Vol.2, No.1, 8-15 (2010)
doi:10.4236/health.2010.21002
SciRes
Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
Health
The role of intracellular sodium (Na+) in the regulation of
calcium (Ca2+)-mediated signaling and toxicity
Xian-Min Yu, Bradley R. Groveman, Xiao-Qian Fang, Shuang-Xiu Lin
Department of Biomedical Sciences, College of Medicine, Florida State University, Tallahassee, USA; xianmin.yu@med.fsu.edu
Received 15 October 2009; revised 1 December 2009; accepted 2 December 2009.
ABSTRACT
It is known that activated N-methyl-D-aspartate
receptors (NMDARs) are a major route of ex-
cessive calcium ion (Ca2+) entry in central neu-
rons, which may activate degradative processes
and thereby cause cell death. Therefore, NMD-
ARs are now recognized to play a key role in the
development of many diseases associated with
injuries to the central nervous system (CNS).
However, it remains a mystery how NMDAR ac-
tivity is recruited in the cellular processes
leading to excitotoxicity and how NMDAR activ-
ity can be controlled at a physiological level.
The sodium ion (Na+) is the major cation in ex-
tracellular space. With its entry into the cell, Na+
can act as a critical intracellular second mes-
senger that regulates many cellular functions.
Recent data have shown that intracellular Na+
can be an important signaling factor underlying
the up-regulation of NMDARs. While Ca2+ influx
during the activation of NMDARs down-regu-
lates NMDAR activity, Na+ influx provides an
essential positive feedback mechanism to over-
come Ca2+-induced inhibition and thereby po-
tentiate both NMDAR activity and inward Ca2+
flow. Extensive investigations have been con-
ducted to clarify mechanisms underlying Ca2+-
mediated signaling. This review focuses on the
roles of Na+ in the regulation of Ca2+-mediated
NMDAR signaling and toxicity.
Keywords: NMDA Receptors; Sodium and Calcium
Influx; Sodium and Calcium Signaling; Excitability;
Toxicity
1. INTRODUCTION
Cytoplasmic Ca2+ is the most common signaling factor in
all types of cells. Normal intracellular Ca2+ concentration
([Ca2+]i) is approximately 40,000-fold lower than ex-
tracellular [Ca2+], which ranges from 1 to 2 mM [1,2].
Ca2+ ions enter neurons via various pathways including
voltage-gated Ca2+ channels, ligand-gated Ca2+ channels
and the Ca2+ exchangers [2,3]. It is known that activated
NMDAR channels are a major route of excessive Ca2+
entry in neurons [4-10]. While excessive intracellular
Ca2+ may activate degradative processes and thereby
cause toxic effects [2,10-12], NMDAR channel activity
may be inhibited by intracellular Ca2+ through: 1) -
actinin/cytoskeleton dissociation from the NR1 subunit of
NMDARs [13]; 2) calmodulin activation [13-16], and 3)
activation of phosphatases, such as calcineurin which
dephosphorylates NMDARs [17-19]. The Ca2+-induced
down- regulation of NMDARs is considered an important
negative feedback mechanism to control NMDAR activ-
ity [20-23]. Based on these findings we questioned: How
do excessive amounts of Ca2+ get into neurons through
NMDARs if NMDARs are inhibited by Ca2+ influx?
Na+ is the major cation in the extracellular space, and it
can enter cells through a variety of routes including
permeation through ligand- (e.g., glutamate) and voltage-
gated cation channels, uptake via membrane exchangers
and gradient-driven co-transporters [24]. NMDAR chan-
nels are highly permeable to both Na+ and Ca2+. Short
burst or tetanic stimulation of afferents that induces
synaptic LTP increases [Na+] up to 40 or 100 mM in
spines and adjacent dendrites [25,26]. These increases
can essentially be prevented by the blockade of NMDARs,
indicating that they are mainly mediated by Na+ entry
through NMDARs [25,26].
Our initial studies demonstrated that intracellular Na+
is an up-regulator of NMDARs, such that raising [Na+]i or
activating Na+ permeable channels may increase NMD-
AR-mediated currents [27-29]. We then identified that in
hippocampal neurons an increase of 5 ± 1 mM in [Na+]i
represents a threshold required to mask the down- regu-
lation of NMDARs induced by Ca2+ influx. Further in-
creases in Na+ influx not only significantly enhance Ca2+
influx induced by the activation of NMDARs, but also
overcome the Ca2+- dependent inhibition of NMDARs
[27,30]. This review focuses on the roles of Na+ in the
development of tissue injury and in the regulation of
Ca2+-mediated NMDAR signaling and toxicity.
X. M. Yu et al. / HEALTH 2 (2010) 8-15
SciRes Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
9
2. Na+ IN THE PROCESS OF
TISSUE INJURY
A significant increase in [Na+]i is a characteristic event
associated with tissue injury [31-37]. Application of
voltage-gated Na+ channel blockers reduce both Na+
entry and apoptotic neuronal death [32] whereas in-
creases of Na+ entry by application of the voltage-gated
Na+ channel activator, veratridine, induce neuronal apop-
tosis and caspase-3 activation [32,33]. There is a report
showing that during anoxia Na+ entry can occur through
either Gd3+-sensitive channels or via Na+/K+/2Cl- co-
transporters in cultured hippocampal neurons [38], im-
plying that multiple pathways for Na+ entry may be ac-
tivated during tissue injury.
It is known that Na+ influx into the cell is accompanied
by chloride ions (Cl-) and water, which can lead to acute
neuronal swelling and damage [4,39]. Previous studies
have shown that Na+ entry may cause an increase in cy-
tosolic Ca2+ through either Na+/Ca2+ exchangers or acti-
vation of voltage-gated Ca2+ channels [40,41], thereby
activating Ca2+-dependent signaling mechanisms. More-
over, Na+ entry via Na+/H+ exchange may cause changes
in intracellular pH, and thereby regulate many cellular
functions including enzyme activity, neuronal growth and
death [38,42-46]. A recent study showed that Na+ influx
plays an important role in the onset of anti-Fas-induced
apoptosis and that blocking Na+ influx may rescue pro-
grammed cell death in Jurkat cells [47]. Cox and col-
leagues reported that the binding of agonists to opioid
receptors on guinea pig cortical neuron membranes is
significantly reduced by increases in [Na+]i of 10-30 mM
[48]. Maximal inhibition of -, - and -opioid receptor
binding by Na+ is approximately 60%, 70% and 20%,
respectively [48]. Co-occurrence of Na+,K+-ATPase
dysfunction and Na+ influx causes α-amino-3-hydroxyl-
5-methyl-4-isoxa-zole-propionate receptor (AMPAR) pro-
teolysis and a rapid reduction of AMPAR cell-surface
expression [49]. Na+-mediated K+ channels such as Slo
gene-encoded K+ channels [50-54], are widely distributed
throughout the nervous system and are involved in both
the regulation of the after-potential following action po-
tentials [51,55], and the protection of neurons from hy-
poxic stimulation [51,52,54].
While the details of the mechanisms remain to be
clarified, significant pharmacological data have demon-
strated the protective effects of blocking Na+ influx dur-
ing injuries to the nervous tissue. The blockade of volt-
age-gated Na+ channels can prevent neurons from trau-
matic spinal cord injury [33,56-60] and the loss of white
matter [30,56,60,61], concurrently reducing the sensiti-
zation associated with pain [62-65] and preventing sei-
zures during kindling development [66]. The inhibition of
Na+/H+ exchange attenuates ischemia-induced cell death
[67,68]. As a result, a major focus of pharmaceutical
research has been on the search for effective therapeutic
approaches that target voltage-gated Na+ channels
[33,36].
3. ROLES OF Na+ IN THE REGULATION OF
Ca2+-MEDIATED NMDAR SIGNALING
AND TOXICITY
3.1. Calcium Influx through Activated
NMDARs is Regulated by Na+ Influx
Activated NMDARs are highly permeable to both Na+
and Ca2+ [20,22,23]. Prolonged increases of intracellular
Ca2+ during NMDAR activation may act as a negative
feedback mechanism controlling NMDAR activity
[20,22,23]. In light of our findings demonstrating that: 1)
intracellular Na+ up-regulates NMDAR channel gating
and 2) multiple types of receptor/channels such as AM-
PARs, voltage-gated Na+ channels, non-selective cation
channels and remote NMDARs may regulate NMDAR
activity through a Na+-dependent mechanism [27,28], we
investigated how NMDARs are regulated when both Ca2+
and Na+ flow into neurons during the same time period
through activated NMDARs [27,30]. Recordings were
conducted in the cell-attached single-channel configura-
tion. In this recording model, recorded surface NMDARs
are isolated by a recording electrode from the bath envi-
ronment and therefore cannot be directly stimulated by
bath-applied agents. We recorded the activity of surface
NMDARs before and after activation of remote
NMDARs (outside the patch) induced by bath application
of NMDA or L-aspartate [30]. To prevent toxic effects
which may be induced by application of NMDA or L-as-
partate, a standard extracellular solution in which NaCl
and KCl were replaced by Na2SO4 and Cs2SO4, was
utilized [30]. Consistent with previous findings [17,28,39,
69,70], no damage of neurons bathed with this standard
solution was observed following NMDA or aspartate
application. NMDAR single-channel activity was evoked
with 10 µM NMDA and 3 µM glycine included in the
standard extracellular solution filling the recording elec-
trodes.
We found that bath application of NMDAR agonists
may change NMDAR channel activity recorded in
cell-attached patches in a concentration-dependent man-
ner. While a significant increase in NMDAR channel
gating occurred during L-aspartate (>100 µM) applica-
tion to neurons bathed with the standard extracellular
solution, the activity of NMDARs was inhibited in neu-
rons when Na+ influx was blocked by replacing ex-
tracellular Na+ with Cs+ or N-methyl-D-glutamine
(NMDG) [28,30].
We measured the ratio of fluorescence at 346 nm ver-
sus 380nm for the Na+-sensitive dye, sodium-binding
benzofuran isophthalate (SBFI), and the Ca2+-sensitive
X. M. Yu et al. / HEALTH 2 (2010) 8-15
SciRes Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
10
dye, Fura-2, in the soma region of neurons. When the Na+
gradient across the cell membrane was decreased by
reducing extracellular Na+ concentration ([Na+]e) to 20
mM and the Na+ ionophore, monensin (10 M) was in-
cluded in the extracellular solution, basal [Ca2+]i and
[Na+]i of neurons were approximately 84 nM and 16 mM,
respectively. Under this condition bath application of
L-aspartate increased [Ca2+]i by 66 nM, decreased [Na+]i
by 5.8 mM and inhibited NMDAR activity [30]. On av-
erage, the overall channel open probability and mean
open time were reduced to 64% and 77% of controls. The
burst and cluster lengths were also significantly reduced.
These inhibitory effects produced by the bath application
of L-aspartate were prevented by either application of
APV or removal of Ca2+ from extracellular solution,
indicating that the activation of remote NMDARs may
also down-regulate recorded NMDAR activity through
Ca2+ influx [30]. Thus, it is demonstrated that NMDARs
can be up- and down-regulated by influxes of Na+ and
Ca2+, respectively.
We then measured changes of [Na+]i and [Ca2+]i in
neurons bathed with extracellular solution containing a
[Na+] of 10, 20 or 145 mM before and during the activa-
tion of NMDARs induced by bath application of
L-aspartate. We found that with an increase in [Na+]e, the
activation NMDARs produced increases in [Na+]i as ex-
pected, but also increased [Ca2+]i. Excluding the effect of
Ca2+ influx-induced Ca2+ release (CICR) from intracel-
lular stores, the increase in [Ca2+]i of neurons bathed with
extracellular solution containing 145 mM Na+ was still
significantly higher than that found in neurons bathed
with extracellular solution containing 10 mM Na+ [28,30 ].
When [Na+]e was reduced to 10 mM, the activation of
NMDARs produced increases in [Na]i and [Ca2+]i by
around 0.8 mM and 35 nM, respectively. Under this
condition, the activation of remote NMDARs inhibited
NMDAR activity recorded in cell-attached patches [30].
When [Na+]e was increased to 20 mM, NMDAR activa-
tion produced a 5 mM increase in [Na+]i and a 50 nM
increase in [Ca2+]i, but no change in the activity of re-
corded NMDARs [30]. Similarly, increasing [K+]e by 30
mM in an extracellular solution containing 170 mM Na+
and 1 M TTX produced increases in [Na+]i and [Ca2+]i
by around 7 mM and 48 nM, respectively, but again
showed no change in the activity of NMDARs recorded
in cell-attached patches either [28,30 ]. Thus, an increase
in [Na+]i of approximately 5 mM appeared to be a critical
concentration for masking the inhibitory effects induced
by Ca2+ influx on NMDARs in cultured hippocampal
neurons [30]. Since a modest increase of [Ca2+]i by ap-
proximately 35 nM inhibited NMDAR activity when
[Na+]e was reduced to 10 mM [30], it was possible that
Na+ influx not only enhanced Ca2+ influx but also masked
the inhibitory effects of Ca2+.
To confirm this hypothesis, we recorded NMDAR
single-channel activity before and during the activation of
remote NMDARs in cell-attached patches with pipettes
filled with a Ca2+-free extracellular solution containing
200 mM Na+ from neurons that had been pre-treated with
BAPTA-AM (10 M for 4 hrs) and bathed with the same
Ca2+-free extracellular solution, or with pipettes filled
with extracellular solution containing 0.3 or 1.2 mM Ca2+
from neurons bathed with the extracellular solution con-
taining the same amount of Ca2+, respectively. We found
that the activation of remote NMDARs produced a simi-
lar up-regulation of NMDAR channel activity when local
and bath [Ca2+] was set at 0, 0.3 and 1.2 mM, implying
again that the effects of Ca2+ influx in the regulation of
NMDARs by remote NMDARs are overcome by Na+
under normal condition [30]. Furthermore, removal of
extracellular Ca2+ did not produce any effect on the
up-regulation of NMDARs by remote NMDARs in neu-
rons bathed with the standard extracellular solution con-
taining 200 mM Na+ [28,30]. Thus, we conclude that Ca2+
influx through activated NMDARs is regulated by Na+
influx, and that the effect of Na+, which overcomes
Ca2+-induced inhibition, provides an essential positive
feedback mechanism enhancing both the NMDAR activ-
ity and the inward flow of Ca2+.
3.2. Depletion of Extracellular Ca2+
Enhances Na+ Influx and Thereby
Causes NMDAR-Mediated Toxicity
Based on the findings that glutamate concentration may
increase in both humans [71,72] and animals after nervous
system injury [73], and that application of NMDAR an-
tagonists may protect neurons from excitotoxic injuries in
both humans [74,75] and animal [39,74,76], it has been
believed that NMDAR-mediated excitoxicity plays a ke- y
role in the development of neuronal death associated with
stroke/traumatic CNS injury. However, it remained unclear
how NMDARs were recruited to cause neurotoxicity.
We examined the effects of extracellular Ca2+ depletion
and reperfusion, which may occur in stroke patients, on
cultured hippocampal neurons [27,77]. Neurons were
bathed initially with an extracellular solution containing:
140 mM NaCl, 5 mM CsCl, 1.8 mM CaCl2, 33 mM
glucose, 25 mM HEPES; pH: 7.35; osmolarity: 310-320
mOsm. The reduction of [Ca2+]e from 1.8 mM to 0.5 or 0
mM caused a significant increase in Caspase-3 activity
and morphological changes in neurons such as swelling,
beading, and/or process disintegration. Significantly less
formazan was observed in 3-(4,5-dimethylthiazol-2-yl)
2,5-diphenyl-tetrazolium bromide (MTT) assays in which
neurons were treated with the extracellular solution con-
taining 0.5 or 0 mM Ca2+, indicating a change in mito-
chondrial function associated with neuronal injury
[78-81]. Unexpectedly, application of NMDAR antago-
nists APV (100 M) and MK801 (2 M) significantly
prevented the above mentioned changes in neurons only
X. M. Yu et al. / HEALTH 2 (2010) 8-15
SciRes Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
11
when the drugs were applied concomitant to the reduction
of [Ca2+]e from 1.8 to 0 mM. No protective effects of the
drugs could be found when they were applied during Ca2+
reperfusion or when [Ca2+]e was reduced from 1.8 to 0.5
mM [77]. These findings suggest that the depletion of
extracellular Ca2+ may evoke NMDAR-mediated neuro-
toxicity [77], and also raised the questions of how and
when NMDAR activity is recruited to induce neuronal
injury following the removal of extracellular Ca2+.
To address this question, we recorded NMDAR sin-
gle-channel activity before and during a depletion of
extracellular Ca2+ from 1.8 to 1.3, 0.5 or 0 mM in cell-
attached patches from cultured hippocampal neurons. To
prevent cell damage during the reduction of extracellular
Ca2+, the Cl- in the standard extracellular solution was
replaced by SO4
2- [28,30,39,77]. Bath application of a
low [Ca2+] solution to neurons caused a parallel shift of
the current-voltage (I/V) relationship in NMDAR sin-
gle-channels recorded in cell-attached patches, which
indicates that there is a cell-depolarization, but no change
in single-channel conductance [30,77]. In order to ac-
count for this, the holding potential was re-adjusted to
maintain a 70 mV patch-potential from the reversal po-
tential of recorded channels. We found that a depletion of
extracellular Ca2+ from 1.8 to 1.3 or 0.5 mM did not in-
duce any significant change in the activity of recorded
channels until [Ca2+]e was reduced from 1.8 to 0 mM
[28,30,77]. The channel activity could be subsequently
abolished with application of the NMDAR antagonist,
MK801, confirming that a [Ca2+]e reduction from 1.8 to 0
mM produces increases in NMDAR activity [28,30,77].
Since the concomitant blockade of NMDARs to the re-
duction of [Ca2+]e from 1.8 to 0.5 mM may actually in-
crease the number of injured neurons [77], the up-regu-
lation of NMDARs appears to be essential in triggering of
toxicity mediated by NMDARs and the application of
NMDAR antagonists in the Ca2+ reperfusion model may
be protective only when NMDARs are recruited.
To identify the mechanisms by which the removal of
extracellular Ca2+ results in the up-regulation of NMD-
ARs, we measured [Ca2+]i and [Na+]i in cultured hippo-
campal neurons before and during reductions of ex-
tracellular Ca2+. A [Ca2+]e reduction-dependent decrease
in [Ca2+]i and increase in [Na+]i were observed [77]. A
depletion of extracellular Ca2+ from 1.8 to 0 mM pro-
duced sufficient increases in [Na+]i capable of enhancing
NMDAR activity [28,30,77]. Furthermore, we found that
the up-regulation of NMDAR activity induced by ex-
tracellular Ca2+ depletion was prevented by the blockade
of Na+ influx [77].
Previous studies showed that the removal of extracel-
lular Ca2+ to 0 mM may increase NMDAR single-channel
conductance [20,23,82], and that reducing intracellular
Ca2+ may reduce the Ca2+-dependent inhibition of
NMDARs and thereby enhance NMDAR channel activity
[13-20,23]. Therefore, it is possible that NMDAR gating
may be enhanced by the removal of extracellular Ca2+
through Na+ and/or Ca2+-dependent mechanisms.
The ensemble currents produced by the summation of
consecutive super-clusters were compared before (1.8
mM) and after reducing [Ca2+]e to 0 mM. We found that
the removal of extracellular Ca2+ may significantly in-
crease the decay time of ensemble currents and that this
effect can be abolished by blocking Na+ influx. This
suggests that the removal of extracellular Ca2+ may affect
NMDAR-mediated whole-cell responses through the
action of Na+ [77].
Large reductions in [Ca2+]e have been found during
instances of high neuronal activity [83-85], the develop-
ment of seizures [86], hypoglycemic coma [87], and
periods of hypoxia and ischemia [88,89]. Ca2+-depletion
has also been reported to induce cell injury and death [90].
Thus, the Na+-dependent enhancement of NMDAR activ-
ity induced by depletion of extracellular Ca2+ may be an
important mechanism underlying the development of neu-
rotoxicity in the CNS.
3.3. Na+ Regulation of Ca2+ Homeostasis
Under resting conditions [Ca2+]i in neurons is normally
maintained at 10–100 nM, and is tightly regulated by both
Ca2+ influx and efflux across the membrane. [Ca2+]i can
be increased by Ca2+ entry through Ca2+ channels (in-
cluding ligand- and voltage-gated Ca2+ channels and
non-selective cation channels) located on the plasma
membrane and by CICR from the endoplasmic reticulum
(ER) upon binding of inositol trisphosphate (IP3) to the
inositol trisphosphate receptor (IP3R). Ca2+-mediated
injury is usually acute and rapid [91]. Disturbances of
Ca2+ homeostasis in the cytoplasm, ER, or mitochondria
can be harmful to cells [3]. Since Ca2+ stores are closely
connected within the cells and interact with each other,
dysregulation of one compartment is usually followed by
responses from the others. Together, they may overwhelm
the cell’s capacity to maintain overall homeostasis and
kill the cell [3].
In the plasma membrane Ca2+-ATPase and the Na+/
Ca2+ exchanger act to transport cytosolic Ca2+ to the ex-
tracellular space. The Na+/Ca2+ exchanger has a low af-
finity for Ca2+ but a high velocity; as such, it removes
Ca2+ only when cytosolic concentrations are high. The
Ca2+-ATPase, has a high affinity for Ca2+ and pumps out
Ca2+ even at low cytosolic concentrations [3,92,93]. In
resting cells, [Ca2+] in the mitochondrial matrix is around
100 nM. When cytosolic [Ca2+] rises, Ca2+ can enter the
mitochondria through a uniporter and thereby regulate
Ca2+ signals [94]. In mitochondria the Na+/Ca2+ ex-
changer extrudes Ca2+ [3,94,95]. However, the activity of
the Na+/Ca2+ exchanger may be reversed on the influx of
Na+ [96]. This reversal in Na+/Ca2 exchange is observed
under pathological conditions [3]. If the mitochondrial
X. M. Yu et al. / HEALTH 2 (2010) 8-15
SciRes Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
12
Na+/Ca2+ exchanger is overwhelmed by Ca2+ entry, the
Ca2+ levels in the mitochondrial matrix may increase
enough to trigger a mitochondrial permeability transition.
The sustained transitions may cause mitochondrial de-
polarization, inhibition of ATP production, and cell death
[97-99]. In the nucleus Ca2+ is involved in the gene
transcription and DNA metabolism [100-102]. Unlike in
the mitochondria, nuclear Ca2+ is found to be rapidly
equilibrated with cytosolic Ca2+. This may occur by dif-
fusion across nuclear pores [103] and/or Ca2+ channels in
the nuclear envelope [104].
CICR during NMDAR activation has been reported
[20,105,106]. We observed that when extracellular solu-
tion contained more Na+, NMDAR activation produced
greater increases in both [Na+]i and [Ca2+]i [30]. Fur-
thermore, in neurons bathed with extracellular solution
containing 145 mM Na+, NMDAR activation-induced
increases in [Ca2+]i were significantly reduced from 100
30 nM (n = 5) to 62 8 nM (n = 8) with thapsigargin (0.1
µM) treatment [30], which depletes intracellular stores of
Ca2+ by blocking Ca2+ re-uptake. In the absence of thap-
sigargin, NMDAR activation only produced a 35 8 nM
(n=8) increase in [Ca2+]i in neurons bathed with ex-
tracellular solution containing 10 mM Na+ [30]. The
blockade of Ca2+ influx by removal of extracellular Ca2+
abolished the NMDAR activation-induced increase in
[Ca2+]i, (data not shown) [2,107]. The increase in [Ca2+]i
induced by Ca2+ release from intracellular stores during
NMDAR activation in neurons bathed with extracellular
solution containing 10 mM Na+ was significantly reduced
when compared with that in neurons bathed with ex-
tracellular solution containing 145 mM Na+ [30]. These
data suggest that CICR from intracellular stores during
NMDAR activation may be regulated by intracellular Na+.
Stys and colleagues provided direct evidence showing
that intra-axonal Ca2+ release during ischemia in rat optic
nerves is mainly dependent on Na+ influx. This Na+ ac-
cumulation stimulates three distinct intra-axonal sources
of Ca2+: 1) the mitochondrial Na+/Ca2+ exchanger driven
in the Na+ import/Ca2+ export mode; 2) positive modula-
tion of ryanodine receptors; and 3) promotion of IP3
generation by phospholipase C [108].
4. QUESTIONS AND FUTURE STUDIES
Na+ entry is a key factor that initiates fast action poten-
tials and shapes sub-threshold electrical properties to
thereby regulate neuronal excitability and neuronal dis-
charge activity [109-113]. Present data have shown that: 1)
intracellular Na+ up-regulates NMDARs; 2) via increas-
ing intracellular Na+, multiple types of receptor/channels
such as AMPA receptors, voltage-gated Na+ channels and
non-selective cation channels, may regulate NMDAR
activity; 3) Na+ influx may enhance Ca2+ influx, mask the
Ca2+-dependent inhibition of NMDARs and significantly
alter Ca2+ homeostasis.
Based on combined investigations of protein crystal
structures in-vitro and functions in cells, Na+ binding
motifs have been characterized in a number of proteins
such as thrombin, Na+/K+-ATPase and various neuro-
transmitter transporters. Thrombin is a serine protease,
the activity of which is regulated by Na+ binding. The
sequence, CDRDGKYG, in the Na+ binding loop is
highly conserved in thrombin from 11 different species
[114]. Investigations into the crystal structure of a bacte-
rial homologue of the Na+/Cl- dependent transporters
from Aquifex aeolicus revealed that there are two Na+
binding sites, named Na1 and Na2 [115]. Na+/ K+-ATPase
is found to have three Na+ biding sites. Na1 is formed
entirely by the side chain oxygen atoms of residues on
three helices in the transmembrane regions (TM) 5, 6 and
8. Na2 is formed almost ‘‘on’’ the TM4 helix with three
main chain carbonyls plus four side chain oxygen atoms
(Asp 811 and Asp 815 on TM6 and Glu 334 on TM4).
The Na3 binding site is contiguous to Na1. The carbonyls
of Gly 813 and Thr 814 (TM6), the hydroxyl of Tyr 778
(TM5), and the carboxyl of Glu 961 (TM9) contribute to
the Na3 binding site [116,117].
To date there is no evidence of a similar amino acid
sequence corresponding to a Na+ binding site, as seen in
these Na+ binding proteins, present in NMDAR subunit
proteins (Yu, unpublished data). Molecular mechanisms
underlying the regulation of NMDARs and Ca2+ signaling
by intracellular Na+ remain unclear. Investigations aiming
to identify critical Na+ targeting site(s) in the regulation of
NMDARs and Ca2+ homeostasis are essentially needed
for understanding activity-dependent neuroplasticity in
the CNS.
5. ACKNOWLEDGMENTS
This work was supported by grant NIH (1R01 NS053567) to X.-M.Y.
REFERENCES
[1] Berridge, M.J., Lipp, P. and Bootman, M.D. (2000) The
versatility and universality of calcium signaling. Nat Rev
Mol Cell Biol, 1, 11-21.
[2] Clapham, D.E. (1995) Calcium signaling. Cell, 80,
259-268.
[3] Dong, Z., Saikumar, P., Weinberg, J.M. and Venkatacha-
lam, M.A. (2006) Calcium in cell injury and death. Ann.
Rev Pathol., 1, 405-434.
[4] Choi, D.W. (1995) Calcium: Still center-stage in hy-
poxic-ischemic neuronal death. Trends. Neurosci., 18,
58-60.
[5] Forder, J.P. and Tymianski, M. (2009) Postsynaptic
mechanisms of excitotoxicity: Involvement of postsy-
naptic density proteins, radicals, and oxidant molecules.
Neuroscience, 158, 293-300.
[6] Lipton, S.A. (2006) Paradigm shift in neuroprotection by
X. M. Yu et al. / HEALTH 2 (2010) 8-15
SciRes Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
13
NMDA receptor blockade: Memantine and beyond. Nat.
Rev. Drug Discov., 5, 160-170.
[7] Mody, I. and MacDonald, J.F. (1995) NMDA recep-
tor-dependent excitotoxicity: The role of intracellular
Ca2+ release. Trends. Pharmacol. Sci., 16, 356-359.
[8] Soriano, F.X. and Hardingham, G.E. (2007) Compart-
mentalized NMDA receptor signaling to survival and
death. J. Physiol., 584, 381-387.
[9] Tymianski, M. and Tator, C.H. (1996) Normal and abnor-
mal calcium homeostasis in neurons: A basis for the
pathophysiology of traumatic and ischemic central nervous
system injury. Neurosurgery, 38, 1176-1195.
[10] Zipfel, G.J., Babcock, D.J., Lee, J.M. and Choi, D.W.
(2000) Neuronal apoptosis after CNS injury: The roles of
glutamate and calcium. J. Neurotrauma., 17, 857-869.
[11] Bredesen, D.E. (2000) Apoptosis: Overview and signal
transduction pathways [In Process Citation]. J. Neuro-
trauma, 17, 801-810.
[12] Fiskum, G. (2000) Mitochondrial participation in
ischemic and traumatic neural cell death [In Process Ci-
tation]. J. Neurotrauma., 17, 843-855.
[13] Krupp, J.J., Vissel, B., Thomas, C.G., Heinemann, S.F.
and Westbrook, G.L. (1999) Interactions of calmodulin
and alpha-actinin with the NR1 subunit modulate Ca2+-
dependent inactivation of NMDA receptors. J. Neurosci.,
19, 1165-1178.
[14] Ehlers, M.D., Zhang, S., Bernhardt, J.P. and Huganir, R.L.
(1996) Inactivation of NMDA Receptors by direct inter-
action of calmodulin with the NR1 subunit. Cell, 84,
745-755.
[15] Wechsler, A. and Teichberg, V.I. (1998) Brain spectrin
binding to the NMDA receptor is regulated by phosphory-
lation, calcium and calmodulin. EMBO J., 17, 3931-3939.
[16] Zhang, S., Ehlers, M.D., Bernhardt, J.P., Su, C.T. and
Huganir, R.L. (1998) Calmodulin mediates calcium de-
pendent inactivation of N-methyl-D-aspartate receptors.
Neuron, 21, 443-453.
[17] Lieberman, D.N. and Mody, I. (1994) Regulation of
NMDA channel function by endogenous Ca2+- dependent
phosphatase. Nature, 369, 235-239.
[18] Mulkey, R.M., Endo, S., Shenolikar, S. and Malenka, R.C.
(1994) Involvement of a calcineurin/inhibitor-1 phos-
phatase cascade in hippocampal long-term depression.
Nature, 369, 486-488.
[19] Tong, G., Shepherd, D. and Jahr, C.E. (1995) Synaptic
desensitization of NMDA receptors by calcineurin. Sci-
ence, 267, 1510-1512.
[20] Dingledine, R., Borges, K., Bowie, D. and Traynelis, S.F.
(1999) The glutamate receptor ion channels. Pharmacol.
Rev., 51, 7-61.
[21] Kyrozis, A., Albuquerque, C., Gu, J. and MacDermott,
A.B. (1996) Ca(2+)-dependent inactivation of NMDA
receptors: Fast kinetics and high Ca2+ sensitivity in rat
dorsal horn neurons. J. Physiol. (Lond), 495, 449-463.
[22] Mayer, M.L. and Westbrook, G.L. (1987) The physiology
of excitatory amino acids in the vertebrate central nervous
system. Prog. Neurobiol., 28, 197-276.
[23] McBain, C.J. and Mayer, M.L. (1994) N-Methyl-D-aspar-
tic acid receptor structure and function. Physiol. Rev., 74,
723-760.
[24] Nicholls, D. and Attwell, D. (1990) The release and up-
take of excitatory amino acids. Trends in Pharmacologi-
cal Sciences, 11, 462-468.
[25] Rose, C.R. (2002) Na+ signals at central synapses. Neu-
roscientist., 8, 532-539.
[26] Rose, C.R. and Konnerth, A. (2001) NMDA recep-
tor-mediated Na+ signals in spines and dendrites. J.
Neurosci., 21, 4207-4214.
[27] Yu, X.M. (2006) The role of intracellular sodium (Na+) in
the regulation of NMDA receptor-mediated channel ac-
tivity and toxicity. Mol Neurobiol., 3, 63-79.
[28] Yu, X.M. and Salter, M.W. (1998) Gain control of NMDA-
receptor currents by intracellular sodium. Nature, 396,
469-474.
[29] Yu, X.M. and Salter, M.W. (1999) Src, a molecular switch
governing gain control of synaptic transmission mediated
by N-methyl-D-aspartate receptors. Proc. Natl. Acad. Sci.
USA, 96, 7697-7704.
[30] Xin, W.K. et al. (2005) A functional interaction of sodium
and calcium in the regulation of NMDA receptor activity
by remote NMDA receptors. J. Neurosci., 25, 139-148.
[31] Ballard-Croft, C., Carlson, D., Maass, D.L. and Horton,
J.W. (2004) Burn trauma alters calcium transporter pro-
tein expression in the heart. J. Appl. Physiol., 97,
1470-1476.
[32] Banasiak, K.J., Burenkova, O. and Haddad, G.G. (2004)
Activation of voltage-sensitive sodium channels during
oxygen deprivation leads to apoptotic neuronal death.
Neuroscience, 126, 31-44.
[33] Baptiste, D.C. and Fehlings, M.G. (2007) Update on the
treatment of spinal cord injury. Prog. Brain Res., 161,
217-233.
[34] Bauer, R., Walter, B., Fritz, H. and Zwiener, U. (1999)
Ontogenetic aspects of traumatic brain edema: Facts and
suggestions. Exp. Toxicol Pathol., 51, 143-150.
[35] Friedman, J.E. and Haddad, G.G. (1994) Anoxia induces
an increase in intracellular sodium in rat central neurons in
vitro. Brain Res., 663, 329-334.
[36] Schwartz, G. and Fehlings, M.G. (2002) Secondary injury
mechanisms of spinal cord trauma: A novel therapeutic
approach for the management of secondary pathophysi-
ology with the sodium channel blocker riluzole. Prog.
Brain Res., 137, 177-190.
[37] Strichartz, G., Rando, T. and Wang, G.K. (1987) An inte-
grated view of the molecular toxinology of sodium
channel gating in excitable cells. Ann. Rev. Neurosci., 10,
237-67.
[38] Sheldon, C., Diarra, A., Cheng, Y.M. and Church, J. (2004)
Sodium influx pathways during and after anoxia in rat
hippocampal neurons. J. Neurosci., 24, 11057-11069.
[39] Choi, D.W. (1993) NMDA receptors and AMPA/kainate
receptors mediate parallel injury in cerebral cortical cul-
tures subjected to oxygen-glucose deprivation. Prog.
Brain Res., 96, 137-43.
[40] Blaustein, M.P., Fontana, G. and Rogowski, R.S. (1996)
The Na(+)-Ca2+ exchanger in rat brain synaptosomes:
Kinetics and regulation. Ann. N. Y. Acad. Sci., 779,
300-17.
[41] Koch, R.A. and Barish, M.E. (1994) Perturbation of intra-
cellular calcium and hydrogen ion regulation in cultured
mouse hippocampal neurons by reduction of the sodium ion
concentration gradient. J. Neurosci., 14, 2585-2593.
[42] Baxter, K.A. and Church, J. (1996) Characterization of
acid extrusion mechanisms in cultured fetal rat hippo-
campal neurones. J. Physiol. (Lond), 493, 457-470.
X. M. Yu et al. / HEALTH 2 (2010) 8-15
SciRes Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
14
[43] Boonstra, J. et al. (1983) Ionic responses and growth
stimulation induced by nerve growth factor and epidermal
growth factor in rat pheochromocytoma (PC12) cells. J.
Cell Biol., 97, 92-98.
[44] Moolenaar, W.H., Defize, L.H. and de Laat, S.W. (1986)
Ionic signaling by growth factor receptors. J. Exp. Biol.,
124, 359-73.
[45] Moolenaar, W.H., Tsien, R.Y., van der Saag, P.T. and de
Laat, S.W. (1983) Na+/H+ exchange and cytoplasmic pH
in the action of growth factors in human fibroblasts. Na-
ture, 304, 645-648.
[46] Sin, W.C. et al. (2009) Regulation of early neurite
morphogenesis by the Na+/H+ exchanger NHE1. J.
Neurosci., 29, 8946-8959.
[47] Bortner, C.D. and Cidlowski, J.A. (2003) Uncoupling cell
shrinkage from apoptosis reveals that Na+ influx is re-
quired for volume loss during programmed cell death. J.
Biol Chem, 278, 39176-39184.
[48] Werling, L.L., Brown, S.R., Puttfarcken, P. and Cox, B.M.
(1986) Sodium regulation of agonist binding at opioid re-
ceptors. II. Effects of sodium replacement on opioid bind-
ing in guinea pig cortical membranes. Mol Pharmacol, 30,
90-95.
[49] Zhang, D. et al. (2009) Na, K-ATPase activity regulates
AMPA receptor turnover through proteasome-mediated
proteolysis. J. Neurosci., 29, 4498-4511.
[50] Bhattacharjee, A. et al. (2003) Slick (Slo2.1), a rap-
idly-gating sodium-activated potassium channel inhibited
by ATP. J. Neurosci, 23, 11681-11691.
[51] Bhattacharjee, A. and Kaczmarek, L.K. (2005) For K(+)
channels, Na(+) is the new Ca(2+). Trends Neurosci, 28,
422-428.
[52] Dryer, S.E. (2003) Molecular identification of the Na+-
activated K+ channel. Neuron, 37, 727-728.
[53] Niu, X.W. and Meech, R.W. (2000) Potassium inhibition
of sodium-activated potassium (K(Na)) channels in guinea-
pig ventricular myocytes. J. Physiol, 526(Pt 1), 81-90.
[54] Yuan, A. et al. (2003) The sodium-activated potassium
channel is encoded by a member of the Slogene family.
Neuron, 37, 765-773.
[55] Liu, X. and Stan, L.L. (2004) Sodium-activated potassium
conductance participates in the depolarizing afterpotential
following a single action potential in rat hippocampal
CA1 pyramidal cells. Brain Res., 1023, 185-192.
[56] Agrawal, S.K. and Fehlings, M.G. (1996) Mechanisms of
secondary injury to spinal cord axons in vitro: Role of
Na+, Na(+)-K(+)-ATPase, the Na(+)-H+exchanger, and
the Na(+)-Ca2+ exchanger. J. Neurosci., 16, 545-552.
[57] Agrawal, S.K. and Fehlings, M.G. (1997) The effect of the
sodium channel blocker QX-314 on recovery after acute
spinal cord injury. J. Neurotrauma., 14, 81-88.
[58] Fehlings, M.G. and Agrawal, S. (1995) Role of sodium in
the pathophysiology of secondary spinal cord injury.
Spine, 20, 2187-2191.
[59] Hains, B.C., Saab, C.Y., Lo, A.C. and Waxman, S.G.
(2004) Sodium channel blockade with phenytoin protects
spinal cord axons, enhances axonal conduction, and im-
proves functional motor recovery after contusion SCI. Exp.
Neurol., 188, 365-377.
[60] Teng, Y.D. and Wrathall, J.R. (1997) Local blockade of
sodium channels by tetrodotoxin ameliorates tissue loss
and long-term functional deficits resulting from experi-
mental spinal cord injury. J. Neurosci., 17, 4359-4366.
[61] Cummins, T.R. and Waxman, S.G. (1997) Downregula-
tion of tetrodotoxin-resistant sodium currents and up-
regulation of a rapidly repriming tetrodotoxin-sensitive
sodium current in small spinal sensory neurons after nerve
injury. J. Neurosci., 17, 3503-3514.
[62] Appelgren, L., Janson, M., Nitescu, P. and Curelaru, I.
(1996) Continuous intracisternal and high cervical in-
trathecal bupivacaine analgesia in refractory head and
neck pain [see comments]. Anesthesiology, 84, 256-272.
[63] Cox, J.J. et al. (2006) An SCN9A channelopathy causes
congenital inability to experience pain. Nature, 444, 894-
898.
[64] Dib-Hajj, S., Black, J.A., Cummins, T.R. and Waxman,
S.G. (2002) NaN/Nav1.9: A sodium channel with unique
properties. Trends Neurosci., 25, 253-259.
[65] Waxman, S.G. and Hains, B.C. (2006) Fire and phantoms
after spinal cord injury: Na+ channels and central pain.
Trends Neurosci., 29, 207-215.
[66] Rogawski, M.A. and Loscher, W. (2004) The neurobiology
of antiepileptic drugs. Nat. Rev. Neurosci., 5, 553-564.
[67] Mentzer, R.M., Lasley, R.D., Jessel, A. and Karmazyn, M.
(2003) Intracellular sodium hydrogen exchange inhibition
and clinical myocardial protection. Ann. Thorac. Surg.,
75, S700-S708.
[68] Vornov, J.J., Thomas, A.G. and Jo, D. (1996) Protective
effects of extracellular acidosis and blockade of so-
dium/hydrogen ion exchange during recovery from
metabolic inhibition in neuronal tissue culture. J. Neuro-
chem., 67, 2379-2389.
[69] Kohr, G., De Koninck, Y. and Mody, I. (1993) Properties of
NMDA receptor channels in neurons acutely isolated from
epileptic (kindled) rats. J. Neurosci., 13, 3612-3627.
[70] Yu, X.M., Askalan, R., Keil, G.J.I. and Salter, M.W. (1997)
NMDA channel regulation by channel-associated protein
tyrosine kinase src. Science, 275, 674-678.
[71] Baker, A.J., Moulton, R.J., MacMillan, V.H. and Shedden,
P. M. (1993) Excitatory amino acids in cerebrospinal fluid
following traumatic brain injury in humans. J. Neurosurg.,
79, 369-372.
[72] Persson, L. and Hillered, L. (1992) Chemical monitoring
of neurosurgical intensive care patients using intracerebral
microdialysis. J. Neurosurg., 76, 72-80.
[73] Benveniste, H., Drejer, J., Schousboe, A. and Diemer,
N.H. (1984) Elevation of the extracellular concentrations
of glutamate and aspartate in rat hippocampus during
transient cerebral ischemia monitored by intracerebral
microdialysis. J. Neurochem., 43, 1369-1374.
[74] Chen, H.-S.V. and Lipton, S.A. (2005) Pharmacological
implications of two distinct mechanisms of interaction of
memantine with NMDA-gated channels. J. Pharmacol.
Exp. Ther., 314, 961-971.
[75] Lipton, S.A. (2004) Paradigm shifts in NMDA receptor
antagonist drug development: Molecular mechanism of
uncompetitive inhibition by memantine in the treatment of
Alzheimer’s disease and other neurologic disorders. J.
Alzheimers. Dis., 6, S61-S74.
[76] Lipton, P. (1999) Ischemic cell death in brain neurons.
Physiol Rev., 79, 1431-1568.
[77] Xin, W.K. et al. (2005) The removal of extracellular
calcium: A novel mechanism underlying the recruitment
of N-methyl-d-aspartate (NMDA) receptors in neurotox-
icity. Eur. J. Neurosci., 21, 622-636.
X. M. Yu et al. / HEALTH 2 (2010) 8-15
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/ Openly accessible at
15
[78] Giardina, S.F. and Beart, P.M. (2002) Kainate receptor-
mediated apoptosis in primary cultures of cerebellar gran-
ule cells is attenuated by mitogen-activated protein and
cyclin-dependent kinase inhibitors. Br. J. Pharmacol., 135,
1733-1742.
[79] Ohyashiki, T., Satoh, E., Okada, M., Takadera, T. and Sa-
hara, M. (2002) Nerve growth factor protects against alu-
minum-mediated cell death. Toxicology, 176, 195-207.
[80] Sawyer, T.W. (1995) Practical applications of neuronal
tissue culture in in vitro toxicology. Clin. Exp. Pharmacol.
Physiol., 22, 295-296.
[81] Wang, X., Mori, T., Sumii, T. and Lo, E.H. (2002) He-
moglobin-induced cytotoxicity in rat cerebral cortical
neurons: Caspase activation and oxidative stress. St ro k e ,
33, 1882-1888.
[82] Westbrook, G.L., Krupp, J.J. and Vissel, B. (1997) Cy-
toskeletal interactions with glutamate receptors at central
synapses. Soc. Gen. Physiol Ser., 52, 163-175.
[83] Hardingham, N.R. et al. (2006) Extracellular calcium
regulates postsynaptic efficacy through group 1 me-
tabotropic glutamate receptors. J. Neurosci., 26, 6337-6345.
[84] Rusakov, D.A. and Fine, A. (2003) Extracellular Ca2+
depletion contributes to fast activity-dependent modulation
of synaptic transmission in the brain. Neuron, 37, 287-297.
[85] Vassilev, P.M., Mitchel, J., Vassilev, M., Kanazirska, M. and
Brown, E.M. (1997) Assessment of frequency-dependent
alterations in the level of extracellular Ca2+ in the synaptic
cleft. Biophys. J., 72, 2103-2116.
[86] Heinemann, U. and Hamon, B. (1986) Calcium and epi-
leptogenesis. Exp. Brain Res., 65, 1-10.
[87] Ekholm, A., Kristian, T. and Siesjo, B.K. (1995) Influence
of hyperglycemia and of hypercapnia on cellular calcium
transients during reversible brain ischemia. Exp. Brain
Res., 104, 462-466.
[88] Harris, R.J., Symon, L., Branston, N.M. and Bayhan, M.
(1981) Changes in extracellular calcium activity in cere-
bral ischaemia. J. Cereb. Blood Flow Metab., 1, 203-209.
[89] Nicholson, C., Bruggencate, G.T., Steinberg, R. and Stockle,
H. (1977) Calcium modulation in brain extracellular mi-
croenvironment demonstrated with ion-selective micropi-
pette. Proc. Natl. Acad. Sci. USA, 74, 1287-1290.
[90] Smith, M.T., Thor, H. and Orrenius, S. (1981) Toxic in-
jury to isolated hepatocytes is not dependent on extracel-
lular calcium. Science, 213, 1257-1259.
[91] Trump, B.F. and Berezesky, I.K. (1995) Calcium-mediated
cell injury and cell death. FASEB J., 9, 219-228.
[92] Carafoli, E., Santella, L., Branca, D. and Brini, M. (2001)
Generation, control, and processing of cellular calcium
signals. Crit Rev Biochem. Mol Biol., 36, 107-260.
[93] Berridge, M.J., Bootman, M.D. and Roderick, H.L. (2003)
Calcium signaling: Dynamics, homeostasis and remod-
eling. Nat Rev Mol Cell Biol, 4, 517-529.
[94] Duchen, M.R. (2000) Mitochondria and calcium: From
cell signaling to cell death. J. Physiol., 529 (Pt 1), 57-68.
[95] Gunter, T.E., Yule, D.I., Gunter, K.K., Eliseev, R.A. and
Salter, J.D. (2004) Calcium and mitochondria. FEBS Lett.,
567, 96-102.
[96] Hilgemann, D.W., Collins, A. and Matsuoka, S. (1992)
Steady-state and dynamic properties of cardiac so-
dium-calcium exchange: Secondary modulation by cyto-
plasmic calcium and ATP. J. Gen. Physiol., 100, 933-961.
[97] Bernardi, P. and Rasola, A. (2007) Calcium and cell death:
The mitochondrial connection. Subcell. Biochem., 45,
481-506.
[98] Crompton, M. (1999) The mitochondrial permeability
transition pore and its role in cell death. J. Biochem., 341
(Pt 2), 233-249.
[99] Lemasters, J.J., Theruvath, T.P., Zhong, Z. and Nieminen,
A.L. (2009) Mitochondrial calcium and the permeability
transition in cell death. Biochim. Biophys. Acta, 1787,
1395-1401.
[100] Hardingham, G.E., Cruzalegui, F.H., Chawla, S. and
Bading, H. (1998) Mechanisms controlling gene expression
by nuclear calcium signals. Cell Calcium, 23, 131-134.
[101] McKenzie, G.J. et al. (2005) Nuclear Ca2+ and CaM
kinase IV specify hormonal- and Notch-responsiveness. J.
Neurochem., 93, 171-185.
[102] Papadia, S., Stevenson, P., Hardingham, N.R., Bading, H.
and Hardingham, G.E. (2005) Nuclear Ca2+ and the camp
response element-binding protein family mediate a late
phase of activity-dependent neuroprotection. J. Neurosci.,
25, 4279-4287.
[103] Allbritton, N.L., Oancea, E., Kuhn, M.A. and Meyer, T.
(1994) Source of nuclear calcium signals. Proc. Natl. Acad.
Sci. USA, 91, 12458-12462.
[104] Mazzanti, M., DeFelice, L.J., Cohen, J. and Malter, H.
(1990) Ion channels in the nuclear envelope. Nature, 343,
764-767.
[105] Sattler, R. and Tymianski, M. (2000) Molecular mecha-
nisms of calcium-dependent excitotoxicity. J. Mol. Med.,
78, 3-13.
[106] Verkhratsky, A. (2002) The endoplasmic reticulum and
neuronal calcium signaling. Cell Calcium, 32, 393-404.
[107] Verkhratsky, A.J. and Petersen, O.H. (1998) Neuronal
calcium stores. Cell Calcium, 24, 333-343.
[108] Nikolaeva, M.A., Mukherjee, B. and Stys, P.K. (2005)
Na+-dependent sources of intra-axonal Ca2+ release in rat
optic nerve during in vitro chemical ischemia. J Neurosci,
25, 9960-9967.
[109] Cantrell, A.R. and Catterall, W.A. (2001) Neuromodula-
tion of Na+ channels: An unexpected form of cellular
plasticity. Nat. Rev Neurosci., 2, 397-407.
[110] Catterall, W.A., Goldin, A.L. and Waxman, S.G. (2003)
International union of pharmacology. XXXIX. compen-
dium of voltage-gated ion channels: Sodium channels.
Pharmacol Rev., 55, 575-578.
[111] Goldin, A.L. (2001) Resurgence of sodium channel re-
search. Ann. Rev. Physiol., 63, 871-894.
[112] Hille, B. (1992) Ionic channels of excitable membranes.
Sinauer, Sunderland.
[113] Waxman, S.G., Dib-Hajj, S., Cummins, T.R. and Black,
J.A. (2000) Sodium channels and their genes: Dynamic
expression in the normal nervous system, dysregulation in
disease states (1). Brain Res., 886, 5-14.
[114] Di Cera, E. et al. (1995) The Na [IMAGE] binding site of
thrombin. J. Biol. Chem., 270, 22089-22092.
[115] Yamashita, A., Singh, S.K., Kawate, T., Jin, Y. and
Gouaux, E. (2005) Crystal structure of a bacterial homo-
logue of Na+/Cl: Dependent neurotransmitter transporters.
Nature, 437, 215-223.
[116] Li, C., Capendeguy, O., Geering, K. and Horisberger, J.D.
(2005) A third Na+-binding site in the sodium pump. Proc.
Natl. Acad. Sci. USA, 102, 12706-12711.
[117] Ogawa, H. and Toyoshima, C. (2002) Homology model-
ing of the cation binding sites of Na+K+-ATPase. Proc.
Natl. Acad. Sci. USA, 99, 15977-15982.