Abstract
It is known that activated N-methyl-D-aspartate receptors (NMDARs) are a major route of excessive calcium ion (Ca2+) entry in central neurons, which may activate degradative processes and thereby cause cell death. Therefore, NMDARs 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 activity is recruited in the cellular processes leading to excitotoxicity and how NMDAR activity can be controlled at a physiological level. The sodium ion (Na+) is the major cation in extracellular space. With its entry into the cell, Na+ can act as a critical intracellular second messenger 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-regulates NMDAR activity, Na+ influx provides an essential positive feedback mechanism to overcome Ca2+-induced inhibition and thereby potentiate both NMDAR activity and inward Ca2+ flow. Extensive investigations have been conducted 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
I. 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 extracellular [Ca2+], which ranges from 1 to 2 mM 14,27. Ca2+ ions enter neurons via various pathways including voltage-gated Ca2+ channels, ligand-gated Ca2+ channels and the Ca2+ exchangers 27,34. It is known that activated NMDAR channels are a major route of excessive Ca2+ entry in neurons 26,41,62,69,91,96,117. While excessive intracellular Ca2+ may activate degradative processes and thereby cause toxic effects 20,27,40,117, NMDAR channel activity may be inhibited by intracellular Ca2+ through: (i) α-actinin/cytoskeleton dissociation from the NR1 subunit of NMDARs 55, (ii) calmodulin activation 37,55,104,116, and (iii) activation of phosphatases, such as calcineurin which dephosphorylates NMDARs 59,72,94. The Ca2+-induced down-regulation of NMDARs is considered an important negative feedback mechanism to control NMDAR activity 33,56,64,66. 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 73. NMDAR channels 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 82,83. These increases can essentially be prevented by the blockade of NMDARs, indicating that they are mainly mediated by Na+ entry through NMDARs 82,83.
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 NMDAR-mediated currents 110,112,113. We then identified that an increase of 5 ± 1 mM in [Na+]i represents a threshold required to mask the down-regulation of NMDARs induced by Ca2+ influx. Further increases in Na+ influx not only significantly enhance Ca2+ influx induced by the activation of NMDARs, but also overcome the Ca2+-dependent inhibition of NMDARs 107,110. 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.
II. Na+ IN THE PROCESS OF TISSUE-INJURY
A significant increase in [Na+]i is a characteristic event associated with tissue injury 6–9,42,87,92. Application of voltage-gated Na+ channel blockers reduce both Na+ entry and apoptotic neuronal death 7 whereas increases of Na+ entry by application of the voltage-gated Na+ channel activator, veratridine, induce neuronal apoptosis and caspase-3 activation 7,8. 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 88, implying that multiple pathways for Na+ entry may be activated 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 25,26. Previous studies have shown that Na+ entry may cause an increase in cytosolic Ca2+ through either Na+/Ca2+ exchangers or activation of voltage-gated Ca2+ channels 17,53, thereby activating Ca2+-dependent signaling mechanisms. Moreover, 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 10,18,70,71,88,89. 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 programmed cell death in Jurkat cells 19. Cox and colleagues 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 105. Maximal inhibition of μ-, δ- and κ-opioid receptor binding by Na+ is approximately 60%, 70% and 20%, respectively 105. Co-occurrence of Na,K-ATPase dysfunction and Na+ influx causes α-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate receptor (AMPAR) proteolysis and a rapid reduction of AMPAR cell-surface expression 115. Na+-mediated K+ channels such as Slo gene-encoded K+ channels 15,16,35,76,114, are widely distributed throughout the nervous system and are involved in both the regulation of the after-potential following action potentials 16,63, and the protection of neurons from hypoxic stimulation 16,35,114.
While the details of the mechanisms remain to be clarified, significant pharmacological data have demonstrated the protective effects of blocking Na+ influx during injuries to the nervous tissue. The blockade of voltage-gated Na+ channels can prevent neurons from traumatic spinal cord injury 1,2,8,39,46,92,93 and the loss of white matter 1,8,30,93, concurrently reducing the sensitization associated with pain 4,28,32,103 and preventing seizures during kindling development 81. The inhibition of Na+-H+ exchange attenuates ischemia-induced cell death 68,100. As a result, a major focus of pharmaceutical research has been the search for effective therapeutic approaches that target voltage-gated Na+ channels 8,87.
III. THE ROLES OF Na+ IN THE REGULATION OF Ca2+-MEDIATED NMDAR SIGNALING AND TOXICITY
Calcium influx through activated NMDARs is regulated by Na+ influx
Activated NMDARs are highly permeable to both Na+ and Ca2+ 33,64,66. Prolonged increases of intracellular Ca2+ during NMDAR activation may act as a negative feedback mechanism controlling NMDAR activity 33,64,66. In light of our findings demonstrating that: 1) intracellular Na+ up-regulates NMDAR channel gating and 2) multiple types of receptor/channels such as AMPARs, voltage-gated Na+ channels, non-selective cation channels and remote NMDARs may regulate NMDAR activity through a Na+-dependent mechanism 110,112, we investigated how NMDARs are regulated when both Ca2+ and Na+ flow into neurons during the same time period through activated NMDARs 107,110. Recordings were conducted in the cell-attached single-channel configuration. In this recording model, recorded surface NMDARs are isolated by a recording electrode from the bath environment 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 107. To prevent toxic effects which may be induced by application of NMDA or aspartate, a standard extracellular solution in which NaCl and KCl were replaced by Na2SO4 and Cs2SO4, was utilized 107. Consistent with previous findings 25,54,59,111,112, 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 electrodes.
We found that bath application of NMDAR agonists may change NMDAR channel activity recorded in cell-attached patches in a concentration-dependent manner. While a significant increase in NMDAR channel gating occurred during L-aspartate (>100 µM) application to neurons bathed with the standard extracellular solution, the activity of NMDARs was inhibited in neurons when Na+ influx was blocked by replacing extracellular Na+ with Cs+ or N-methyl-D-glutamine (NMDG) 107,112.
We measured the ratio of fluorescence at 346 nm versus 380nm for the Na+-sensitive dye, sodium-binding benzofuran isophthalate (SBFI), and the Ca2+-sensitive 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 µM) was included 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 107. On average, 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 107. 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 activation 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 expected, but also increased [Ca2+]i. Excluding the effect of Ca2+ influx-induced Ca2+ release (CICR) from intracellular 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+ 107,110. 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 107. When [Na+]e was increased to 20 mM, NMDAR activation produced a 5 mM increase in [Na+]i and a 50 nM increase in [Ca2+]i, but no change in the activity of recorded NMDARs 107. Similarly, increasing [K+]e by 30 mM in an extracellular solution containing 170 mM Na+ and 1 µM 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 107,112. 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 107. Since a modest increase of [Ca2+]i by approximately 35 nM inhibited NMDAR activity when [Na+]e was reduced to 10 mM 107, 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 µM 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 containing the same amount of Ca2+, respectively. We found that the activation of remote NMDARs produced a similar 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 107. Furthermore, removal of extracellular Ca2+ did not produce any effect on the up-regulation of NMDARs by remote NMDARs in neurons bathed with the standard extracellular solution containing 200 mM Na+ 107,112. 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 activity and the inward flow of Ca2+.
Depletion of extracellular Ca2+ enhances Na+ influx and thereby causes NMDAR-mediated toxicity
Based on findings that glutamate concentration may increase in both humans 5,80 and animals after nervous system injury 11, and that application of NMDAR antagonists may protect neurons from excitotoxic injuries in both humans 24,61 and animal 24,25,60, it has been believed that NMDAR-mediated excitoxicity plays a key role in the development of neuronal death associated with stroke/traumatic CNS injury. However, it remained unclear how NMDARs are recruited to cause neurotoxicity.
We examined the effects of extracellular Ca2+ depletion and reperfusion, which may occur in stroke patients, on cultured hippocampal neurons 108,110. 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 containing 0.5 or 0 mM Ca2+, indicating a change in mitochondrial function associated with neuronal injury 43,78,86,101. Unexpectedly, application of NMDAR antagonists APV (100 µM) and MK801 (2 µM) significantly prevented the above mentioned changes in neurons only 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 108. These findings suggest that the depletion of extracellular Ca2+ may evoke NMDAR-mediated neurotoxicity 108, 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 single-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 SO42–25,107,108,112. Bath application of a low [Ca2+] solution to neurons caused a parallel shift of the current-voltage (I/V) relationship in NMDAR single-channels recorded in cell-attached patches, which indicates that there is a cell-depolarization, but no change in single-channel conductance 107,108. In order to account for this, the holding potential was re-adjusted to maintain a 70 mV patch-potential from the reversal potential of recorded channels. We found that a depletion of extracellular Ca2+ from 1.8 to 1.3 or 0.5 did not induce any significant change in the activity of recorded channels until [Ca2+]e was reduced from 1.8 to 0 mM 107,108,112. 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 107,108,112. Since the concomitant blockade of NMDARs to the reduction of [Ca2+]e from 1.8 to 0.5 mM may actually increase the number of injured neurons 108, the up-regulation of NMDARs appears to be essential in the 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 NMDARs, we measured [Ca2+]i and [Na+]i in cultured hippocampal neurons before and during reductions of extracellular Ca2+. A [Ca2+]e reduction-dependent decrease in [Ca2+]i and increase in [Na+]i were observed 108. A depletion of extracellular Ca2+ from 1.8 to 0 mM produced sufficient increases in [Na+]i capable of enhancing NMDAR activity 107,108,112. Furthermore, we found that the up-regulation of NMDAR activity induced by extracellular Ca2+ depletion was prevented by the blockade of Na+ influx 108.
Previous studies showed that the removal of extracellular Ca2+ to 0 mM may increase NMDAR single-channel conductance 33,66,106, and that reducing intracellular Ca2+ may reduce the Ca2+-dependent inhibition of NMDARs and thereby enhance NMDAR channel activity 33,37,55,59,66,72,94,104,116. 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 increase 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+ 108.
Large reductions in [Ca2+]e have been found during instances of high neuronal activity 48,84,97, the development of seizures 50, hypoglycemic coma 38, and periods of hypoxia and ischemia 49,74. Ca2+-depletion has also been reported to induce cell injury and death 90. Thus, the Na+-dependent enhancement of NMDAR activity induced by depletion of extracellular Ca2+ may be an important mechanism underlying the development of neurotoxicity in the CNS.
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 (including 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 95. Disturbances of Ca2+ homeostasis in the cytoplasm, ER, or mitochondria can be harmful to cells 34. 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 34.
In the plasma membrane Ca2+-ATPase and the Na+/Ca2+ exchanger act to transport cytosolic Ca2+ to the extracellular space. The Na+/Ca2+ exchanger has a low affinity 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 13,22,34. 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 36. In mitochondria the Na+/Ca2+ exchanger extrudes Ca2+ 34,36,45. However, the activity of the Na+/Ca2+ exchanger may be reversed on the influx of Na+ 51. This reversal in Na+/Ca2 exchange is observed under pathological conditions 34. If the mitochondrial 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 depolarization, inhibition of ATP production, and cell death 12,29,57. In the nucleus Ca2+ is involved in the gene transcription and DNA metabolism 47,67,79. Unlike in the mitochondria, nuclear Ca2+ is found to be rapidly equilibrated with cytosolic Ca2+. This may occur by diffusion across nuclear pores 3 and/or Ca2+ channels in the nuclear envelope 65.
CICR during NMDAR activation has been reported 33,85,98. We observed that when extracellular solution contained more Na+, NMDAR activation produced greater increases in both [Na+]i and [Ca2+]i 107. Furthermore, 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 107, which depletes intracellular stores of Ca2+ by blocking Ca2+ re-uptake. In the absence of thapsigargin, NMDAR activation only produced a 35 ± 8 nM (n = 8) increase in [Ca2+]i in neurons bathed with extracellular solution containing 10 mM Na+ 107. The blockade of Ca2+ influx by removal of extracellular Ca2+ abolished the NMDAR activation-induced increase in [Ca2+]i, (data not shown) 27,99. 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 extracellular solution containing 145 mM Na+ 107. 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+ accumulation stimulates three distinct intra-axonal sources of Ca2+: (1) the mitochondrial Na+/Ca2+ exchanger driven in the Na+ import/Ca2+ export mode, (2) positive modulation of ryanodine receptors, and (3) promotion of IP3 generation by phospholipase C 75.
IV. QUESTIONS AND FUTURE STUDIES
Na+ entry is a key factor that initiates fast action potentials and shapes sub-threshold electrical properties to thereby regulate neuronal excitability and neuronal discharge activity 21,23,44,52,102. Present data have shown that: 1) intracellular Na+ up-regulates NMDARs; 2) via increasing 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 neurotransmitter 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 31. Investigations into the crystal structure of a bacterial homologue of the Na+/Cl− dependent transporters from Aquifex aeolicus revealed that there are two Na+ binding sites, named Na1 and Na2 109. 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 58,77.
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.
Acknowledgments
This work was supported by grant NIH (1R01 NS053567) to X.-M.Y.
Reference List
- 1.Agrawal SK, Fehlings MG. 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. 1996;16:545–552. doi: 10.1523/JNEUROSCI.16-02-00545.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Agrawal SK, Fehlings MG. The effect of the sodium channel blocker QX-314 on recovery after acute spinal cord injury. J. Neurotrauma. 1997;14:81–88. doi: 10.1089/neu.1997.14.81. [DOI] [PubMed] [Google Scholar]
- 3.Allbritton NL, Oancea E, Kuhn MA, Meyer T. Source of nuclear calcium signals. Proc. Natl. Acad. Sci. U. S. A. 1994;91:12458–12462. doi: 10.1073/pnas.91.26.12458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Appelgren L, Janson M, Nitescu P, Curelaru I. Continuous intracisternal and high cervical intrathecal bupivacaine analgesia in refractory head and neck pain. Anesthesiology. 1996;84:256–272. doi: 10.1097/00000542-199602000-00003. [see comments] [DOI] [PubMed] [Google Scholar]
- 5.Baker AJ, Moulton RJ, MacMillan VH, Shedden PM. Excitatory amino acids in cerebrospinal fluid following traumatic brain injury in humans. J Neurosurg. 1993;79:369–372. doi: 10.3171/jns.1993.79.3.0369. [DOI] [PubMed] [Google Scholar]
- 6.Ballard-Croft C, Carlson D, Maass DL, Horton JW. Burn trauma alters calcium transporter protein expression in the heart. J Appl. Physiol. 2004;97:1470–1476. doi: 10.1152/japplphysiol.01149.2003. [DOI] [PubMed] [Google Scholar]
- 7.Banasiak KJ, Burenkova O, Haddad GG. Activation of voltage-sensitive sodium channels during oxygen deprivation leads to apoptotic neuronal death. Neuroscience. 2004;126:31–44. doi: 10.1016/S0306-4522(03)00425-1. [DOI] [PubMed] [Google Scholar]
- 8.Baptiste DC, Fehlings MG. Update on the treatment of spinal cord injury. Prog. Brain Res. 2007;161:217–233. doi: 10.1016/S0079-6123(06)61015-7. [DOI] [PubMed] [Google Scholar]
- 9.Bauer R, Walter B, Fritz H, Zwiener U. Ontogenetic aspects of traumatic brain edema--facts and suggestions. Exp. Toxicol Pathol. 1999;51:143–150. doi: 10.1016/S0940-2993(99)80088-8. [DOI] [PubMed] [Google Scholar]
- 10.Baxter KA, Church J. Characterization of acid extrusion mechanisms in cultured fetal rat hippocampal neurones. J. Physiol. (Lond) 1996;493:457–470. doi: 10.1113/jphysiol.1996.sp021396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Benveniste H, Drejer J, Schousboe A, Diemer NH. Elevation of the extracellular concentrations of glutamate and aspartate in rat hippocampus during transient cerebral ischemia monitored by intracerebral microdialysis. J Neurochem. 1984;43:1369–1374. doi: 10.1111/j.1471-4159.1984.tb05396.x. [DOI] [PubMed] [Google Scholar]
- 12.Bernardi P, Rasola A. Calcium and cell death: the mitochondrial connection. Subcell. Biochem. 2007;45:481–506. doi: 10.1007/978-1-4020-6191-2_18. [DOI] [PubMed] [Google Scholar]
- 13.Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol. 2003;4:517–529. doi: 10.1038/nrm1155. [DOI] [PubMed] [Google Scholar]
- 14.Berridge MJ, Lipp P, Bootman MD. The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol. 2000;1:11–21. doi: 10.1038/35036035. [DOI] [PubMed] [Google Scholar]
- 15.Bhattacharjee A, et al. Slick (Slo2.1), a rapidly-gating sodium-activated potassium channel inhibited by ATP. J Neurosci. 2003;23:11681–11691. doi: 10.1523/JNEUROSCI.23-37-11681.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Bhattacharjee A, Kaczmarek LK. For K(+) channels, Na(+) is the new Ca(2+) Trends Neurosci. 2005;28:422–428. doi: 10.1016/j.tins.2005.06.003. [DOI] [PubMed] [Google Scholar]
- 17.Blaustein MP, Fontana G, Rogowski RS. The Na(+)-Ca2+ exchanger in rat brain synaptosomes. Kinetics and regulation. Ann. N. Y. Acad. Sci. 1996;779(300-17):300–317. doi: 10.1111/j.1749-6632.1996.tb44803.x. [DOI] [PubMed] [Google Scholar]
- 18.Boonstra J, et al. Ionic responses and growth stimulation induced by nerve growth factor and epidermal growth factor in rat pheochromocytoma (PC12) cells. J. Cell Biol. 1983;97:92–98. doi: 10.1083/jcb.97.1.92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Bortner CD, Cidlowski JA. Uncoupling cell shrinkage from apoptosis reveals that Na+ influx is required for volume loss during programmed cell death. J Biol Chem. 2003;278:39176–39184. doi: 10.1074/jbc.M303516200. [DOI] [PubMed] [Google Scholar]
- 20.Bredesen DE. Apoptosis: overview and signal transduction pathways. J Neurotrauma. 2000;17:801–810. doi: 10.1089/neu.2000.17.801. [In Process Citation] [DOI] [PubMed] [Google Scholar]
- 21.Cantrell AR, Catterall WA. Neuromodulation of Na+ channels: an unexpected form of cellular plasticity. Nat. Rev Neurosci. 2001;2:397–407. doi: 10.1038/35077553. [DOI] [PubMed] [Google Scholar]
- 22.Carafoli E, Santella L, Branca D, Brini M. Generation, control, and processing of cellular calcium signals. Crit Rev Biochem. Mol Biol. 2001;36:107–260. doi: 10.1080/20014091074183. [DOI] [PubMed] [Google Scholar]
- 23.Catterall WA, Goldin AL, Waxman SG. International Union of Pharmacology. XXXIX. Compendium of voltage-gated ion channels: sodium channels. Pharmacol Rev. 2003;55:575–578. doi: 10.1124/pr.55.4.7. [DOI] [PubMed] [Google Scholar]
- 24.Chen H-SV, Lipton SA. Pharmacological implications of two distinct mechanisms of interaction of memantine with NMDA-gated channels. J. Pharmacol. Exp. Ther. 2005;314:961–971. doi: 10.1124/jpet.105.085142. [DOI] [PubMed] [Google Scholar]
- 25.Choi DW. NMDA receptors and AMPA/kainate receptors mediate parallel injury in cerebral cortical cultures subjected to oxygen-glucose deprivation. Prog. Brain Res. 1993;96(137-43):137–143. doi: 10.1016/s0079-6123(08)63263-x. [DOI] [PubMed] [Google Scholar]
- 26.Choi DW. Calcium: still center-stage in hypoxic-ischemic neuronal death. Trends. Neurosci. 1995;18:58–60. [PubMed] [Google Scholar]
- 27.Clapham DE. Calcium signaling. Cell. 1995;80:259–268. doi: 10.1016/0092-8674(95)90408-5. [DOI] [PubMed] [Google Scholar]
- 28.Cox JJ, et al. An SCN9A channelopathy causes congenital inability to experience pain. Nature. 2006;444:894–898. doi: 10.1038/nature05413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Crompton M. The mitochondrial permeability transition pore and its role in cell death. Biochem. J. 1999;341(Pt 2):233–249. [PMC free article] [PubMed] [Google Scholar]
- 30.Cummins TR, Waxman SG. Downregulation of tetrodotoxin-resistant sodium currents and upregulation of a rapidly repriming tetrodotoxin-sensitive sodium current in small spinal sensory neurons after nerve injury. J. Neurosci. 1997;17:3503–3514. doi: 10.1523/JNEUROSCI.17-10-03503.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Di Cera E, et al. The Na[IMAGE] Binding Site of Thrombin. J. Biol. Chem. 1995;270:22089–22092. doi: 10.1074/jbc.270.38.22089. [DOI] [PubMed] [Google Scholar]
- 32.Dib-Hajj S, Black JA, Cummins TR, Waxman SG. NaN/Nav1.9: a sodium channel with unique properties. Trends Neurosci. 2002;25:253–259. doi: 10.1016/s0166-2236(02)02150-1. [DOI] [PubMed] [Google Scholar]
- 33.Dingledine R, Borges K, Bowie D, Traynelis SF. The glutamate receptor ion channels. Pharmacol. Rev. 1999;51:7–61. [PubMed] [Google Scholar]
- 34.Dong Z, Saikumar P, Weinberg JM, Venkatachalam MA. Calcium in cell injury and death. Annu. Rev Pathol. 2006;1:405–434. doi: 10.1146/annurev.pathol.1.110304.100218. [DOI] [PubMed] [Google Scholar]
- 35.Dryer SE. Molecular identification of the Na+-activated K+ channel. Neuron. 2003;37:727–728. doi: 10.1016/s0896-6273(03)00119-3. [DOI] [PubMed] [Google Scholar]
- 36.Duchen MR. Mitochondria and calcium: from cell signalling to cell death. J. Physiol. 2000;529(Pt 1):57–68. doi: 10.1111/j.1469-7793.2000.00057.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Ehlers MD, Zhang S, Bernhardt JP, Huganir RL. Inactivation of NMDA Receptors by direct interaction of calmodulin with the NR1 subunit. Cell. 1996;84:745–755. doi: 10.1016/s0092-8674(00)81052-1. [DOI] [PubMed] [Google Scholar]
- 38.Ekholm A, Kristian T, Siesjo BK. Influence of hyperglycemia and of hypercapnia on cellular calcium transients during reversible brain ischemia. Exp. Brain Res. 1995;104:462–466. doi: 10.1007/BF00231980. [DOI] [PubMed] [Google Scholar]
- 39.Fehlings MG, Agrawal S. Role of sodium in the pathophysiology of secondary spinal cord injury. Spine. 1995;20:2187–2191. doi: 10.1097/00007632-199510001-00002. [DOI] [PubMed] [Google Scholar]
- 40.Fiskum G. Mitochondrial participation in ischemic and traumatic neural cell death. J. Neurotrauma. 2000;17:843–855. doi: 10.1089/neu.2000.17.843. [In Process Citation] [DOI] [PubMed] [Google Scholar]
- 41.Forder JP, Tymianski M. Postsynaptic mechanisms of excitotoxicity: Involvement of postsynaptic density proteins, radicals, and oxidant molecules. Neuroscience. 2009;158:293–300. doi: 10.1016/j.neuroscience.2008.10.021. [DOI] [PubMed] [Google Scholar]
- 42.Friedman JE, Haddad GG. Anoxia induces an increase in intracellular sodium in rat central neurons in vitro. Brain Res. 1994;663:329–334. doi: 10.1016/0006-8993(94)91281-5. [DOI] [PubMed] [Google Scholar]
- 43.Giardina SF, Beart PM. Kainate receptor-mediated apoptosis in primary cultures of cerebellar granule cells is attenuated by mitogen-activated protein and cyclin-dependent kinase inhibitors. Br J Pharmacol. 2002;135:1733–1742. doi: 10.1038/sj.bjp.0704636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Goldin AL. Resurgence of sodium channel research. Annu. Rev Physiol. 2001;63:871–894. doi: 10.1146/annurev.physiol.63.1.871. [DOI] [PubMed] [Google Scholar]
- 45.Gunter TE, Yule DI, Gunter KK, Eliseev RA, Salter JD. Calcium and mitochondria. FEBS Lett. 2004;567:96–102. doi: 10.1016/j.febslet.2004.03.071. [DOI] [PubMed] [Google Scholar]
- 46.Hains BC, Saab CY, Lo AC, Waxman SG. Sodium channel blockade with phenytoin protects spinal cord axons, enhances axonal conduction, and improves functional motor recovery after contusion SCI. Exp. Neurol. 2004;188:365–377. doi: 10.1016/j.expneurol.2004.04.001. [DOI] [PubMed] [Google Scholar]
- 47.Hardingham GE, Cruzalegui FH, Chawla S, Bading H. Mechanisms controlling gene expression by nuclear calcium signals. Cell Calcium. 1998;23:131–134. doi: 10.1016/s0143-4160(98)90111-7. [DOI] [PubMed] [Google Scholar]
- 48.Hardingham NR, et al. Extracellular calcium regulates postsynaptic efficacy through group 1 metabotropic glutamate receptors. J. Neurosci. 2006;26:6337–6345. doi: 10.1523/JNEUROSCI.5128-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Harris RJ, Symon L, Branston NM, Bayhan M. Changes in extracellular calcium activity in cerebral ischaemia. J. Cereb. Blood Flow Metab. 1981;1:203–209. doi: 10.1038/jcbfm.1981.21. [DOI] [PubMed] [Google Scholar]
- 50.Heinemann U, Hamon B. Calcium and epileptogenesis. Exp. Brain Res. 1986;65:1–10. doi: 10.1007/BF00243826. [DOI] [PubMed] [Google Scholar]
- 51.Hilgemann DW, Collins A, Matsuoka S. Steady-state and dynamic properties of cardiac sodium-calcium exchange. Secondary modulation by cytoplasmic calcium and ATP. J. Gen. Physiol. 1992;100:933–961. doi: 10.1085/jgp.100.6.933. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Hille B. Ionic channels of excitable membranes. Sunderland: Sinauer; 1992. [Google Scholar]
- 53.Koch RA, Barish ME. Perturbation of intracellular calcium and hydrogen ion regulation in cultured mouse hippocampal neurons by reduction of the sodium ion concentration gradient. J. Neurosci. 1994;14:2585–2593. doi: 10.1523/JNEUROSCI.14-05-02585.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Kohr G, De Koninck Y, Mody I. Properties of NMDA receptor channels in neurons acutely isolated from epileptic (kindled) rats. J. Neurosci. 1993;13:3612–3627. doi: 10.1523/JNEUROSCI.13-08-03612.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Krupp JJ, Vissel B, Thomas CG, Heinemann SF, Westbrook GL. Interactions of calmodulin and alpha-actinin with the NR1 subunit modulate Ca2+-dependent inactivation of NMDA receptors. J. Neurosci. 1999;19:1165–1178. doi: 10.1523/JNEUROSCI.19-04-01165.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Kyrozis A, Albuquerque C, Gu J, MacDermott AB. Ca(2+)-dependent inactivation of NMDA receptors: fast kinetics and high Ca2+ sensitivity in rat dorsal horn neurons. J. Physiol. (Lond) 1996;495:449–463. doi: 10.1113/jphysiol.1996.sp021606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Lemasters JJ, Theruvath TP, Zhong Z, Nieminen AL. Mitochondrial calcium and the permeability transition in cell death. Biochim. Biophys. Acta. 2009;1787:1395–1401. doi: 10.1016/j.bbabio.2009.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Li C, Capendeguy O, Geering K, Horisberger JD. A third Na+-binding site in the sodium pump. Proc. Natl. Acad. Sci. U. S. A. 2005;102:12706–12711. doi: 10.1073/pnas.0505980102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Lieberman DN, Mody I. Regulation of NMDA channel function by endogenous Ca2+- dependent phosphatase. Nature. 1994;369:235–239. doi: 10.1038/369235a0. [DOI] [PubMed] [Google Scholar]
- 60.Lipton P. Ischemic cell death in brain neurons. Physiol Rev. 1999;79:1431–1568. doi: 10.1152/physrev.1999.79.4.1431. [DOI] [PubMed] [Google Scholar]
- 61.Lipton SA. Paradigm shift 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. 2004;6:S61–S74. doi: 10.3233/jad-2004-6s610. [DOI] [PubMed] [Google Scholar]
- 62.Lipton SA. Paradigm shift in neuroprotection by NMDA receptor blockade: memantine and beyond. Nat. Rev. Drug Discov. 2006;5:160–170. doi: 10.1038/nrd1958. [DOI] [PubMed] [Google Scholar]
- 63.Liu X, Stan LL. Sodium-activated potassium conductance participates in the depolarizing afterpotential following a single action potential in rat hippocampal CA1 pyramidal cells. Brain Res. 2004;1023:185–192. doi: 10.1016/j.brainres.2004.07.017. [DOI] [PubMed] [Google Scholar]
- 64.Mayer ML, Westbrook GL. The physiology of excitatory amino acids in the vertebrate central nervous system. Prog. Neurobiol. 1987;28:197–276. doi: 10.1016/0301-0082(87)90011-6. [DOI] [PubMed] [Google Scholar]
- 65.Mazzanti M, DeFelice LJ, Cohen J, Malter H. Ion channels in the nuclear envelope. Nature. 1990;343:764–767. doi: 10.1038/343764a0. [DOI] [PubMed] [Google Scholar]
- 66.McBain CJ, Mayer ML. N-Methyl-D-aspartic acid receptor structure and function. Physiol. Rev. 1994;74:723–760. doi: 10.1152/physrev.1994.74.3.723. [DOI] [PubMed] [Google Scholar]
- 67.McKenzie GJ, et al. Nuclear Ca2+ and CaM kinase IV specify hormonal- and Notch-responsiveness. J. Neurochem. 2005;93:171–185. doi: 10.1111/j.1471-4159.2005.03010.x. [DOI] [PubMed] [Google Scholar]
- 68.Mentzer RM, Jr, Lasley RD, Jessel A, Karmazyn M. Intracellular sodium hydrogen exchange inhibition and clinical myocardial protection. Ann. Thorac. Surg. 2003;75:S700–S708. doi: 10.1016/s0003-4975(02)04700-8. [DOI] [PubMed] [Google Scholar]
- 69.Mody I, MacDonald JF. NMDA receptor-dependent excitotoxicity: the role of intracellular Ca2+ release. Trends. Pharmacol. Sci. 1995;16:356–359. doi: 10.1016/s0165-6147(00)89070-7. [DOI] [PubMed] [Google Scholar]
- 70.Moolenaar WH, Defize LH, de Laat SW. Ionic signalling by growth factor receptors. J. Exp. Biol. 1986;124(359-73):359–373. doi: 10.1242/jeb.124.1.359. [DOI] [PubMed] [Google Scholar]
- 71.Moolenaar WH, Tsien RY, van der Saag PT, de Laat SW. Na+/H+ exchange and cytoplasmic pH in the action of growth factors in human fibroblasts. Nature. 1983;304:645–648. doi: 10.1038/304645a0. [DOI] [PubMed] [Google Scholar]
- 72.Mulkey RM, Endo S, Shenolikar S, Malenka RC. Involvement of a calcineurin/inhibitor-1 phosphatase cascade in hippocampal long-term depression. Nature. 1994;369:486–488. doi: 10.1038/369486a0. [DOI] [PubMed] [Google Scholar]
- 73.Nicholls D, Attwell D. The release and uptake of excitatory amino acids. Trends in Pharmacological Sciences. 1990;11:462–468. doi: 10.1016/0165-6147(90)90129-v. [DOI] [PubMed] [Google Scholar]
- 74.Nicholson C, Bruggencate GT, Steinberg R, Stockle H. Calcium modulation in brain extracellular microenvironment demonstrated with ion-selective micropipette. Proc. Natl. Acad. Sci. U. S. A. 1977;74:1287–1290. doi: 10.1073/pnas.74.3.1287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Nikolaeva MA, Mukherjee B, Stys PK. Na+-dependent sources of intra-axonal Ca2+ release in rat optic nerve during in vitro chemical ischemia. J Neurosci. 2005;25:9960–9967. doi: 10.1523/JNEUROSCI.2003-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Niu XW, Meech RW. Potassium inhibition of sodium-activated potassium (K(Na)) channels in guinea-pig ventricular myocytes. J Physiol. 2000;526(Pt 1):81–90. doi: 10.1111/j.1469-7793.2000.00081.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Ogawa H, Toyoshima C. Homology modeling of the cation binding sites of Na+K+-ATPase. Proc. Natl. Acad. Sci. U. S. A. 2002;99:15977–15982. doi: 10.1073/pnas.202622299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Ohyashiki T, Satoh E, Okada M, Takadera T, Sahara M. Nerve growth factor protects against aluminum-mediated cell death. Toxicology. 2002;176:195–207. doi: 10.1016/s0300-483x(02)00139-7. [DOI] [PubMed] [Google Scholar]
- 79.Papadia S, Stevenson P, Hardingham NR, Bading H, Hardingham GE. Nuclear Ca2+ and the cAMP response element-binding protein family mediate a late phase of activity-dependent neuroprotection. J. Neurosci. 2005;25:4279–4287. doi: 10.1523/JNEUROSCI.5019-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Persson L, Hillered L. Chemical monitoring of neurosurgical intensive care patients using intracerebral microdialysis. J Neurosurg. 1992;76:72–80. doi: 10.3171/jns.1992.76.1.0072. [DOI] [PubMed] [Google Scholar]
- 81.Rogawski MA, Loscher W. The neurobiology of antiepileptic drugs. Nat Rev Neurosci. 2004;5:553–564. doi: 10.1038/nrn1430. [DOI] [PubMed] [Google Scholar]
- 82.Rose CR. Na+ signals at central synapses. Neuroscientist. 2002;8:532–539. doi: 10.1177/1073858402238512. [DOI] [PubMed] [Google Scholar]
- 83.Rose CR, Konnerth A. NMDA receptor-mediated Na+ signals in spines and dendrites. J Neurosci. 2001;21:4207–4214. doi: 10.1523/JNEUROSCI.21-12-04207.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Rusakov DA, Fine A. Extracellular Ca2+ depletion contributes to fast activity-dependent modulation of synaptic transmission in the brain. Neuron. 2003;37:287–297. doi: 10.1016/s0896-6273(03)00025-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Sattler R, Tymianski M. Molecular mechanisms of calcium-dependent excitotoxicity. J Mol Med. 2000;78:3–13. doi: 10.1007/s001090000077. [DOI] [PubMed] [Google Scholar]
- 86.Sawyer TW. Practical applications of neuronal tissue culture in in vitro toxicology. Clin Exp. Pharmacol Physiol. 1995;22:295–296. doi: 10.1111/j.1440-1681.1995.tb02000.x. [DOI] [PubMed] [Google Scholar]
- 87.Schwartz G, Fehlings MG. Secondary injury mechanisms of spinal cord trauma: a novel therapeutic approach for the management of secondary pathophysiology with the sodium channel blocker riluzole. Prog. Brain Res. 2002;137:177–190. doi: 10.1016/s0079-6123(02)37016-x. [DOI] [PubMed] [Google Scholar]
- 88.Sheldon C, Diarra A, Cheng YM, Church J. Sodium influx pathways during and after anoxia in rat hippocampal neurons. J Neurosci. 2004;24:11057–11069. doi: 10.1523/JNEUROSCI.2829-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Sin WC, et al. Regulation of early neurite morphogenesis by the Na+/H+ exchanger NHE1. J. Neurosci. 2009;29:8946–8959. doi: 10.1523/JNEUROSCI.2030-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Smith MT, Thor H, Orrenius S. Toxic injury to isolated hepatocytes is not dependent on extracellular calcium. Science. 1981;213:1257–1259. doi: 10.1126/science.7268433. [DOI] [PubMed] [Google Scholar]
- 91.Soriano FX, Hardingham GE. Compartmentalized NMDA receptor signalling to survival and death. J. Physiol. 2007;584:381–387. doi: 10.1113/jphysiol.2007.138875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Strichartz G, Rando T, Wang GK. An integrated view of the molecular toxinology of sodium channel gating in excitable cells. Annu. Rev. Neurosci. 1987;10(237-67):237–267. doi: 10.1146/annurev.ne.10.030187.001321. [DOI] [PubMed] [Google Scholar]
- 93.Teng YD, Wrathall JR. Local blockade of sodium channels by tetrodotoxin ameliorates tissue loss and long-term functional deficits resulting from experimental spinal cord injury. J. Neurosci. 1997;17:4359–4366. doi: 10.1523/JNEUROSCI.17-11-04359.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Tong G, Shepherd D, Jahr CE. Synaptic desensitization of NMDA receptors by calcineurin. Science. 1995;267:1510–1512. doi: 10.1126/science.7878472. [DOI] [PubMed] [Google Scholar]
- 95.Trump BF, Berezesky IK. Calcium-mediated cell injury and cell death. FASEB J. 1995;9:219–228. doi: 10.1096/fasebj.9.2.7781924. [DOI] [PubMed] [Google Scholar]
- 96.Tymianski M, Tator CH. Normal and abnormal calcium homeostasis in neurons: a basis for the pathophysiology of traumatic and ischemic central nervous system injury. Neurosurgery. 1996;38:1176–1195. doi: 10.1097/00006123-199606000-00028. [DOI] [PubMed] [Google Scholar]
- 97.Vassilev PM, Mitchel J, Vassilev M, Kanazirska M, Brown EM. Assessment of frequency-dependent alterations in the level of extracellular Ca2+ in the synaptic cleft. Biophys. J. 1997;72:2103–2116. doi: 10.1016/S0006-3495(97)78853-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Verkhratsky A. The endoplasmic reticulum and neuronal calcium signalling. Cell Calcium. 2002;32:393–404. doi: 10.1016/s0143416002001896. [DOI] [PubMed] [Google Scholar]
- 99.Verkhratsky AJ, Petersen OH. Neuronal calcium stores. Cell Calcium. 1998;24:333–343. doi: 10.1016/s0143-4160(98)90057-4. [DOI] [PubMed] [Google Scholar]
- 100.Vornov JJ, Thomas AG, Jo D. Protective effects of extracellular acidosis and blockade of sodium/hydrogen ion exchange during recovery from metabolic inhibition in neuronal tissue culture. J. Neurochem. 1996;67:2379–2389. doi: 10.1046/j.1471-4159.1996.67062379.x. [DOI] [PubMed] [Google Scholar]
- 101.Wang X, Mori T, Sumii T, Lo EH. Hemoglobin-induced cytotoxicity in rat cerebral cortical neurons: caspase activation and oxidative stress. Stroke. 2002;33:1882–1888. doi: 10.1161/01.str.0000020121.41527.5d. [DOI] [PubMed] [Google Scholar]
- 102.Waxman SG, Dib-Hajj S, Cummins TR, Black JA. Sodium channels and their genes: dynamic expression in the normal nervous system, dysregulation in disease states(1) Brain Res. 2000;886:5–14. doi: 10.1016/s0006-8993(00)02774-8. [DOI] [PubMed] [Google Scholar]
- 103.Waxman SG, Hains BC. Fire and phantoms after spinal cord injury: Na+ channels and central pain. Trends Neurosci. 2006;29:207–215. doi: 10.1016/j.tins.2006.02.003. [DOI] [PubMed] [Google Scholar]
- 104.Wechsler A, Teichberg VI. Brain spectrin binding to the NMDA receptor is regulated by phosphorylation, calcium and calmodulin. EMBO J. 1998;17:3931–3939. doi: 10.1093/emboj/17.14.3931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Werling LL, Brown SR, Puttfarcken P, Cox BM. Sodium regulation of agonist binding at opioid receptors. II. Effects of sodium replacement on opioid binding in guinea pig cortical membranes. Mol Pharmacol. 1986;30:90–95. [PubMed] [Google Scholar]
- 106.Westbrook GL, Krupp JJ, Vissel B. Cytoskeletal interactions with glutamate receptors at central synapses. Soc. Gen. Physiol Ser. 1997;52:163–175. [PubMed] [Google Scholar]
- 107.Xin WK, et al. A functional interaction of sodium and calcium in the regulation of NMDA receptor activity by remote NMDA receptors. J Neurosci. 2005;25:139–148. doi: 10.1523/JNEUROSCI.3791-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Xin WK, et al. The removal of extracellular calcium: a novel mechanism underlying the recruitment of N-methyl-d-aspartate (NMDA) receptors in neurotoxicity. Eur. J. Neurosci. 2005;21:622–636. doi: 10.1111/j.1460-9568.2005.03888.x. [DOI] [PubMed] [Google Scholar]
- 109.Yamashita A, Singh SK, Kawate T, Jin Y, Gouaux E. Crystal structure of a bacterial homologue of Na+/Cl--dependent neurotransmitter transporters. Nature. 2005;437:215–223. doi: 10.1038/nature03978. [DOI] [PubMed] [Google Scholar]
- 110.Yu X-M. The role of intracellular sodium (Na+) in the regulation of NMDA receptor-mediated channel activity and toxicity. Mol Neurobiol. 2006;3:63–79. doi: 10.1385/MN:33:1:063. [DOI] [PubMed] [Google Scholar]
- 111.Yu X-M, Askalan R, Keil GJI, Salter MW. NMDA channel regulation by channel-associated protein tyrosine kinase Src. Science. 1997;275:674–678. doi: 10.1126/science.275.5300.674. [DOI] [PubMed] [Google Scholar]
- 112.Yu X-M, Salter MW. Gain control of NMDA-receptor currents by intracellular sodium. Nature. 1998;396:469–474. doi: 10.1038/24877. [DOI] [PubMed] [Google Scholar]
- 113.Yu X-M, Salter MW. Src, a molecular switch governing gain control of synaptic transmission mediated by N-methyl-D-aspartate receptors. Proc. Natl. Acad. Sci. U. S. A. 1999;96:7697–7704. doi: 10.1073/pnas.96.14.7697. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Yuan A, et al. The sodium-activated potassium channel is encoded by a member of the Slo gene family. Neuron. 2003;37:765–773. doi: 10.1016/s0896-6273(03)00096-5. [DOI] [PubMed] [Google Scholar]
- 115.Zhang D, et al. Na,K-ATPase activity regulates AMPA receptor turnover through proteasome-mediated proteolysis. J. Neurosci. 2009;29:4498–4511. doi: 10.1523/JNEUROSCI.6094-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Zhang S, Ehlers MD, Bernhardt JP, Su CT, Huganir RL. Calmodulin mediates calcium-dependent inactivation of N-methyl-D-aspartate receptors. Neuron. 1998;21:443–453. doi: 10.1016/s0896-6273(00)80553-x. [DOI] [PubMed] [Google Scholar]
- 117.Zipfel GJ, Babcock DJ, Lee JM, Choi DW. Neuronal apoptosis after CNS injury: the roles of glutamate and calcium. J. Neurotrauma. 2000;17:857–869. doi: 10.1089/neu.2000.17.857. [DOI] [PubMed] [Google Scholar]