH produces 1,3-BPG, which bound GAPDH and 3-PGK plays a crucial role in swiftly refill-
is then converted to 3-PG and ATP by the GAPDH-bound ing vesicles with the major excitatory neurotransmitter Glu
3-PGK. This conversion is coupled, since GAPDH and 3-PGK (and possibly other neurotransmitters) in the presynaptic
exist in a complex (Fig. 2). Bernhard and co-workers (46, 47) nerve terminal. According to our hypothesis depicted in Fig. 14,
have demonstrated direct transfer of 1,3-BPG from GAPDH to GAP produced via glucose metabolism is converted to 1,3-BPG
3-PGK in muscle tissue. Inhibition of vesicle-bound GAPDH at the expense of NAD and Pi by synaptic vesicle-bound
activity by iodoacetate led to a reduction of ATP synthesis as GAPDH; 1,3-BPG transfers its high energy phosphate to ADP,
well as to an attenuation of Glu uptake into synaptic vesicles catalyzed by GAPDH-bound 3-PGK, forming ATP at the sur-
face of vesicles. This ATP is utilized by vesicular H⫹-pump

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


(Fig. 5). This was also observed in synaptosomes (Figs. 11 and
12). Iodoacetate decreased the amount of exocytotically re- ATPase to generate an electrochemical proton gradient across
leased Glu (Fig. 13). Thus, it is likely that, in synaptosomes, the the vesicle membrane, providing the driving force for Glu
ATP generated from the glycolytic intermediates provides the transport into synaptic vesicles. ADP produced by ATPase
energy required for the majority of the Glu taken up into would be rapidly recycled to ATP by the GAPDH-3PGK system
synaptic vesicles. Oligomycin inhibition of mitochondrial ATP in the presence of glycolysis, eliminating its potential inhibi-
synthesis had little effect on vesicular Glu accumulation in tory effect on Glu uptake. Acute depletion of glucose could lead
to a reduction of ATP produced via glycolysis within the micro-
synaptosomes (Figs. 11 and 12). Moreover, substitution of
environment of the presynaptic nerve terminal. This would
pyruvate for glucose, which increased ATP levels in synapto-
diminish Glu transport into the vesicle, resulting in a de-
somes, failed to elevate vesicular Glu content (Fig. 12). These
creased amount of Glu released, thereby attenuating Glu-me-
observations indicate that mitochondrially produced ATP con-
diated synaptic transmission. Moreover, in the prolonged ab-
tributes minimally to Glu transport into synaptic vesicles in
sence of glycolysis, excess ADP would be accumulated, im-
synaptosomes, suggesting the critical importance of glycolyti-
pairing the ATP-dependent vesicular uptake system, further
cally produced ATP in vesicular Glu accumulation in the pre-
contributing to reduction in vesicular Glu content and release.
synaptic nerve terminal.
The mechanism outlined above could offer insight into our
In cardiac myocytes, there is evidence that glycolysis-derived
molecular understanding of hypoglycemia-induced, abnormal
ATP is preferentially used to prevent ATP-sensitive K⫹ chan-
electrophysiological activity and resulting clinical symptoms as
nels from opening, thereby maintaining cellular membrane
well as into the essential requirement for glucose metabolism
potential (56). In the skeletal muscle, evidence suggests that
in normal synaptic transmission.
fast twitch fibers of the white vastus utilize ATP produced by
glycolysis, whereas the slow twitch fibers of the soleus largely Acknowledgments—We are grateful to Dr. Kiyokazu Ogita (Setsunan
use ATP synthesized by mitochondria. The former has a high University, Hirakata, Japan) for technical advice in the superfusion
release assay, Dr. Koji Hirata for synaptic vesicle preparation, and
glycolytic capacity and a low oxidative capacity, in contrast to
Oriental Yeast Co., Ltd., Tokyo, for the kind gift of affinity-purified
the latter, which is opposite (57). GAPDH activity is 3– 4 times polyclonal antibodies to recombinant PGM type B. We also thank Mary
higher in the white quadriceps than in the soleus (58). GAPDH Roth for excellent assistance in preparation of the manuscript.
is a component associated with the skeletal muscle junctional
“foot,” which links the transverse tubule and terminal cister- REFERENCES
nae, forming the triad, the site of excitation-contraction cou- 1. Siesjo, B. K. (1978) Brain Energy Metabolism, pp. 101–130, John Wiley & Sons,
Inc., New York
pling (59, 60). Heilmeyer et al. (60) have demonstrated that the 2. McNay, E. C., Fries, T. M., and Gold, P. E. (2000) Proc. Natl. Acad. Sci. U. S. A.
junctional triad is capable of producing ATP from GAP, NAD, 97, 2881–2885
3. Lewis, L. D., Ljunggren, B., Ratcheson, R. A., and Siesjo, B. K. (1974) J. Neu-
Pi, and ADP, indicating that localized glycolytic ATP synthesis rochem. 23, 673– 679
can occur at the site where an energy supply is immediately 4. Dirks, B., Hanke, H., Krieglstein, J., Stock, R., and Wickop, G. (1980) J. Neu-
needed. These observations are consistent with the concept rochem. 35, 311–317
5. Ghajar, J. B. G., Plum, F., and Duffy, T. E. (1982) J. Neurochem. 38, 397– 409
that rapid, energy-consuming cellular processes rely on glyco- 6. Bachelard, H. S., Cox, D. W. G., and Drower, J. (1984) J. Physiol. 352, 91–102
lytically produced ATP. 7. Fleck, M. W., Henze, D. A., Barrionuevo, G., and Palmer, A. M. (1993) J. Neu-
rosci. 13, 3944 –3955
ATP generated at the surface of the synaptic vesicle may 8. Cox, D. W. G., and Bachelard, H. S. (1982) Brain Res. 239, 527–534
play an essential role in ensuring prompt Glu refilling of syn- 9. Cox, D. W. G., Morris, P. G., Feeney, J., and Bachelard, H. S. (1983) Biochem.
J. 212, 365–370
aptic vesicles in neurons. Mitochondria alone may not entirely 10. Kanatani, T., Mizuno, K., and Okada, Y. (1995) Experientia 51, 213–216
be able to meet the ATP requirement for maintaining normal 11. Shoji, S. (1992) Synapse 12, 322–332
neurotransmission, an extremely rapid cellular process. Syn- 12. Spuler, A., Endres, W., and Grafe, P. (1988) Exp. Neurol. 100, 248 –252
13. Ueda, T., and Plagens, D. G. (1987) Proc. Natl. Acad. Sci. U. S. A. 84,
aptic vesicles release their contents and then must be rapidly 1229 –1233
refilled in order to continuously support normal neurotrans- 14. Morino, H., Fischer-Bovenkerk, C., Kish, P. E., and Ueda, T. (1991) J. Neuro-
chem. 56, 1049 –1057
mission (61). Energy demand for vesicular uptake would be 15. Watkins, J. C., and Evans, R. H. (1981) Annu. Rev. Pharmacol. Toxicol. 21,
dynamic and reflect nerve activity. The synaptic vesicle’s func- 165–204
5940 GAPDH and Vesicular Glu Accumulation
16. Cotman, C. W., Foster, A., and Lanthorn, T. (1981) in Glutamate as a Neuro- 39. Fairbanks, G., Steck, S. L., and Wallach, D. F. H. (1971) Biochemistry 10,
transmitter (DiChiara, G., and Gessa, G. L., eds) pp. 1–27, Raven Press, 2606 –2617
New York 40. Laemmli, U. K. (1970) Nature 227, 680 – 685
17. Fonnum, F. (1984) J. Neurochem. 42, 1–11 41. Ueda, T., Maeno, H., and Greengard, P. (1973) J. Biol. Chem. 248, 8295– 8305
18. Ueda, T. (1986) in Excitatory Amino Acids (Roberts, P. J., Storm-Mathisen, J., 42. Bole, D. G., Hirata, K., and Ueda, T. (2002) Neurosci. Lett. 322, 17–20
and Bradford, H. F., eds) pp. 173–195, Macmillan, London, UK 43. Ogita, K., Hirata, K., Bole, D. G., Yoshida, S., Tamura, Y., Leckenby, A. M.,
19. Collingridge, G. L., and Bliss, T. V. P. (1987) Trends Neurosci. 10, 288 –293 and Ueda, T. (2001) J. Neurochem. 77, 34 – 42
20. Cotman, C. W., Monaghan, D. T., and Ganong, A. H. (1988) Annu. Rev. 44. Berenski, L. M., Kim, C. J., and Jung, C. Y. (1990) J. Biol. Chem. 265,
Neurosci. 11, 61– 80 15449 –15454
21. Nicholls, D. G. (1989) J. Neurochem. 52, 331–341 45. Lanzetta, P. A., Alvarez, L. J., Reinach, P. S., and Candia, O. A. (1979) Analyt.
22. Maycox, P. R., Hell, J. W., and Jahn, R. (1990) Trends Neurosci. 13, 83– 87 Biochem. 100, 95–97
23. Özkan, E. D., and Ueda, T. (1998) Jpn. J. Pharmacol. 77, 1–10 46. Srivastava, D. K., and Bernhard, S. A. (1986) Curr. Top. Cell. Reg. 28, 1– 68
24. Reimer, R. J., Fremeau, R. T. Jr., Bellocchio, E. E., and Edwards, R. H. (2001) 47. Weber, J. P., and Bernhard, S. A. (1982) Biochemistry 21, 4184 – 4194
Curr. Opin. Cell Biol. 13, 417– 421 48. Fischer-Bovenkerk, C., Kish, P. E., and Ueda, T. (1988) J. Neurochem. 51,
25. Takamori, S., Rhee, J. S., Rosenmund, C., and Jahn, R. (2000) Nature 407, 1054 –1059
189 –194
49. Fykse, E. M., Christensen, H., and Fonnum, F. (1989) J. Neurochem. 52,
26. Otis, T. S. (2001) Neuron 29, 11–14
946 –951
27. Naito, S., and Ueda, T. (1983) J. Biol. Chem. 258, 696 – 699
50. Kish, P. E., and Ueda, T. (1989) Methods Enzymol. 174, 9 –25
28. Naito, S., and Ueda, T. (1985) J. Neurochem. 44, 99 –109
51. Winter, H. C., and Ueda, T. (1993) Neurochem. Res. 18, 79 – 85
29. Maycox, P. R., Deckwerth, T., Hell, J. W., and Jahn, R. (1988) J. Biol. Chem.
263, 15423–15428 52. Krimsky, I., and Racker, E. (1955) Science 122, 319 –321
30. Hell, J. W., Maycox, P. R., and Jahn, R. (1990) J. Biol. Chem. 265, 2111–2117 53. Bloch, W., MacQuarrie, R. A., and Bernhard, S. A. (1971) J. Biol. Chem. 246,
31. Tabb, J. S., and Ueda, T. (1991) J. Neurosci. 11, 1822–1828 780 –790
32. Tabb, J. S., Kish, P. E., Van Dyke, R., and Ueda, T. (1992) J. Biol. Chem. 267, 54. Schlaefer, M., Volknandt, W., and Zimmermann, H. (1994) J. Neurochem. 63,
15412–15418 1924 –1931
33. Wolosker, H., Reis, M., Assreuy, J., and de Meis, L. (1996) J. Neurochem. 66, 55. Rogalski-Wilk, A. A., and Cohen, R. S. (1997) Cell. Mol. Neurobiol. 17, 51–70
1943–1948 56. Weiss, J. N., and Lamp, S. T. (1987) Science 238, 67– 69
34. Bellocchio, E. E., Reimer, R. J., Fremeau, R. T., Jr., and Edwards, R. H. (2000) 57. Holloszy, J. O., and Booth, F. W. (1976) Annu. Rev. Physiol. 38, 273–291
Science 289, 957–960 58. Baldwin, K. M., Winder, W. W., Terjung, R. L., and Holloszy, J. O. (1973)
35. Özkan, E. D., Lee, F. S., and Ueda, T. (1997) Proc. Natl. Acad. Sci. U. S. A. 94, Am. J. Physiol. 225, 962–966
4137– 4142 59. Brunschwig, J. P., Brandt, N., Caswell, A. H., and Lukeman, D. S. (1982)
36. Ueda, T., Greengard, P., Berzins, K., Cohen, R. S., Blomberg, F., Grab, D. J., J. Cell Biol. 93, 533–542

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


and Siekevitz, P. (1979) J. Cell Biol. 83, 308 –319 60. Heilmeyer, L. M. G., Jr., Han, J.-W., Thieleczek, R., Varsanyi, M., and Mayr,
37. Dunkley, P. R., Heath, J. W., Harrison, S. M., Jarvie, P. E., Glenfield, P. J., and G. W. (1990) Mol. Cell. Biochem. 99, 111–116
Rostas, A. P. (1988) Brain Res. 441, 59 –71 61. Pyle, J. L., Kavalali, E. T., Piedras-Renteria, E. S., and Tsien, R. W. (2000)
38. Bradford, M. M. (1976) Anal. Biochem. 72, 248 –254 Neuron 28, 221–231
MEMBRANE TRANSPORT STRUCTURE
FUNCTION AND BIOGENESIS:
Glycolysis and Glutamate Accumulation
into Synaptic Vesicles: ROLE OF
GLYCERALDEHYDE PHOSPHATE
DEHYDROGENASE AND
3-PHOSPHOGLYCERATE KINASE

Atsushi Ikemoto, David G. Bole and


Tetsufumi Ueda
J. Biol. Chem. 2003, 278:5929-5940.
doi: 10.1074/jbc.M211617200 originally published online December 17, 2002

Downloaded from http://www.jbc.org/ by guest on March 4, 2015


Access the most updated version of this article at doi: 10.1074/jbc.M211617200

Find articles, minireviews, Reflections and Classics on similar topics on the JBC Affinity Sites.

Alerts:
• When this article is cited
• When a correction for this article is posted

Click here to choose from all of JBC's e-mail alerts

This article cites 58 references, 18 of which can be accessed free at


http://www.jbc.org/content/278/8/5929.full.html#ref-list-1

You might also like