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 46 毫秒
1.
In order to study the regulatory mechanisms of phospholipase C-gamma (PLC-gamma) via the intrinsic SH2/SH3 region (Z region), two recombinant Z proteins, rP45Z and rP38Z, derived from rat PLC-gamma 1 and PLC-gamma 2, respectively, were purified from the inclusion bodies of Escherichia coli. We examined their direct effects on phosphoinositide hydrolysis induced by four different PLC isoforms purified from bovine brain and thymus, and found that both of these Z proteins suppress the enzyme activity of all four PLC isoforms in a dose-dependent manner. This suppressive effect is very potent and stoichiometric. The kinetics studies indicate that the suppression is non-competitive. This suppression is eliminated by treatment with proteases but is not affected by heat treatment at 95 degrees C for 15 min, indicating that the primary structure might be important for the action of Z proteins. Comparative studies suggested that two Z proteins but not Src and phosphatidylinositol 3-kinase possess, adjacent to their SH2 and SH3 motifs, a phospholipase C inhibitor (PCI) region that strongly suppresses their phosphatidylinositol 4,5-bisphosphate (PIP2)-hydrolyzing activity. A series of synthetic peptides identical with the sequence of the proposed PCI region, including an octamer, YRKMRLRY, inhibited PIP2 hydrolysis induced by four different phospholipase C isoforms. These results demonstrate that both types of phospholipase C-gamma contain the PCI sequence which is responsible for the inhibition of PIP2 hydrolysis, indicating that phospholipase C-gamma is a self-regulating enzyme.  相似文献   

2.
A gene encoding phosphoinositide-specific phospholipase C (PLC), designated ML-PLCδ, was cloned from mud loach (Misgurnus mizolepis) liver. A complete cDNA encoding ML-PLCδ was isolated by screening the cDNA library of mud loach liver and using the 5′-rapid amplification of cDNA ends (RACE) method. The full-length ML-PLCδ gene contains an open reading frame of 2325 base pairs encoding a 774 amino acid protein with a molecular mass of 88,072 Da; this corresponds to the size of the protein expressed in Escherichia coli BL21 (DE3) using pET28a vector. It contains all of the characteristic domains found in mammalian PLCδ isozymes (PH domain, EF-hands, X–Y catalytic region, and a C2 domain). A homology search revealed that ML-PLCδ shares relatively high sequence identity with mammalian PLCδ1 (51–52%) and catfish PLCδ (64%). The recombinant ML-PLCδ protein expressed as a histidine-tagged fusion protein in E. coli was purified to apparent homogeneity by Ni2+-NTA affinity chromatography. The recombinant ML-PLCδ showed a concentration-dependent PLC activity to phosphatidylinositol 4,5-bis-phosphate (PIP2) and its activity was Ca2+-dependent, which was similar to mammalian PLCδ isozymes.  相似文献   

3.
In RBL-2H3 tumor mast cells, cross-linking the high affinity IgE receptor (FcεRI) with antigen activates cytosolic tyrosine kinases and stimulates Ins(1,4,5)P3 production. Using immune complex phospholipase assays, we show that FcεRI cross-linking activates both PLCγ1 and PLCγ2. Activation is accompanied by the increased phosphorylation of both PLCγ isoforms on serine and tyrosine in antigen-treated cells. We also show that the two PLCγ isoforms have distinct subcellular localizations. PLCγ1 is primarily cytosolic in resting RBL-2H3 cells, with low levels of plasma membrane association. After antigen stimulation, PLCγ1 translocates to the plasma membrane where it associates preferentially with membrane ruffles. In contrast, PLCγ2 is concentrated in a perinuclear region near the Golgi and adjacent to the plasma membrane in resting cells and does not redistribute appreciably after FcεRI cross-linking. The activation of PLCγ1, but not of PLCγ2, is blocked by wortmannin, a PI 3-kinase inhibitor previously shown to block antigen-stimulated ruffling and to inhibit Ins(1,4,5)P3 synthesis. In addition, wortmannin strongly inhibits the antigen-stimulated phosphorylation of both serine and tyrosine residues on PLCγ1 with little inhibition of PLCγ2 phosphorylation. Wortmannin also blocks the antigen-stimulated translocation of PLCγ1 to the plasma membrane. Our results implicate PI 3-kinase in the phosphorylation, translocation, and activation of PLCγ1. Although less abundant than PLCγ2, activated PLCγ1 may be responsible for the bulk of antigen-stimulated Ins(1,4,5)P3 production in RBL-2H3 cells.  相似文献   

4.
Clostridium perfringens alpha-toxin elicits various immune responses such as the release of cytokines, chemokines, and superoxide via the GM1a/TrkA complex. Alpha-toxin possesses phospholipase C (PLC) hydrolytic activity that contributes to signal transduction in the pathogenesis of gas gangrene. Little is known about the relationship between lipid metabolism and TrkA activation by alpha-toxin. Using live-cell fluorescence microscopy, we monitored transbilayer movement of diacylglycerol (DAG) with the yellow fluorescent protein-tagged C1AB domain of protein kinase C-γ (EYFP-C1AB). DAG accumulated at the marginal region of the plasma membrane in alpha toxin-treated A549 cells, which also exhibited GM1a clustering and TrkA phosphorylation. Annexin V binding assays showed that alpha-toxin induced the exposure of phosphatidylserine on the outer leaflet of the plasma membrane. However, H148G, a variant toxin which binds cell membrane and has no enzymatic activity, did not induce DAG translocation, GM1a clustering, or TrkA phosphorylation. Alpha-toxin also specifically activated endogenous phospholipase Cγ-1 (PLCγ-1), a TrkA adaptor protein, via phosphorylation. U73122, an endogenous PLC inhibitor, and siRNA for PLCγ-1 inhibited the formation of DAG and release of IL-8. GM1a accumulation and TrkA phosphorylation in A549 cells treated with alpha-toxin were also inhibited by U73122. These results suggest that the flip-flop motion of hydrophobic lipids such as DAG leads to the accumulation of GM1a and TrkA. We conclude that the formation of DAG by alpha-toxin itself (first step) and activation of endogenous PLCγ-1 (second step) leads to alterations in membrane dynamics, followed by strong phosphorylation of TrkA.  相似文献   

5.
As previously described, WRK1 plasma membrane possesses a vasopressin-sensitive phospholipase C [G. Guillon et al., FEBS Lett. 196, 155–159]. In the present study, we examined the sensitivity of this enzyme to guanylnucleotides. GTPγS induced a time- and dose-dependent stimulation of Ins(1,4,5)P3 and Ins(1,4)P2 accumulation. No accumulation of InsP1, Ins(1,3,4)P3 or Ins(1,3,4,5)P4 occured under similar conditions. Gpp(NH)p produced the same effect but was less potent. GTP and a nonhydrolyzable analogue of ATP, App(NH)p, were without effect. Calcium also stimulated the phospholipase C activity in a time- and dose-dependent manner. In the absence of calcium, the activity of GTPγS was considerably reduced. Physiological calcium concentrations (between 10−8 and 10−7M), allowed maximal GTPγS stimulation of phospholipase C activity. In this system, the presence of vasopressin alone did not generate inositol phosphate accumulation. However, this hormone: (i) reduced the lag-time observed during GTPγS stimulation, (ii) increased the sensitivity of phospholipase C to GTPγS, and (iii) did not modify the stimulation of phospholipase C induced by maximal doses of GTPγS. Unlike sodium fluoride, GTPγS elicited an irreversible activation of phospholipace C. Calcium, GTPγS and sodium fluoride stimulated the phospholipase C activity via mechanisms sharing a common step, since their maximal effects were not additive. Cholera toxin treatment, known to produce complete ADP-ribosylation of ‘αs’ subunits, partially reduced the basal and the maximal GTPγS-mediated stimulation of phospholipase C activity as well as that caused by vasopressin. This inhibition was not mimicked by treatment with either forskolin or pertussi toxin.  相似文献   

6.
Microinjection of a truncated form of the c-kit tyrosine kinase present in mouse spermatozoa (tr-kit) activates mouse eggs parthenogenetically, and tr-kit– induced egg activation is inhibited by preincubation with an inhibitor of phospholipase C (PLC) (Sette, C., A. Bevilacqua, A. Bianchini, F. Mangia, R. Geremia, and P. Rossi. 1997. Development [Camb.]. 124:2267–2274). Co-injection of glutathione-S-transferase (GST) fusion proteins containing the src-homology (SH) domains of the γ1 isoform of PLC (PLCγ1) competitively inhibits tr-kit– induced egg activation. A GST fusion protein containing the SH3 domain of PLCγ1 inhibits egg activation as efficiently as the whole SH region, while a GST fusion protein containing the two SH2 domains is much less effective. A GST fusion protein containing the SH3 domain of the Grb2 adaptor protein does not inhibit tr-kit–induced egg activation, showing that the effect of the SH3 domain of PLCγ1 is specific. Tr-kit–induced egg activation is also suppressed by co-injection of antibodies raised against the PLCγ1 SH domains, but not against the PLCγ1 COOH-terminal region. In transfected COS cells, coexpression of PLCγ1 and tr-kit increases diacylglycerol and inositol phosphate production, and the phosphotyrosine content of PLCγ1 with respect to cells expressing PLCγ1 alone. These data indicate that tr-kit activates PLCγ1, and that the SH3 domain of PLCγ1 is essential for tr-kit–induced egg activation.  相似文献   

7.
In rat luteal cells labeled with (3H]oleic acid, PGF-stimulated phospholipase D (PLD) activation was investigated. The PLD activity was detected by measuring the accumulation of [3H]phosphatidylethanol (PtdEt) in the presence of ethanol. PGF stimulated PtdEt accumulation at concentrations of more than 100 nM in the presence of ethanol. However, PtdEt accumulation did not change in the absence of ethanol. PGF (1 μM) increased PtdEt accumulation after 1 min, and the accumulation reached a plateau by 2–3 min. These results indicate that PGF activates PLD in rat luteal cells. U-73122, a phospholipase C (PLC) inhibitor, and staurosporine, a protein kinase C (PKC) inhibitor, did not inhibit PGF-stimulated [3H]PtdEt accumulation. These results suggest that PGF-induced PLD activation is different from PLC-PKC systems. We reported previously that PGF stimulated the release of arachidonic acid. The effects of indomethacin, nordihydroguaiaretic acid (NDGA), and 5,8,11,14-eicosatetraynoic acid (ETYA), inhibitors of arachidonic acid metabolism, on PGF-stimulated PtdEt accumulation were examined. Pretreatment with indomethacin enhanced PGF-induced PtdEt accumulation. In contrast, pretreatment with NDGA and ETYA inhibited PGF-induced PtdEt accumulation. It is suggested that PGF-stimulated PLD activation is mediated via lipoxygenase products.  相似文献   

8.
Summary The objectives of this study were to determine the genetic basis of the electrophoretic differences of human plasma protein C inhibitors (PCI) from 977 individuals. Three discrete antibodies were produced against the PCI purified from human plasma and peptides that corresponded to the N-terminal 15 amino acid residues and the C-terminal 15 residues of human PCI, the chemical structures of which were determined by cDNA sequence analysis. The combined techniques of polyacrylamide gel isoelectric focusing and immunoblotting with these three different antibodies resolved the plasma PCI into several isoprotein bands, with a pH range of 6–7. These PCI isoproteins, however, were not stained by anti-human kallikrein, anti-human protein C or anti-human urokinase antibodies. Therefore, each of the PCI bands, which were detected by immunoblotting with the anti-PCI antibody and the two different anti-peptide antibodies, were derived from free PCI, and not an inactive PCI species. Two common phenotypes, designated PCI 1 and 1–2, were recognized, and family studies showed that they represented homozygosity or heterozygosity for two autosomal codominant alleles, PCI *1 and PCI *2. A population study of plasma samples collected from 977 Japanese individuals indicated that the frequencies of the PCI *1 and PCI *2 alleles were 0.988 and 0.012, respectively.  相似文献   

9.
A sensitive and specific assay of human plasma for the determination of (5α,7β,16β)-16[(4-chlorophenyl)oxy]-4,7-dimethyl-4-aza-andronstan-3-one (I), a selective inhibitor of human type 1 5α-reductase, has been developed. The method is based on high-performance liquid chromatography (HPLC) with tandem mass spectrometric (MS–MS) detection. The analyte (I) and internal standard, Proscar (II), were isolated from the basified biological matrix using a liquid–liquid extraction with methyl-tert.-butyl ether (MTBE). The organic extract was evaporated to dryness, the residue was reconstituted in mobile phase and injected into the HPLC system. The MS–MS detection was performed on a PE Sciex API III Plus tandem mass spectrometer using a heated nebulizer interface. Multiple reaction monitoring using the precursor→product ion combinations of m/z 430→114 and 373→305 was used to quantify I and internal standard (II), respectively. The assay was validated in the concentration range of 0.5 to 500 ng/ml in human plasma. The precision of the assay, expressed as coefficient of variation (C.V.), was less than 7% over the entire concentration range, with adequate assay specificity and accuracy. The HPLC–MS–MS method provided sufficient sensitivity to completely map the 24 h pharmacokinetic time-course following a single 0.5 mg dose of I.  相似文献   

10.
Microbial transformation of isosteviol oxime (ent-16-E-hydroxyiminobeyeran-19-oic acid) (2) with Aspergillus niger BCRC 32720 and Absidia pseudocylindrospora ATCC 24169 yielded several compounds. In addition to bioconverting the d-ring to lactone and lactam moieties, 4α-carboxy-13α-hydroxy-13,16-seco-ent-19-norbeyeran-16-oic acid 13,16-lactone (7) and 4α-carboxy-13α-amino-13,16-seco-ent-19-norbeyeran-16-oic acid 13,16-lactam (10), one known compound, ent-1β,7α-dihydroxy-16-oxo-beyeran-19-oic acid (6), and five new compounds, ent-7α-hydroxy-16-E-hydroxyiminobeyeran-19-oic acid (3), ent-1β,7α-dihydroxy-16-E-hydroxyiminobeyeran-19-oic acid (4), ent-1β-hydroxy-16-E-hydroxyiminobeyeran-19-oic acid (5), ent-8β-cyanomethyl-13-methyl-12-podocarpen-19-oic acid (8), and ent-8β-cyanomethyl-13-methyl-13-podocarpen-19-oic acid (9), were isolated from the microbial transformation of 2. Elucidation of the structures of these isolated compounds was primarily based on 1D and 2D NMR, and HRESIMS data, and 35 were further confirmed by X-ray crystallographic analyses. Additionally, the inhibitory effects of all of these compounds were evaluated on NF-κB and AP-1 activation in LPS-stimulated RAW 264.7 macrophages. Among the compounds tested, 5 and 10 significantly inhibited NF-κB activation, with 5 showing equal potency to dexamethasone; 3 and 69 significantly inhibited AP-1 activation, particularly 8, which showed more inhibitory activity than dexamethasone.  相似文献   

11.
The human genes encoding α1-antitrypsin (α1AT, gene symbol PI), corticosteroid-binding globulin (CBG), α1-antichymotrypsin (AACT), and protein C inhibitor (PCI) are related by descent, and they all map to human chromosome 14q32.1. This serine protease inhibitor (serpin) gene cluster also contains an antitrypsin-related sequence (ATR, gene symbol PIL), but the precise molecular organization of this region has not been defined. In this report we describe the generation and characterization of an 370-kb cosmid contig that includes all five serpin genes. Moreover, a newly described serpin, kallistatin (KAL, gene symbol PI4), was also mapped within the region. Gene order within this interval is cen–CBG–ATR–α1AT–KAL–PCI–AACT–tel. The genes occupy 320 kb of genomic DNA, and they are organized into two discrete subclusters of three genes each that are separated by 170 kb. The distal subcluster includes KAL, PCI, and AACT; it occupies 63 kb of DNA, and all three genes are transcribed in a proximal-to-distal orientation. Within the subcluster, there is 12 kb of intergenic DNA between KAL and PCI and 19 kb between PCI and AACT. The proximal subcluster includes α1AT, ATR, and CBG; it occupies 90 kb of genomic DNA, with 12 kb of DNA between α1AT and ATR and 40 kb between ATR and CBG. These genes are all transcribed in a distal-to-proximal orientation. This represents the first detailed physical map of the serpin gene cluster on 14q32.1.  相似文献   

12.
Recent evidence has revived interest in the idea that phosphoinositides (PIs) may play a role in signal transduction in mammalian olfactory receptor neurons (ORNs). To provide direct evidence that odorants indeed activate PI signaling in ORNs, we used adenoviral vectors carrying two different fluorescently tagged probes, the pleckstrin homology (PH) domains of phospholipase Cδ1 (PLCδ1) and the general receptor of phosphoinositides (GRP1), to monitor PI activity in the dendritic knobs of ORNs in vivo. Odorants mobilized PI(4,5)P2/IP3 and PI(3,4,5)P3, the substrates and products of PLC and PI3K. We then measured odorant activation of PLC and PI3K in olfactory ciliary-enriched membranes in vitro using a phospholipid overlay assay and ELISAs. Odorants activated both PLC and PI3K in the olfactory cilia within 2 s of odorant stimulation. Odorant-dependent activation of PLC and PI3K in the olfactory epithelium could be blocked by enzyme-specific inhibitors. Odorants activated PLC and PI3K with partially overlapping specificity. These results provide direct evidence that odorants indeed activate PI signaling in mammalian ORNs in a manner that is consistent with the idea that PI signaling plays a role in olfactory transduction.  相似文献   

13.
In this study, we purified and characterized the voltage-dependent anion channel (VDAC) from the Torpedo electric organ. Using immunogold labeling, VDAC was colocalized with the voltage-gated Ca2+ channel in the synaptic plasma membrane. By immunoblot analysis, five protein bands in synaptosomes isolated from the Torpedo electric organ cross reacted with two monoclonal anti-VDAC antibody. No more than about 7 to 10% mitochondrial contains could be detected in any synaptosomal membrane preparation tested. This was estimated by comparing the specific activity in mitochondria and synaptosomes of succinate–cytochrome-c oxidoreductase and antimycin-insensitive NADH–cytochrome-c oxidoreductase activities; mitochondrial inner and outer membrane marker enzymes, respectively. [14C]DCCD (dicyclohexylcarbodiimide), which specifically label mitochondrial VDAC, labeled four 30–35 kDa protein bands that were found to interact with the anti-VDAC antibody. The distribution of the Torpedo VDAC protein bands was different among membranes isolated from various tissues. VDAC was purified from synaptosomes and a separation between two of the proteins was obtained. The two purified proteins were characterized by their single channel activity and partial amino acid sequences. Upon reconstitution into a planar lipid bilayer, the purified VDACs showed voltage-dependent channel activity with properties similar to those of purified mitochondrial VDAC. Amino acid sequence of four peptides, derived from VDAC band II, exhibited high homology to sequences present in human VDAC1 (98%), VDAC2 (91.8%), and VDAC3 (90%), while another peptide, derived from VDAC band III, showed lower homology to either VDAC1 (88.4%) or VDAC2 (79%). Two more peptides show high homology to the sequence present in mouse brain VDAC3 (100 and 78%). In addition, we demonstrate the translocation of ATP into synaptosomes, which is inhibited by DCCD and by the anion transport inhibitor DIDS. The possible function of VDAC in the synaptic plasma membrane is discussed.  相似文献   

14.
Agonist stimulation of G protein-coupled receptors (GPCRs) can transactivate epidermal growth factor receptors (EGFRs), but the precise mechanisms for this transactivation have not been defined. Key to this process is the protease-mediated “shedding” of membrane-tethered ligands, which then activate EGFRs. The specific proteases and the events involved in GPCR-EGFR transactivation are not fully understood. We have tested the hypothesis that transactivation can occur by a membrane-delimited process: direct increase in the activity of membrane type-1 matrix metalloprotease (MMP14, MT1-MMP) by heterotrimeric G proteins, and in turn, the generation of heparin-binding epidermal growth factor (HB-EGF) and activation of EGFR. Using membranes prepared from adult rat cardiac myocytes and fibroblasts, we found that MMP14 activity is increased by angiotensin II, phenylephrine, GTP, and guanosine 5′-O-[γ-thio]triphosphate (GTPγS). MMP14 activation by GTPγS occurs in a concentration- and time-dependent manner, does not occur in response to GMP or adenosine 5′-[γ-thio]triphosphate (ATPγS), and is not blunted by inhibitors of Src, PKC, phospholipase C (PLC), PI3K, or soluble MMPs. This activation is specific to MMP14 as it is inhibited by a specific MMP14 peptide inhibitor and siRNA knockdown. MMP14 activation by GTPγS is pertussis toxin-sensitive. A role for heterotrimeric G protein βγ subunits was shown by using the Gβγ inhibitor gallein and the direct activation of recombinant MMP14 by purified βγ subunits. GTPγS-stimulated activation of MMP14 also results in membrane release of HB-EGF and the activation of EGFR. These results define a previously unrecognized, membrane-delimited mechanism for EGFR transactivation via direct G protein activation of MMP14 and identify MMP14 as a heterotrimeric G protein-regulated effector.  相似文献   

15.
Summary The role of phospholipids in the binding of [3H] tetrodotoxin to garfish olfactory nerve axon plasma membrane was studied by the use of purified phospholipases. Treatment of the membranes with low concentrations of either phospholipase A2 (Crotalus adamanteus andNaja naja) or phospholipase C (Bacillus cereus andClostridium perfringens) resulted in a marked reduction in tetrodotoxin binding activity. A 90% reduction in the activity occurred with about 45% hydrolysis of membrane phospholipids by phospholipase A2, and with phospholipase C the lipid hydrolysis was about 60–70% for a 70–80% reduction in the binding activity. Phospholipase C fromB. cereus andCl. perfringens had similar inhibitory effects. Bovine serum albumin protected the tetrodotoxin binding activity of the membrane from the inhibitory effect of phospholipase A2 but not from that of phospholipase C. In the presence of albumin about 25% of the membrane phospholipids remained unhydrolyzed by phospholipase A2. It is suggested that these unhydrolyzed phospholipids are in a physical state different from the rest of the membrane phospholipids and that these include the phospholipids which are directly related to the tetrodotoxin binding component. It is concluded that phospholipids form an integral part of the tetrodotoxin binding component of the axon membrane and that the phospholipase-caused inhibition of the binding activity is due to effects resulting from alteration of the phospholipid components.  相似文献   

16.
A second trypsin inhibitor (DMTI-II) was purified from the seed of Dimorphandra mollis (Leguminosae-Mimosoideae) by ammonium sulfate precipitation (30–60%), gel filtration, and ion-exchange and affinity chromatography. A molecular weight of 23 kDa was estimated by gel filtration on a Superdex 75 column SDS-PAGE under reduced conditions showed that DMTI-II consisted of a single polypeptide chain, although isoelectric focusing revealed the presence of three isoforms. The dissociation constant of 1.7 × 10–9 M with bovine trypsin indicated a high affinity between the inhibitor and this enzyme. The inhibitory activity was stable over a wide pH range and in the presence of DTT. The N-terminal sequence of DMTI-II showed a high degree of homology with other Kunitz-type inhibitors.  相似文献   

17.
A series of m-alkyl α,α,α-trifluoroacetophenones (1–5) was synthesized and evaluated as inhibitors of acetylcholinesterase from Torpedo california. All ketones (1–5) were found to be potent inhibitors of the enzyme; m-t-butyl α,α,α-trifluoroacetophenone (4) was the most potent inhibitor with a Ki value of 3.7 pM.  相似文献   

18.
Protein C inhibitor (PCI) is a serpin with broad protease reactivity. It binds glycosaminoglycans and certain phospholipids that can modulate its inhibitory activity. PCI can penetrate through cellular membranes via binding to phosphatidylethanolamine. The exact mechanism of PCI internalization and the intracellular role of the serpin are not well understood. Here we showed that testisin, a glycosylphosphatidylinositol-anchored serine protease, cleaved human PCI and mouse PCI (mPCI) at their reactive sites as well as at sites close to their N terminus. This cleavage was observed not only with testisin in solution but also with cell membrane-anchored testisin on U937 cells. The cleavage close to the N terminus released peptides rich in basic amino acids. Synthetic peptides corresponding to the released peptides of human PCI (His1–Arg11) and mPCI (Arg1–Ala18) functioned as cell-penetrating peptides. Because intact mPCI but not testisin-cleaved mPCI was internalized by Jurkat T cells, a truncated mPCI mimicking testisin-cleaved mPCI was created. The truncated mPCI lacking 18 amino acids at the N terminus was not taken up by Jurkat T cells. Therefore our model suggests that testisin or other proteases could regulate the internalization of PCI by removing its N terminus. This may represent one of the mechanisms regulating the intracellular functions of PCI.  相似文献   

19.
Phosphoinositide-specific phospholipase C (PLC) control the levels of their substrate phosphatidylinositol 4,5-bisphosphate (PI(4,5)P2), and its hydrolysis products diacylglycerol (DAG) and Ins(1,4,5)P3, second messengers key to growth control and cell movement. The former is modulated by breakdown of plasma membrane and nuclear phosphoinositides, while the latter is mediated by phosphoinositide-driven remodeling of the actin cytoskeleton. The roles of PLC in the etiology and progression of breast carcinoma, however, are poorly understood. Previous studies reported a correlation between PLCβ2 expression and breast tumor grade, making PLCβ2 a potential marker for clinical outcome (Bertagnolo et al., 2006). While over-expression of PLCβ2 is not sufficient to induce transformation of normal breast epithelial cells, it appears to play a role in promoting cell migration (Bertagnolo et al., 2007).Here we examined the expression of this and other PLC mRNA (β1–β4, δ1, δ3 and δ4, γ1 and γ2) in normal breast epithelial lines, MCF-10A, and compared that pattern to breast tumor lines MDA-MB-231 and to T47D, using real-time relative-quantification PCR. Our results show that PLCγ1, γ2 and δ1 and δ3 are more highly expressed in the transformed cell lines compared to MCF-10A when normalized to mRNA encoding various house keeping proteins; whereas PLCβ2 mRNA levels were considerably lower than other PLC subtypes, including PLCβ1 in the metastatic lines. Examination of PLC mRNA levels from normal and cancerous human breast tissue showed a similar pattern of expression, however, when staging or tumor size was considered, PLCδ1 and δ3 expression were positively correlated.To test whether PLCδ1 or δ3 played any role in tumor cell proliferation or cell migration, we transfected cells with siRNA specifically targeting these isoforms. RNAi mediated knockdown of either PLC isoform, reduced proliferation of the MDA-MB-231 cells. Morphological changes including cell rounding, and surface blebbing and nuclear fragmentation were observed. These changes were accompanied by reductions in cell migration activities. On the other hand, PLCδ1 knockdown failed to cause comparable morphological changes in the normal MCF-10A line, but did reduce cell proliferation and migration. Taken together, these data are consistent with the idea that PLCδ1 and δ3 isoforms support the growth and migration of normal and neoplastic mammary epithelial cells in vitro.  相似文献   

20.
Plasma β-amyloid protein (Aβ) isoforms are considered potential biomarkers for Alzheimer's disease (AD) and dementia. The relation between plasma and cerebrospinal fluid (CSF) levels of Aβ isoforms remains unclear. In order to identify possible correlations between Aβ levels in plasma and CSF we determined Aβ levels in time-linked plasma and CSF samples. Aβ concentrations in plasma (Aβ1–42 and AβN–42) and CSF (Aβ1–42) samples from 49 AD patients, 47 non-Alzheimer's disease dementia (NONAD) patients, 39 MCI patients and 29 controls were determined using a multi-parameter fluorimetric bead-based immunoassay using xMAP® technology (for plasma) and a conventional single-parameter ELISA (for CSF). Plasma Aβ1–42 concentrations did not correlate with CSF Aβ1–42 concentrations in the total study population, or in the different diagnostic groups. No correlations between plasma AβN–42 and CSF Aβ1–42 levels were found either. The CSF/serum albumin index did not show any significant differences between AD, NONAD, MCI and controls.These results suggest that the Aβ levels in plasma are independent of the Aβ levels in CSF both in dementia and controls. The fact that CSF and plasma Aβ do not correlate in patients as well as controls and no significant differences in plasma Aβ1–42 or AβN–42 between patients and controls can be detected hampers the diagnostic utility of the plasma Aβ levels as biomarkers for dementia.  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