首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 203 毫秒
1.
Diphosphoinositol pentakisphosphate (InsP7) and bis-diphosphoinositol tetrakisphosphate contain pyrophosphate bonds. InsP7 is formed from inositol hexakisphosphate (InsP6) by a family of three inositol hexakisphosphate kinases (InsP6K). In this study we establish one of the InsP6Ks, InsP6K2, as a physiologic mediator of cell death. Overexpression of wild-type InsP6K2 augments the cytotoxic actions of multiple cell stressors in diverse cell lines, whereas transfection with a dominant negative InsP6K2 decreases cell death. During cell death, InsP6 kinase activity is enhanced, and intracellular InsP7 level is augmented. Deletion of InsP6K2 but not the other forms of InsP6K diminishes cell death, suggesting that InsP6K2 is the major InsP6 kinase involved in cell death. Cytotoxicity is associated with a translocation of InsP6K2 from nuclei to mitochondria, whereas the intracellular localization of the other isoforms of the enzyme does not change. The present study provides compelling evidence that endogenous InsP6K2, by generating InsP7, provides physiologic regulation of the apoptotic process.  相似文献   

2.
The inositol pyrophosphate disphosphoinositol pentakisphosphate (PP-InsP(3)/InsP(7)) is formed in mammals by two recently cloned inositol hexakiphosphate kinases, InsP(6)K1 and InsP(6)K2 (Saiardi, A., Erdjument-Bromage, H., Snowman, A. M., Tempst, P., and Snyder, S. H. (1999) Curr. Biol. 9, 1323-1326). We now report the identification, cloning, and characterization of a third InsP(7) forming enzyme designated InsP(6)K3. InsP(6)K3 displays 50 and 45% sequence identity to InsP(6)K1 and InsP(6)K2, respectively, with a smaller mass (46 kDa) and a more basic character than the other two enzymes. InsP(6)K3 is most enriched in the brain where its localization resembles InsP(6)K1 and InsP(6)K2. Intracellular disposition discriminates the three enzymes with InsP(6)K2 being exclusively nuclear, InsP(6)K3 predominating in the cytoplasm, and InsP(6)K1 displaying comparable nuclear and cytosolic densities.  相似文献   

3.
Luo HR  Saiardi A  Nagata E  Ye K  Yu H  Jung TS  Luo X  Jain S  Sawa A  Snyder SH 《Neuron》2001,31(3):439-451
Diphosphoinositol-pentakisphosphate (InsP7) and bis-diphosphoinositol tetrakisphosphate (InsP8) possess pyrophosphate bonds. InsP7 is formed from inositol hexakisphosphate (InsP6) by recently identified InsP6 kinases designated InsP6K1 and InsP6K2. We now report the identification, cloning, and characterization of a novel protein, GRAB (guanine nucleotide exchange factor for Rab3A), which interacts with both InsP6K1 and Rab3A, a Ras-like GTPase that regulates synaptic vesicle exocytosis. GRAB is a physiologic GEF (guanine nucleotide exchange factor) for Rab3A. Consistent with a role of Rab3A in synaptic vesicle exocytosis, GRAB regulates depolarization-induced release of dopamine from PC12 cells and nicotinic agonist-induced hGH release from bovine adrenal chromaffin cells. The association of InsP6K1 with GRAB fits with a role for InsP7 in vesicle exocytosis.  相似文献   

4.
Inositol pyrophosphate diphosphoinositol pentakisphosphate is ubiquitously present in mammalian cells and contains highly energetic pyrophosphate bonds. We have previously reported that inositol hexakisphosphate kinase type 2 (InsP(6)K2), which converts inositol hexakisphosphate to inositol pyrophosphate diphosphoinositol pentakisphosphate, mediates apoptotic cell death via its translocation from the nucleus to the cytoplasm. Here, we report that InsP(6)K2 is localized mainly in the cytoplasm of lymphoblast cells from patients with Huntington disease (HD), whereas this enzyme is localized in the nucleus in control lymphoblast cells. The large number of autophagosomes detected in HD lymphoblast cells is consistent with the down-regulation of Akt in response to InsP(6)K2 activation. Consistent with these observations, the overexpression of InsP(6)Ks leads to the depletion of Akt phosphorylation and the induction of cell death. These results suggest that InsP(6)K2 activation is associated with the pathogenesis of HD.  相似文献   

5.
InsP(6) is abundant in cereals and legumes. InsP(6) and lower inositol phosphates, in particular InsP(3), participate in important intracellular processes. In addition, InsP(6) possess significant health benefits, such as anti-cancer effect, kidney stones prevention, lowering serum cholesterol. Because of the insensitivity of existing methods for determination of non-radiolabeled inositol phosphates, little is known about the natural occurrence, much less on the concentrations of InsP(6) and InsP(3) in biological samples. Using gas chromatography-mass detection analysis of HPLC chromatographic fractions, we report a measurement of unlabeled total InsP(3) and InsP(6) (a) as they occur within cells culture, tissues, and plasma, and (b) their changes depending on the presence of exogenous InsP(6). When rats were fed on a purified diet in which InsP(6) was undetectable (AIN-76A) the levels of InsP(6) in brain were 3.35 +/- 0.57 (SE) micromol.kg(-1) and in plasma 0.023 +/- 0.008 (SE) micromol.l(-1). The presence of InsP(6) in diet dramatically influenced its levels in brain and in plasma. When rats were given an InsP(6)-sufficient diet (AIN-76A + 1% InsP(6)), the levels of InsP(6) were about 100-fold higher in brain tissues (36.8 +/- 1.8 (SE)) than in plasma (0.29 +/- 0.02 (SE)); InsP(6) concentrations were 8.5-fold higher than total InsP(3) concentrations in either plasma (0.033 +/- 0.012 (SE)) and brain (4.21 +/- 0.55 (SE)). When animals were given an InsP(6)-poor diet (AIN-76A only), there was a 90% decrease in InsP(6) content in both brain tissue and plasma (p < 0.001); however, there was no change in the level of total InsP(3). In non-stimulated malignant cells (MDA-MB 231 and K562) the InsP(6) contents were 16.2 +/- 9.1 (SE) micromol.kg(-1) for MDA-MB 231 cells and 15.6 +/- 2.7 (SE) for K 562 cells. These values were around 3-fold higher than those of InsP(3) (4.8 +/- 0.5 micromol.kg(-1) and 6.9 +/- 0.1 (SE) for MDA-MB 231 and K562 cells respectively). Treatment of malignant cells with InsP(6) resulted in a 2-fold increase in the intracellular concentrations of total InsP(3) (9.5 +/- 1.3 (SE) and 10.8 +/- 1.0 (SE) micromol.kg(-1) for MDA-MB 231 and K562 cells respectively, p < 0.05), without changes in InsP(6) levels. These results indicate that exogenous InsP(6) directly affects its physiological levels in plasma and brain of normal rats without changes on the total InsP(3) levels. Although a similar fluctuation of InsP(6) concentration was not seen in human malignant cell lines following InsP(6) treatment, an increased intracellular levels of total InsP(3) was clearly observed.  相似文献   

6.
Vibrio cholerae secretes a large virulence-associated multifunctional autoprocessing RTX toxin (MARTX(Vc)). Autoprocessing of this toxin by an embedded cysteine protease domain (CPD) is essential for this toxin to induce actin depolymerization in a broad range of cell types. A homologous CPD is also present in the large clostridial toxin TcdB and recent studies showed that inositol hexakisphosphate (Ins(1,2,3,4,5,6)P(6) or InsP(6)) stimulated the autoprocessing of TcdB dependent upon the CPD (Egerer, M., Giesemann, T., Jank, T., Satchell, K. J., and Aktories, K. (2007) J. Biol. Chem. 282, 25314-25321). In this work, the autoprocessing activity of the CPD within MARTX(Vc) is similarly found to be inducible by InsP(6). The CPD is shown to bind InsP(6) (K(d), 0.6 microm), and InsP(6) is shown to stimulate intramolecular autoprocessing at both physiological concentrations and as low as 0.01 microm. Processed CPD did not bind InsP(6) indicating that, subsequent to cleavage, the activated CPD may shift to an inactive conformation. To further pursue the mechanism of autoprocessing, conserved residues among 24 identified CPDs were mutagenized. In addition to cysteine and histidine residues that form the catalytic site, 2 lysine residues essential for InsP(6) binding and 5 lysine and arginine residues resulting in loss of activity at low InsP(6) concentrations were identified. Overall, our data support a model in which basic residues located across the CPD structure form an InsP(6) binding pocket and that the binding of InsP(6) stimulates processing by altering the CPD to an activated conformation. After processing, InsP(6) is shown to be recycled, while the cleaved CPD becomes incapable of further binding of InsP(6).  相似文献   

7.
We have established a new system for studying phytic acid, myo-inositol hexakisphosphate (InsP(6)) synthesis in suspension-cultured cells of Catharanthus. InsP(6) and other intermediates of myo-inositol (Ins) phosphate metabolism were measured using an ion chromatography method. The detection limit for InsP(6) was less than 50 nM, which was sufficient to analyze Ins phosphates in living cells. Synthesis of Ins phosphates was induced by incubation in high inorganic phosphate medium. InsP(6) was mainly accumulated in vacuoles and was enhanced when cells were grown in high concentration of inorganic phosphates with the cations K(+), Ca(2+), or Zn(2+). However, there was a strong tendency for InsP(6) to accumulate in the vacuole in the presence of Ca(2+) and in nonvacuolar compartments when supplied with Zn(2+), possibly due to precipitation of InsP(6) with Zn(2+) in the cytosol. A vesicle transport inhibitor, brefeldin A, stimulated InsP(6) accumulation. The amounts of both Ins(3)P(1) myo-inositol monophosphate synthase, a key enzyme for InsP(6) synthesis, and Ins(1,4,5)P(3) kinase were unrelated to the level of accumulation of InsP(6). The mechanisms for InsP(6) synthesis and localization into vacuoles in plant cells are discussed.  相似文献   

8.
Luo HR  Huang YE  Chen JC  Saiardi A  Iijima M  Ye K  Huang Y  Nagata E  Devreotes P  Snyder SH 《Cell》2003,114(5):559-572
Inositol phosphates are well-known signaling molecules, whereas the inositol pyrophosphates, such as diphosphoinositol pentakisphosphate (InsP7/IP7) and bis-diphosphoinositol tetrakisphosphate (InsP8/IP8), are less well characterized. We demonstrate physiologic regulation of Dictyostelium chemotaxis by InsP7 mediated by its competition with PtdIns(3,4,5)P3 for binding pleckstrin homology (PH) domain-containing proteins. Chemoattractant stimulation triggers rapid and sustained elevations in InsP7/InsP8 levels. Depletion of InsP7 and InsP8 by deleting the gene for InsP6 kinase (InsP6K/IP6K), which converts inositol hexakisphosphate (InsP6/IP6) to InsP7, causes rapid aggregation of mutant cells and increased sensitivity to cAMP. Chemotaxis is mediated by membrane translocation of certain PH domain-containing proteins via specific binding to PtdIns(3,4,5)P3. InsP7 competes for PH domain binding with PtdIns(3,4,5)P3 both in vitro and in vivo. InsP7 depletion enhances PH domain membrane translocation and augments downstream chemotactic signaling activity.  相似文献   

9.
Saiardi et al. (Saiardi, A., Erdjument-Bromage, H., Snowman, A., Tempst, P., and Snyder, S. H. (1999) Curr. Biol. 9, 1323-1326) previously described the cloning of a kinase from yeast and two kinases from mammals (types 1 and 2), which phosphorylate inositol hexakisphosphate (InsP(6)) to diphosphoinositol pentakisphosphate, a "high energy" candidate regulator of cellular trafficking. We have now studied the significance of InsP(6) kinase activity in Saccharomyces cerevisiae by disrupting the kinase gene. These ip6kDelta cells grew more slowly, their levels of diphosphoinositol polyphosphates were 60-80% lower than wild-type cells, and the cells contained abnormally small and fragmented vacuoles. Novel activities of the mammalian and yeast InsP(6) kinases were identified; inositol pentakisphosphate (InsP(5)) was phosphorylated to diphosphoinositol tetrakisphosphate (PP-InsP(4)), which was further metabolized to a novel compound, tentatively identified as bis-diphosphoinositol trisphosphate. The latter is a new substrate for human diphosphoinositol polyphosphate phosphohydrolase. Kinetic parameters for the mammalian type 1 kinase indicate that InsP(5) (K(m) = 1.2 micrometer) and InsP(6) (K(m) = 6.7 micrometer) compete for phosphorylation in vivo. This is the first time a PP-InsP(4) synthase has been identified. The mammalian type 2 kinase and the yeast kinase are more specialized for the phosphorylation of InsP(6). Synthesis of the diphosphorylated inositol phosphates is thus revealed to be more complex and interdependent than previously envisaged.  相似文献   

10.
Luo HR  Saiardi A  Yu H  Nagata E  Ye K  Snyder SH 《Biochemistry》2002,41(8):2509-2515
Diphosphoinositol pentakisphosphate (InsP(7)) and bis-diphosphoinositol tetrakisphosphate (InsP(8)) contain energetic pyrophosphate groups, occur throughout animal and plant kingdoms, and are synthesized by a recently cloned family of inositol hexakisphosphate kinases (InsP(6)Ks). We report that these inositol pyrophosphates mediate homologous DNA recombination in yeast S. cerevisae. Hyperrecombination, caused by altered protein kinase C1 (PKC1), is lost in yeast with deletion of yeast InsP(6)K (yInsP(6)K) and can be restored selectively by catalytically active yeast or mammalian InsP(6)Ks. Inositol pyrophosphates are required for two forms of hyperrecombination that differ in mechanism, suggesting some generalities for actions of inositol pyrophosphates in recombination.  相似文献   

11.
12.
The ionic mechanism of inositol trisphosphate (InsP3)-induced Ca2+ release was investigated in microsomes (microsomal fractions) isolated from rat brain. InsP3 stimulated Ca2+ release from microsomes incubated in media containing 100 mM-KCl. The InsP3-induced Ca2+ release was insensitive to a variety of Ca2+-channel blockers; however, the K+-channel blockers tetraethylammonium chloride (TEA; 1 mM) and 9-tetraethylammonium chloride (9-TEA; 1 mM) blocked InsP3-induced Ca2+ release. Moreover, addition of InsP3 increased 86Rb+ influx into the microsomes. The influx of 86Rb+ also was sensitive to TEA and 9-TEA. The above results suggest that InsP3-induced Ca2+ release requires an opposite flow of K+ ions, and modulation of K+ channels by TEA and 9-TEA may underlie the inhibition of InsP3-induced Ca2+ release from brain microsomes by these agents.  相似文献   

13.
When [3H]inositol-prelabelled N1E-115 cells were stimulated with carbamylcholine (CCh) (100 microM), high K+ (60 mM), and prostaglandin E1 (PGE1) (10 microM), a transient increase in [3H]inositol pentakisphosphate (InsP5) accumulation was observed. The accumulation reached its maximum level at 15 s and had declined to the basal level at 2 min. CCh, high K+, and PGE1 also caused accumulations of [3H]inositol 1,4,5-trisphosphate [Ins(1,4,5)P3], [3H]inositol 1,3,4,6-tetrakisphosphate [Ins(1,3,4,6)P4], and [3H]inositol hexakisphosphate (InsP6). Muscarine and CCh induced accumulations of [3H]Ins(1,4,5)P3, [3H]-Ins(1,3,4,6)P4, [3H]InsP5, and [3H]InsP6 with a similar potency and exerted these maximal effects at 100 microM, whereas nicotine failed to do so at 1 mM. With a slower time course, CCh, high K+, and PGE1 caused accumulations of [3H]-inositol 1,3,4-trisphosphate [Ins(1,3,4)P3] and [3H]inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4]. In an N1E-115 cell homogenate, [3H]Ins(1,4,5)P3, [3H]Ins(1,3,4,5)P4, and [3H]Ins(1,3,4)P3 were converted to [3H]InsP5 through [3H]-Ins(1,3,4,6)P4. The above results indicate that Ins(1,3,4,6)P4, InsP5, and InsP6 are rapidly formed by several kinds of stimulants in N1E-115 cells.  相似文献   

14.
When [3H]inositol-prelabeled cultured bovine adrenal chromaffin cells were stimulated with high K+ (56 mM) and nicotine (10 microM), a large and transient increase in [3H]inositol 1,3,4,5,6-pentakisphosphate (InsP5) accumulation was observed. The accumulation reached the maximum level at 15 s and then declined to the basal level at 2 min. The time course of accumulation of InsP5 was parallel to that of [3H]inositol 1,4,5-trisphosphate (Ins(1,4,5)P3). Angiotensin II (Ang II) (10 microM) rapidly accumulated InsP5, but the level was sustained for 2 min. With a slower time course and a lesser amount than InsP5, high K+, nicotine, and Ang II caused an accumulation of [3H]inositol 1,3,4,5-tetrakisphosphate and [3H]inositol hexakisphosphate. Veratridine (100 microM), maitotoxin (10 ng/ml), ATP (30 microM), platelet-derived growth factor (10 ng/ml), and endothelin (10 ng/ml) also induced the InsP5 accumulation. High K+, nicotine, veratridine, and maitotoxin induced an increase in 45Ca2+ uptake, whereas Ang II, ATP, platelet-derived growth factor, and endothelin did not cause 45Ca2+ uptake. Nifedipine, a calcium channel antagonist, inhibited the high K(+)-induced InsP5 accumulation but failed to affect the Ang II-induced InsP5 accumulation. In an EGTA-containing and Ca2(+)-depleted medium, the high K(+)-induced InsP5 accumulation was completely inhibited, whereas the InsP5 accumulation induced by Ang II was not significantly inhibited. 12-O-tetradecanoylphorbol-13-acetate inhibited partially the Ang II-induced InsP5 accumulation but failed to inhibit the high K(+)-induced accumulation. In those experiments, the changes of InsP5 accumulation were closely correlated to those of Ins(1,4,5)P3. In the chromaffin cell homogenate, [3H] Ins(1,4,5)P3 was converted eventually to [3H]InsP5 through [3H]inositol 1,3,4,6-tetrakisphosphate. Taken together, the above results suggest that InsP5 is rapidly formed by a variety of stimulants and that the formation of InsP5 may occur through two mechanisms, i.e. Ca2+ uptake-dependent and Ca2+ uptake-independent ones in cultured adrenal chromaffin cells.  相似文献   

15.
16.
We have demonstrated that the human homolog of the rat inositol phosphate multikinase is an inositol 1,3,4,6-tetrakisphosphate 5-kinase (InsP(4) 5-kinase). The cDNA of the human gene contained a putative open reading frame of 1251 bp encoding 416 amino acids with 83.6% identity compared with the rat protein. The substrate specificity of the recombinant human protein demonstrated preference for Ins(1,3,4,6)P(4) with a catalytic efficiency (V(max)/K(m)) 43-fold greater than that of Ins(1,3,4,5)P(4) and 2-fold greater than that of Ins(1,4,5)P(3). The apparent V(max) was 114 nmol of Ins(1,3,4,5,6)P(5) formed/min/mg of protein, and the apparent K(m) was 0.3 microm Ins(1,3,4,6)P(4). The functional homolog in yeast is Ipk2p, and ipk2-null yeast strains do not synthesize Ins(1,3,4,5,6)P(5) or InsP(6). Synthesis of these compounds was restored by transformation with wild-type yeast IPK2 but not with human InsP(4) 5-kinase. Thus the human gene does not complement for the loss of the yeast gene because yeast cells do not contain the substrate Ins(1,3,4,6)P(4), and the reaction of the human protein with Ins(1,3,4,5)P(4) is insufficient to effect rescue or synthesis of InsP(5) and InsP(6). Therefore the major activity of human InsP(4) 5-kinase is phosphorylation at the D-5 position, and the pathways for synthesis of Ins(1,3,4,5,6)P(5) in yeast versus humans are different.  相似文献   

17.
Inositol 1,4,5-trisphosphate (InsP3) receptor immunoreactivity in SH-SY5Y human neuroblastoma cells was monitored with a monoclonal antibody raised against the mouse cerebellar InsP3 receptor. Recognition of a protein corresponding to the InsP3 receptor (molecular mass, approximately 275 kDa) was inhibited markedly following exposure of cells to 0.1 mM carbachol. This effect was half-maximal and maximal at approximately 2 and approximately 6 h, respectively; was blocked by atropine; but was not mimicked by thapsigargin, K+, or phorbol 12-myristate 13-acetate. However, the decrease in immunoreactivity following exposure of cells to carbachol for 5 h was blocked if the extracellular Ca2+ concentration was reduced from 1.3 mM to 200 nM. This manipulation also reduced markedly carbachol-induced increases in InsP3 concentration at 5 h. These data indicate that chronic muscarinic stimulation of phosphoinositide hydrolysis reduces InsP3 receptor concentration in SH-SY5Y cells, perhaps via a mechanism that involves prolonged elevation of InsP3 levels.  相似文献   

18.
The enzyme(s) responsible for the production of inositol hexakisphosphate (InsP(6)) in vertebrate cells are unknown. In fungal cells, a 2-kinase designated Ipk1 is responsible for synthesis of InsP(6) by phosphorylation of inositol 1,3,4,5,6-pentakisphosphate (InsP(5)). Based on limited conserved sequence motifs among five Ipk1 proteins from different fungal species, we have identified a human genomic DNA sequence on chromosome 9 that encodes human inositol 1,3,4,5,6-pentakisphosphate 2-kinase (InsP(5) 2-kinase). Recombinant human enzyme was produced in Sf21 cells, purified, and shown to catalyze the synthesis of InsP(6) or phytic acid in vitro. The recombinant protein converted 31 nmol of InsP(5) to InsP(6)/min/mg of protein (V(max)). The Michaelis-Menten constant for InsP(5) was 0.4 microM and for ATP was 21 microM. Saccharomyces cerevisiae lacking IPK1 do not produce InsP(6) and show lethality in combination with a gle1 mutant allele. Here we show that expression of the human InsP(5) 2-kinase in a yeast ipk1 null strain restored the synthesis of InsP(6) and rescued the gle1-2 ipk1-4 lethal phenotype. Northern analysis on human tissues showed expression of the human InsP(5) 2-kinase mRNA predominantly in brain, heart, placenta, and testis. The isolation of the gene responsible for InsP(6) synthesis in mammalian cells will allow for further studies of the InsP(6) signaling functions.  相似文献   

19.
The yeast and Drosophila pathways leading to the production of inositol hexakisphosphate (InsP(6)) have been elucidated recently. The in vivo pathway in humans has been assumed to be similar. Here we show that overexpression of Ins(1,3,4)P(3) 5/6-kinase in human cell lines results in an increase of inositol tetrakisphosphate (InsP(4)) isomers, inositol pentakisphosphate (InsP(5)) and InsP(6), whereas its depletion by RNA interference decreases the amounts of these inositol phosphates. Expression of Ins(1,3,4,6)P(4) 5-kinase does not increase the amount of InsP(5) and InsP(6), although its depletion does block InsP(5) and InsP(6) production, showing that it is necessary for production of InsP(5) and InsP(6). Expression of Ins(1,3,4,5,6)P(5) 2-kinase increases the amount of InsP(6) by depleting the InsP(5) in the cell, and depletion of 2-kinase decreases the amount of InsP(6) and causes an increase in InsP(5). These results are consistent with a pathway that produces InsP(6) through the sequential action of Ins(1,3,4)P(3) 5/6-kinase, Ins(1,3,4,6)P(4) 5-kinase, and Ins(1,3,4,5,6)P5 2-kinase to convert Ins(1,3,4)P(3) to InsP(6). Furthermore, the evidence implicates 5/6-kinase as the rate-limiting enzyme in this pathway.  相似文献   

20.
Inositol hexakisphosphate (InsP6) is a most abundant inositol polyphosphate that changes simultaneously with inositol 1,4,5-trisphosphate in depolarized neurons. However, the role of InsP6 in neuronal signaling is unknown. Mass assay reveals that the basal levels of InsP6 in several brain regions tested are similar. InsP6 mass is significantly elevated in activated brain neurons and lowered by inhibition of neuronal activity. Furthermore, the hippocampus is most sensitive to electrical challenge with regard to percentage accumulation of InsP6. In hippocampal neurons, InsP6 stimulates adenylyl cyclase (AC) without influencing cAMP phosphodiesterases, resulting in activation of protein kinase A (PKA) and thereby selective enhancement of voltage-gated L-type Ca2+ channel activity. This enhancement was abolished by preincubation with PKA and AC inhibitors. These data suggest that InsP6 increases L-type Ca2+ channel activity by facilitating phosphorylation of PKA phosphorylation sites. Thus, in hippocampal neurons, InsP6 serves as an important signal in modulation of voltage-gated L-type Ca2+ channel activity.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号