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.
Pasteurella multocida toxin, either native or recombinant (rPMT), is an extremely effective mitogen for Swiss 3T3 cells and acts at picomolar concentrations (Rozengurt, E., Higgins, T. E., Chanter, N., Lax, A. J., and Staddon, J. M. (1990) Proc. Natl. Acad. Sci. U. S. A. 87, 123-127). Here, we show that similar concentrations of rPMT markedly stimulated the phosphorylation of an acidic 80-kDa protein in [32P]Pi-labeled Swiss 3T3 cells. Co-migration on one- and two-dimensional gels and phosphopeptide analysis indicated that this phosphoprotein was indistinguishable from 80K, a known protein kinase C substrate. In parallel cultures, the stimulation of 80K phosphorylation by rPMT (5-10-fold) was comparable to that induced by bombesin or phorbol dibutyrate (PBt2). However, the increase in phosphorylation by rPMT occurred after a pronounced lag period (1-3 h, depending upon the concentration of rPMT) in contrast to the relatively immediate stimulation by PBt2 or bombesin. Early, but not late, addition of either PMT antiserum or the lysosomotrophic agent methylamine selectively inhibited 80K phosphorylation in response to rPMT. 80K phosphorylation persisted after removal of free toxin and was not inhibited by cycloheximide. It appears that rPMT enters the cells via an endocytotic pathway to initiate and perpetuate events leading to 80K phosphorylation. rPMT, like PBt2, also stimulated the phosphorylation of 87-kDa and 33-kDa proteins in Swiss 3T3 cells. Phosphorylation of the 80K and 87-kDa proteins by rPMT or PBt2 were greatly attenuated in cells depleted of protein kinase C. In contrast, phosphorylation of the 33-kDa protein by rPMT, but not by PBt2, persisted in the absence of protein kinase C. rPMT, like bombesin, caused a translocation of protein kinase C to the cellular particulate fraction. The toxin enhanced the cellular content of diacylglycerol. rPMT also caused a time- and dose-dependent decrease in the binding of 125I-epidermal growth factor to its receptor which was blocked by methylamine and dependent only in part upon the presence of protein kinase C. We conclude that rPMT stimulates protein kinase C-dependent and -independent protein phosphorylation in Swiss 3T3 cells.  相似文献   

2.
The effects of bombesin and insulin, separately and in combination, have been studied in Swiss mouse 3T3 cells. Bombesin caused a rapid transfer of 3H from the lipid inositol pool of prelabeled cells into inositol phosphates. Label in inositol tetrakisphosphate (InsP4) and in Ins1,4,5P3 and Ins1,3,4P3 rose within 10 sec of stimulation and that in Ins1,4P2, another InsP2 and InsP1, more slowly. Insulin, which had little effect on its own, increased the turnover of inositol lipids due to acute bombesin stimulation and also enhanced the DNA synthesis evoked by prolonged bombesin treatment. The results suggest that bombesin acting as a growth factor, uses inositol lipids as part of its transduction mechanism and that insulin acts synergistically to enhance both inositol phosphate formation and DNA synthesis.  相似文献   

3.
Treatment of Swiss 3T3 cells with a subsaturating concentration of recombinant Pasteurella multocida toxin (rPMT) markedly potentiated the production of inositol phosphates induced by bombesin, vasopressin, and endothelin but not by platelet-derived growth factor (PDGF) (AA and BB homodimers). Similarly, the neuropeptides but not PDGF caused a shift in the dose-dependent increase in inositol phosphates induced by rPMT. The rate of accumulation of inositol phosphates induced by bombesin was increased 2-fold by rPMT treatment while that of PDGF was unaffected. rPMT treatment also enhanced bombesin-induced inositol(1,4,5)trisphosphate, the direct product of phosphatidylinositol 4,5-bisphosphate hydrolysis. In contrast, treatment of cells with rPMT had no effect on the tyrosine phosphorylation of phospholipase C gamma. Depletion of protein kinase C increased rPMT-induced inositol phosphates in a manner similar to that observed for bombesin but not PDGF. Thus, rPMT selectively potentiates neuropeptide-mediated inositol phosphate production. The action of rPMT on phosphatidylinositol 4,5-bisphosphate hydrolysis persisted in streptolysin O-permeabilized cells. Addition of guanosine 5'-O-(beta-thiodiphosphate) to permeabilized cells markedly reduced rPMT-induced inositol phosphates in a time- and dose-dependent manner. rPMT also increased the sensitivity of phospholipase C for free calcium. Our results strongly suggest that the action of rPMT facilitates the coupling of G protein to phospholipase C.  相似文献   

4.
PRIP-1 was isolated as a novel inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] binding protein with a domain organization similar to phospholipase C-delta1 (PLC-delta1) but lacking the enzymatic activity. Further studies revealed that the pleckstrin homology (PH) domain of PRIP-1 is the region responsible for binding Ins(1,4,5)P3. In this study we aimed to clarify the role of PRIP-1 at the physiological concentration in Ins(1,4,5)P3-mediated Ca2+ signaling, as we had previously used COS-1 cells overexpressing PRIP-1 (Takeuchi et al., 2000, Biochem J 349:357-368). For this purpose we employed PRIP-1 knock out (PRIP-1-/-) mice generated previously (Kanematsu et al., 2002, EMBO J 21:1004-1011). The increase in free Ca2+ concentration in response to purinergic receptor stimulation was lower in primary cultured cortical neurons prepared from PRIP-1-/- mice than in those from wild type mice. The relative amounts of [3H]Ins(1,4,5)P3 measured in neurons labeled with [3H]inositol was also lower in cells from PRIP-1-/- mice. In contrast, PLC activities in brain cortex samples from PRIP-1-/- mice were not different from those in the wild type mice, indicating that the hydrolysis of Ins(1,4,5)P3 is enhanced in cells from PRIP-1-/- mice. In vitro analyses revealed that type1 inositol polyphosphate 5-phosphatase physically interacted with a PH domain of PRIP-1 (PRIP-1PH) and its enzyme activity was inhibited by PRIP-1PH. However, physical interaction with these two proteins did not appear to be the reason for the inhibition of enzyme activity, indicating that binding of Ins(1,4,5)P3 to the PH domain prevented its hydrolyzation. Together, these results indicate that PRIP-1 plays an important role in regulating the Ins(1,4,5)P3-mediated Ca2+ signaling by modulating type1 inositol polyphosphate 5-phosphatase activity through binding to Ins(1,4,5)P3.  相似文献   

5.
To investigate the mechanisms by which inositol phosphates regulate cytosolic free Ca2+ concentration ([Ca2+]c), we injected Xenopus oocytes with inositol phosphates and measured Ca2+-activated Cl- currents as an assay of [Ca2+]c. Inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) injection (0.1-10.0 pmol) induced an initial transient Cl- current (I1) followed by a second more prolonged Cl- current (I2). Both currents were Ca2+-dependent, but the source of Ca2+ was different. Release of intracellular Ca2+ stores produced I1, whereas influx of extracellular Ca2+ produced I2; Ca2+-free bathing media and inorganic calcium channel blockers (Mn2+, Co2+) did not alter I1 but completely and reversibly inhibited I2. Injection of the Ins(1,4,5)P3 metabolite, inositol 1,3,4,5-tetrakisphosphate (Ins(1,3,4,5)P4) (0.2-10.0 pmol) generated a Ca2+-dependent Cl- current with superimposed current oscillations that resulted from release of intracellular Ca2+, not Ca2+ influx. Injection of the Ins(1,3,4,5)P4 metabolite, inositol 1,3,4-trisphosphate (10.0 pmol), or the synthetic inositol trisphosphate isomer, inositol 2,4,5-trisphosphate (1.0-10.0 pmol), mimicked the effect of Ins(1,4,5)P3, stimulating an I1 resulting from release of intracellular Ca2+ and an I2 resulting from influx of extracellular Ca2+. The results indicate that several inositol trisphosphate isomers stimulate both release of intracellular Ca2+ and influx of extracellular Ca2+. Ins(1,3,4,5)P4 also stimulated release of intracellular Ca2+, but it was neither sufficient nor required for Ca2+ influx.  相似文献   

6.
D-Myo-inositol 1,4,5-trisphosphate (Ins[1,4-,5]P3) inhibits rat heart sarcolemmal Ca(2+)-ATPase activity (T. H. Kuo, Biochem. Biophys. Res. Commun. 152: 1111, 1988). We have studied the effect and mechanism of action of Ins(1,4,5)P3 and related inositol phosphates on human red cell membrane Ca(2+)-ATPase (EC 3.6.1.3) activity in vitro. At 10(-6) M, Ins(1,4,5)P3 and D-myo-inositol 4,5-bisphosphate (Ins[4,5]P2) inhibited human erythrocyte membrane Ca(2+)-ATPase activity in vitro by 42 and 31%, respectively. D-Myo-inositol 1,3,4,5-tetrakisphosphate, D-myo-inositol 1,4-bisphosphate, and D-myo-inositol 1-phosphate were not inhibitory. Enzyme inhibition by Ins(1,4,5)P3 was blocked by heparin. Exogenous purified calmodulin also stimulated red cell membrane Ca(2+)-ATPase activity; this stimulation was inhibited by Ins(1,4,5)P3. Ins(4,5)P2 and Ins(1,4,5)P3, but not Ins(1,4)P2, inhibited the binding of [125I]calmodulin to red cell membranes. Thus, specific inositol phosphates reduce plasma membrane Ca(2+)-ATPase activity and enhancement of the latter in vitro by purified calmodulin. The mechanism of these effects may in part relate to inhibition by inositol phosphates of binding of calmodulin to erythrocyte membranes.  相似文献   

7.
In bovine adrenal microsomes, Ins(1,4,5)P3 binds to a specific high-affinity receptor site (Kd = 11 nM) with low affinity for two other InsP3 isomers, Ins(1,3,4)P3 and Ins(2,4,5)P3. In the same subcellular fractions Ins(1,4,5)P3 was also the most potent stimulus of Ca2+ release of all the inositol phosphates tested. Of the many inositol phosphates recently identified in angiotensin-II-stimulated adrenal glomerulosa and other cells, Ins(1,3,4,5)P4 has been implicated as an additional second messenger that may act in conjunction with Ins(1,4,5)P3 to elicit Ca2+ mobilization. In the present study, an independent action of Ins(1,3,4,5)P4 was observed in bovine adrenal microsomes. Heparin, a sulphated polysaccharide which binds to Ins(1,4,5)P3 receptors in several tissues, inhibited both the binding of radiolabelled Ins(1,4,5)P3 and its Ca2(+)-releasing activity in adrenal microsomes. In contrast, heparin did not inhibit the mobilization of Ca2+ by Ins(1,3,4,5)P4, even at doses that abolished the Ins(1,4,5)P3 response. Such differential inhibition of the Ins(1,4,5)P3- and Ins(1,3,4,5)P4-induced Ca2+ responses by heparin indicates that Ins(1,3,4,5)P4 stimulates the release of Ca2+ from a discrete intracellular store, and exerts this action via a specific receptor site that is distinct from the Ins(1,4,5)P3 receptor.  相似文献   

8.
We determined the temporal relationship between the formation of inositol phosphates and increase in cytosolic [Ca2+] elicited by bombesin, vasopressin and platelet-derived growth factor (PDGF) in quiescent Swiss 3T3 cells. These responses were measured under identical conditions. Bombesin caused a rapid increase in inositol 1,4,5-trisphosphate which coincided with the increase in cytosolic [Ca2+]. This was followed by a slower but marked increase in inositol 1,3,4-trisphosphate and inositol-bisphosphate. Vasopressin elicited a similar sequence of events. In sharp contrast, highly purified porcine PDGF induced increases in cytosolic [Ca2+] and inositol 1,4,5-trisphosphate that were temporally uncoupled: detectable inositol polyphosphate formation occurred after Ca2+ mobilization from intracellular stores. The same temporal dissociation was observed when a recombinant v-sis product was used instead of porcine PDGF. However, PDGF was as effective as bombesin in stimulating the formation of inositol phosphates after 5-10 min of incubation. The data suggest that PDGF increases cytosolic [Ca2+] via a different signal transduction pathway from that utilized by bombesin and vasopressin. These findings have important implications for understanding the signal transduction pathway activated by PDGF.  相似文献   

9.
Despite no evidence for desensitization of phospholipase C-coupled gonadotropin-releasing hormone (GnRH) receptors, we previously reported marked suppression of GnRH-mediated Ca(2+) responses in alphaT3-1 cells by pre-exposure to GnRH. This suppression could not be accounted for solely by reduced inositol 1,4,5-trisphosphate (Ins(1,4,5)P(3)) responses, thereby implicating uncoupling of Ins(1,4,5)P(3) production and Ca(2+) mobilization (McArdle, C. A., Willars, G. B., Fowkes, R. C., Nahorski, S. R., Davidson, J. S., and Forrest-Owen, W. (1996) J. Biol. Chem. 271, 23711-23717). In the current study we demonstrate that GnRH causes a homologous and heterologous desensitization of Ca(2+) signaling in alphaT3-1 cells that is coincident with a rapid (t((12)) < 20 min), marked, and functionally relevant loss of type I Ins(1,4,5)P(3) receptor immunoreactivity and binding. Furthermore, using an alphaT3-1 cell line expressing recombinant muscarinic M(3) receptors we show that the unique resistance of the GnRH receptor to rapid desensitization contributes to a fast, profound, and sustained loss of Ins(1,4,5)P(3) receptor immunoreactivity. These data highlight a potential role for rapid Ins(1,4,5)P(3) receptor down-regulation in homologous and heterologous desensitization and in particular suggest that this mechanism may contribute to the suppression of the reproductive system that is exploited in the major clinical applications of GnRH analogues.  相似文献   

10.
Inositol phosphates: proliferation, metabolism and function   总被引:21,自引:0,他引:21  
After the initial discovery of receptor-linked generation of inositol(1,4,5)trisphosphate (Ins(1,4,5)P3) it was generally assumed that Ins(1,4,5)P3 and its proposed breakdown products inositol(1,4)bisphosphate (Ins(1,4)P2) and Ins1P, along with cyclic inositol monophosphate, were the only inositol phosphates found in significant amounts in animal cells. Since then, three levels of complexity have been introduced. Firstly, Ins(1,4,5)P3 can be phosphorylated to Ins(1,3,4,5)P4, and the subsequent metabolism of these two compounds has been found to be intricate and probably different between various tissues. The functions of Ins(1,4,5)P3 and Ins(1,3,4,5)P4 are almost certainly to regulate cytosolic Ca2+ concentrations, but the reasons for the labyrinth of the metabolic pathways after their deactivation by a specific 5-phosphatase remain obscure. Secondly, inositol pentakis- and hexakisphosphates have been found in many animal cells other than avian erythrocytes. It has been shown that their synthesis pathway is entirely separate from the inositol phosphates discussed above, both in terms of many of the isomers involved and probably in the subcellular localization; some possible functions of InsP5 and InsP6 are discussed here. Thirdly, cyclic inositol polyphosphates have been reported in stimulated tissues; the evidence for their occurrence in vivo and their possible physiological significance are also discussed.  相似文献   

11.
In previous works, we synthesized a series of inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) analogs, with a substituent on the second carbon of the inositol ring. Using these analogs, the Ins(1,4,5)P3 affinity media were also synthesized (Hirata, M., Watanabe, Y., Ishimatsu, T., Yanaga, F., Koga, T., and Ozaki, S. (1990) Biochem. Biophys. Res. Commun. 168, 379-386). When the cytosol fraction from the rat brain was applied to an Ins(1,4,5)P3 affinity column, an eluate with a 2 M NaCl solution was found to have remarkable Ins(1,4,5)P3-binding activity. The active fraction was further fractionated with gel filtration chromatography, and two proteins with an apparent molecular mass of 130 or 85 kDa were found to be Ins(1,4,5)P3-binding proteins but with no Ins(1,4,5)P3 metabolizing activities. Partial amino acid sequences determined after proteolysis and reversed-phase chromatography revealed that the protein with an apparent molecular mass of 85 kDa is the delta-isozyme of phospholipase C and that of 130 kDa has no sequence the same as the Ins(1,4,5)P3-recognizing proteins hitherto examined. Ins(1,4,5)P3 at concentrations greater than 1 microM strongly inhibited 85-kDa phospholipase C delta activity, without changing its dependence on the concentrations of free Ca2+ and H+. Among inositol phosphates examined, Ins(3,4,5,6)P4 inhibited the binding of [3H]Ins(1,4,5)P3 to the 130-kDa protein at much the same concentrations as seen with Ins(1,4,5)P3. This report seems to be the first evidence for the presence of soluble Ins(1,4,5)P3-binding proteins in the rat brain, one of which is the delta isozyme of phospholipase C.  相似文献   

12.
The binding of [3H]Ins(1,4,5)P3 to bovine adrenocortical microsomes has been shown to be rapid, reversible and saturable. The microsomal preparation contained a single population of high affinity sites (KD = 6.82+/-2.3 nM, Bmax = 370+/-38 fmol/mg protein). The binding site was shown to exhibit positional specificity with respect to inositol trisphosphate binding, i.e. Ins(2,4,5)P3 was able to compete with [3H]Ins(1,4,5)P3 whereas Ins(1,3,4)P3 was not. Ins(1,3,4,5)P4 showed a similar affinity for the receptor as Ins(2,4,5)P3 whereas the other inositol phosphates tested, ATP, GTP and 2,3-DPG, were poor competitors. [3H]Ins(1,4,5)P3-binding was independent of free Ca2+ concentrations. The adrenocortical microsomal preparation has been incorporated into an assay which has been used to determine the basal and vasopressin-stimulated content of neutralised acid extracts of rat hepatocytes. Intracellular concentrations of Ins(1,4,5)P3 were calculated to be 0.22+/-0.15 microM basal and 2.53+/-1.8 microM at peak stimulation. This assay provides a simple, specific and quantitative method for the measurement of Ins(1,4,5)P3 concentrations in the picomolar range.  相似文献   

13.
The action of carbachol on the generation of inositol trisphosphate and tetrakisphosphate isomers was investigated in dog-thyroid primary cultured cells radiolabelled with [3H]inositol. The separation of the inositol phosphate isomers was performed by reverse-phase high pressure liquid chromatography. The structure of inositol phosphates co-eluting with inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] and inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4] standards was determined by enzymatic degradation using a purified Ins(1,4,5)P3/Ins(1,3,4,5)P4 5-phosphatase. The data indicate that Ins(1,3,4,5)P4 was the only [3H]inositol phosphate which co-eluted with a [32P]Ins(1,3,4,5)P4 standard, whereas 80% of the [3H]InsP3 co-eluting with an Ins(1,4,5)P3 standard was actually this isomer. In the presence of Li+, carbachol led to rapid increases in [3H]Ins(1,4,5)P4. The level of Ins(1,4,5)P3 reached a peak at 200% of the control after 5-10 s of stimulation and fell to a plateau that remained slightly elevated for 2 min. The level of Ins(1,3,4,5)P4 reached its maximum at 20s. The level of inositol 1,3,4-trisphosphate [Ins(1,3,4)P3] increased continuously for 2 min after the addition of carbachol. Inositol-phosphate generation was also investigated under different pharmacological conditions. Li+ largely increased the level of Ins(1,3,4)P3 but had no effect on Ins(1,4,5)P3 and Ins(1,3,4,5)P4. Forskolin, which stimulates dog-thyroid adenylate cyclase and cyclic-AMP accumulation, had no effect on the generation of inositol phosphates. The absence of extracellular Ca2+ largely decreased the level of Ins(1,3,4,5)P4 as expected considering the Ca2(+)-calmodulin sensitivity of the Ins(1,4,5)P3 3-kinase. Staurosporine, an inhibitor of protein kinase C, increased the levels of Ins(1,4,5)P3, Ins(1,3,4,5)P4 and Ins(1,3,4)P3. This supports a negative feedback control of diacyglycerol on Ins(1,4,5)P3 generation.  相似文献   

14.
Multiple inositol polyphosphate phosphatase (MIPP) is an enzyme that, in vitro, has the interesting property of degrading higher inositol polyphosphates to the Ca2+ second messenger, inositol 1,4,5-trisphosphate (Ins(1,4,5)P3), independently of inositol lipid breakdown. We hypothesized that a truncated cytosolic form of the largely endoplasmic reticulum-confined MIPP (cyt-MIPP) could represent an important new tool in the investigation of Ins(1,4,5)P3-dependent intracellular Ca2+ homeostasis. To optimize our ability to judge the impact of cyt-MIPP on intracellular Ca2+ concentration ([Ca2+]i) we chose a poorly responsive beta-cell line (HIT M2.2.2) with an abnormally low [Ca2+]i. Our results show for the first time in an intact mammalian cell that cyt-MIPP expression leads to a significant enhancement of Ins(1,4,5)P3 concentration. This is achieved without a significant interference from other cyt-MIPP-derived inositol phosphates. Furthermore, the low basal [Ca2+]i of these cells was raised to normal levels (35 to 115 nm) when they expressed cyt-MIPP. Noteworthy is that the normal feeble glucose-induced Ca2+ response of HIT M2.2.2 cells was enhanced dramatically by mechanisms related to this increase in basal [Ca2+]i. These data support the use of cyt-MIPP as an important tool in investigating Ins(1,4,5)P3-dependent Ca2+ homeostasis and suggest a close link between Ins(1,4,5)P3 concentration and basal [Ca2+]i, the latter being an important modulator of Ca2+ signaling in the pancreatic beta-cell.  相似文献   

15.
The ability of epidermal growth factor (EGF) and angiotensin II to stimulate production of inositol trisphosphate and mobilize intracellular Ca2+ in hepatocytes was compared using quin2 fluorescence to monitor changes in Ca2+ levels and high performance liquid chromatography to resolve the inositol trisphosphate (InsP3) isomers. Both EGF and angiotensin II stimulated an increase in free intracellular Ca2+ concentration ([Ca2+]i) as well as a rapid increase in the production of inositol 1,4,5-trisphosphate (Ins(1,4,5)P3). Concentrations of angiotensin II which gave a rise in [Ca2+]i equivalent to that seen with maximal doses of EGF produced an equivalent increase in Ins(1,4,5)P3 formation. Both EGF and angiotensin II stimulated the formation of the Ins(1,3,4)P3 and inositol 1,3,4,5-tetrakisphosphate isomers. The formation of the Ins(1,3,4)P3 isomer lagged behind production of Ins(1,4,5)P3 but eventually reached higher levels in the cell. The initial rise in [Ca2+]i and InsP3 levels stimulated by EGF and angiotensin II was not affected by reducing the external Ca2+ concentration below 30 nM with an excess of [ethylenebis(oxyethylenenitrilo)] tetraacetic acid. Treatment of hepatocytes for 30-180 s with 1 micrograms/ml phorbol 12-myristate 13-acetate prior to the addition of EGF blocked the EGF-stimulated production of Ins(1,4,5)P3 and the increase in [Ca2+]i. Phorbol 12-myristate 13-acetate attenuated the production of Ins(1,4,5)P3 generated by angiotensin II over the concentration range of 10(-10) to 10(-8) M; however, the Ca2+ signal was only inhibited at the 10(-10) M dose of angiotensin II. Treatment of rats with pertussis toxin for 72 h prior to isolating hepatocytes blocked the ability of EGF to increase Ins(1,4,5)P3 and Ins(1,3,4)P3 but did not inhibit the ability of any concentration of angiotensin II to stimulate formation of InsP3 or inositol tetrakisphosphate. The observation that pertussis toxin selectively abolishes EGF-stimulated inositol lipid breakdown suggests that EGF and angiotensin II use different mechanisms to activate phospholipase C in hepatocytes.  相似文献   

16.
The release of Ca2+ from intracellular stores is triggered by the second messenger inositol (1,4,5)-trisphosphate (Ins(1,4,5)P3). The regulation of this process is critically important for cellular homeostasis. Ins(1,4,5)P3 is rapidly metabolised, either to inositol (1,4)-bisphosphate (Ins(1,4)P2) by inositol polyphosphate 5-phosphatases or to inositol (1,3,4,5)-tetrakisphosphate (Ins(1,3,4,5)P4) by one of a family of inositol (1,4,5)P3 3-kinases (IP3-3Ks). Three isoforms of IP3-3K have now been identified in mammals; they have a conserved C-terminal catalytic domain, but divergent N-termini. This review discusses the metabolism of Ins(1,4,5)P3, compares the IP3-3K isoforms and addresses potential mechanisms by which their activity might be regulated.  相似文献   

17.
An enzyme which catalyses the ATP-dependent phosphorylation of inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] was purified approx. 180-fold from rat brain cytosol by (NH4)2SO4 precipitation, chromatography through hydroxyapatite, anion-exchange fast protein liquid chromatography and gel-filtration chromatography. Gel filtration on Sepharose 4B CL gives an Mr of 200 x 10(3) for the native enzyme. The inositol tetrakisphosphate (InsP4) produced by the enzyme has the chromatographic, chemical and metabolic properties of Ins(1,3,4,5)P4. Ins(1,4,5)P3 3-kinase displays simple Michaelis-Menten kinetics for both its substrates, having Km values of 460 microM and 0.44 microM for ATP and Ins(1,4,5)P3 respectively. When many of the inositol phosphates known to occur in cells were tested, only Ins(1,4,5)P3 was a substrate for the enzyme; the 2,4,5-trisphosphate was not phosphorylated. Inositol 4,5-bisphosphate and glycerophosphoinositol 4,5-bisphosphate were phosphorylated much more slowly than Ins(1,4,5)P3. CTP, GTP and adenosine 5'-[gamma-thio]triphosphate were unable to substitute for ATP. When assayed under conditions of first-order kinetics, Ins(1,4,5)P3 kinase activity decreased by about 40% as the [Ca2+] was increased over the physiologically relevant range. This effect was insensitive to the presence of calmodulin and appeared to be the result of an increase in the Km of the enzyme for Ins(1,4,5)P3. Preincubation with ATP and the purified catalytic subunit of cyclic AMP-dependent protein kinase did not affect the rate of phosphorylation of Ins(1,4,5)P3 when the enzyme was assayed at saturating concentrations of Ins(1,4,5)P3 or at concentrations close to its Km for this substrate.  相似文献   

18.
The Ca2(+)-mobilizing second messenger D-myo-inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) is converted to the putative messenger D-myo-inositol 1,3,4,5-tetrakisphosphate by Ins(1,4,5)P3 3-kinase. We found that cAMP-dependent protein kinase and protein kinase C phosphorylate, and thereby modulate, the activity of Ins(1,4,5)P3 3-kinase. cAMP-dependent kinase introduced a stoichiometric amount of phosphate at serine 109 of the 53-kDa polypeptide and caused a 1.8-fold increase in Vmax, whereas the protein kinase C-dependent phosphorylation reduced the Vmax to one-fourth of that of the unphosphorylated enzyme. Upon prolonged incubation, protein kinase C introduced phosphate at multiple sites in Ins(1,4,5)P3 3-kinase, and the resulting inactivation of the enzyme appeared to be well-correlated with the simultaneous phosphorylation of two major sites, serine 109 and serine 175. The Km for Ins(1,4,5)P3 was not affected significantly after phosphorylation by either protein kinase. We propose, therefore, that the phosphorylation of Ins(1,4,5)P3 3-kinase by cAMP-dependent kinase and protein kinase C constitutes mechanisms of cross-talk between cellular signaling pathways that use various second messengers such as inositol phosphates, diacylglycerol, Ca2+, and cAMP.  相似文献   

19.
In human HeLa carcinoma cells, histamine causes a dose-dependent formation of inositol phosphates, production of diacylglycerol and a transient rise in intracellular [Ca2+]. These responses are completely blocked by the H1-receptor antagonist pyrilamine. In streptolysin-O-permeabilized cells, formation of inositol phosphates by histamine is strongly potentiated by guanosine 5'-[gamma-thio]triphosphate and inhibited by guanosine 5'-[beta-thio]diphosphate, suggesting the involvement of a GTP-binding protein. Histamine stimulates the rapid but transient formation of Ins(1,4,5)P3, Ins(1,3,4)P3 and InsP4. InsP accumulates in a much more persistent manner, lasting for at least 30 min. Studies with streptolysin-O-permeabilized cells indicate that InsP accumulation results from dephosphorylation of Ins(1,4,5)P3, rather than direct hydrolysis of PtdIns. The rise in intracellular [Ca2+] is biphasic, with a very fast release of Ca2+ from intracellular stores, that parallels the Ins(1,4,5)P3 time course, followed by a more prolonged phase of Ca2+ influx. In individual cells, histamine causes a rapid initial hyperpolarization of the plasma membrane, which can be mimicked by microinjected Ins(1,4,5)P3. Histamine-induced hyperpolarization is followed by long-lasting oscillations in membrane potential, apparently owing to periodic activation of Ca2+-dependent K+ channels. These membrane-potential oscillations can be mimicked by microinjection of guanosine 5'-[gamma-thio]triphosphate, but are not observed after microinjection of Ins(1,4,5)P3. We conclude that H1-receptors in HeLa cells activate a PtdInsP2-specific phospholipase C through participation of a specific G-protein, resulting in long-lasting oscillations of cytoplasmic free Ca2+.  相似文献   

20.
In previous studies it has been shown that both bradykinin and histamine increase the formation of 3H-labeled inositol phosphates in adrenal chromaffin cells prelabelled with [3H]inositol and that both these agonists stimulate release of catecholamines by a mechanism dependent on extracellular calcium. Here, we have used mass assays of inositol 1,4,5-trisphosphate [Ins(1,4,5)P3] and inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4] to investigate changes in levels of these two candidates as second messengers in response to stimulation with bradykinin and histamine. Bradykinin increased the mass of Ins(1,3,4,5)P4 despite the failure in earlier studies with [3H]inositol-labelled cells to observe a bradykinin-mediated increase in content of [3H]InsP4. Bradykinin elicited a very rapid increase in level of Ins(1,4,5)P3, which was maximal at 5-10 s and then rapidly decreased to a small but sustained elevation at 2 min. The bradykinin-elicited Ins(1,3,4,5)P4 response increased to a maximum at 30-60 s and at 2 min was still elevated severalfold above basal levels. Histamine, which produced a larger overall total inositol phosphate response in [3H]inositol-loaded cells, produced significantly smaller Ins(1,4,5)P3 and Ins(1,3,4,5)P4 responses compared with bradykinin. The bradykinin stimulation of Ins(1,4,5)P3 accumulation was partially dependent on a high (1.8 mM) extracellular Ca2+ concentration, whereas the Ins(1,3,4,5)P4 response was almost completely lost when the extracellular Ca2+ concentration was reduced to 100 nM. Changes in the inositol polyphosphate second messengers are compared with the time course of bradykinin-stimulated increases in free intracellular Ca2+ concentrations and noradrenaline release.  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