Skip to main content
Erschienen in: Molecular Pain 1/2007

Open Access 01.12.2007 | Research

Inward currents induced by ischemia in rat spinal cord dorsal horn neurons

verfasst von: Meng Chen, Yuan-Xiang Tao, Jianguo G Gu

Erschienen in: Molecular Pain | Ausgabe 1/2007

download
DOWNLOAD
print
DRUCKEN
insite
SUCHEN

Abstract

Hypoxia and ischemia occur in the spinal cord when blood vessels of the spinal cord are compressed under pathological conditions such as spinal stenosis, tumors, and traumatic spinal injury. Here by using spinal cord slice preparations and patch-clamp recordings we investigated the influence of an ischemia-simulating medium on dorsal horn neurons in deep lamina, a region that plays a significant role in sensory hypersensitivity and pathological pain. We found that the ischemia-simulating medium induced large inward currents in dorsal horn neurons recorded. The onset of the ischemia-induced inward currents was age-dependent, being onset earlier in older animals. Increases of sensory input by the stimulation of afferent fibers with electrical impulses or by capsaicin significantly speeded up the onset of the ischemia-induced inward currents. The ischemia-induced inward currents were abolished by the glutamate receptor antagonists CNQX (20 μM) and APV (50 μM). The ischemia-induced inward currents were also substantially inhibited by the glutamate transporter inhibitor TBOA (100 μM). Our results suggest that ischemia caused reversal operation of glutamate transporters, leading to the release of glutamate via glutamate transporters and the subsequent activation of glutamate receptors in the spinal dorsal horn neurons.
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​1744-8069-3-10) contains supplementary material, which is available to authorized users.

Competing interests

The author(s) declare that they have no competing interests.

Background

Glutamate is the principle neurotransmitter that mediates sensory transmission in the spinal cord dorsal horn. Under physiological conditions, glutamate is released synaptically by primary afferent fibers, descending terminals from supraspinal regions, and excitatory interneurons in the spinal cord dorsal horn [1]. The synaptically released glutamate is rapidly taken up through glutamate transporters located at presynaptic terminals, postsynaptic cells, and on the surrounding glia cells [25]. These transporters keep extracellular glutamate at low levels to ensure high fidelity sensory transmission, to limit nonspecific neuronal excitation and hyperactivity, and to prevent excitatory toxicity [3, 6].
Increased glutamate concentrations in extracellular spaces can occur as a consequence of CNS tissue injury, which in turn can produce neuronal hyperactivity and secondary neuronal tissue damage due to excitatory toxicity [7]. It has been shown that extracellular glutamate levels increased significantly in the brain following ischemic and hypoxic injury [8, 9]. In the spinal cord, ischemia and hypoxia can occur under a number of pathological conditions including traumatic spinal cord injury, tumors within the spinal cord, spinal stenosis, cardiac arrest, massive hemorrhagic shock, and surgical procedures [1014]. These conditions often cause spinal blood vessel compression, resulting in spinal cord ischemia and hypoxia. Similar to the brain, spinal cord ischemia and hypoxia also can result in the increases of extracellular glutamate levels to cause neuronal excitatory toxicity in the spinal cord. When these pathological processes occur in the dorsal horn of the spinal cord, sensory functions may be significantly altered to result in pathological pain states.
Leak of glutamate from damaged cells and release of glutamate from synaptic sites were thought to contribute to the elevation of extracellular glutamate concentrations under pathological conditions. However, studies have suggested that a change of glutamate transporter function plays a critical role in the sustained elevation of extracellular glutamate levels during ischemia and hypoxia [9, 15]. Under physiological conditions, glutamate transporters co-transport one glutamate molecule and 3 Na+ ions into the cell to maintain the concentration gradient of micromolar extracellular glutamate against millimolar intracellular glutamate [16, 17]. This active transport function is supported by the transmembrane ion gradients established by Na+-K+ ATPase [16, 17]. Under pathological conditions, for example, during brain ischemia and hypoxia, ATP is depleted and Na+-K+ ATPase function is impaired. This subsequently results in the loss of transmembrane ion gradients and thereby reducing the driving force for the active uptake of glutamate from extracelular glutamate [15]. In fact, studies using brain tissues suggested that the depletion of intracellular energy not only compromises glutamate uptake, but also can result in glutamate release through glutamate transporter system due to the reversal operation of the glutamate transporters [9].
In the present study, we tested the hypothesis that ischemic condition results in the reversal operation of glutamate transport system to cause glutamate release and subsequent excitation of sensory neurons in the spinal cord dorsal horn. The study may have implications in pathological pain states associated with ischemic and hypoxic conditions in the spinal cord dorsal horn [5].

Results

The ischemic condition was produced by the bath application of a modified Kreb's solution that did not contain glucose and was bubbled with N2 gas to deoxygenate the solution. The bath solution also contained 1 mM sodium cyanide to block glycolysis and oxidative phosphorylation [9]. When this ischemia-simulating medium [18, 19] was perfused to the spinal cord slice preparations, we recorded large inward currents (ischemia-induced inward currents) from lamina V neurons of the spinal cord dorsal horn (Fig 1A). The onset time of the ischemia-induced inward currents showed large variations. When animals at the age of 6 days old were used, the onset of the ischemia-induced inward currents was at 22 ± 1 min (n = 6, Figure 1A&1B). The ischemia-induced inward currents reached peak levels rapidly and then gradually decayed to baseline levels in recordings when the spinal cord slices were prepared from these young animals (Figure 1A). The onset time of the ischemia-induced inward current became shorter when older animals were used. For animals at the age of 10 days old, the onset of the ischemia-induced inward currents was at 18.5 ± 0.5 min (n = 6). When animals at the age of 20 days old were used, the onset time was 8 ± 0.4 min (n = 6), 3 times shorter than the onset time for animals at the age of 6 days old (Figure 1). We observed that, in older animals, the ischemia-induced inward currents usually did not return to baseline and membrane seals on patch-clamp electrodes were eventually lost (Figure 1A). The age-dependence in the onset of the ischemia-induced inward currents was observed in the range from 4 days to 20 days old (Figure 1B); animals over 20 days old were not examined in this study. The peak amplitude of the ischemia-induced inward currents, on the other hand, was not found to be age-dependence. Overall, the peak amplitude was 720 ± 38 pA (n = 43)
We tested whether sensory input affects the ischemia-induced inward currents. Sensory input into spinal cord dorsal horn was elicited by electrical stimulation of primary afferent fibers at the intensity of 100 μA and frequency of 100 Hz. In this set of experiments, all the spinal cord slices were prepared from animals at the age of 9 days old. When primary afferent fibers were not stimulated electrically, the onset of the ischemia-induced inward currents was at 20 ± 0.6 min (n = 6) following the application of ischemic solution (Figure 2A&2C). On the other hand, the onset time was 16 ± 0.3 (n = 4) min when primary afferent fibers were repeatedly stimulated at the frequency of 100 Hz (Figure 2B&2C), significantly shorter than the onset time when primary afferent fibers were not stimulated (P < 0.05, Figure 2C). There was no significant differences in the peak current amplitude between the experiments without electrical stimulation (1400 ± 350 pA, n = 6) and the experiments with electrical stimulation (1359 ± 345 pA, n = 4).
We tested whether stimulation of TRPV1-expressing nociceptive afferent fibers with capsaicin also affect ischemia-induced inward currents. In this set of experiments, all the spinal cord slices were prepared from animals at the age of 7 days. In experiments when capsaicin was not applied, ischemia-induced inward currents had onset time of 22 ± 1.8 min (n = 6) following the application of the ischemic Kreb's solution (Figure 3A&3C). When capsaicin (2 μM) was included in the ischemic solution, the onset time of the inward currents was 18 ± 0.2 min (n = 4; Fig 3B&3C), significantly shorter than the onset time of the experiments when capsaicin was absent (P < 0.05, Figure 3C). There was no significant differences in the peak current amplitude between the experiments without capsaicin stimulation (805 ± 395 pA, n = 6) and the experiments with capsaicin stimulation (765 ± 120 pA, n = 4).
We explored mechanisms by which ischemia induced inward currents. In this set of experiments, all animals used were at the age of 10 days old. The ischemia-induced inward currents had an onset time of ~17 min (Figure 4A, also see Figure 1B) with peak currents of 646 ± 150 pA (n = 4). In the presence of glutamate receptor antagonists CNQX (20 μM, for non-NMDA receptors) and APV (50 μM, for NMDA receptors), ischemia-induced inward currents were almost completely abolished (16 ± 2 pA, n = 6, p > 0.05; Figure 4B&4D). In the presence of DL-threo-β-benzyloxyaspartate (TBOA, 100 μM), a potent glutamate transporter inhibitor, the ischemia-induced inward currents were also largely inhibited (Figure 4C&4D). The ischemia-induced inward currents were 77 ± 36 pA (n = 9) in the presence of TBOA, only 12% of the ischemia-induced inward currents in those experiments when TBOA was absent (Figure 4D).
We determined whether ischemic conditions have a significant effect on inward currents evoked by exogenously applied glutamate. In this set of experiments, glutamate (500 μM, 1 min) was first applied for 1 min under the normal Kreb's bath condition. We then switched the normal bath solution to ischemic bath solution. Following a 10-min perfusion of ischemic bath solution, a second application of glutamate (500 μM, 1 min) was tested. This application was made prior to the onset of the ischemia-induced inward currents (Figure 5A). We found that the inward current evoked by the second glutamate application was 250 ± 65% of the inward current induced by the first glutamate application (p < 0.05, n = 6) (Figure 5A&5C). In contrast, when tissues were maintained at normal Kreb's bath solution, the inward current elicited by the second application of glutamate was usually slightly smaller than the current elicited by the first glutamate application (78 ± 8%, p > 0.05, n = 6, Figure 5B&5C).

Discussion

In this study, we show that ischemia induced significant inward currents in dorsal horn neurons of the spinal cord and that the inward currents could be completely blocked by ionotropic glutamate receptor inhibitors as well as by glutamate transporter inhibitors. This observation is consistent with a previous study using brain slice preparations showing that glutamate transporter inhibitors blocked ischemia-induced inward currents [9]. These findings suggest that ischemia may cause reversal operation of glutamate transporters, leading to the release of glutamate via glutamate transporters and the subsequent activation of glutamate receptors in the neurons of the spinal dorsal horn and brain. Our study, however, cannot exclude other mechanisms that may contribute to ischemia-induced membrane depolarization as shown in hippocampal neurons [18, 19].
We observed that the onset of ischemia-induced inward currents took many minutes. This delay in the onset suggests that intracellular energy depletion is a slow process under our experimental conditions and transporter reversal does not occur before intracellular energy is substantially depleted. Interestingly, we have found that the onset times of ischemia-induced currents are age-dependent, being shorter in older animals and longer in younger ones. One explanation for the age-dependence is that the rates of intracellular energy depletion may be different between younger and older animals under ischemic conditions. The onset times of ischemia-induced currents were shortened when primary afferent fibers were stimulated electrically or with the noxious stimulant capsaicin. This is mort likely due to fast energy depletion when primary afferent fibers were stimulated. In addition to the demonstration of ischemia-induced inward currents, we have also showed that inward currents evoked by exogenous glutamate were significantly larger under ischemia condition than under normal condition. This is probably because the actual concentrations of exogenous glutamate that reached the recorded neurons were higher under ischemia condition than under normal condition. This result suggest that prior to the reversal operation of glutamate transporters, glutamate uptake is severely compromised under ischemia condition. Glutamate transporters are expressed on both neuronal and glia cells in the spinal cord dorsal horn [4, 5]. To date, five subtypes of glutamate transporters have been cloned: GLAST (EAAT1), GLT-1 (EAAT2), EAAC-1 (EAAT3), EAAT4 and EAAT5 [6]. GLAST, and GLT-1 are predominantly localized in astrocytes [2, 20, 21], while EAAC-1, EAAT4, and EAAT5 appears to be mostly neuronal [2225]. While both neuronal and glial glutamate transporters actively participate in the uptake of extracellular glutamate [26, 27]. Glutamate-induced excitatory toxicity under ischemia conditions appears to be mainly due to the impairment of glial glutamate transporters [3, 28, 29].
Previous studies have demonstrated that TBOA blocks glutamate uptake under physiological conditions [30]. We showed that ischemia-induced glutamate release was inhibited by glutamate transporter inhibitor TBOA in the present study. These results together suggest that TBOA bi-directionally blocks glutamate transporters. Effects of TBOA on sensory behaviors have previously been studied in both normal animals and animals with pathological pain conditions. In normal animals, intrathecal injection of TBOA was shown to induce nociceptive behaviors, such as licking, shaking, and caudally directed biting [31]. These effects were thought due to the block of glutamate uptake by TBOA, which subsequently results in the elevation of extracellular glutamate levels to cause hyperactivity in the spinal cord dorsal horn neurons [31]. Interestingly, under pathological pain conditions, glutamate transporter inhibitors were found to produce anti-nociceptive effects. For example, glutamate transporter inhibitors were shown to attenuate the induction of allodynia induced by PGE2, PGF2a, and NMDA [32], to reduced formalin-induced nociceptive responses, and to attenuate Complete Freund's adjuvant (CFA)-evoked thermal hyperalgesia [5, 33]. In addition, transient knockdown of spinal GLT-1 led to significant reduction of nociceptive behavior in the formalin model [5, 33]. It has been proposed that pathological pain conditions may cause a depletion of intracellular energy in the spinal cord dorsal horn, which subsequently reverses glutamate transporters to release glutamate and to produce hyperactivity in the spinal cord dorsal horn neurons [5]. Therefore, by blocking glutamate release from its transporters, glutamate transporter inhibitors may execute an anti-nociceptive effect under pathological conditions [5]. The present in vitro study supports the rationales for the use of glutamate transporter inhibitors.

Methods

Principles of laboratory animal care (NIH publication No. 86-23, revised 1985) were followed in all the experiments described in the present study. Spinal cord slice preparations and patch-clamp recordings were described in details in our previous studies [34]. In brief, Sprague Dawley rats at the postnatal age of 4–21 days were used. Transverse spinal cord slices were prepared from lumbar enlargement of the spinal cords. The thickness of each slice was 400 μm. The spinal cord slices were maintained in a basket submerged in ~200 ml Krebs solution at 24°C. The Krebs solution contained (in mM): NaCl 117, KCl 3.6, CaCl2 2.5, MgCl2 1.2, NaH2PO4 1.2, NaHCO3 25 and glucose 11; the solution was saturated with 95 % O2 and 5% CO2 and had pH of 7.4. In each experiment, a spinal cord slice was transferred to a recording chamber and placed on the stage of an upright IR-DIC microscope. The slice was perfused at ~10 ml/min with the Krebs solution and all experiments were performed at room temperature (24°C). To induce ischemia, the slices were perfused by a ischemic bath solution, which is modified Kreb's solution that had no glucose (replaced with 7 mM sucrose) and contained 1 mM sodium cyanide [9], and the bath solution was deoxygenated by continuously bubbling with 100% nitrogen.
Individual neurons were identified under an IR-DIC microscope with a 40× water immersion objective. Whole cell patch-clamp recordings were made in deep lamina (lamina V) neurons with electrodes that were filled with an internal solution contained (mM): Cs2SO4 110, TEA-Cl 5, CaCl2 0.5, MgCl2 2, EGTA 5, HEPES 5, pH 7.3. The resistance of electrodes was ~5 MΩ when filled with the internal solution. The access resistance was below 30 MΩ and was not compensated. Signals was amplified and filtered at 2 kHz (Axopatch 200B) and sampled at 5 kHz using pCLAMP 7.0 (Axon Instruments). In all recordings, neurons were voltage-clamped at -30 mV. In some experiments, primary afferent fibers were stimulated by focal electrical stimulation at the dorsal root entry zone or by chemical stimulation with capsaicin. For the focal electrical stimulation, the stimulation intensity was 100 μA and the stimulation frequency was 100 Hz. For chemical stimulation, capsaicin at the concentration of 2 μM was bath applied to stimulate capsaicin-sensitive nociceptive afferent fibers. Pharmacological tests, including the tests of CNQX (20 μM), APV (50 μM), and TBOA (100 μM) were performed by applications of these compound through bath solution. Unless otherwise indicated, all recordings were performed in the presence of 20 μM bicucullin and 2 μM strychnine.
Data were presented as mean ± SEM. Student's t-tests were used for statistical analysis and significance was considered at the level of the p < 0.05.
Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://​creativecommons.​org/​licenses/​by/​2.​0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Competing interests

The author(s) declare that they have no competing interests.
Literatur
1.
Zurück zum Zitat Yoshimura M, Jessell T: Amino acid-mediated EPSPs at primary afferent synapses with substantia gelatinosa neurones in the rat spinal cord. J Physiol 1990, 430: 315–35.PubMedCentralPubMedCrossRef Yoshimura M, Jessell T: Amino acid-mediated EPSPs at primary afferent synapses with substantia gelatinosa neurones in the rat spinal cord. J Physiol 1990, 430: 315–35.PubMedCentralPubMedCrossRef
2.
Zurück zum Zitat Rothstein JD, Martin L, Levey AI, Dykes-Hoberg M, Jin L, Wu D, Nash N, Kuncl RW: Localization of neuronal and glial glutamate transporters. Neuron 1994, 13: 713–725. 10.1016/0896-6273(94)90038-8PubMedCrossRef Rothstein JD, Martin L, Levey AI, Dykes-Hoberg M, Jin L, Wu D, Nash N, Kuncl RW: Localization of neuronal and glial glutamate transporters. Neuron 1994, 13: 713–725. 10.1016/0896-6273(94)90038-8PubMedCrossRef
3.
Zurück zum Zitat Anderson CM, Swanson RA: Astrocyte glutamate transport: review of properties, regulation, and physiological functions. Glia 2000, 32: 1–14. 10.1002/1098-1136(200010)32:1<1::AID-GLIA10>3.0.CO;2-WPubMedCrossRef Anderson CM, Swanson RA: Astrocyte glutamate transport: review of properties, regulation, and physiological functions. Glia 2000, 32: 1–14. 10.1002/1098-1136(200010)32:1<1::AID-GLIA10>3.0.CO;2-WPubMedCrossRef
4.
Zurück zum Zitat Tao F, Liaw WJ, Zhang B, Yaster M, Rothstein JD, Johns RA, Tao YX: Evidence of neuronal excitatory amino acid carrier 1 expression in rat dorsal root ganglion neurons and their central terminals. Neuroscience 2004, 123: 1045–51. 10.1016/j.neuroscience.2003.11.026PubMedCrossRef Tao F, Liaw WJ, Zhang B, Yaster M, Rothstein JD, Johns RA, Tao YX: Evidence of neuronal excitatory amino acid carrier 1 expression in rat dorsal root ganglion neurons and their central terminals. Neuroscience 2004, 123: 1045–51. 10.1016/j.neuroscience.2003.11.026PubMedCrossRef
5.
Zurück zum Zitat Tao YX, Gu JG, Stephens RL: Role of spinal cord glutmate transporter during normal sensory transmission and pathological pain states. Molecular Pain 2005, 1: 30. 10.1186/1744-8069-1-30PubMedCentralPubMedCrossRef Tao YX, Gu JG, Stephens RL: Role of spinal cord glutmate transporter during normal sensory transmission and pathological pain states. Molecular Pain 2005, 1: 30. 10.1186/1744-8069-1-30PubMedCentralPubMedCrossRef
6.
Zurück zum Zitat Danbolt NC: Glutamate uptake. Prog Neurobiol 2001, 65: 1–105. 10.1016/S0301-0082(00)00067-8PubMedCrossRef Danbolt NC: Glutamate uptake. Prog Neurobiol 2001, 65: 1–105. 10.1016/S0301-0082(00)00067-8PubMedCrossRef
7.
Zurück zum Zitat Choi DW: Glutamate neurotoxicity and diseases of the nervous system. Neuron 1998, 1: 623–634. 10.1016/0896-6273(88)90162-6CrossRef Choi DW: Glutamate neurotoxicity and diseases of the nervous system. Neuron 1998, 1: 623–634. 10.1016/0896-6273(88)90162-6CrossRef
8.
Zurück zum Zitat Phillis JW, Smith-Barbour M, Perkins LM, O'Regan MH: Characterization of glutamate, aspartate, and GABA release from ischemic rat cerebral cortex. Brain Res Bull 1994, 34: 457–466. 10.1016/0361-9230(94)90019-1PubMedCrossRef Phillis JW, Smith-Barbour M, Perkins LM, O'Regan MH: Characterization of glutamate, aspartate, and GABA release from ischemic rat cerebral cortex. Brain Res Bull 1994, 34: 457–466. 10.1016/0361-9230(94)90019-1PubMedCrossRef
9.
Zurück zum Zitat Rossi DJ, Oshima T, Attwell D: Glutamate release in severe brain ischaemia is mainly by reversed uptake. Nature 2000,403(6767):316–21. 10.1038/35002090PubMedCrossRef Rossi DJ, Oshima T, Attwell D: Glutamate release in severe brain ischaemia is mainly by reversed uptake. Nature 2000,403(6767):316–21. 10.1038/35002090PubMedCrossRef
10.
Zurück zum Zitat Carlson GD, Gorden C: Current developments in spinal cord injury research. Spine J 2002,2(2):116–28. 10.1016/S1529-9430(01)00029-8PubMedCrossRef Carlson GD, Gorden C: Current developments in spinal cord injury research. Spine J 2002,2(2):116–28. 10.1016/S1529-9430(01)00029-8PubMedCrossRef
11.
Zurück zum Zitat Bal S, Celiker R, Palaoglu S, Cila A: F wave studies of neurogenic intermittent claudication in lumbar spinal stenosis. Am J Phys Med Rehabil 2006,85(2):135–40. 10.1097/01.phm.0000197586.91860.a1PubMedCrossRef Bal S, Celiker R, Palaoglu S, Cila A: F wave studies of neurogenic intermittent claudication in lumbar spinal stenosis. Am J Phys Med Rehabil 2006,85(2):135–40. 10.1097/01.phm.0000197586.91860.a1PubMedCrossRef
12.
Zurück zum Zitat Kudo Y, Ohtaki H, Dohi K, Yin L, Nakamachi T, Endo S, Yofu S, Hiraizumi Y, Miyaoka H, Shioda S: Neuronal damage in rat brain and spinal cord after cardiac arrest and massive hemorrhagic shock. Crit Care Med 2006,34(11):2820–6. 10.1097/01.CCM.0000242522.48734.64PubMedCrossRef Kudo Y, Ohtaki H, Dohi K, Yin L, Nakamachi T, Endo S, Yofu S, Hiraizumi Y, Miyaoka H, Shioda S: Neuronal damage in rat brain and spinal cord after cardiac arrest and massive hemorrhagic shock. Crit Care Med 2006,34(11):2820–6. 10.1097/01.CCM.0000242522.48734.64PubMedCrossRef
13.
Zurück zum Zitat Novy J, Carruzzo A, Maeder P, Bogousslavsky J: Spinal cord ischemia: clinical and imaging patterns, pathogenesis, and outcomes in 27 patients. Arch Neurol 2006,63(8):1113–20. 10.1001/archneur.63.8.1113PubMedCrossRef Novy J, Carruzzo A, Maeder P, Bogousslavsky J: Spinal cord ischemia: clinical and imaging patterns, pathogenesis, and outcomes in 27 patients. Arch Neurol 2006,63(8):1113–20. 10.1001/archneur.63.8.1113PubMedCrossRef
14.
Zurück zum Zitat Traul DE, Shaffrey ME, Schiff D: Part I: spinal-cord neoplasms-intradural neoplasms. Lancet Oncol 2007, 8: 35–45. 10.1016/S1470-2045(06)71009-9PubMedCrossRef Traul DE, Shaffrey ME, Schiff D: Part I: spinal-cord neoplasms-intradural neoplasms. Lancet Oncol 2007, 8: 35–45. 10.1016/S1470-2045(06)71009-9PubMedCrossRef
15.
Zurück zum Zitat Jabaudon D, Scanziani M, Gähwiler BH, Gerber U: Acute decrease in net glutamate uptake during energy deprivation. Proc Natl Acad Sci USA 2000, 97: 5610–5615. 10.1073/pnas.97.10.5610PubMedCentralPubMedCrossRef Jabaudon D, Scanziani M, Gähwiler BH, Gerber U: Acute decrease in net glutamate uptake during energy deprivation. Proc Natl Acad Sci USA 2000, 97: 5610–5615. 10.1073/pnas.97.10.5610PubMedCentralPubMedCrossRef
16.
Zurück zum Zitat Zerangue N, Kavanaugh MP: Flux coupling in a neuronal glutamate transporter. Nature 1996, 383: 634–7. 10.1038/383634a0PubMedCrossRef Zerangue N, Kavanaugh MP: Flux coupling in a neuronal glutamate transporter. Nature 1996, 383: 634–7. 10.1038/383634a0PubMedCrossRef
17.
Zurück zum Zitat Levy LM, Warr O, Attwell D: Stoichiometry of the glial glutamate transporter GLT-1 expressed inducibly in a Chinese hamster ovary cell line selected for low endogenous Na + -dependent glutamate uptake. J Neurosci 1998, 18: 9620–9628.PubMed Levy LM, Warr O, Attwell D: Stoichiometry of the glial glutamate transporter GLT-1 expressed inducibly in a Chinese hamster ovary cell line selected for low endogenous Na + -dependent glutamate uptake. J Neurosci 1998, 18: 9620–9628.PubMed
18.
Zurück zum Zitat Tanaka E, Yamamoto S, Kudo Y, Mihara S, Higashi H: Mechanisms underlying the rapid depolarization produced by deprivation of oxygen and glucose in rat hippocampal CA1 neurons in vitro. J Neurophysiol 1997, 78: 891–902.PubMed Tanaka E, Yamamoto S, Kudo Y, Mihara S, Higashi H: Mechanisms underlying the rapid depolarization produced by deprivation of oxygen and glucose in rat hippocampal CA1 neurons in vitro. J Neurophysiol 1997, 78: 891–902.PubMed
19.
Zurück zum Zitat Yamamoto S, Tanaka E, Shoji Y, Kudo Y, Inokuchi H, Higashi H: Factors that reverse the persistent depolarization produced by deprivation of oxygen and glucose in rat hippocampal CA1 neurons in vitro. J Neurophysiol 1997, 78: 903–11.PubMed Yamamoto S, Tanaka E, Shoji Y, Kudo Y, Inokuchi H, Higashi H: Factors that reverse the persistent depolarization produced by deprivation of oxygen and glucose in rat hippocampal CA1 neurons in vitro. J Neurophysiol 1997, 78: 903–11.PubMed
20.
Zurück zum Zitat Danbolt NC, Storm-Mathisen J, Kanner BI: An [Na + + K + ] coupled L-glutamate transporter purified from rat brain is located in glial cell processes. Neuroscience 1992, 51: 295–310. 10.1016/0306-4522(92)90316-TPubMedCrossRef Danbolt NC, Storm-Mathisen J, Kanner BI: An [Na + + K + ] coupled L-glutamate transporter purified from rat brain is located in glial cell processes. Neuroscience 1992, 51: 295–310. 10.1016/0306-4522(92)90316-TPubMedCrossRef
21.
Zurück zum Zitat Lehre KP, Levy LM, Ottersen OP, Storm-Mathisen J, Danbolt NC: Differential expression of two glial glutamate transporters in the rat brain: quantitative and immunocytochemical observations. J Neurosci 1995, 15: 1835–53.PubMed Lehre KP, Levy LM, Ottersen OP, Storm-Mathisen J, Danbolt NC: Differential expression of two glial glutamate transporters in the rat brain: quantitative and immunocytochemical observations. J Neurosci 1995, 15: 1835–53.PubMed
22.
Zurück zum Zitat Berger UV, Hediger MA: Comparative analysis of glutamate transporter expression in rat brain using differential double in situ hybridization. Anat Embryol 1998, 198: 13–30. 10.1007/s004290050161PubMedCrossRef Berger UV, Hediger MA: Comparative analysis of glutamate transporter expression in rat brain using differential double in situ hybridization. Anat Embryol 1998, 198: 13–30. 10.1007/s004290050161PubMedCrossRef
23.
Zurück zum Zitat Kugler P, Schmitt A: Glutamate transporter EAAC1 is expressed in neurons and glial cells in the rat nervous system. Glia 1999, 27: 129–42. 10.1002/(SICI)1098-1136(199908)27:2<129::AID-GLIA3>3.0.CO;2-YPubMedCrossRef Kugler P, Schmitt A: Glutamate transporter EAAC1 is expressed in neurons and glial cells in the rat nervous system. Glia 1999, 27: 129–42. 10.1002/(SICI)1098-1136(199908)27:2<129::AID-GLIA3>3.0.CO;2-YPubMedCrossRef
24.
Zurück zum Zitat Dehnes Y, Chaudhry FA, Ullensvang K, Lehre KP, Storm-Mathisen J, Danbolt NC: The glutamate transporter EAAT4 in rat cerebellar Purkinje cells: a glutamate-gated chloride channel concentrated near the synapse in parts of the dendritic membrane facing astroglia. J Neurosci 1998, 18: 3606–19.PubMed Dehnes Y, Chaudhry FA, Ullensvang K, Lehre KP, Storm-Mathisen J, Danbolt NC: The glutamate transporter EAAT4 in rat cerebellar Purkinje cells: a glutamate-gated chloride channel concentrated near the synapse in parts of the dendritic membrane facing astroglia. J Neurosci 1998, 18: 3606–19.PubMed
25.
Zurück zum Zitat Arriza JL, Eliasof S, Kavanaugh MP, Amara SG: Excitatory amino acid transporter 5, a retinal glutamate transporter coupled to a chloride conductance. Proc Natl Acad Sci USA 1997, 94: 4155–60. 10.1073/pnas.94.8.4155PubMedCentralPubMedCrossRef Arriza JL, Eliasof S, Kavanaugh MP, Amara SG: Excitatory amino acid transporter 5, a retinal glutamate transporter coupled to a chloride conductance. Proc Natl Acad Sci USA 1997, 94: 4155–60. 10.1073/pnas.94.8.4155PubMedCentralPubMedCrossRef
26.
Zurück zum Zitat Mennerick S, Shen W, Xu W, Benz A, Tanaka K, Shimamoto K, Isenberg KE, Krause JE, Zorumski CF: Substrate turnover by transporters curtails synaptic glutamate transients. J Neurosci 1999, 19: 9242–51.PubMed Mennerick S, Shen W, Xu W, Benz A, Tanaka K, Shimamoto K, Isenberg KE, Krause JE, Zorumski CF: Substrate turnover by transporters curtails synaptic glutamate transients. J Neurosci 1999, 19: 9242–51.PubMed
27.
Zurück zum Zitat Lievens JC, Bernal F, Forni C, Mahy N, Kerkerian-Le Goff L: Characterization of striatal lesions produced by glutamate uptake alteration: cell death, reactive gliosis, and changes in GLT1 and GADD45 mRNA expression. Glia 2000, 29: 222–32. 10.1002/(SICI)1098-1136(20000201)29:3<222::AID-GLIA4>3.0.CO;2-0PubMedCrossRef Lievens JC, Bernal F, Forni C, Mahy N, Kerkerian-Le Goff L: Characterization of striatal lesions produced by glutamate uptake alteration: cell death, reactive gliosis, and changes in GLT1 and GADD45 mRNA expression. Glia 2000, 29: 222–32. 10.1002/(SICI)1098-1136(20000201)29:3<222::AID-GLIA4>3.0.CO;2-0PubMedCrossRef
28.
Zurück zum Zitat Kanai Y, Hediger MA, Wang Y, Schielke JP, Welty DF: Knockout of glutamate transporters reveals a major role for astroglial transporter in excitotoxicity and clearance of glutamate. Neuron 1996, 16: 675–686. 10.1016/S0896-6273(00)80086-0PubMedCrossRef Kanai Y, Hediger MA, Wang Y, Schielke JP, Welty DF: Knockout of glutamate transporters reveals a major role for astroglial transporter in excitotoxicity and clearance of glutamate. Neuron 1996, 16: 675–686. 10.1016/S0896-6273(00)80086-0PubMedCrossRef
29.
Zurück zum Zitat Mitani A, Tanaka K: Functional changes of glial glutamate transporter GLT-1 during ischemia: an in vivo study in the hippocampal CA1 of normal mice and mutant mice lacking GLT-1. J Neurosci 2003, 23: 7176–7182.PubMed Mitani A, Tanaka K: Functional changes of glial glutamate transporter GLT-1 during ischemia: an in vivo study in the hippocampal CA1 of normal mice and mutant mice lacking GLT-1. J Neurosci 2003, 23: 7176–7182.PubMed
30.
Zurück zum Zitat Shimamoto K, LeBrun B, Yasuda-kamatani Y, Sakaitani M, Shigeri Y, Yumoto N, Nakajima T: DL-threo-β-benzyloxyaspartate, a potent blocker of excitatory amino acid transporters. Mol Pharmacol 1998, 53: 195–201.PubMed Shimamoto K, LeBrun B, Yasuda-kamatani Y, Sakaitani M, Shigeri Y, Yumoto N, Nakajima T: DL-threo-β-benzyloxyaspartate, a potent blocker of excitatory amino acid transporters. Mol Pharmacol 1998, 53: 195–201.PubMed
31.
Zurück zum Zitat Liaw WJ, Stephens RL Jr, Zhang B, Sepkuty J, Cho EE, Johns RA, Rothstein JD, Tao YX: Spinal glutamate uptake is critical for maintaining normal sensory transmission in rat spinal cord. Pain 2005, 115: 60–70. 10.1016/j.pain.2005.02.006PubMedCrossRef Liaw WJ, Stephens RL Jr, Zhang B, Sepkuty J, Cho EE, Johns RA, Rothstein JD, Tao YX: Spinal glutamate uptake is critical for maintaining normal sensory transmission in rat spinal cord. Pain 2005, 115: 60–70. 10.1016/j.pain.2005.02.006PubMedCrossRef
32.
Zurück zum Zitat Minami T, Matsumura S, Okuda-Ashitaka E, Shimamoto K, Sakimura K, Mishina M, Mori H, Ito S: Characterization of the glutamatergic system for induction and maintenance of allodynia. Brain Res 2001, 895: 178–185. 10.1016/S0006-8993(01)02069-8PubMedCrossRef Minami T, Matsumura S, Okuda-Ashitaka E, Shimamoto K, Sakimura K, Mishina M, Mori H, Ito S: Characterization of the glutamatergic system for induction and maintenance of allodynia. Brain Res 2001, 895: 178–185. 10.1016/S0006-8993(01)02069-8PubMedCrossRef
33.
Zurück zum Zitat Niederberger E, Schmidtko A, Rothstein JD, Geisslinger G, Tegeder I: Modulation of spinal nociceptive processing through the glutamate transporter GLT-1. Neuroscience 2003, 116: 81–87. 10.1016/S0306-4522(02)00547-XPubMedCrossRef Niederberger E, Schmidtko A, Rothstein JD, Geisslinger G, Tegeder I: Modulation of spinal nociceptive processing through the glutamate transporter GLT-1. Neuroscience 2003, 116: 81–87. 10.1016/S0306-4522(02)00547-XPubMedCrossRef
34.
Zurück zum Zitat Nakatsuka T, Gu JG: ATP P2X receptor-mediated enhancement of glutamate release and evoked EPSCs in dorsal horn neurons of the rat spinal cord. J Neurosci 2001, 21: 6522–6531.PubMed Nakatsuka T, Gu JG: ATP P2X receptor-mediated enhancement of glutamate release and evoked EPSCs in dorsal horn neurons of the rat spinal cord. J Neurosci 2001, 21: 6522–6531.PubMed
Metadaten
Titel
Inward currents induced by ischemia in rat spinal cord dorsal horn neurons
verfasst von
Meng Chen
Yuan-Xiang Tao
Jianguo G Gu
Publikationsdatum
01.12.2007
Verlag
BioMed Central
Erschienen in
Molecular Pain / Ausgabe 1/2007
Elektronische ISSN: 1744-8069
DOI
https://doi.org/10.1186/1744-8069-3-10

Weitere Artikel der Ausgabe 1/2007

Molecular Pain 1/2007 Zur Ausgabe

Bei schweren Reaktionen auf Insektenstiche empfiehlt sich eine spezifische Immuntherapie

Insektenstiche sind bei Erwachsenen die häufigsten Auslöser einer Anaphylaxie. Einen wirksamen Schutz vor schweren anaphylaktischen Reaktionen bietet die allergenspezifische Immuntherapie. Jedoch kommt sie noch viel zu selten zum Einsatz.

Hinter dieser Appendizitis steckte ein Erreger

23.04.2024 Appendizitis Nachrichten

Schmerzen im Unterbauch, aber sonst nicht viel, was auf eine Appendizitis hindeutete: Ein junger Mann hatte Glück, dass trotzdem eine Laparoskopie mit Appendektomie durchgeführt und der Wurmfortsatz histologisch untersucht wurde.

Ärztliche Empathie hilft gegen Rückenschmerzen

23.04.2024 Leitsymptom Rückenschmerzen Nachrichten

Personen mit chronischen Rückenschmerzen, die von einfühlsamen Ärzten und Ärztinnen betreut werden, berichten über weniger Beschwerden und eine bessere Lebensqualität.

Mehr Schaden als Nutzen durch präoperatives Aussetzen von GLP-1-Agonisten?

23.04.2024 Operationsvorbereitung Nachrichten

Derzeit wird empfohlen, eine Therapie mit GLP-1-Rezeptoragonisten präoperativ zu unterbrechen. Eine neue Studie nährt jedoch Zweifel an der Notwendigkeit der Maßnahme.

Update AINS

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.