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.
《The Journal of cell biology》1996,132(6):1053-1060
Acanthamoeba myosin-IA and myosin-IB are single-headed molecular motors that may play an important role in membrane-based motility. To better define the types of motility that myosin-IA and myosin IB can support, we determined the rate constants for key steps on the myosin-I ATPase pathway using fluorescence stopped-flow, cold-chase, and rapid-quench techniques. We determined the rate constants for ATP binding, ATP hydrolysis, actomyosin-I dissociation, phosphate release, and ADP release. We also determined equilibrium constants for myosin-I binding to actin filaments, ADP binding to actomyosin-I, and ATP hydrolysis. These rate constants define an ATPase mechanism in which (a) ATP rapidly dissociates actomyosin-I, (b) the predominant steady-state intermediates are in a rapid equilibrium between actin-bound and free states, (c) phosphate release is rate limiting and regulated by heavy- chain phosphorylation, and (d) ADP release is fast. Thus, during steady- state ATP hydrolysis, myosin-I is weakly bound to the actin filament like skeletal muscle myosin-II and unlike the microtubule-based motor kinesin. Therefore, for myosin-I to support processive motility or cortical contraction, multiple myosin-I molecules must be specifically localized to a small region on a membrane or in the actin-rich cell cortex. This conclusion has important implications for the regulation of myosin-I via localization through the unique myosin-I tails. This is the first complete transient kinetic characterization of a member of the myosin superfamily, other than myosin-II, providing the opportunity to obtain insights about the evolution of all myosin isoforms.  相似文献   

2.
In the intestinal brush border, the mechanoenzyme myosin-I links the microvillus core actin filaments with the plasma membrane. Previous immunolocalization shows that myosin-I is associated with vesicles in mature enterocytes (Drenckhahn, D., and R. Dermietzel. 1988. J. Cell Biol. 107:1037-1048) suggesting a potential role mediating vesicle motility. We now report that myosin-I is associated with Golgi-derived vesicles isolated from cells that are rapidly assembling brush borders in intestinal crypts. Crypt cells were isolated in hyperosmotic buffer, homogenized, and fractionated using differential- and equilibrium- density centrifugation. Fractions containing 50-100-nm vesicles, a similar size to those observed in situ, were identified by EM and were shown to contain myosin-I as demonstrated by immunoblotting and immunolabel negative staining. Galactosyltransferase, a marker enzyme for trans-Golgi membranes was present in these fractions, as was alkaline phosphatase, which is an apical membrane targeted enzyme. Galactosyltransferase was also present in vesicles immuno-purified with antibodies to myosin-I. Villin, a marker for potential contamination from fragmented microvilli, was absent. Myosin-I was found to reside on the vesicle "outer" or cytoplasmic surface for it was accessible to exogenous proteases and intact vesicles could be immunolabeled with myosin-I antibodies in solution. The bound myosin-I could be extracted from the vesicles using NaCl, KI and Na2CO3, suggesting that it is a vesicle peripheral membrane protein. These vesicles were shown to bundle actin filaments in an ATP-dependent manner. These results are consistent with a role for myosin-I as an apically targeted motor for vesicle translocation in epithelial cells.  相似文献   

3.
Characterization of a second myosin from Acanthamoeba castellanii.   总被引:21,自引:0,他引:21  
We purified a 400,000 molecular weight myosin, myosin-II, from Acanthamoeba castellanii. The sequence of ion exchange chromatography, actomyosin precipitation, actin extraction, and gel permeation chromatography yields per 100 g of cells about 11 mg of myosin-II which is 90 to 96% pure. ATPase activity is highest in the presence of Ca2+, but the enzyme is also active in EDTA provided high concentrations of K+ are present. The molecule consists of two 175,000 molecular weight heavy chains, one or two 17,500 molecular weight light chains, and two 16,500 molecular weight light chains. Myosin-II is rich in acidic residues and contains about 32 residues of cysteine/mol. The sedimentation coefficient is 5.9 S. Intrinsic viscosity is 126 cc/g. By equilibrium ultracentrifugation, the molecular weight averages depended upon the initial loading concentration in a way that suggested a 400,000 molecular weight species is in equilibrium with a 200,000 molecular weight species. By electron microscopy the molecule was seen to have two globular heads at one end of a tail 90 nm long. In KCl solutions of less than 0.25 M, the myosin-II tails self-associate to form the backbone of very small (6.6 x 205 nm) bipolar filaments with central bare zones 97 nm long. Myosin-II binds to actin filaments, forming periodic arrowhead-shaped complexes, but its Mg2+ ATPase activity is activated only 50% or less by actin. When radioactive myosin-II is incubated up to 90 min in unlabeled Acanthamoeba homogenates, it is not degraded into smaller fragments, such as the 190,000 molecular weight myosin-I. Our observations and the detailed enzymatic data presented by Maruta and Korn ((1977) J. Biol. Chem. 252, 6501-6509) argue that the smaller Acanthamoeba myosin-I (Pollard, T. D., and Korn, E. D. (1973) J. Biol. Chem, 248, 4682-2690) does not arise by fragmentation of myosin-II in the homogenate or extract.  相似文献   

4.
Synapsin I, a major neuron-specific phosphoprotein, is localized on the cytoplasmic surface of small synaptic vesicles to which it binds with high affinity. It contains a collagenase-resistant head domain and a collagenase-sensitive elongated tail domain. In the present study, the interaction between synapsin I and phospholipid vesicles has been characterized, and the protein domains involved in these interactions have been identified. When lipid vesicles were prepared from cholesterol and phospholipids using a lipid composition similar to that found in native synaptic vesicle membranes (40% phosphatidylcholine, 32% phosphatidylethanolamine, 12% phosphatidylserine, 5% phosphatidylinositol, 10% cholesterol, wt/wt), synapsin I bound with a dissociation constant of 14 nM and a maximal binding capacity of about 160 fmol of synapsin I/microgram of phospholipid. Increasing the ionic strength decreased the affinity without greatly affecting the maximal amount of synapsin I bound. When vesicles containing cholesterol and either phosphatidylcholine or phosphatidylcholine/phosphatidylethanolamine were tested, no significant binding was detected under any conditions examined. On the other hand, phosphatidylcholine vesicles containing either phosphatidylserine or phosphatidylinositol strongly interacted with synapsin I. The amount of synapsin I maximally bound was directly proportional to the percentage of acidic phospholipids present in the lipid bilayer, whereas the Kd value was not affected by varying the phospholipid composition. A study of synapsin I fragments obtained by cysteine-specific cleavage showed that the collagenase-resistant head domain actively bound to phospholipid vesicles; in contrast, the collagenase-sensitive tail domain, though strongly basic, did not significantly interact. Photolabeling of synapsin I was performed with the phosphatidylcholine analogue 1-palmitoyl-2-[11-[4-[3-(trifluoromethyl)diazirinyl]phenyl] [2-3H]undecanoyl]-sn-glycero-3-phosphocholine; this compound generates a highly reactive carbene that selectively interacts with membrane-embedded domains of membrane proteins. Synapsin I was significantly labeled upon photolysis when incubated with lipid vesicles containing acidic phospholipids and trace amounts of the photoactivatable phospholipid. Proteolytic cleavage of photolabeled synapsin I localized the label to the head domain of the molecule.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

5.
Microtubules (MT) are required for the efficient transport of membranes from the trans-Golgi and for transcytosis of vesicles from the basolateral membrane to the apical cytoplasm in polarized epithelia. MTs in these cells are primarily oriented with their plus ends basally near the Golgi and their minus-ends in the apical cytoplasm. Here we report that isolated Golgi and Golgi-enriched membranes from intestinal epithelial cells possess the actin based motor myosin-I, the MT minus- end-directed motor cytoplasmic dynein and its in vitro motility activator dynactin (p150/Glued). The Golgi can be separated into stacks, possessing features of the Golgi cisternae, and small membranes enriched in the trans-Golgi network marker TGN 38/41. Whereas myosin-I is present on all membranes in the Golgi fraction, dynein is present only on the small membrane fraction. Dynein, like myosin-I, is associated with membranes as a cytoplasmic peripheral membrane protein. Dynein and myosin-I coassociate with membranes that bind to MTs and cross-link actin filaments and MTs in a nucleotide-dependent manner. We propose that cytoplasmic dynein moves Golgi membranes along MTs to the cell cortex where myosin-I provides local delivery through the actin- rich cytoskeleton to the apical membrane.  相似文献   

6.
The Acanthamoeba myosin-IA heavy chain gene encodes a 134-kDa protein with a catalytic domain, three potential light chain binding sites, and a tail with separately folded tail homology (TH) -1, -2, and -3 domains. TH-1 is highly resistant to trypsin digestion despite consisting of 15% lysine and arginine. TH-2/3 is resistant to alpha-chymotrypsin digestion. The peptide link between TH-1 and TH-2/3 is cleaved by trypsin, alpha-chymotrypsin, and endo-AspN but not V8 protease. The CD spectra of TH-2/3 indicate predominantly random structure, turns, and beta-strands but no alpha-helix. The hydrodynamic properties of TH-2/3 (Stokes' radius of 3.0 nm, sedimentation coefficient of 1.8 S, and molecular mass of 21.6 kDa) indicate that these domains are as long as the whole myosin-I tail in reconstructions of electron micrographs. Furthermore, separately expressed and purified TH-1 binds with high affinity to TH-2/3. Thus we propose that TH-1 and TH-2/3 are arranged side by side in the myosin-IA tail. Separate TH-1, TH-2, and TH-2/3 each binds muscle actin filaments with high affinity. Salt inhibits TH-2/3 binding to muscle actin but not amoeba actin filaments. TH-1 enhances binding of TH-2/3 to muscle actin filaments at physiological salt concentration, indicating that TH-1 and TH-2/3 cooperate in actin binding. An intrinsic fluorescence assay shows that TH-2/3 also binds with high affinity to the protein Acan125 similar to the SH3 domain of myosin-IC. Phylogenetic analysis of SH3 sequences suggests that myosin-I acquired SH3 domain after the divergence of the genes for myosin-I isoforms.  相似文献   

7.
Myosin-I is the single-headed member of the myosin superfamily that associates with lipid membranes. Biochemical experiments have shown that myosin-I membrane binding is the result of electrostatic interactions between the basic tail domain and acidic phospholipids. To better understand the dynamics of myosin-I membrane association, we measured the rates of association and dissociation of a recombinant myo1c tail domain (which includes three IQ domains and bound calmodulins) to and from large unilamellar vesicles using fluorescence resonance energy transfer. The apparent second-order rate constant for lipid-tail association in the absence of calcium is fast with nearly every lipid-tail collision resulting in binding. The rate of binding is decreased in the presence of calcium. Time courses of myo1c-tail dissociation are best fit by two exponential rates: a fast component that has a rate that depends on the ratio of acidic phospholipid to myo1c-tail (phosphatidylserine (PS)/tail) and a slow component that predominates at high PS/tail ratios. The dissociation rate of the slow component is slower than the myo1c ATPase rate, suggesting that myo1c is able to stay associated with the lipid membrane during multiple catalytic cycles of the motor. Calcium significantly increases the lifetimes of the membrane-bound state, resulting in dissociation rates 0.001 s(-1).  相似文献   

8.
Small unilamellar phosphatidylserine/phosphatidylcholine liposomes incubated on one side of planar phosphatidylserine bilayer membranes induced fluctuations and a sharp increase in the membrane conductance when the Ca2+ concentration was increased to a threshold of 3--5 mM in 100 mM NaCl, pH 7.4. Under the same ionic conditions, these liposomes fused with large (0.2 micrometer diameter) single-bilayer phosphatidylserine vesicles, as shown by a fluorescence assay for the mixing of internal aqueous contents of the two vesicle populations. The conductance behavior of the planar membranes was interpreted to be a consequence of the structural rearrangement of phospholipids during individual fusion events and the incorporation of domains of phosphatidylcholine into the Ca2+-complexed phosphatidylserine membrane. The small vesicles did not aggregate or fuse with one another at these Ca2+ concentrations, but fused preferentially with the phosphatidylserine membrane, analogous to simple exocytosis in biological membranes. Phosphatidylserine vesicles containing gramicidin A as a probe interacted with the planar membranes upon raising the Ca2+ concentration from 0.9 to 1.2 mM, as detected by an abrupt increase in the membrane conductance. In parallel experiments, these vesicles were shown to fuse with the large phosphatidylserine liposomes at the same Ca2+ concentration.  相似文献   

9.
Myo1e is the widely expressed subclass-1 member of the myosin-I family. We performed a kinetic analysis of a truncated myo1e that consists of the motor and the single IQ motif with a bound calmodulin. We determined the rates and equilibrium constants for the key steps in the ATPase cycle. The maximum actin activated ATPase rate (V(max)) and the actin concentration at half-maximum of V(max) (K(ATPase)) of myo1e are similar to those of the native protein. The K(ATPase) is low (approximately 1 microm), however the affinity of myo1e for actin in the presence of ATP is very weak. A weak actin affinity and a rapid rate of phosphate release result in a pathway under in vitro assay conditions in which phosphate is released while myo1e is dissociated from actin. Actin activation of the ATPase activity and the low K(ATPase) are the result of actin activation of ADP release. We propose that myo1e is tuned to function in regions of high concentrations of cross-linked actin filaments. Additionally, we found that ADP release from actomyo1e is > 10-fold faster than other vertebrate myosin-I isoforms. We propose that subclass-1 myosin-Is are tuned for rapid sliding, whereas subclass-2 isoforms are tuned for tension maintenance or stress sensing.  相似文献   

10.
Small unilamellar phosphatidylserine/phosphatidylcholine liposomes incubated on one side of planar phosphatidylserine bilayer membranes induced fluctuations and a sharp increase in the membrane conductance when the Ca2+ concentration was increased to a threshold of 3–5 mM in 100 mM NaCl, pH 7.4. Under the same ionic conditions, these liposomes fused with large (0.2 μm diameter) single-bilayer phosphatidylserine vesicles, as shown by a fluorescence assay for the mixing of internal aqueous contents of the two vesicle populations. The conductance behavior of the planar membranes was interpreted to be a consequence of the structural rearrangement of phospholipids during individual fusion events and the incorporation of domains of phosphatidylcholine into the Ca2+-complexed phosphatidylserine membrane. The small vesicles did not aggregate or fuse with one another at these Ca2+ concentrations, but fused preferentially with the phosphatidylserine membrane, analogous to simple exocytosis in biological membranes. Phosphatidylserine vesicles containing gramicidin A as a probe interacted with the planar membranes upon raising the Ca2+ concentration from 0.9 to 1.2 mM, as detected by an abrupt increase in the membrane conductance. In parallel experiments, these vesicles were shown to fuse with the large phosphatidylserine liposomes at the same Ca2+ concentration.  相似文献   

11.
To evaluate the role of the hydration layer on the protein surface of actomyosin, we compared the effects of urea and guanidine-HCl on the sliding velocities and ATPase activities of the actin-heavy meromyosin (HMM) system. Both chemicals denature proteins, but only urea perturbs the hydration layer. Both the sliding velocity of actin filaments and actin-activated ATPase activity decreased with increasing urea concentrations. The sliding movement was completely inhibited at 1.0 M urea, while actin filaments were bound to HMM molecules fixed on the glass surface. Guanidine-HCl (0-0.05 M) drastically decreased both the sliding velocity and ATPase activation of acto-HMM complexes. Under this condition, actin filaments almost detached from HMM molecules. In contrast, the ATPase activity of HMM without actin filaments was almost independent of urea concentrations <1.0 M and guanidine-HCl concentrations <0.05 M. An increase in urea concentrations up to 2.0 M partly induced changes in the ternary structure of HMM molecules, while the actin filaments were stable in this concentration range. Hydration changes around such actomyosin complexes may alter both the stability of part of the myosin molecules, and the affinity for force transmission between actin filaments and myosin heads.  相似文献   

12.
Type I myosins are highly conserved actin-based molecular motors that localize to the actin-rich cortex and participate in motility functions such as endocytosis, polarized morphogenesis, and cell migration. The COOH-terminal tail of yeast myosin-I proteins, Myo3p and Myo5p, contains an Src homology domain 3 (SH3) followed by an acidic domain. The myosin-I SH3 domain interacted with both Bee1p and Vrp1p, yeast homologues of human WASP and WIP, adapter proteins that link actin assembly and signaling molecules. The myosin-I acidic domain interacted with Arp2/3 complex subunits, Arc40p and Arc19p, and showed both sequence similarity and genetic redundancy with the COOH-terminal acidic domain of Bee1p (Las17p), which controls Arp2/3-mediated actin nucleation. These findings suggest that myosin-I proteins may participate in a diverse set of motility functions through a role in actin assembly.  相似文献   

13.
(Ca2+ + Mg2+)ATPase (EC 3.6.1.3) was solubilized from human erythrocyte membranes by detergent extraction with Triton N-101 (0.5 mg/mg membrane protein) and purified by calmodulin affinity chromatography. ATPase activity was assayed in mixtures of Triton N-101 and phospholipid, without reconstitution into bilayer vesicles. At low levels of phospholipid (5 micrograms/ml), the ATPase activity was highly sensitive to the detergent concentration, with maximal activity occurring at or near the critical micelle concentration of the detergent. With increased amounts of phospholipid (50 micrograms/ml), detergent concentrations greater than the critical micelle concentration were required for maximal activity. Detergent alone did not support ATPase activity. Sonicated phospholipid in the form of vesicles was equally ineffective. Activity seemed to be dependent on the presence of detergent/phospholipid mixed micelles. The acidic phospholipids, phosphatidylserine and phosphatidylinositol, as well as the commercial phospholipid preparation, Asolectin, gave activities five to eight times greater than the same amount of phosphatidylcholine. Mixtures of phosphatidylserine and phosphatidylcholine produced intermediate ATPase activities, with the maximal value dependent on the phosphatidylserine concentration. Addition of phosphatidylcholine to fixed concentrations of phosphatidylserine caused a rise in activity that was independent of the ratio of the two phospholipids or the total phospholipid concentration. Phosphatidylcholine may therefore be irreplaceable for some aspect of ATPase function. The number of phospholipid molecules present in mixed micelles at maximal ATPase activity was calculated to be near 50. This value implied that the hydrophobic surface of the ATPase molecule must be completely coated by a single layer of phospholipid molecules for maximum activity to occur.  相似文献   

14.
The effects of salts and non-ionic detergents on renal brush borders have been studied. 2 M sodium chloride, iodide or thiocyanate dissociated up to 40% of the protein from the brush borders, destroying the core filaments and resulting in the formation of membrane vesicles; EDTA had a similar effect on structure but released little protein. Triton X-100 and Nonidet P-40 extracted up to 60% of the protein including the major membrane glycoproteins and the enzymes trehalase, maltase and aminopeptidase (microsomal). Triton exhibited a selective effect on lipids removing phosphatidylserine, phosphatidylethanolamine and sphingomyelin but not the bulk of the phosphatidylcholine or cholesterol. The residual structures after Triton extraction comprised the core filaments associated with vesicles of lipid containing alkaline phosphatase and several other proteins. Treatment of these core-vesicle complexes with 2 M sodium chloride dissociated the filaments, releasing the vesicles which could be recovered as a pellicle on centrifugation. It is suggested that the proteins found in the vesicles might serve to interconnect the core filaments with the lipid bilayer.  相似文献   

15.
Summary The interaction of furosemide with different phospholipids was investigated. Its influence on the lipid structure was inferred from its effect on the phase transition properties of lipids and on the conductance of planar bilayer membranes. The thermotropic properties of dipalmitoyl phosphatidylcholine, phosphatidylethanolamine (natural), dipalmitoyl phosphatidylethanolamine, brain sphingomyelin, brain cerebrosides and phosphatidylserine in the presence and absence of furosemide were investigated by differential scanning calorimetry,. The modifying effect of furosemide seems to be strongest on phosphatidylethanolamine (natural) and sphingomyelin bilayers. The propensity of furosemide to decrease the electrical resistance of planar lipid membranes was also studied and it is shown that the drug facilitates the transport of ions. Partition coefficients of furosemide between lipid bilayers and water were measured.Abbreviations DSC differential scanning calorimetry - PLM planar lipid membranes - DPPC dipalmitoyl phosphatidylcholine - DMPC dimyristoyl phosphatidylcholine - PE phosphatidyl ethanol  相似文献   

16.
The binding, conformation and orientation of a hydrophilic vector peptide penetratin in lipid membranes and its state of self-association in solution were examined using circular dichroism (CD), analytical ultracentrifugation and fluorescence spectroscopy. In aqueous solution, penetratin exhibited a low helicity and sedimented as a monomer in the concentration range approximately 50-500 microM. The partitioning of penetratin into phospholipid vesicles was determined using tryptophan fluorescence anisotropy titrations. The apparent penetratin affinity for 20% phosphatidylserine/80% egg phosphatidylcholine vesicles was inversely related to the total peptide concentration implying repulsive peptide-peptide interactions on the lipid surface. The circular dichroism spectra of the peptide when bound to unaligned 20% phosphatidylserine/80% egg phosphatidylcholine vesicles and aligned hydrated phospholipid multilayers were attributed to the presence of both alpha-helical and beta-turn structures. The orientation of the secondary structural elements was determined using oriented circular dichroism spectroscopy. From the known circular dichroism tensor components of the alpha-helix, it can be concluded that the orientation of the helical structures is predominantly perpendicular to the membrane surface, while that of the beta-type carbonyls is parallel to the membrane surface. On the basis of our observations, we propose a novel model for penetratin translocation.  相似文献   

17.
A unique feature of protein networks in living cells is that they can generate their own force. Proteins such as non-muscle myosin II are an integral part of the cytoskeleton and have the capacity to convert the energy of ATP hydrolysis into directional movement. Non-muscle myosin II can move actin filaments against each other, and depending on the orientation of the filaments and the way in which they are linked together, it can produce contraction, bending, extension, and stiffening. Our measurements with differential scanning calorimetry showed that non-muscle myosin II inserts into negatively charged phospholipid membranes. Using lipid vesicles made of DMPG/DMPC at a molar ratio of 1:1 at 10 mg/ml in the presence of different non-muscle myosin II concentrations showed a variation of the main phase transition of the lipid vesicle at around 23 °C. With increasing concentrations of non-muscle myosin II the thermotropic properties of the lipid vesicle changed, which is indicative of protein-lipid interaction/insertion. We hypothesize that myosin tail binds to acidic phospholipids through an electrostatic interaction using the basic side groups of positive residues; the flexible, amphipathic helix then may partially penetrate into the bilayer to form an anchor. Using the stopped-flow method, we determined the binding affinity of non-muscle myosin II when anchored to lipid vesicles with actin, which was similar to a pure actin-non-muscle myosin II system. Insertion of myosin tail into the hydrophobic region of lipid membranes, a model known as the lever arm mechanism, might explain how its interaction with actin generates cellular movement.  相似文献   

18.
Ponticulin, an F-actin binding transmembrane glycoprotein in Dictyostelium plasma membranes, was isolated by detergent extraction from cytoskeletons and purified to homogeneity. Ponticulin is an abundant membrane protein, averaging approximately 10(6) copies/cell, with an estimated surface density of approximately 300 per microns2. Ponticulin solubilized in octylglucoside exhibited hydrodynamic properties consistent with a ponticulin monomer in a spherical or slightly ellipsoidal detergent micelle with a total molecular mass of 56 +/- 6 kD. Purified ponticulin nucleated actin polymerization when reconstituted into Dictyostelium lipid vesicles, but not when a number of commercially available lipids and lipid mixtures were substituted for the endogenous lipid. The specific activity was consistent with that expected for a protein comprising 0.7 +/- 0.4%, by mass, of the plasma membrane protein. Ponticulin in octylglucoside micelles bound F- actin but did not nucleate actin assembly. Thus, ponticulin-mediated nucleation activity was sensitive to the lipid environment, a result frequently observed with transmembrane proteins. At most concentrations of Dictyostelium lipid, nucleation activity increased linearly with increasing amounts of ponticulin, suggesting that the nucleating species is a ponticulin monomer. Consistent with previous observations of lateral interactions between actin filaments and Dictyostelium plasma membranes, both ends of ponticulin-nucleated actin filaments appeared to be free for monomer assembly and disassembly. Our results indicate that ponticulin is a major membrane protein in Dictyostelium and that, in the proper lipid matrix, it is sufficient for lateral nucleation of actin assembly. To date, ponticulin is the only integral membrane protein known to directly nucleate actin polymerization.  相似文献   

19.
Total internal reflection fluorescence microscopy (TIRFM) has been employed to investigate the Ca(2+)-dependent membrane-binding characteristics of fluorescein-labeled bovine prothrombin-fragment 1 (F-BF1). Light scattering measurements demonstrated that F-BF1 bound to small unilamellar phosphatidylserine/phosphatidylcholine (25/75, mol/mol) vesicles with an apparent dissociation constant (1.5 +/- 0.2 microM) similar to that of unlabeled protein (1.1 +/- 0.1 microM). Negatively charged supported planar membranes were constructed by fusing small unilamellar vesicles at quartz surfaces. TIRFM measurements under equilibrium conditions showed that F-BF1 bound to planar membranes with an apparent dissociation constant (0.9 +/- 0.2 microM) approximately equal to that on vesicles. Total internal reflection/fluorescence photobleaching recovery (TIR/FPR) curves for F-BF1 on 25 mol% PS planar surfaces were diffusion-influenced at F-BF1 solution concentrations less than or equal to 5 microM. Fluorescence recovery rates from samples of high F-BF1 concentrations were slowed by increasing the solution viscosity with glycerol, thus providing further support for a diffusion-limited effect at low F-BF1 concentrations. Analysis of the reaction-limited fluorescence recovery curves at F-BF1 solution concentrations greater than or equal to 10 microM gave average association and dissociation kinetic rates of approximately 10(5) M-1 s-1 and approximately 0.1 s-1, respectively. Kinetic association rates increased significantly with increasing PS, whereas kinetic dissociation rates increased only slightly. Fluorescence recovery curves were nonmonoexponential; possible mechanisms for this behavior are described.  相似文献   

20.
Lactadherin, a milk protein, contains discoidin-type lectin domains with homology to the phosphatidylserine-binding domains of blood coagulation factor VIII and factor V. We have found that lactadherin functions, in vitro, as a potent anticoagulant by competing with blood coagulation proteins for phospholipid binding sites [J. Shi and G.E. Gilbert, Lactadherin inhibits enzyme complexes of blood coagulation by competing for phospholipid binding sites, Blood 101 (2003) 2628-2636]. We wished to characterize the membrane-binding properties that correlate to the anticoagulant capacity. We labeled bovine lactadherin with fluorescein and evaluated binding to membranes of composition phosphatidylserine/phosphatidylethanolamine/phosphatidylcholine, 4:20:76 supported by 2 mum diameter glass microspheres. Lactadherin bound saturably with an apparent KD of 3.3+/-0.4 nM in a Ca++ -independent manner. The number of lactadherin binding sites increased proportionally to the phosphatidylserine content over a range 0-2% and less rapidly for higher phosphatidylserine content. Inclusion of phosphatidylethanolamine in phospholipid vesicles did not enhance the apparent affinity or number of lactadherin binding sites. The number of sites was at least 4-fold higher on small unilamellar vesicles than on large unilamellar vesicles, indicating that lactadherin binding is enhanced by membrane curvature. Lactadherin bound to membranes with synthetic dioleoyl phosphatidyl-L-serine but not dioleoyl phosphatidyl-D-serine indicating stereoselective recognition of phosphatidyl-L-serine. We conclude that lactadherin resembles factor VIII and V with stereoselective preference for phosphatidyl-L-serine and preference for highly curved membranes.  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