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 550 毫秒
1.
Annexin VI, a member of a family of Ca(2+)-dependent phospholipid- and membrane-binding proteins, interacts with the Ca(2+)-regulated EF-hand proteins, S100A1 and S100B, and blocks the ability of these two proteins to inhibit the assembly of desmin and glial fibrillary acidic protein (GFAP) into intermediate filaments in a Ca(2+)- and dose-dependent manner. S100A1 and S100B each possess one annexin VI binding site, characterized by an affinity for annexin VI in the submicromolar range. Binding of annexin VI to either S100 protein occurs at a site that appears to differ in some parts from that recognizing desmin and GFAP. As S100A1 and S100B exist in solution as homodimers in which the two monomers are related by a 2-fold symmetry axis, each of the above S100 homodimers likely crosslinks two annexin VI molecules, a situation that appears typical of all the annexin-S100 protein complexes described thus far. However, whereas in the cases of other annexin-S100 complexes the C-terminal extension of the S100 molecule appears indispensable for annexin binding, the annexin VI binding site cannot be restricted to the S100A1 and S100B C-terminal extension. We speculate that the annexin VI site on S100A1/B may only partially overlap to the desmin/GFAP site. In contrast, no effects of annexin V on the ability of S100A1 or S100B to affect the desmin and GFAP assemblies could be documented, although binding of annexin V to S100A1 and S100B could be detected at relatively high Ca2+ concentrations. The present data suggest that annexin VI might regulate S100A1 and S100B activities and vice versa.  相似文献   

2.
Annexins belong to a family of proteins that are characterized by their ability to bind phospholipids in a Ca(2+)-dependent manner that is thought to be involved in a variety of biological processes. The present study determined the localization of annexins in subcellular fractions, nuclei in particular, of cow mammary gland by immunoblot analysis using monoclonal antibodies to annexins I, II, IV, and VI. The analysis revealed that annexins I, II, and IV were present in cytosol, but VI was not. Annexins I and IV were found in the nuclear fraction, but annexin II was only faintly present. Annexin VI was also undetectable in this fraction. Cytosolic annexin I had a molecular mass of 36 kDa. The 36-kDa annexin I was also found in the nuclear fraction. A 38-kDa annexin I was additionally detected in nuclei. The cytosolic and nuclear 36-kDa annexin I and the nuclear 38-kDa annexin I showed different isoelectric points, as revealed by two-dimensional PAGE. Annexin IV from cytosolic and nuclear fractions had similar molecular masses and isoelectric points.  相似文献   

3.
We have previously demonstrated that annexin IV, one of the calcium/phospholipid-binding annexin family proteins, binds to glycosaminoglycans (GAGs) in a calcium-dependent manner (Kojima, K., Yamamoto, K., Irimura, T., Osawa, T., Ogawa, H., and Matsumoto, I. (1996) J. Biol. Chem. 271, 7679-7685). In this study, we investigated the GAG binding specificities of annexins IV, V, and VI by affinity chromatography and solid phase assays. Annexin IV was found to bind in a calcium-dependent manner to all the GAG columns tested. Annexin V bound to heparin and heparan sulfate columns but not to chondroitin sulfate columns. Annexin VI was adsorbed to heparin and heparan sulfate columns in a calcium-independent manner, and to chondroitin sulfate columns in a calcium-dependent manner. An N-terminal half fragment (A6NH) and a C-terminal half fragment (A6CH) of annexin VI, each containing four units, were prepared by digestion with V8 protease and examined for GAG binding activities. A6NH bound to heparin in the presence of calcium but not to chondroitin sulfate C, whereas A6CH bound to heparin calcium-independently and to chondroitin sulfate C calcium-dependently. The results showed that annexin IV, V, and VI have different GAG binding properties. Some annexins have been reported to be detected not only in the cytoplasm but also on the cell surface or in extracellular components. The findings suggest that the some annexins function as recognition elements for GAGs in extracellular space.  相似文献   

4.
Oxidative stress can cause changes in intracellular free calcium concentration ([Ca2+]i) that resemble those occurring under normal cell signaling. In the alveolar macrophage, hydroperoxide-induced elevation of [Ca2+]i modulates the respiratory burst and other important physiologic functions. The source of Ca2+ released by hydroperoxide is intracellular but separate from the endoplasmic reticulum pool released by receptor-mediated stimuli (Hoyal, C. R., Gozal, E., Zhou, H., Foldenauer, K., and Forman, H. J. (1996) Arch. Biochem. Biophys. 326, 166-171). Previous studies in other cells have suggested that mitochondria are a potential source of oxidant-induced [Ca2+]i elevation. In this study we have identified another potential source of hydroperoxide-releasable intracellular calcium, that bound to annexin VI on the inner surface of the plasma membrane. Translocation of annexin VI from the membrane during exposure to t-butyl hydroperoxide matched elevation of [Ca2+]i as a function of time and t-butyl hydroperoxide concentration. The translocation was possibly due to a combination of ATP depletion and oxidative modification of membrane lipids and proteins. A sustained increase in [Ca2+]i occurring > 50 pmol/10(6) cells (50 microM under these conditions) appeared to be a consequence of membrane Ca2+-ATPase dysfunction. These results suggest that exposure to oxidative stress results in early alterations to the plasma membrane and concomitant release of Ca2+ into the cytosol. In addition it suggests a mechanism for participation of annexin VI translocation that may underlie the alterations in macrophage function by oxidative stress.  相似文献   

5.
Annexin V, VI and VII-core (delta1-107) are members of the annexin protein family and bind to acidic phospholipid membranes in a calcium dependent manner. They also show ion channel activity under certain conditions. As annexins bind peripherally to lipid membranes, ion channel formation must consist of at least two steps: An adsorption reaction regulating the binding of annexin to the membrane surface and the opening and closing of the active species controlling the channel activity. By using the baseline current through the patch clamp seal as a probe for unoccupied binding sites at the membrane, we show that the adsorption of annexins to membranes is not only calcium dependent but also strongly voltage dependent. Whereas the free transfer energies at low calcium concentrations are similar for all three annexins, the binding of annexin V becomes much tighter with higher calcium levels, compared to annexin VI and VII-core. This correlates with the finding that annexin VI and VII-core display channel activity much more often than annexin V if one assumes that a high coverage of the membrane surface with annexins stabilizes the bilayer. At higher protein concentrations weaker binding is observed in agreement with the previously reported anti-cooperativity of membrane binding.  相似文献   

6.
Annexins are a unique family of membrane-associated, Ca2+ and phospholipid-binding proteins found in various tissues. Among the 12 isoforms, Annexin II, V and VI exist in heart tissue in the highest amounts. Annexin VI has been shown to affect intracellular Ca2+ cycling and contractility in isolated cardiomyocytes. Annexin V is present in both cardiomyocytes and non-myocyte cell types in the heart and may play a role in the regulation of cellular ion fluxes, organization and secretion, while the cardiac effects of annexin II are unclear. To identify changes in annexin II, V and VI isoforms that might occur in human heart failure, we measured mRNA and protein levels of these three annexins in transplanted left ventricular tissue of 12 patients with end-stage congestive heart failure due to coronary artery disease (CAD, n=6) or idiopathic dilated cardiomyopathy (DCM, n=6) who underwent cardiac transplantation. Normal heart tissue (C, n=6) was used as a control. Northern blot analyses showed a significant decrease (61%) in annexin VI mRNA levels in heart failure patients compared with controls (1.08+/-0.16 v 2.79+/-0.20 A.U.C. unit, determined by laser densitometry, mean+/-s.e.). In contrast, we found a 67% increase (2. 32+/-0.27 v 3.88+/-0.29) in annexin II mRNA levels and a two-fold increase (1.00+/-0.24 v 2.21+/-0.29) in annexin V mRNA levels in cardiomyopathic hearts as compared to normal hearts. Western blot analyses demonstrated a corresponding decrease (46.1%) in annexin VI protein levels in the heart failure group as compared to controls (2. 63+/-0.22 v 4.88+/-0.52), while annexin II protein levels showed a significant 40.7% increase in patients with heart failure compared to those in normal hearts (5.08+/-0.67 v 3.61+/-0.32). Annexin V protein levels were also significantly increased (45%) in heart failure patients compared with normal (2.14+/-0.19 v 1.48+/-0.11). No difference in either annexins II, V or VI mRNA and protein levels were found between CAD and DCM patients. We conclude that human end-stage heart failure is associated with a down regulation of annexin VI and up regulation of annexin II and V proteins. Coordinate changes were observed in steady-state mRNA levels. These results suggest that these annexin isoforms may contribute to the regulation of intracellular Ca2+ homeostasis in the cardiomyopathic heart.  相似文献   

7.
In cell culture, human osteoblasts and the osteosarcoma cell line MG-63 express annexins I, II, IV, V and VI. Small proportions of annexins IV and V are lost from MG-63 cells into the culture medium in a sedimentable form. however, the bulk of these annexins is intracellular. In non-confluent cells 3 days after passaging, annexin IV and annexin V are strongly present throughout the nucleus and are also present in the cytoplasm. On elevation of the intracellular calcium concentration with the lonophore ionomycin, the intranuclear pools of annexin IV in 38 +/- 4% of cells and annexin V in 70 +/- 5% of cells show relocation to the nuclear membrane within 40 s. Extracellular ATP, which causes a transient increase in the cytosolic free calcium concentration by acting at P2-purinoceptors, also causes relocation of the intranuclear pool of annexin IV in 22 +/- 4% of cells and of annexin V in 38 +/- 8% of cells. After stimulation no significant reversal of the relocation is observed. Elevation of intracellular calcium with ionophore and ATP also causes relocation of the cytoplasmic pools of annexins IV and V. The results support a role for annexins at cellular membranes in response to elevation of cytosolic calcium levels.  相似文献   

8.
The crystal structure of a calcium-bound form of bovine annexin VI has been determined with X-ray diffraction data to 2.9 A by molecular replacement. Six Ca2+ ions were found, five in AB loops, one in a DE loop. Two loops (II-AB, which binds calcium, and V-AB, which does not) have conformations that differ significantly from those in calcium-free, human recombinant annexin VI. There are only small differences between the calci- and the apo-annexin VI in the rest of the molecule. Calcium by itself does not promote a major conformational change.  相似文献   

9.
To determine whether rat osteoblasts synthesize proteins of the annexin family and to evaluate the extent to which glucocorticoids modulate the expression of annexins by these cells, osteoblasts were grown in primary cultures in the absence or presence of dexamethasone, and the expression of annexins was evaluated by immunoblotting using polyclonal antibodies against human annexins. Four different annexins (I, II, V, and VI) were found to be expressed by rat osteoblasts. The expression of annexin I, but not the other annexins studied, was increased in osteoblasts cultured in the presence of dexamethasone (173 +/- 33% increase comparing untreated cells and cells treated for 10 days with 5 x 10(-7) M dexamethasone). Increased expression of annexin I was observed after the third day of exposure to dexamethasone and rose thereafter until day 10; annexin I expression increased with dexamethasone concentrations above 10(-10) M throughout the range of concentrations studied. The increase in annexin I protein was associated with an increase in annexin I mRNA and was completely blocked by the concomitant addition of the glucocorticoid receptor antagonist RU 38486. The increase in annexin I content following dexamethasone treatment was associated with an increase in alkaline phosphatase activity and PTH-induced cAMP stimulation, whereas phospholipase A2 activity in the culture medium was reduced to undetectable levels. The finding that four annexins are expressed in rat osteoblasts in primary culture raises the possibility that these proteins could play an important role in bone formation by virtue of their ability to bind calcium and phospholipids, serve as Ca2+ channels, interact with cytoskeletal elements, and/or regulate phospholipase A2 activity. In addition, the dexamethasone-induced increase in annexin I may represent a mechanism by which glucocorticoids modify osteoblast function.  相似文献   

10.
Previous studies have suggested that the NG2 proteoglycan interacts with type VI collagen. We have further characterized this interaction using a solid phase binding assay in which purified NG2 was shown to bind to pepsin-solubilized type VI collagen. In addition, NG2 bound a recombinant alpha2 (VI) collagen chain but did not appreciably bind to the recombinant alpha1 (VI) chain or the N-terminal domain of alpha3 (VI) (N9-N2). Binding of NG2 to type VI collagen was shown to be concentration-dependent and saturable and to depend mainly on the NG2 core protein, since chondroitinase-treated NG2 bound the collagen as well as undigested samples. In addition, the binding studies revealed several other possible ligands for NG2, including type II collagen, type V collagen, tenascin, and laminin. Binding of the proteoglycan to these molecules was also shown to be mediated by domains contained within the NG2 core protein. The ability of NG2 to bind to these extracellular matrix molecules was compared with that of the chondroitin sulfate proteoglycan decorin, revealing an almost identical binding pattern of the two proteoglycans to the different collagen types. In addition, decorin was found to effectively inhibit the ability of NG2 to bind to collagen, thus suggesting that the two proteoglycans may bind to some of the same regions on the collagen substrates. In contrast, decorin did not bind tenascin and was ineffective in inhibiting the binding of NG2 to tenascin or laminin, indicating that NG2 may bind these two molecules using a separate domain that is distinct from its collagen binding region.  相似文献   

11.
With the use of ATP analogues, we have found that porcine liver annexin (Anx) IV can be covalently labelled with 8-azido[gamma-32P]-ATP in the presence of Ca2+ (Kd 4.2 microM) and that the labelling is prevented by asolectin/cholesterol liposomes or chelation of calcium ions. On the other hand, non-covalent binding of 2'-(or 3')-O-(2,4,6-trinitrophenyl)adenosine 5'-triphosphate (TNP-ATP) to AnxIV occurs optimally in the presence of liposomes and Ca2+ (Kd 7 microM). These observations were further confirmed by the results of intrinsic fluorescence quenching of AnxIV with various nucleotides, suggesting the existence of a relationship between Ca(2+)-, phospholipid- and ATP-binding sites within the annexin molecule. The interaction of AnxIV with nucleotides does not significantly affect its in vitro properties concerning the binding to phosphatidylserine (PS) monolayers.  相似文献   

12.
Many parallel processes occur during the final stages of apoptosis. It is not clear which of these processes occur in all or most models of apoptosis and which occur only in some. In addition, the temporal relationship of these events is not always well understood. Correlated flow cytometric measurements were used to address these questions. Several models of apoptosis were studied, including thymocytes treated with dexamethasone. MOLT-4 cells treated with etoposide, U937 cells treated with anti-Fas, HL-60 cells treated with camptothecin, Raji cells grown in low serum, and aged neutrophils. All models showed a decrease in LDS-751 and fluorescein diacetate (FDA) staining, an increase in staining with dihydrorhodamine 123 (dhR123) or dihydroethidium, and an acidification of the cytoplasm. In each model, these changes were highly correlated, appearing simultaneously as multiparameter measurements. Changes in membrane status detected with merocyanin 540 (MC540) and annexin V behaved differently. A population with LDS-751 and FDA changes but without annexin V or MC540 changes could be demonstrated in some models. Several models did not show any change in annexin V binding, and HL-60 did not show a change in MC540 binding during apoptosis. The loss of cell surface antigens (CD45 and CD16) from aged neutrophils occurred in the entire LDS-751 and FDA dim population, even though other membrane changes (including the appearance of annexin V binding sites) were only apparent in a subset of these cells. These results suggest a model for the ordering of some of the terminal processes in apoptosis, with annexin V and MC540 changes trailing other events in apoptosis. These results confirm the need for caution in using a single-parameter measurement as an indicator of apoptosis for any new model being studied.  相似文献   

13.
Annexin II belongs to a family of calcium-dependent, phospholipid binding proteins. Annexin II was first identified as an intracellular protein and attributed intracellular functions. Although it lacks a signal peptide and its mechanism of secretion is unknown, extracellular annexin II has recently been found in several tissues as both soluble and membrane-bound protein. Cell-surface annexin II has been identified as a receptor for a number of polypeptide ligands. Extracellular annexin II may be important in several biological processes, such as fibrinolysis, cell adhesion, ligand-mediated cell signaling and virus infection. These processes provide several possibilities for therapeutic approaches targeting extracellular annexin II, and future research should further illuminate the biology of this molecule.  相似文献   

14.
Two site-directed mutants of Escherichia coli DNA helicase II (UvrD) were constructed to examine the functional significance of motif VI in a superfamily I helicase. Threonine 604 and arginine 605, representing two of the most highly conserved residues in motif VI, were replaced with alanine, generating the mutant alleles uvrD-T604A and uvrD-R605A. Genetic complementation studies indicated that UvrD-T604A, but not UvrD-R605A, functioned in methyl-directed mismatch repair and UvrABC-mediated nucleotide excision repair. Both mutant enzymes were purified and single-stranded DNA (ssDNA)-stimulated ATP hydrolysis, duplex DNA unwinding, and ssDNA binding were studied in the steady-state and compared to wild-type UvrD. UvrD-T604A exhibited a serious defect in ssDNA binding in the absence of nucleotide. However, in the presence of a non-hydrolyzable ATP analog, DNA binding was only slightly compromised. Limited proteolysis experiments suggested that UvrD-T604A had a "looser" conformation and could not undergo conformational changes normally associated with ATP binding/hydrolysis and DNA binding. UvrD-R605A, on the other hand, exhibited nearly normal DNA binding but had a severe defect in ATP hydrolysis (kcat=0.063 s-1 compared to 162 s-1 for UvrD). UvrD-T604A exhibited a much less severe decrease in ATPase activity (kcat=8.8 s-1). The Km for ATP for both mutants was not significantly changed. The results suggest that residues within motif VI of helicase II are essential for multiple biochemical properties associated with the enzyme and that motif VI is potentially involved in conformational changes related to the coupling of ATPase and DNA binding activities.  相似文献   

15.
It is commonly believed that MgATP2- is the substrate of F1-ATPases and ATP4- acts as a competitive inhibitor. However, the velocity equation for such competitive inhibition is equivalent to that for a rapid equilibrium ordered binding mechanism in which ATP4- adds first and the binding of Mg2+ is dependent on the formation of the E x ATP4- complex. According to this ordered-binding model, solution formed MgATP2- is not recognized by the ATPase as a direct substrate, and the high-affinity binding of Mg2+ to the E x ATP4- complex is the key reaction towards the formation of the ternary complex. These models (and others) were tested with an F1- ATPase, isolated from Halobacterium saccharovorum, by evaluating the rate of ATP hydrolysis as a function of free [ATP4-] or free [Mg2+]. The rates were asymmetrical with respect to increasing [ATP4-] versus increasing [Mg2+]. For the ordered-binding alternative, a series of apparent dissociation constants were obtained for ATP4-(K(A)aPP), which decreased as [Mg2+] increased. From this family of K(A)aPP the true K(A) was retrieved by extrapolation to [Mg2+] = 0 and was found to be 0.2 mM. The dissociation constants for Mg2+, established from these experiments, were also apparent (K(B)aPP) and dependent on [ATP4-] as well as on the pH. The actual K(B) was established from a series of K(B)aPP by extrapolating to [ATP4-] = infinity and to the absence of competing protons, and was found to be 0.0041 mM. The pKa of the protonable group for Mg2+ binding is 8.2. For the competitive inhibition alternative, rearrangement of the constants and fitting to the velocity equation gave an actual binding constant for MgATP2- (K(EAB)) of 0.0016 mM and for ATP4- (K(EA)) of 0.2 mM. Decision between the two models has far-reaching mechanistic implications. In the competitive inhibition model MgATP2- binds with high affinity, but Mg2+ cannot bind once the E x ATP4- complex is formed, while in the ordered-binding model binding of Mg2+ requires that ATP4- adds first. The steric constraints evident in the diffraction structure of the ATP binding site in the bovine mitochondrial F-ATPase [Abrahams, J. P., Leslie, A. G. W., Lutter, R. & Walker, J. E. (1994) Nature 370, 621-628] tend to favor the ordered-binding model, but the final decision as to which kinetic model is valid has to be from further structural studies. If the ordered-binding model gains more experimental support, a revision of the current concepts of unisite catalysis and negative cooperativity of nucleotide binding will be necessary.  相似文献   

16.
By means of selective extraction in a Ca(2+)-chelating medium and immunoblotting, four annexins (I, II, V, and VI) were identified in both isolated rat renal glomeruli and rat glomerular mesangial cells. Upon 32P labeling of these cells in culture, annexin I was immunoprecipitated using a specific polyclonal antibody and was found to incorporate radioactivity in a constitutive manner. However, as with epidermal growth factor (200 ng/ml), addition of angiotensin II (10(-7) M), arginine-vasopressin (10(-7) M), or endothelin I (10(-7) M) resulted in a 2-3-fold stimulation of annexin I phosphorylation. The basal phosphorylation as well as the stimulating effect of angiotensin II were also detected by immunoblotting annexin extracts using an antiphosphotyrosine antibody. In addition, among various phosphotyrosyl proteins isolated from EGTA extracts by adsorption onto an anti-phosphotyrosine antibody, annexin I was specifically recognized by Western blotting using a monoclonal anti-annexin I antibody, and displayed the same increase upon cell stimulation with angiotensin II. Moreover, thin layer chromatographic analysis of phosphoamino acids present in immunoprecipitated [32P]annexin I showed an exclusive labeling of phosphotyrosine residue(s). Finally, the effect of angiotensin II was detectable after 10 min, maximal at 6 h, and present until 12 h of incubation. Using 12-h stimulation, tyrosine phosphorylation of annexin I displayed a maximum at 10(-7) to 10(-6) M angiotensin II. These data report for the first time the stimulation of annexin I tyrosine phosphorylation by biologically active peptides acting via receptors belonging to the superfamily of seven hydrophobic domain, G-protein-linked receptors, which lack an intrinsic protein tyrosine kinase. This suggests a possible role of annexin I in the mitogenic effect of angiotensin II, arginine-vasopressin, and endothelin I, which was previously observed on rat glomerular mesangial cells as well as on other cells.  相似文献   

17.
Previously we reported that annexin VI is required for the budding of clathrin-coated pits from human fibroblast plasma membranes in vitro. Here we show that annexin VI bound to the NH2-terminal 28-kD portion of membrane spectrin is as effective as cytosolic annexin VI in supporting coated pit budding. Annexin VI-dependent budding is accompanied by the loss of approximately 50% of the spectrin from the membrane and is blocked by the cysteine protease inhibitor N-acetyl-leucyl-leucyl-norleucinal (ALLN). Incubation of fibroblasts in the presence of ALLN initially blocks the uptake of low density lipoprotein (LDL), but the cells recover after 1 h and internalize LDL with normal kinetics. The LDL internalized under these conditions, however, fails to migrate to the center of the cell and is not degraded. ALLN-treated cells have twice as many coated pits and twofold more membrane clathrin, suggesting that new coated pits have assembled. Annexin VI is not required for the budding of these new coated pits and ALLN does not inhibit. Finally, microinjection of a truncated annexin VI that inhibits budding in vitro has the same effect on LDL internalization as ALLN. These findings suggest that fibroblasts are able to make at least two types of coated pits, one of which requires the annexin VI-dependent activation of a cysteine protease to disconnect the clathrin lattice from the spectrin membrane cytoskeleton during the final stages of budding.  相似文献   

18.
Matrix vesicles (MVs), structures that accumulate Ca2+ during the initiation of mineral formation in growing bone, are rich in annexin V. When MVs are fused with planar phospholipid bilayers, a multiconductance Ca2+ channel is formed, with activity essentially identical to that observed when annexin V is delivered to the bilayer with phosphatidylserine liposomes. Ca2+ currents through this channel, from either MV or annexin V liposomes, are blocked by Zn2+, as is Ca2+ uptake by MV incubated in synthetic cartilage lymph. Blockage by Zn2+ was most effective when applied to the side containing the MV or liposomes. ATP and GTP differentially modulated the activity of this channel: ATP increased the amplitude of the current and the number of conductance states; GTP dramatically reduced the number of events and conductance states, leading to well-defined Ca2+ channel activity from either MV or the annexin V liposomes. In the distinctive effects of ATP, GTP, and Zn2+ on the Ca2+ channel activity observed in both the MV and the liposome systems, the common factor was the presence of annexin V. From this we conclude that Ca2+ entry into MV results from the presence of annexin V in these membrane-enclosed structures.  相似文献   

19.
We have previously shown that annexin I, a member of a family of calcium-dependent phospholipid and membrane binding proteins, interacts with profilin with high specificity and affinity. This finding further suggests that annexin I is involved through profilin in the regulation of membrane-cytoskeleton organization. We have investigated the consequences of a complex formed by these two proteins on the functions of both profilin and annexin I. Annexin I is able to modify the inhibitory effect of profilin on actin polymerization. This action is partial and the mechanism involved appears to be complex. On the other hand, the association between annexin I and profilin is sufficiently strong to inhibit the self-association of annexin I. The binding capacity of annexin I to liposomes containing phosphatidylserine, which mimics annexin I binding to membranes, is also decreased by profilin. This binding is nevertheless restored when phosphatidylinositol 4,5-biphosphate (PtdInsP2) is included in the liposomes. Finally, the capacity of annexin I to aggregate liposomes is also modified. It is worthwhile mentioning that the liposomes-binding and liposomes-aggregating activities of annexin I are independently regulated. The cell localization and functions of annexin I and profilin suggest that interaction between these two proteins may be directly implicated in the regulation of membrane-cytoskeleton. The phospholipid composition of membranes may be one of the modulating factors.  相似文献   

20.
Annexin VI has been isolated from rat liver endosomes and affinity purified antibodies have been produced. By Western blotting, in rat liver subcellular fractions, anti-annexin VI was demonstrated to recognise a 68 kDa band in the three endosomal fractions. In the present study, immunogold labeling of ultrathin Lowicryl sections of rat liver has been used to get insights into the ultrastructural hepatocyte localization. Although at the immunofluorescence level the staining seemed located at the apical, canalicular plasma membrane, domain of the hepatocytes, the electron microscopy revealed that 80% of the labeling, with the anti-annexin VI antibody was specifically localized not at the plasma membrane but in the close subapical endocytic compartment surrounding the bile canalicular plasma membrane of the hepatocyte. Double immunogold labeling with an anti peptide antibody to Rab5 and anti-annexin VI showed that 80% of the Rab5 positive apical endosomes were also labeled with anti-annexin VI antibodies. However, there was no significant colocalization of annexin VI and structures labeled with antibodies to the polymeric immunoglobulin receptor. The results suggest that annexin VI could be involved in regulating the functioning of this apical compartment in the hepatocyte.  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