Literature
首页医源资料库在线期刊分子药理学杂志2007年第69卷第11期

Regulation of Glutathione Synthesis via Interaction between Glutamate Transport-Associated Protein 3-18 (GTRAP3-18) and Excitatory Amino Acid Carrier-1 (EAAC1

来源:《分子药理学杂志》
摘要:【关键词】RegulationRegulationofthecysteinetransporterknownasexcitatoryaminoacidcarrier-1(EAAC1)forintracellularglutathione(GSH)contentwasinvestigatedusinghumanembryonickidney(HEK)293cellsasamodelsystem。GSHcontentwassignificantlyreducedbyL-aspartate--hyd......

点击显示 收起

【关键词】  Regulation

    Regulation of the cysteine transporter known as excitatory amino acid carrier-1 (EAAC1) for intracellular glutathione (GSH) content was investigated using human embryonic kidney (HEK) 293 cells as a model system. GSH content was significantly reduced by L-aspartate--hydroxamate (50–250 µM), an inhibitor of both EAAC1 and GLT1, both of which are transporters to take up cysteine, whereas dihydrokainate (1–100 µM), a specific inhibitor of GLT1, failed to do so. This indicates that EAAC1 is involved in GSH content in HEK293 cells. We examined the effect of glutamate transport-associated protein 3-18 (GTRAP3-18), which is capable of interacting with EAAC1. The GSH content decreased when the GTRAP3-18 protein level at the plasma membrane was increased by methyl--cyclodextrin (250 µM), rendering the cells more vulnerable to oxidative stress. Intracellular GSH increased when the GTRAP3-18 protein level at the plasma membrane was decreased by antisense oligonucleotides, rendering the cells more resistant to oxidative stress. Furthermore, we found that the increase in GSH content produced by stimulating protein kinase C, a translocator and activator of EAAC1, was inhibited by an increase in cell surface GTRAP3-18 protein. These results show GTRAP3-18 to negatively and dominantly regulate cellular GSH content via interaction with EAAC1 at the plasma membrane.

    GSH helps maintain the sulfhydryl groups of proteins in the reduced state and the iron heme in the ferrous state. It also serves as a reducing agent for glutaredoxin, DNA, toxic peroxides, reactive oxygen species, and free radicals. GSH is synthesized from glutamate, glycine, and cysteine. As for intracellular cysteine availability for GSH synthesis, two different mechanisms exist. One mechanism, which is found in astrocytes and immature neurons, is uptake of cystine, which is then converted to cysteine (Cho and Bannai, 1990; Murphy et al., 1990). The other mechanism, which is found in mature neurons being unable to take up cystine (Sagara et al., 1993), is direct uptake of cysteine (Sagara et al., 1993; Dringen, 2000). In cultured cells compared with tissues, cystine uptake might be down-regulated (McBean, 2002). In fact, human embryonic kidney (HEK) 293 cells have markedly low ability to take up cystine (Shih and Murphy, 2001) compared with brain tissues (Pacchioni et al., 2007).

    In mature neurons, which can not take up cystine, cysteine is the rate-limiting factor for GSH synthesis (Dringen et al., 1999; Dringen, 2000; Dringen and Hirrlinger, 2003). Cell culture studies suggest excitatory amino acid carrier-1 (EAAC1) to be a neuronal cysteine transporter (Shanker et al., 2001; Chen and Swanson, 2003; Himi et al., 2003). Moreover, Aoyama et al. (2006) demonstrated EAAC1 to be an essential transporter of cysteine needed for GSH synthesis, as evidenced by EAAC1 gene-deficient mice displaying both a low level of neuronal GSH and vulnerability to oxidative stresses. EAAC1 is a member of the family of sodium-dependent excitatory amino acid transporters (EAATs). EAAC1 is widely expressed in neurons in the mature brain (Rothstein et al., 1994), but its contribution to glutamate reuptake from the synaptic cleft is so minuscule (reviewed in Danbolt, 2001) that its major function may be cysteine transport. The structural requirements of EAAC1 for glutamate and cysteine transport seem to be different, because point mutations in the EAAC1 primary structure result in clear dissociation of the transport capability for each amino acid (Bendahan et al., 2000). GTRAP3-18 (glutamate transport-associated protein for EAAC1) is a membrane-associated protein that interacts with EAAC1 (Lin et al., 2001) and negatively modulates EAAC1-mediated glutamate reuptake in vitro as well as in vivo (Lin et al., 2001; Butchbach et al., 2002, 2003). However, because differential structural conformation of EAAC1 is required for cysteine and glutamate transport, whether GTRAP3-18 is capable of regulating cysteine uptake and intracellular GSH content remains to be established.

    To answer this question, we used HEK293 cells, which express only EAAC1 among EAATs (Lin et al., 2001) and have little ability to take up cystine. We demonstrated GTRAP3-18 to dominantly and negatively determine the intracellular GSH content.

    Materials. L-Aspartate--hydroxamate (LAH), methyl--cyclodextrin (MeCD), 4,9,12,13,20-pentahydroxytiglia-1,6-dien-3-one 12-tetradecanoate 13-acetate (PMA), 4,9,12,13,20-pentahydroxytiglia-1,6-dien-3-one 12-tetradecanoate 13-acetate (4-PMA), quisqualate, and anti-actin antibody were purchased from Sigma-Aldrich (St. Louis, MO). DL-threo--benzyloxyaspartate (TBOA), and dihydrokainate (DHK) are from TOCRIS (Bristol, UK). Anti-EAAC1 antibody was obtained from Alpha Diagnostic International (San Antonio, TX) and anti-GTRAP3-18 antibody was from Trans Genic Inc (Hyogo, Japan).

    Cell Culture. HEK293 cells were grown in minimum essential medium supplemented with 10% fetal calf serum at 37°C under 5% CO2 in air.

    Detection of GSH. GSH concentration in HEK293 cells was determined using ThioGlo-1 (Calbiochem, San Diego, CA), a maleimide reagent that produces a highly fluorescent adduct upon reaction with thiol groups. GSH content was estimated from the fluorescence response via the interaction of ThioGlo-1 mainly with intracellular GSH. Cells were incubated at 37°C for 30 min with 10 µM ThioGlo-1. After washing with phosphate-buffered saline to remove excess nonreacted ThioGlo-1, the level of fluorescence was measured using a Multimode Detector DTX800 (Beckman Coulter, Fullerton, CA).

    Transfection of GTRAP3-18 Antisense Oligonucleotides. HEK293 cells were transiently transfected with sense (GTGAACCTTGCCCCGCTC) or antisense (GAGCGGGGCAAGGTTCAC) GTR-AP3-18 oligonucleotides using SuperFect (QIAGEN, Valencia, CA), as described previously (Watabe et al., 2004).

    Immunoblot Analysis. Immunoblotting was performed as described previously (Watabe et al., 1996). Cells were lysed in buffer containing SDS and mercaptoethanol, and the cell lysate was then boiled. Denatured proteins were separated on polyacrylamide gel and transferred to a polyvinylidene difluoride membrane (GE Healthcare, Chalfont St. Giles, Buckinghamshire, UK). The membrane was incubated with a blocking solution (2% bovine serum albumin dissolved in phosphate-buffered saline containing 0.2% Tween 20) for 1 h at room temperature and incubated with a first antibody dissolved in blocking solution overnight at 4°C. After washing, the membrane was incubated for 1 h with horseradish-linked secondary antibody. Immunoreactive proteins were detected with an enhanced chemiluminescence system (GE Healthcare). Band intensities were measured using Scion Image release beta 4.0.3 (Scion Corporation, Frederick, MD).

    Cellular Cholesterol Assay. The cells were extracted by sonication with 1% Triton X-100 in chloroform. After centrifugation at 10,000g for 10 min, organic phase was collected and dried. Dried lipids were used for measurement. Cholesterol assay was measured using a cholesterol quantitation kit (BioVision, Mountain View, CA) according to the manufacturer's directions, except that cholesterol esterase was omitted from reaction mixture.

    Quantification of DNA Fragmentation. DNA fragmentation was measured using a Cell Death Detection ELISAPLUS kit (Roche Diagnostics, Indianapolis, IN) as described previously (Watabe and Nakaki, 2004).

    Immunofluorescence Microscopy. As described previously (Watabe et al., 2000), cells were washed with phosphate-buffered saline and fixed with 3.7% formaldehyde for 20 min. Cells were permeabilized with phosphate-buffered saline containing 0.2% Triton X-100 for 5 min and then washed three times with phosphate-buffered saline. Incubation with primary antibody was carried out for 1 h at room temperature. Excess antibody was washed out three times with phosphate-buffered saline. This was followed by incubation with an appropriate fluorophore-labeled secondary antibody for 1 h at room temperature in an area shielded from light. After washing out the excess antibody three times with phosphate-buffered saline, coverslips were mounted using a ProLong Antifade Kit (Invitrogen, Carlsbad, CA). Fluorescent images were obtained using a Zeiss fluorescence microscope (Zeiss, Oberkochen, Germany) and an inverted laser-scanning fluorescent microscope MRC-1024 using Laser Sharp 2000 software (Bio-Rad Laboratories, Tokyo, Japan).

    Fig. 1. Effects of inhibitors of EAAT and cystine transporter on GSH content in HEK293 cells. After the cells had been treated with TBOA (A), DHK (B), LAH (C), or quisqualate (D) at the indicated concentrations for 3 h, GSH assay was performed. *, p < 0.05 compared with control.

    Cell Surface Biotinylation. Labeling of proteins on the plasma membrane was accomplished by cell surface biotinylation using a Pinpoint Cell Surface Protein Isolation Kit (Pierce, Rockford, IL) in accordance with the manufacturer's instructions.

    Statistics. Values are mean ± S.E. from three experiments. Statistical analysis of the data was performed using analysis of variance followed by Fisher's test. A p value < 0.05 was considered significant.

    We first ascertained whether GSH content is dependent on EAAC1 in HEK293 cells using EAAT inhibitors. The nonspecific EAAT inhibitor TBOA dose dependently decreased the intracellular GSH content (Fig. 1A). The GLT-1 specific inhibitor DHK failed to decrease the intracellular GSH content at a concentration sufficient to inhibit GLT-1-mediated glutamate uptake (Fig. 1B). LAH, which inhibits both GLT-1 and EAAC1, decreased the intracellular GSH content (Fig. 1C). Furthermore, the cystine transporter inhibitor quisqualate failed to decrease the intracellular GSH content (Fig. 1D). Therefore, EAAC1 mediates cysteine uptake for GSH synthesis in the cells.

    Fig. 2. Effects of MeCD and GTRAP3-18 antisense oligonucleotides on GTRAP3-18 expression and GSH content in HEK293 cells. A, after cell lysate preparation, immunoblot analysis was performed using anti-GTRAP3-18 –specific antibody. B, the cells were treated with MeCD at the indicated concentrations or transiently transfected with GTRAP3-18 antisense or sense oligonucleotides at the indicated concentrations. Two days later, immunoblot analysis was performed. C, the cells were transiently transfected with 10 µM GTRAP3-18 sense (S) or antisense (As) oligonucleotides, or treated with 250 µMMeCD. Two days later, immunoblot analysis was performed. D, the cells were transiently transfected with 10 µM GTRAP3-18 sense (S) or antisense (As) oligonucleotides and then treated with 250 µM MeCD. Two days later, GSH assay was performed. *, p < 0.05 compared with control. E, after the cells had been transiently transfected with 10 µM GTRAP3-18 sense (Sense) or antisense (Antisense) oligonucleotides, cells were treated with 250 µM MeCD. Two days later, GSH assay was performed.

    We next examined the effect of intracellular GTRAP3-18 level on cellular GSH content. To increase GTRAP3-18 expression, the cells were treated with MeCD, which increases GTRAP3-18 expression (Butchbach et al., 2003). As shown in Fig. 2A, GTRAP3-18 expression was examined using anti-GTRAP3-18 –specific antibody. MeCD increased the level of endogenous GTRAP3-18 protein in cells, as expected (Fig. 2B, left). To decrease GTRAP3-18 expression, we used GTRAP3-18 antisense oligonucleotides (Lin et al., 2001) and observed a specific reduction in the level of endogenous GTRAP3-18 protein (Fig. 2B, right). The EAAC1 protein level was not affected by MeCD or GTRAP3-18 antisense oligonucleotide treatment (Fig. 2C). Under these experimental conditions, intracellular GSH was increased concomitantly with a reduction in GTRAP3-18 protein level, whereas GSH content was decreased concomitantly with an increase in GTRAP3-18 protein level (Fig. 2D). When MeCD-increased GTRAP3-18 was decreased by GTRAP3-18 antisense oligonucleotides, the GSH level was restored (Fig. 2E). MeCD has a high affinity for cholesterol and has been shown to promote the efflux of cholesterol from cells (Kilsdonk et al., 1995). To examine whether the efflux of cholesterol caused the decrease in the GSH content, we quantitated the cholesterol content. The cholesterol content was slightly reduced by MeCD at the concentration that increased GTRAP3-18 amount (Fig. 3A). However, GTRAP3-18 antisense oligonucleotides, which restored the GSH level decreased by MeCD, did not affect the cholesterol content compared with GTRAP3-18 sense oligonucleotides (Fig. 3B). This result indicates that the GSH level was not reduced by cholesterol efflux but by MeCD-elevated GTRAP3-18. Moreover, an increased level of GTRAP3-18 rendered the cells more vulnerable to oxidative stress, such as hydrogen peroxide (Fig. 4A), and a decreased level of GTRAP3-18 rendered the cells more resistant to hydrogen peroxide (Fig. 4B). These results show that GTRAP3-18 negatively regulates the intracellular GSH content, which in turn affects susceptibility to oxidative stresses such as hydrogen peroxide.

    Fig. 3. Effects of MeCD and GTRAP3-18 antisense oligonucleotides on cholesterol content in HEK293 cells. A, after the cells had been treated with MeCD at the indicated concentrations for 2 days, cholesterol assay was performed. B, after the cells had been transiently transfected with 10 µM GTRAP3-18 sense (Sense) or antisense (Antisense) oligonucleotides, cells were treated with 250 µMMeCD. Two days later, cholesterol assay was performed.

    Fig. 4. Effects of MeCD and GTRAP3-18 antisense oligonucleotides on hydrogen peroxide-induced apoptosis in HEK293 cells. A, after the cells had been treated with 250 µM MeCD for 2 days, they were treated with hydrogen peroxide at the indicated concentrations for 24 h and DNA fragmentation assay was performed. B, the cells were transiently transfected with 10 µM GTRAP3-18 sense (S) or antisense (As) oligonucleotides. Two days later, they were treated with 75 µM hydrogen peroxide for 24 h, and DNA fragmentation assay was performed. *, p < 0.05 compared with control.

    Immunocytochemical analysis revealed that GTRAP3-18 in control cells was present in both the plasma membrane and the intracellular compartment, and colocalized with EAAC1 (Fig. 5). In MeCD-treated cells, GTRAP3-18 immunoreactivity was augmented in both the cell membrane and the intracellular compartment, whereas MeCD did not alter the expression of EAAC1 protein (Fig. 5A). Cell surface EAAC1-associated GTRAP3-18 was increased by MeCD. We ascertained that cell surface EAAC1-associated GTRAP3-18 was increased by MeCD using another technique, a cell surface biotinylation assay. There was an increase in not only nonsurface (nonbiotinylated) but also surface (biotinylated) GTRAP3-18 level by MeCD (Fig. 6). Protein kinase C activation is known to positively regulate cell surface expression of EAAC1 and activation of glutamate uptake by EAAC1 (González et al., 2002, 2003; Fournier et al., 2004). Therefore, we examined the effect of protein kinase C activation on GTRAP3-18 expression and GSH level in MeCD-treated cells. PMA, a protein kinase C activator, induced an increase in cell surface EAAC1 level (Figs. 5 and 6). GSH content was elevated concomitantly with the increase in surface EAAC1 level by PMA, whereas 4-PMA, which is inactive on protein kinase C, did not increase the GSH content (Fig. 7B). Because the inhibition of EAAC1 activity by LAH suppressed the GSH content by elevated PMA, the PMA-elevated GSH content was mediated through EAAC1 activity (Fig. 7, C and D). Treatment of MeCD-treated cells with PMA induced a large increase in cell-surface-colocalized EAAC1 and GTRAP3-18 (Figs. 5, A and B, and 6). It is noteworthy, however, that the PMA-induced increase in GSH content was inhibited by the MeCD-induced increase in cell surface GTRAP3-18 protein (Fig. 7).

    Fig. 5. Immunofluorescence microscopy showing changes in EAAC1 and GTRAP3-18 localization by MeCD and PMA treatment of HEK293 cells. The cells were treated with 250 µMMeCD for 2 days and then with 1 µM PMA for 3 h. Intracellular localization of EAAC1 (green) and GTRAP3-18 (red) was examined by immunofluorescence microscopy using a fluorescent microscope (A) and a confocal laser-scanning fluorescent microscope (B).

    Fig. 6. Biotinylation showing changes in EAAC1 and GTRAP3-18 on cell surface by MeCD and PMA treatment of HEK293 cells. Cells were treated with 250 µMMeCD for 2 days and then with 1 µM PMA for 3 h. After biotinylation of intact cells, immunoblotting of biotinylated (cell surface) and nonbiotinylated (intracellular) fractions were performed using each specific antibody. Actin was measured to determine the degree of intracellular protein labeling by the biotin reagent.

    Fig. 7. Changes in GSH content by MeCD and PMA treatment of HEK293 cells. A, after the cells had been treated with MeCD at the indicated concentrations for 2 days and with 1 µM PMA for 3 h, GSH assay was performed. B, after the cells had been treated with 250 µMMeCD for 2 days and with PMA or 4-PMA at the indicated concentrations for 3 h, GSH assay was performed. C, after pretreatment with LAH at the indicated concentrations, the cells were treated with 1 µM PMA for 3 h and GSH assay was performed. D, after pretreatment with 250 µMLAH, the cells were treated with PMA at the indicated concentrations for 3 h, and GSH assay was performed.

    After the recent discovery of EAAC1 as a major neuroprotective molecule (Aoyama et al., 2006), we explored the mechanisms underlying the regulation of EAAC1 activity determining GSH content. Our results have demonstrated a cellular protein, GTRAP3-18, to dominantly and negatively regulate EAAC1 activity and determine the intracellular GSH content. In the brain, EAAC1 is the primary neuronal transporter among members of the EAAT family (Rothstein et al., 1994). To explore the mechanisms regulating EAAC1 activity on GSH content, we attempted to construct an in vitro neuronal EAAC1 model system. We judged that the use of a primary neuronal culture system would be unsuitable for our purpose because the expression not only of EAAC1 but also other EAATs, which are not expressed in neurons, is induced by preparation of a primary culture of neurons (Himi et al., 2003). Therefore, we used HEK293 cells, because they stably express only EAAC1 among members of the EAAT family (Lin et al., 2001), have markedly low ability to take up cystine via the cystine transporter (Shih and Murphy, 2001), are a cell line derived from humans, and are thus suitable as an in vitro neuronal EAAC1 model system. We used MeCD to increase GTRAP3-18 expression. MeCD increases endogenous GTRAP3-18 in HEK293 cells (Butchbach et al., 2003). Our study showed that MeCD reduced the GSH content via an increase in GTRAP3-18. Moreover, this increased GTRAP3-18 rendered the cells more vulnerable to oxidative stress. MeCD has high affinity for cholesterol and has been shown to promote efflux of cholesterol from the cell (Kilsdonk et al., 1995). There is no evidence for an association of EAAC1 with cholesterol efflux (Butchbach et al., 2004), and the mechanisms underlying the MeCD-induced increase in GTRAP3-18 are poorly understood. We examined the change in cholesterol content under our experimental condition and found that the reduction of cholesterol content by MeCD was slight. Previous reports suggest that the MeCD concentration required to increase GTRAP3-18 is lower than that required to cause cholesterol depletion (Kilsdonk et al., 1995; Butchbach et al., 2004). Moreover, we showed that GSH was reduced by MeCD-increased GTRAP3-18 but not by cholesterol efflux. On the other hand, we used an antisense oligonucleotide technique to down-regulate GTRAP3-18 expression rather than short interfering RNA, which can produce serious "off target" consequences. GTRAP3-18 antisense oligonucleotides specifically reduced endogenous GTRAP3-18 protein level and concomitantly increased intracellular GSH. This decreased level of GTRAP3-18 rendered the cells more resistant to hydrogen peroxide. These results clearly show GTRAP3-18 to negatively regulate the intracellular GSH content, which in turn affects susceptibility to oxidative stresses such as hydrogen peroxide.

    Morphine has long been known to diminish intracellular GSH (Roberts et al., 1987), but the mechanisms underlying this GSH depletion are unknown. A murine homolog of GTRAP3-18 was identified as a factor that is up-regulated in the basomedial amygdala in repeatedly morphine-administered mice (Ikemoto et al., 2002). Based on our results, the morphine-induced GSH depletion is possibly caused by induction of GTRAP3-18 on the plasma membrane, resulting in negative regulation of EAAC1.

    Each EAAT family member is chemically modified by various stimuli. Oxygen radicals and hydrogen peroxide induce persistent inhibition of EAATs, probably via direct interaction with the transport process (Volterra et al., 1994). Nitric oxide generators such as sodium nitroprusside and S-nitroso-N-acetylpenicillamine decrease glutamate uptake into the synaptosome (Pogun et al., 1994). Peroxynitrite, formed by the combination of superoxide anion and nitric oxide, inhibits glutamate uptake by neuronal transporter EAAC1 (Trotti et al., 1996). However, the functional significance of these chemical modifications of EAAC1 remains unknown. Another modification of EAAC1 involves phosphorylation. In a tumor cell line, the cell surface expression and activity of EAAC1 seem to be regulated by several phosphorylation pathways. A protein kinase C-mediated pathway is known to positively regulate cell surface expression and activation of glutamate uptake by EAAC1 (Danbolt, 2001; González et al., 2002, 2003; Fournier et al., 2004; Huang et al., 2006). In particular, Huang et al. (2006) reported EAAC1 to be regulated by protein kinase C. Protein kinase C belongs to a classic subtype activated by diacylglycerol, which is produced by phospholipase C (Newton, 2001). Gq-coupled receptors, among many G protein-coupled receptors such as 1 adrenergic receptors, M1 muscarinic receptors, and H1 histaminergic receptors, cause activation of phospholipase C via the  subunit of Gq, which is activated by its ligand binding (Zhou et al., 1994). Because Hsu et al. (2005) reported that epinephrine increased the GSH level, activation of EAAC1 via phosphorylation by protein kinase C is possibly caused by activation of these Gq-coupled receptors, resulting in positive regulation of EAAC1 and GSH synthesis.

    EAAC1 is mainly localized in the intracellular compartment, with approximately 20% in the plasma membrane (Nieoullon et al., 2006). Phosphorylation by protein kinase C induces translocation of EAAC1 from the intracellular compartment to the plasma membrane and expression of its function as an amino acid transporter. On the other hand, GTRAP3-18 is also present mainly in the intracellular compartment and partially at the plasma membrane via binding to EAAC1 (Lin et al., 2001). Therefore, we examined the effect of protein kinase C activation on both GTRAP3-18 expression and GSH level in MeCD-treated cells. Confirming the findings of Lin et al. (2001), GTRAP3-18 in control cells was present in both the plasma membrane and the intracellular compartment, and colocalized with EAAC1. We further demonstrated that PMA, a protein kinase C activator, induced an increase in cell surface EAAC1 level and a concomitant increase in GSH content. This result suggests that protein kinase C up-regulates not only glutamate but also cysteine uptake by EAAC1. Moreover, treatment of MeCD-treated cells with PMA induced a large increase in cell-surface-colocalized EAAC1 and GTRAP3-18 and decreased the GSH content. It is noteworthy that the PMA-induced increase in GSH content was inhibited by the MeCD-induced increase in plasma membrane GTRAP3-18 protein. GTRAP3-18 associated with EAAC1 in the plasma membrane dominantly and negatively regulated cysteine up-take for GSH synthesis, and determined intracellular GSH content even if protein kinase C, which activates EAAC1, was activated. The phosphorylation of serine 465 in EAAC1 by protein kinase C is reportedly important for both the increase in EAAC1 activity and redistribution to the plasma membrane (Huang et al., 2006). Lin et al. (2001) reported that GTRAP3-18 was identified by a yeast two-hybrid screen system using the C-terminal intracellular domain (arginine 438 –phenylalanine 524) of EAAC1 (Kanai and Hediger, 1992; Lin et al., 2001; Yernool et al., 2004). Therefore, serine 465 of EAAC1, which is phosphorylated by protein kinase C, is located within the binding domain for GTRAP3-18. This, together with our results, indicates that GTRAP3-18 inhibits EAAC1 activity by masking the serine 465 residue, which is the site of phosphorylation by protein kinase C. Therefore, it is possible that a putative inhibitory compound against GTRAP3-18 would be an efficient GSH-increasing agent.

    Because the GSH content in discrete brain areas is reduced in patients with Parkinson's or Alzheimer's disease (Dringen and Hirrlinger, 2003), GTRAP3-18 is a potential therapeutic target for increasing the neuronal GSH level. The discovery of a GTRAP3-18 inhibitory compound that increases neuronal GSH would contribute to developing novel therapeutic strategies to protect neurons in patients with neurodegenerative disorders, including Parkinson's and Alzheimer's diseases.

    ABBREVIATIONS: GSH, glutathione; HEK, human embryonic kidney; EAAC1, excitatory amino acid carrier-1; EAAT, excitatory amino acid transporter; GTRAP3-18, glutamate transport-associated protein for EAAC1; LAH, L-aspartate--hydroxamate; MeCD, methyl--cyclodextrin; PMA, 4,9,12,13,20-pentahydroxytiglia-1,6-dien-3-one 12-tetradecanoate 13-acetate; 4-PMA, 4,9,12,13,20-pentahydroxytiglia-1,6-dien-3-one 12-tetradecanoate 13-acetate; TBOA, DL-threo--benzyloxyaspartate; DHK, dihydrokainate.

【参考文献】
  Aoyama K, Suh SW, Hamby AM, Liu J, Chan WY, Chen Y, and Swanson RA (2006) Neuronal glutathione deficiency and age-dependent neurodegeneration in the EAAC1 deficient mouse. Nat Neurosci 9: 119–126.

Bendahan A, Armon A, Madani N, Kavanaugh MP, and Kanner BI (2000) Arginine 447 plays a pivotal role in substrate interactions in a neuronal glutamate transporter. J Biol Chem 275: 37436–37442.[Abstract/Free Full Text]

Butchbach ME, Guo H, and Lin CL (2003) Methyl-beta-cyclodextrin but not retinoic acid reduces EAAT3-mediated glutamate uptake and increases GTRAP3-18 expression. J Neurochem 84: 891–894.

Butchbach ME, Lai L, and Lin CL (2002) Molecular cloning, gene structure, expression profile and functional characterization of the mouse glutamate transporter (EAAT3) interacting protein GTRAP3-18. Gene 292: 81–90.

Butchbach ME, Tian G, Guo H, and Lin CL (2004) Association of excitatory amino acid transporters, especially EAAT2, with cholesterol-rich lipid raft microdomains: importance for excitatory amino acid transporter localization and function. J Biol Chem 279: 34388–34396.[Abstract/Free Full Text]

Chen Y and Swanson RA (2003) The glutamate transporters EAAT2 and EAAT3 mediate cysteine uptake in cortical neuron cultures. J Neurochem 84: 1332–1339.

Cho Y and Bannai S (1990) Uptake of glutamate and cysteine in C-6 glioma cells and in cultured astrocytes. J Neurochem 55: 2091–2097.

Danbolt NC (2001) Glutamate uptake. Prog Neurobiol 65: 1–105.

Dringen R (2000) Metabolism and functions of glutathione in brain. Prog Neurobiol 62: 649–671.

Dringen R and Hirrlinger J (2003) Glutathione pathways in the brain. Biol Chem 384: 505–516.

Dringen R, Pfeiffer B, and Hamprecht B (1999) Synthesis of the antioxidant glutathione in neurons: supply by astrocytes of CysGly as precursor for neuronal glutathione. J Neurosci 19: 562–569.[Abstract/Free Full Text]

Fournier KM, González MI, and Robinson MB (2004) Rapid trafficking of the neuronal glutamate transporter, EAAC1: evidence for distinct trafficking pathways differentially regulated by protein kinase C and platelet-derived growth factor. J Biol Chem 279: 34505–34513.[Abstract/Free Full Text]

González MI, Bannerman PG, and Robinson MB (2003) Phorbol myristate acetate-dependent interaction of protein kinase Calpha and the neuronal glutamate transporter EAAC1. J Neurosci 23: 5589–5593.[Abstract/Free Full Text]

González MI, Kazanietz MG, and Robinson MB (2002) Regulation of the neuronal glutamate transporter excitatory amino acid carrier-1 (EAAC1) by different protein kinase C subtypes. Mol Pharmacol 62: 901–910.[Abstract/Free Full Text]

Himi T, Ikeda M, Yasuhara T, Nishida M, and Morita I (2003) Role of neuronal glutamate transporter in the cysteine uptake and intracellular glutathione levels in cultured cortical neurons. J Neural Transm 110: 1337–1348.

Hsu DZ, Wang ST, Deng JF, and Liu MY (2005) Epinephrine protects against severe acute gastric bleeding in rats: role of nitric oxide and glutathione. Shock 23: 253–257.

Huang Y, Feng X, Sando JJ, and Zuo Z (2006) Critical role of serine 465 in isoflurane-induced increase of cell-surface redistribution and activity of glutamate transporter type 3. J Biol Chem 281: 38133–38138.[Abstract/Free Full Text]

Ikemoto MJ, Inoue K, Akiduki S, Osugi T, Imamura T, Ishida N, and Ohtomi M (2002) Identification of addicsin/GTRAP3-18 as a chronic morphine-augmented gene in amygdala. Neuroreport 13: 2079–2084.

Kanai Y and Hediger MA (1992) Primary structure and functional characterization of a high-affinity glutamate transporter. Nature 360: 467–471.

Kilsdonk EP, Yancey PG, Stoudt GW, Bangerter FW, Johnson WJ, Phillips MC, and Rothblat GH (1995) Cellular cholesterol efflux mediated by cyclodextrins. J Biol Chem 270: 17250–17256.[Abstract/Free Full Text]

Lin CI, Orlov I, Ruggiero AM, Dykes-Hoberg M, Lee A, Jackson M, and Rothstein JD (2001) Modulation of the neuronal glutamate transporter EAAC1 by the interacting protein GTRAP3-18. Nature 410: 84–88.

McBean GJ (2002) Cerebral cystine uptake: a tale of two transporters. Trends Pharmacol Sci 23: 299–302.

Murphy TH, Schnaar RL, and Coyle JT (1990) Immature cortical neurons are uniquely sensitive to glutamate toxicity by inhibition of cystine uptake. Faseb J 4: 1624–1633.

Newton AC (2001) Protein kinase C: structural and spatial regulation by phosphorylation, cofactors, and macromolecular interactions. Chem Rev 101: 2353–2364.

Nieoullon A, Canolle B, Masmejean F, Guillet B, Pisano P, and Lortet S (2006) The neuronal excitatory amino acid transporter EAAC1/EAAT3: does it represent a major actor at the brain excitatory synapse? J Neurochem 98: 1007–1018.

Pacchioni AM, Vallone J, Melendez RI, Shih A, Murphy TH, and Kalivas PW (2007) Nrf2 gene deletion fails to alter psychostimulant-induced behavior or neurotoxicity. Brain Res 1127: 26–35.

Pogun S, Dawson V, and Kuhar MJ (1994) Nitric oxide inhibits 3H-glutamate transport in synaptosomes. Synapse 18: 21–26.

Roberts SM, Skoulis NP, and James RC (1987) A centrally-mediated effect of morphine to diminish hepatocellular glutathione. Biochem Pharmacol 36: 3001–3005.

Rothstein JD, Martin L, Levey AI, Dykes-Hoberg M, Jin L, Wu D, Nash N, and Kuncl RW (1994) Localization of neuronal and glial glutamate transporters. Neuron 13: 713–725.

Sagara JI, Miura K, and Bannai S (1993) Maintenance of neuronal glutathione by glial cells. J Neurochem 61: 1672–1676.

Shanker G, Allen JW, Mutkus LA, and Aschner M (2001) The uptake of cysteine in cultured primary astrocytes and neurons. Brain Res 902: 156–163.

Shih AY and Murphy TH (2001) xCt cystine transporter expression in HEK293 cells: pharmacology and localization. Biochem Biophys Res Commun 282: 1132–1137.

Trotti D, Rossi D, Gjesdal O, Levy LM, Racagni G, Danbolt NC, and Volterra A (1996) Peroxynitrite inhibits glutamate transporter subtypes. J Biol Chem 271: 5976–5979.[Abstract/Free Full Text]

Volterra A, Trotti D, Tromba C, Floridi S, and Racagni G (1994) Glutamate uptake inhibition by oxygen free radicals in rat cortical astrocytes. J Neurosci 14: 2924–2932.

Watabe M, Hishikawa K, Takayanagi A, Shimizu N, and Nakaki T (2004) Caffeic acid phenethyl ester induces apoptosis by inhibition of NFkappaB and activation of Fas in human breast cancer MCF-7 cells. J Biol Chem 279: 6017–6026.[Abstract/Free Full Text]

Watabe M, Machida K, and Osada H (2000) MT-21 is a synthetic apoptosis inducer that directly induces cytochrome c release from mitochondria. Cancer Res 60: 5214–5222.[Abstract/Free Full Text]

Watabe M, Masuda Y, Nakajo S, Yoshida T, Kuroiwa Y, and Nakaya K (1996) The cooperative interaction of two different signaling pathways in response to bufalin induces apoptosis in human leukemia U937 cells. J Biol Chem 271: 14067–14072.[Abstract/Free Full Text]

Watabe M and Nakaki T (2004) Rotenone induces apoptosis via activation of bad in human dopaminergic SH-SY5Y cells. J Pharmacol Exp Ther 311: 948–953.[Abstract/Free Full Text]

Yernool D, Boudker O, Jin Y, and Gouaux E (2004) Structure of a glutamate transporter homologue from Pyrococcus horikoshii. Nature 431: 811–818.

Zhou CJ, Akhtar RA, and Abdel-Latif AA (1994) Identification of phosphoinositide-specific phospholipase C-beta 1 and GTP-binding protein, Gq alpha, in bovine iris sphincter membranes: characteristics of the phospholipase and its coupling to cholinergic muscarinic receptors. Exp Eye Res 59: 377–384.


作者单位:Department of Pharmacology, Teikyo University School of Medicine, Tokyo, Japan

作者: 2009-8-25
医学百科App—中西医基础知识学习工具
  • 相关内容
  • 近期更新
  • 热文榜
  • 医学百科App—健康测试工具