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 31 毫秒
1.
We have recently demonstrated that in quiescent fibroblasts protein kinase C (PKC) epsilon(95) is phosphorylated at Ser(729), Ser(703), and Thr(566) and that upon passage of quiescent cells phosphorylation at Ser(729) is lost, giving rise to PKCepsilon(87). Ser(729) may be rephosphorylated later, suggesting cycling between PKCepsilon(87) and PKCepsilon(95). Here we show that the dephosphorylation at Ser(729) is insensitive to okadaic acid, calyculin, ascomycin C, and cyclosporin A, suggesting that dephosphorylation at this site is not mediated through protein phosphatases 1, 2A or 2B. We demonstrate that this dephosphorylation at Ser(729) requires serum and cell readhesion and is sensitive to rapamycin, PD98059, chelerythrine, and Ro-31-8220. These results suggest that the phosphorylation status of Ser(729) in the hydrophobic domain at Ser(729) is regulated independently of the phosphorylation status of other sites in PKCepsilon, by a mTOR-sensitive phosphatase. The mitogen-activated protein kinase pathway and PKC are also implicated in regulating the dephosphorylation at Ser(729).  相似文献   

2.
The importance of activation loop phosphorylation in the regulation of protein kinase D (PKD/protein kinase C (PKC) mu) activity has become controversial. In order to clarify the mechanism(s) of PKD activation, we developed a novel phosphospecific antibody recognizing phosphorylated Ser(748) in PKD (pS748). Western blot analysis with the pS748 antibody, carried out with a variety of PKD forms and in a variety of cell types including full-length PKD transfected in COS-7 and HEK 293 cells, a green fluorescent protein-PKD fusion protein transfected in either Swiss 3T3 fibroblasts or Madin-Darby canine kidney epithelial cells, and endogenous PKD expressed in A20 lymphocytes and Rat-1 fibroblasts, indicated that Ser(748) phosphorylation was absent from unstimulated cells. In contrast, dramatic increases in Ser(748) phosphorylation were induced by phorbol esters, bombesin, or cross-linking of B lymphocyte antigen receptors or by cotransfection with active PKCepsilon or PKCeta. Western analysis using a second phosphospecific antibody, which primarily recognizes PKD phosphorylated at Ser(744), revealed that Ser(744) phosphorylation accompanies Ser(748) phosphorylation during PKD activation in vivo. Ser(744)/Ser(748) phosphorylation requires PKC but not PKD activity, indicative of transphosphorylation. Our results provide new experimental evidence indicating that activation loop phosphorylation at Ser(744) and Ser(748) occurs during PKD activation in vivo and support the notion of a PKC-PKD phosphorylation cascade.  相似文献   

3.
Persistent activation of protein kinase D (PKD) via protein kinase C (PKC)-mediated signal transduction is accompanied by phosphorylation at Ser(744) and Ser(748) located in the catalytic domain activation loop, but whether PKC isoforms directly phosphorylate these residues, induce PKD autophosphorylation, or recruit intermediate upstream kinase(s) is unclear. Here, we explore the mechanism whereby PKC activates PKD in response to cellular stimuli. We first assessed in vitro PKC-PKD transphosphorylation and PKD activation. A PKD738-753 activation loop peptide was well phosphorylated by immunoprecipitated PKC isoforms, consistent with similarities between the loop and their known substrate specificities. A similar peptide with glutamic acid replacing Ser(748) was preferentially phosphorylated by PKCepsilon, suggesting that PKD containing phosphate at Ser(748) is rapidly targeted by this isoform at Ser(744). When incubated in the presence of phosphatidylserine, phorbol 12,13-dibutyrate and ATP, intact PKD slowly autophosphorylated in the activation loop but only at Ser(748). In contrast, addition of purified PKCepsilon to the incubation mixture induced rapid Ser(744) and Ser(748) phosphorylation, concomitant with persistent 2-3-fold increases in PKD activity, measured using reimmunoprecipitated PKD to phosphorylate an exogenous peptide, syntide-2. We also further examined pleckstrin homology domain-mediated PKD regulation to determine its relationship with activation loop phosphorylation. The high constitutive activity of the pleckstrin homology (PH) domain deletion mutant PKD-deltaPH was not abrogated by mutation of Ser(744) and Ser(748) to alanines, suggesting that one function of activation loop phosphorylation in the PKD activation mechanism is to relieve autoinhibition by the PH domain. These studies provide evidence of a direct PKCepsilon-PKD phosphorylation cascade and provide additional insight into the activation mechanism.  相似文献   

4.
We demonstrate that GFP-PKCepsilon concentrates at a perinuclear site in living fibroblasts and that cell passage induces rapid translocation of PKCepsilon to the periphery where it appears to colocalise with F-actin. When newly passaged cells have adhered and are proliferating again, GFP-PKCepsilon returns to its perinuclear site. GFP-PKCepsilon co-localises with wheat germ agglutinin suggesting that it is associated with the Golgi at the perinuclear site. In support, PKCepsilon is detected in a Golgi-enriched fraction in pre-passage cells but is lost from the fraction after passage. PKCepsilon at the perinuclear Golgi site is phosphorylated at Ser729 but cell passage induces the loss of the phosphate at this site as reported previously [England et al. (2001) J. Biol. Chem. 276, 10437-10442]. PKCepsilon S729A, S729E and S729T mutants, which are not recognised by a specific antiphosphoPKCepsilon (Ser729) antibody, do not concentrate at a perinuclear/Golgi site in proliferating fibroblasts. This suggests that both phosphorylation and serine rather than threonine are needed at position 729 to locate PKCepsilon at its perinuclear/Golgi site. Phorbol ester induced translocation of PKCepsilon to the nucleus also requires dephosphorylation at Ser729; after translocation nuclear PKCepsilon lacks a phosphate at Ser729. Sulphation and secretion of glycosaminoglycan (GAG) chains from fibroblasts increases on passage and returns to basal as cells proliferate showing that cell passage influences secretory events at the Golgi. The results indicate that Ser729 phosphorylation plays a role in determining PKCepsilon localisation in fibroblasts.  相似文献   

5.
K Cheng  W W Chan  R Arias  A Barreto  B Butler 《Life sciences》1992,51(25):1957-1967
In GH3 cells and other clonal rat pituitary tumor cells, TRH has been shown to mediate its effects on prolactin release via a rise of cytosolic Ca2+ and activation of protein kinase C. In this study, we examined the role of protein kinase C in TRH-stimulated prolactin release from female rat primary pituitary cell culture. Both TRH and PMA stimulated prolactin release in a dose-dependent manner. When present together at maximal concentrations, TRH and PMA produced an effect which was slightly less than additive. Pretreatment of rat pituitary cells with 10(-6) M PMA for 24 hrs completely down-regulated protein kinase C, since such PMA-pretreated cells did not release prolactin in response to a second dose of PMA. Interestingly, protein kinase C down-regulation had no effect on TRH-induced prolactin release from rat pituitary cells. In contrast, PMA-pretreated GH3 cells did not respond to a subsequent stimulation by either PMA or TRH. Pretreatment of rat pituitary cells with TRH (10(-7) M, 24 hrs) inhibited the subsequent response to TRH, but not PMA. Forskolin, an adenylate cyclase activator, stimulated prolactin release by itself and in a synergistic manner when incubated together with TRH or PMA. The synergistic effects of forskolin on prolactin release was greater in the presence of PMA than TRH. Down-regulation of protein kinase C by PMA pretreatment abolished the synergistic effect produced by PMA and forskolin but had no effect on those generated by TRH and forskolin. sn-1,2-Dioctanylglycerol (DOG) pretreatment attenuated the subsequent response to DOG and PMA but not TRH. The effect of TRH, but not PMA, on prolactin release required the presence of extracellular Ca2+. In conclusion, the mechanism by which TRH causes prolactin release from rat primary pituitary cells is different from that of GH3 cells; the former is a protein kinase C-independent process whereas the latter is at least partially dependent upon the activation of protein kinase C.  相似文献   

6.
Thyrotropin-releasing hormone (TRH) increases rapidly two potential intracellular signals, inositol trisphosphate (IP3) and free cytosolic calcium ([Ca2+]i), for stimulated prolactin release and synthesis in GH4C1 rat pituitary cells. We have examined the temporal relationships between TRH-enhanced formation of inositol phosphates and TRH-elevated [Ca2+]i. TRH-enhanced IP3 content was closely paralleled by the initial phase of TRH-elevated [Ca2+]i. To investigate receptor-effector coupling for these rapid actions of TRH, we examined their dependence on receptor number in five GH4C1 variant strains containing 0-2.6 X 10(5) receptor sites/cell. We found that receptor number (up to 1.7 X 10(5)/cell) was limiting for TRH-enhanced IP3 formation as well as for both the initial burst and plateau phases of TRH-elevated [Ca2+]i. The ED50 for rapid (5 s) TRH-stimulated IP3 formation was higher than for other sustained TRH actions in these cells, and we postulated that the initial TRH receptor interactions occur with rapid dissociation kinetics. To test this hypothesis, we performed rapid dilution experiments following a 1-s stimulation and found that TRH-stimulated IP3 formation decreased within 4 s of dilution and disappeared within 60 s at which time fresh TRH could restimulate IP3 formation. We conclude that receptor occupancy is the limiting step for TRH-stimulated IP3 formation and elevated [Ca2+]i and that maximal TRH action requires multiple rapid interactions between TRH and its receptor.  相似文献   

7.
8.
9.
In pituitary GH3B6 cells, signaling involving the protein kinase C (PKC) multigene family can self-organize into a spatiotemporally coordinated cascade of isoform activation. Indeed, thyrotropin-releasing hormone (TRH) receptor activation sequentially activated green fluorescent protein (GFP)-tagged or endogenous PKCbeta1, PKCalpha, PKCepsilon, and PKCdelta, resulting in their accumulation at the entire plasma membrane (PKCbeta and -delta) or selectively at the cell-cell contacts (PKCalpha and -epsilon). The duration of activation ranged from 20 s for PKCalpha to 20 min for PKCepsilon. PKCalpha and -epsilon selective localization was lost in the presence of G?6976, suggesting that accumulation at cell-cell contacts is dependent on the activity of a conventional PKC. Constitutively active, dominant-negative PKCs and small interfering RNAs showed that PKCalpha localization is controlled by PKCbeta1 activity and is calcium independent, while PKCepsilon localization is dependent on PKCalpha activity. PKCdelta was independent of the cascade linking PKCbeta1, -alpha, and -epsilon. Furthermore, PKCalpha, but not PKCepsilon, is involved in the TRH-induced beta-catenin relocation at cell-cell contacts, suggesting that PKCepsilon is not the unique functional effector of the cascade. Thus, TRH receptor activation results in PKCbeta1 activation, which in turn initiates a calcium-independent but PKCbeta1 activity-dependent sequential translocation of PKCalpha and -epsilon. These results challenge the current understanding of PKC signaling and raise the question of a functional dependence between isoforms.  相似文献   

10.
To study phosphorylation of the endogenous type I thyrotropin-releasing hormone receptor in the anterior pituitary, we generated phosphosite-specific polyclonal antibodies. The major phosphorylation site of receptor endogenously expressed in pituitary GH3 cells was Thr(365) in the receptor tail; distal sites were more phosphorylated in some heterologous models. beta-Arrestin 2 reduced thyrotropin-releasing hormone (TRH)-stimulated inositol phosphate production and accelerated internalization of the wild type receptor but not receptor mutants where the critical phosphosites were mutated to Ala. Phosphorylation peaked within seconds and was maximal at 100 nm TRH. Based on dominant negative kinase and small interfering RNA approaches, phosphorylation was mediated primarily by G protein-coupled receptor kinase 2. Phosphorylated receptor, visualized by immunofluorescence microscopy, was initially at the plasma membrane, and over 5-30 min it moved to intracellular vesicles in GH3 cells. Dephosphorylation was rapid (t((1/2)) approximately 1 min) if agonist was removed while receptor was at the surface. Dephosphorylation was slower (t((1/2)) approximately 4 min) if agonist was withdrawn after receptor had internalized. After agonist removal and dephosphorylation, a second pulse of agonist caused extensive rephosphorylation, particularly if most receptor was still on the plasma membrane. Phosphorylated receptor staining was visible in prolactin- and thyrotropin-producing cells in rat pituitary tissue from untreated rats and much stronger in tissue from animals injected with TRH. Our results show that the TRH receptor can rapidly cycle between a phosphorylated and nonphosphorylated state in response to changing agonist concentrations and that phosphorylation can be used as an indicator of receptor activity in vivo.  相似文献   

11.
Xiao L  Zhao Q  Du Y  Yuan C  Solaro RJ  Buttrick PM 《Biochemistry》2007,46(23):7054-7061
Cardiac myosin binding protein C (cMyBPC) phosphorylation is essential for normal cardiac function. Although PKC was reported to phosphorylate cMyBPC in vitro, the relevant PKC isoforms and functions of PKC-mediated cMyBPC phosphorylation are unknown. We recently reported that a transgenic mouse model with cardiac-specific overexpression of PKCepsilon (PKCepsilon TG) displayed enhanced sarcomeric protein phosphorylation and dilated cardiomyopathy. In the present study, we have investigated cMyBPC phosphorylation in PKCepsilon TG mice. Western blotting and two-dimensional gel electrophoresis demonstrated a significant increase in cMyBPC serine (Ser) phosphorylation in 12-month-old TG mice compared to wild type (WT). In vitro PKCepsilon treatment of myofibrils increased the level of cMyBPC Ser phosphorylation in WT mice to that in TG mice, whereas treatment of TG myofibrils with PKCepsilon showed only a minimal increase in cMyBPC Ser phosphorylation. Three peptide motifs of cMyBPC were identified as the potential PKCepsilon consensus sites including a 100% matched motif at Ser302 and two nearly matched motifs at Ser811 and Ser1203. We treated synthetic peptides corresponding to the sequences of these three motifs with PKCepsilon and determined phosphorylation by mass spectrometry and ELISA assay. PKCepsilon induced phosphorylation at the Ser302 site but not at the Ser811 or Ser1203 sites. A S302A point mutation in the Ser302 peptide abolished the PKCepsilon-dependent phosphorylation. Taken together, our data show that the Ser302 on mouse cMyBPC is a likely PKCepsilon phosphorylation site both in vivo and in vitro and may contribute to the dilated cardiomyopathy associated with increased PKCepsilon activity.  相似文献   

12.
In spite of its central roles in cell cycle progression, senescence, and aging, knowledge about the posttranslational regulation of P16 (also known as INK4A and MTS1) remains limited. While it has been reported that P16 could be phosphorylated at Ser7, Ser8, Ser140, and Ser152, the corresponding kinases have not been identified yet. Here we report that IKKβ, a primary kinase for IκBα phosphorylation, is involved in P16 phosphorylation. Immunoprecipitation and kinase assays showed that IKKβ specifically binds to P16 and phosphorylates P16 at Ser8 in WI38 cells. Biochemical characterization of phosphomimetic Ser → Glu P16 mutants demonstrated that phosphorylation at Ser8 of P16 brings about a significant loss of its cyclin-dependent kinase (CDK) 4-inhibitory activity while P16 retains structurally and functionally intact upon phosphorylation at Ser7, Ser140, and Ser152. Our results reveal the novel role of IKKβ in P16 phosphorylation and broaden our understanding of the regulation of P16.  相似文献   

13.
TRH regulates PRL secretion and synthesis in GH4C1 rat pituitary cells. TRH responses are associated with activation of protein kinase C (PKC) isozymes and elevation of cytosolic calcium. To determine which PKC isozymes are involved in TRH-directed responses, we evaluated the effect of TRH on GH cell alpha-, beta-, delta-, and epsilon-PKC isozymes. Immunoblot analysis demonstrated that TRH caused rapid redistribution of all isozymes to a Triton X-100-insoluble (i.e. cytoskeletal) fraction. Corollary immunocytofluorescence studies demonstrated that redistributed PKCs accumulate in cell peripheries. Exocytosis involves reorganization of the cytoskeleton, therefore, each of the GH cell PKCs is appropriately located to phosphorylate proteins important for cytoskeleton organization. To determine the relative contributions of calcium and PKC signal transduction pathways in mediating TRH responses, the effects of potassium depolarization (which increases cytosolic calcium) and phorbol dibutyrate (which activates all PKC isozymes without increasing calcium) were compared. The data indicate that TRH-mediated reorganization of vinculin proceeds via a calcium-mediated pathway, whereas fragmentation of actin filaments proceeds via a PKC-dependent pathway. Selective down-modulation of epsilon-PKC with prolonged TRH-treatment was used to demonstrate that epsilon-PKC is not necessary for certain TRH-stimulated biological responses.  相似文献   

14.
15.
MAPKAP kinase 2 (MK2) is required for tumor necrosis factor synthesis. Tristetraprolin (TTP) binds to the 3'-untranslated region of tumor necrosis factor mRNA and regulates its fate. We identified in vitro and in vivo phosphorylation sites in TTP using nanoflow high pressure liquid chromatography microelectrospray ionization tandem mass spectrometry and novel methods for direct digestion of TTP bound to affinity matrices (GSH-beads or anti-Myc linked to magnetic beads). MK2Delta3B, activated in Escherichia coli by p38alpha, phosphorylates TTP in vitro at major sites Ser(52) and Ser(178) (>10-fold in abundance) as well as at several minor sites that were detected after enriching for phosphopeptides with immobilized metal affinity chromatography. MK2 phosphorylation of TTP creates a functional 14-3-3 binding site. In cells, TTP was phosphorylated at Ser(52), Ser(178), Thr(250), and Ser(316) and at SP sites in a cluster (Ser(80)/Ser(82)/Ser(85)). Anisomycin treatment of NIH 3T3 cells increased phosphorylation of Ser(52) and Ser(178). Overexpression of MK2 sufficed to increase phosphorylation of Ser(52) and Ser(178) but not Ser(80)/Ser(82)/Ser(85) or Thr(250). Thus, Ser(52) and Ser(178) are putative MK2 sites in vivo. Identified phosphosite(s) may be biologic switches controlling mRNA stability and translation.  相似文献   

16.
The properties of the dephospho and in vitro phosphorylated forms of recombinant sorghum phosphoenolpyruvate carboxylase have been compared with those of the authentic dark (dephospho) and light (phospho) leaf enzyme forms and two mutant enzymes in which the phosphorylatable serine residue (Ser8) has been changed by site-directed mutagenesis to Cys (S8C) or Asp (S8D). Kinetic analysis of the purified recombinant, mutant, and leaf enzyme forms at pH 8.0 indicated virtually identical Vmax, apparent Km (phosphoenolpyruvate), and half-maximal activation (glucose 6-P) values of about 44 units/mg, 1.1 mM, and 0.23 mM, respectively. In contrast, the Ser8, S8C, and dark leaf enzymes were about 3-fold more sensitive to inhibition by L-malate at pH 7.3 than the Ser8-P, S8D, and light leaf enzyme forms. These comparative results indicate that: (i) Ser8 is an important determinant in the regulation of sorghum phosphoenolpyruvate carboxylase activity by negative (L-malate), but not positive (glucose 6-phosphate) metabolite effectors, (ii) phosphorylation of this target residue can be functionally mimicked by Asp, but not Cys, and (iii) negative charge contributes to the effect of regulatory phosphorylation on this C4-photosynthesis enzyme.  相似文献   

17.
RGS5 is a member of regulators of G protein signaling (RGS) proteins that attenuate heterotrimeric G protein signaling by functioning as GTPase-activating proteins (GAPs). We investigated phosphorylation of RGS5 and the resulting change of its function. In 293T cells, transiently expressed RGS5 was phosphorylated by endogenous protein kinases in the basal state. The phosphorylation was enhanced by phorbol 12-myristate 13-acetate (PMA) and endothelin-1 (ET-1), and suppressed by protein kinase C (PKC) inhibitors, H7, calphostin C and staurosporine. These results suggest involvement of PKC in phosphorylation of RGS5. In in vitro experiments, PKC phosphorylated recombinant RGS5 protein at serine residues. RGS5 protein phosphorylated by PKC showed much lower binding capacity for and GAP activity toward Galpha subunits than did the unphosphorylated RGS5. In cells expressing RGS5, the inhibitory effect of RGS5 on ET-1-induced Ca(2+) responses was enhanced by staurosporine. Mass spectrometric analysis of the phosphorylated RGS5 revealed that Ser166 was one of the predominant phosphorylation sites. Substitution of Ser166 by aspartic acid abolished the binding capacity to Galpha subunits and the GAP activity, and markedly reduced the inhibitory effect on ET-1-induced Ca(2+) responses. These results indicate that phosphorylation at Ser166 of RGS5 by PKC causes loss of the function of RGS5 in G protein signaling. Since this serine residue is conserved in RGS domains of many RGS proteins, the phosphorylation at Ser166 by PKC might act as a molecular switch and have functional significance.  相似文献   

18.
Polyphosphoinositide hydrolysis was examined in membranes from thyrotropin-releasing hormone (TRH)-responsive GH3 pituitary cells. [3H]Inositol phosphates (IP2 and IP3) were generated upon incubation of membranes from [3H]inositol-labeled cells indicating the presence of a membrane-associated polyphosphoinositide phosphodiesterase (PPI PDE). Membrane PPI PDE activity was found to be stimulated by TRH and by GTP-gamma-S in Ca2+-modulated manner. In addition, TRH-stimulated PPI hydrolysis was potentiated by GTP. These results demonstrate direct in vitro effects of a hormone on PPI turnover and suggest the involvement of a GTP-binding component in transmembrane signalling by TRH.  相似文献   

19.
Depolarization of membrane potential by high external K+ activates Ca2+ influx via voltage-dependent Ca2+ channels in GH4C1 cells (Tan, K.-N., and Tashjian, A. H., Jr. (1983) J. Biol. Chem. 258, 418-426). The involvement of this channel in thyrotropin-releasing hormone (TRH) action on prolactin (PRL) release was assessed by comparing the pharmacological characteristics of TRH-induced PRL release with PRL release due to high K+. Two components of TRH-stimulated PRL release were detected. The major component (approximately equal to 75%) was dependent on external Ca2+ concentration and was inhibited by voltage-dependent Ca2+ channel blockers in a manner quantitatively similar to high K+-stimulated PRL release. The minor component (approximately equal to 25%) of TRH-stimulated PRL release was insensitive to voltage-dependent Ca2+ channel blockers and could occur in the presence of low external Ca2+ (10(-5)-10(-7) M). Neither voltage-dependent Ca2+ channel blockers nor depletion of medium Ca2+ prevented the action of TRH on mobilizing cell-associated 45Ca2+ from GH4C1 cells. Divalent cations that permeate voltage-dependent Ca2+ channels (Sr2+ and Ba2+) substituted for Ca2+ in supporting high K+- and TRH-stimulated PRL release while Mg2+, a nonpermeant cation, did not. We conclude that TRH stimulates PRL release by increasing [Ca2+]i through at least two mechanisms: one requires only low [Ca2+]o, the second involves Ca2+ influx via voltage-dependent Ca2+ channels. This latter mechanism accounts for approximately equal to 75% of maximum TRH-induced PRL release.  相似文献   

20.
The effect of decreasing the concentration of receptors for thyrotropin-releasing hormone (TRH) on the surface of cloned rat pituitary (GH3) cells on TRH-stimulated inositol trisphosphate (Ins-P3) formation was investigated. Incubation of cells with dibutyryl cAMP (Bt2cAMP) for 16 h caused a decrease in [3H] TRH binding to intact cells to a minimum level 37 +/- 9.1% of control. Scatchard analysis of the concentration dependency of [3H]TRH binding showed that the effect of Bt2cAMP was to lower the receptor concentration without affecting its affinity for TRH. Similar decreases in [3H]TRH binding were found in cells incubated with 8-bromo-cAMP, cholera toxin, and sodium butyrate and, as shown previously, with TRH. In cells incubated with 1 mM Bt2cAMP for 16 h, but not for 1 h, the maximum TRH-induced increase in Ins-P3 was inhibited to 25 +/- 3.2% of that in control cells. Inhibition of TRH-induced Ins-P3 formation was also observed in cells treated with 8-bromo-cAMP, cholera toxin, and sodium butyrate for 16 h, and with TRH for 48 h. Inhibition of TRH-induced Ins-P3 formation and lowering of TRH receptor concentration caused by Bt2cAMP occurred in parallel with increasing doses of Bt2cAMP; at 16 h of exposure, half-maximal effects occurred with 0.3 mM Bt2cAMP. The concentration dependency of TRH-induced Ins-P3 formation was the same in control and Bt2cAMP-treated cells; half-maximal effects occurred with 10 nM TRH. These data demonstrate that decreases in TRH receptor concentration caused by several agents that act via different mechanisms are associated with reduced stimulation of Ins-P3 formation and suggest that the TRH receptor is tightly coupled to stimulation of hydrolysis of phosphatidylinositol 4,5-bisphosphate by a phospholipase C.  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