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 15 毫秒
1.
W Witke  M Schleicher  A A Noegel 《Cell》1992,68(1):53-62
We generated by gene disruption Dictyostelium cells that lacked both the F-actin cross-linking proteins, alpha-actinin and gelation factor. Several major cell functions, such as growth, chemotaxis, phagocytosis, and pinocytosis, were apparently unaltered. However, in all double mutants, development was greatly impaired. After formation of aggregates, cells were very rarely able to form fruiting bodies. This ability was rescued when mutant and wild-type strains were mixed in a ratio of 70 to 30. The developmental program in the mutant was not arrested, since the expression pattern of early and late genes remained unchanged. Development of the mutant was rendered normal when a functional alpha-actinin gene was introduced and expressed, showing the morphogenetic defect to be due to the absence of the two F-actin cross-linking proteins. These findings suggest the existence of a functional network allowing mutual complementation of certain actin-binding proteins.  相似文献   

2.
A monoclonal antibody, mAb 47-19-2, was used to study the subunit topology of the rod-shaped alpha-actinin molecules of Dictyostelium discoideum and to screen for mutants defective in the production of alpha-actinin. Electron microscopy of rotary-shadowed alpha-actinin-antibody complexes showed binding of mAb 47-19-2 to both ends of the alpha-actinin rods and cleavage of the rods into its subunits, indicating that the two subunits of alpha-actinin extend in an anti-parallel mode through the whole length of the rod. The antibody binding sites were located in close proximity to the sites responsible for actin cross-linking, which is consistent with the blocking activity of the antibody. In a mutant, HG1130, no antibody label was detected in colony blots, and by immunoblotting of mutant proteins separated by SDS-PAGE, only trace amounts of alpha-actinin were found. The mutant showed normal binding of antibodies directed against the actin-binding proteins severin and capping protein. The mutation responsible for the alpha-actinin defect was recessive and located on linkage group I of the genetic map of D. discoideum. HG1130 cells grew on bacteria at a normal rate and also axenically like cells of the parent strain AX2. After starvation the mutant cells expressed the contact site A glycoprotein, a marker of the aggregation-competent stage, and reacted chemotactically to cyclic AMP. The aggregation patterns and fruiting bodies of the mutant appeared to be normal. Patching and capping on the surface of HG1130 cells was induced by antibodies against the contact site A glycoprotein.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Reflection interference contrast microscopy combined with digital image processing was applied to study the motion of Dictyostelium discoideum cells in their pre-aggregative state on substrata of different adhesiveness (glass, albumin-covered glass, and freshly cleaved mica). The temporal variations of the size and shape of the cell/substratum contact area and the time course of advancement of pseudopods protruding in contact with the substratum were analyzed. The major goal was to study differences between the locomotion of wild-type cells and strains of triple mutants deficient in two F-actin cross-linking proteins (alpha-actinin and the 120-kDa gelation factor) and one F-actin fragmenting protein (severin). The size of contact area, AC, of both wild-type and mutant cells fluctuates between minimum and maximum values on the order of minutes, pointing toward an intrinsic switching mechanism associated with the mechanochemical control system. The fluctuation amplitudes are much larger on freshly cleaved mica than on glass. Wild-type and mutant cells exhibit remarkable differences on mica but not on glass. These differences comprise the population median of AC and alterations in pseudopod protrusion. AC is smaller by a factor of two or more for all mutants. Pseudopods protrude slower and shorter in the mutants. It is concluded that cell shape and pseudopods are destabilized by defects in the actin-skeleton, which can be overcompensated by strongly adhesive substrata. Several features of amoeboid cell locomotion on substrata can be understood on the basis of the minimum bending energy concept of soft adhering shells and by assuming that adhesion induces local alterations of the composite membrane consisting of the protein/lipid bilayer on the cell surface and the underlying actin-cortex.  相似文献   

4.
The 120-kD gelation factor and alpha-actinin are among the most abundant F-actin cross-linking proteins in Dictyostelium discoideum. Both molecules are homodimers and have extended rod-like configurations that are respectively approximately 35 and 40 nm long. Here we report the complete cDNA sequence of the 120-kD gelation factor which codes for a protein of 857 amino acids. Its calculated molecular mass is 92.2 kD which is considerably smaller than suggested by its mobility in SDS-PAGE. Analysis of the sequence shows a region that is highly homologous to D. discoideum alpha-actinin, chicken fibroblast alpha-actinin, and human dystrophin. This conserved domain probably represents an actin binding site that is connected to the rod-forming part of the molecule via a highly charged stretch of amino acids. Whereas the sequence of alpha-actinin (Noegel, A., W. Witke, and M. Schleicher. 1987. FEBS [Fed. Eur. Biochem. Soc.] Lett. 221:391-396) suggests that the extended rod domain of the molecule is based on four spectrin-like repeats with high alpha-helix potential, the rod domain of the 120-kD gelation factor is constructed from six 100-residue repeats that have a high content of glycine and proline residues and which, in contrast to alpha-actinin, do not appear to have a high alpha-helical content. These repeats show a distinctive pattern of regions that have high beta-sheet potential alternating with short zones rich in residues with a high potential for turns. This observation suggests that each 100-residue motif has a cross-beta conformation with approximately nine sheets arranged perpendicular to the long axis of the molecule. In the high beta-potential zones every second residue is often hydrophobic. In a cross-beta structure, this pattern would result in one side of the domain having a surface rich in hydrophobic side chains which could account for the dimerization of the 120-kD gelation factor subunits.  相似文献   

5.
6.
R S Hock  G Davis  D W Speicher 《Biochemistry》1990,29(40):9441-9451
A method was developed to purify human smooth muscle filamin in high yield and structural domains were defined by using mild proteolysis to dissect the molecule into intermediate-sized peptides. Unique domains were defined and aligned by using high-resolution peptide mapping of iodinated peptides on cellulose plates. The amino- and carboxyl-terminal orientation of these domains within the molecule was determined by amino acid sequence analysis of several aligned peptides. In addition to the three unique domains which were identified, a number of smaller and larger fragments were also characterized and aligned within the intact molecule. These structural domains and related peptides provide a useful set of defined fragments for further elucidation of structure-function relationships. The two known functionally important binding sites of filamin, the self-association site and the actin-binding site, have been localized. Self-association of two monomers in a tail-to-tail orientation involves a small protease-sensitive region near the carboxyl terminal of the intact polypeptide chain. Sedimentation assays indicate that an actin-binding site is located near the blocked amino terminal of the filamin molecule. Sequences derived from large peptides mapping near the amino terminal show homology to the amino-terminal actin-binding site of alpha-actinin (chicken fibroblast and Dictyostelium), Dictyostelium 120-kDa actin gelation factor, beta-spectrin (human red cell and Drosophila), and human dystrophin. This homology is particularly interesting for two reasons. The functional form of filamin is single stranded, in contrast to alpha-actinin and spectrin which are antiparallel double-stranded actin cross-linkers. Also, no homology to the spectrin-like segments which comprise most of the mass of spectrin, alpha-actinin, and dystrophin was found. Instead, the sequence of a domain located near the center of the filamin molecule (tryptic 100-kDa peptide, T100) shows homology to the published internal repeats of the Dictyostelium 120-kDa actin gelation factor. On the basis of these results, a model of human smooth muscle filamin substructure is presented. Also, comparisons of human smooth muscle filamin, avian smooth muscle filamin, and human platelet filamin are reported.  相似文献   

7.
Actin-binding proteins are known to regulate in vitro the assembly of actin into supramolecular structures, but evidence for their activities in living nonmuscle cells is scarce. Amebae of Dictyostelium discoideum are nonmuscle cells in which mutants defective in several actin-binding proteins have been described. Here we characterize a mutant deficient in the 120-kD gelation factor, one of the most abundant F-actin cross-linking proteins of D. discoideum cells. No F-actin cross-linking activity attributable to the 120-kD protein was detected in mutant cell extracts, and antibodies recognizing different epitopes on the polypeptide showed the entire protein was lacking. Under the conditions used, elimination of the gelation factor did not substantially alter growth, shape, motility, or chemotactic orientation of the cells towards a cAMP source. Aggregates of the mutant developed into fruiting bodies consisting of normally differentiated spores and stalk cells. In cytoskeleton preparations a dense network of actin filaments as typical of the cell cortex, and bundles as they extend along the axis of filopods, were recognized. A significant alteration found was an enhanced accumulation of actin in cytoskeletons of the mutant when cells were stimulated with cyclic AMP. Our results indicate that control of cell shape and motility does not require the fine-tuned interactions of all proteins that have been identified as actin-binding proteins by in vitro assays.  相似文献   

8.
Mechanical interactions between desmin and Z-disks, costameres, and nuclei were measured during passive deformation of single muscle cells. Image processing and continuum kinematics were used to quantify the structural connectivity among these structures. Analysis of both wild-type and desmin-null fibers revealed that the costamere protein talin colocalized with the Z-disk protein alpha-actinin, even at very high strains and stresses. These data indicate that desmin is not essential for mechanical coupling of the costamere complex and the sarcomere lattice. Within the sarcomere lattice, significant differences in myofibrillar connectivity were revealed between passively deformed wild-type and desmin-null fibers. Connectivity in wild-type fibers was significantly greater compared to desmin-null fibers, demonstrating a significant functional connection between myofibrils that requires desmin. Passive mechanical analysis revealed that desmin may be partially responsible for regulating fiber volume, and consequently, fiber mechanical properties. Kinematic analysis of alpha-actinin strain fields revealed that knockout fibers transmitted less shear strain compared to wild-type fibers and experienced a slight increase in fiber volume. Finally, linkage of desmin intermediate filaments to muscle nuclei was strongly suggested based on extensive loss of nuclei positioning in the absence of desmin during passive fiber loading.  相似文献   

9.
Summary The Mod A mutation inDictyostelium discoideum results in a post-translational modification which reduces the activity and electrophoretic mobility of a group of lysosomal glycoproteins. To determine whether this mutation might affect protein bound oligosaccharides, metabolically labeled [2]3H-mannose glycopeptides were isolated from wild-type (AX3) and mutant cells (M31) ofDictyostelium discoideum. A group of large, negatively charged glycopeptides are significantly depleted in strain M31 compared to AX3. Cells of each strain double labeled with3H-mannose and35SO4 or32PO4 showed that the large, negatively charged glycopeptides of AX3 contain both sulfate and phosphate while those of M31 are depleted in these groups. The kinetics of35SO4 release from the glycopeptides of each strain suggested that both contained similar sulfated sugar(s), but that M31 glycopeptides contained three-fold less than those of AX3. Acid hydrolysis of32PO4 containing3H-mannose glycopeptides showed the presence of3H-mannose-6-32-P-phosphate in the AX3 hydrolysates while the glycopeptides of M31 contain only 15% as much mannose-6-phosphate as those of AX3.  相似文献   

10.
《Biorheology》1997,34(6):377-385
We report the results from a series of rheological tests of fresh bovine brain tissue. Using a standard Bohlin VOR shear rheometer, shear relaxation and oscillating strain sweep experiments were performed on disks of brain tissue 30 mm in diameter, with a thickness of 1.5–2 mm. The strain sweep experiment showed that the viscoelastic strain limit is of the order of 0.1% strain. Shear relaxation data do not indicate the presence of a long-term elastic modulus, indicating fluid-like behavior. A relaxation spectrum was calculated by inverting the experimental data and used to predict oscillatory response, which agreed well with measured data.  相似文献   

11.
We have employed an interferometric technique for the local measurement of bending modulus, membrane tension, and adhesion energy of motile cells adhering to a substrate. Wild-type and mutant cells of Dictyostelium discoideum were incubated in a flow chamber. The flow-induced deformation of a cell near its adhesion area was determined by quantitative reflection interference contrast microscopy (RICM) and analyzed in terms of the elastic boundary conditions: equilibrium of tensions and bending moments at the contact line. This technique was employed to quantify changes caused by the lack of talin, a protein that couples the actin network to the plasma membrane, or by the lack of cortexillin I or II, two isoforms of the actin-bundling protein cortexillin. Cells lacking either cortexillin I or II exhibited reduced bending moduli of 95 and 160 k(B)T, respectively, as compared to 390 k(B)T, obtained for wild-type cells. No significant difference was found for the adhesion energies of wild-type and cortexillin mutant cells. In cells lacking talin, not only a strongly reduced bending modulus of 70 k(B)T, but also a low adhesion energy one-fourth of that in wild-type cells was measured.  相似文献   

12.
Gelation factor (ABP120) is one of the principal actin-cross-linking proteins of Dictyostelium discoideum. The extended molecule has an N-terminal 250-residue actin-binding domain and a rod constructed from six 100-residue repeats that have an Ig fold. The ability to dimerize is crucial to the actin cross-linking function of gelation factor and is mediated by the rod in which the two chains are arranged in an antiparallel fashion. We report the 2.2 A resolution crystal structure of rod domains 5 and 6, which shows that dimerization is mediated primarily by rod domain 6 and is the result of a double edge-to-edge extension of beta-sheets. Thus, contrary to earlier proposals, the chains of the dimeric gelation factor molecule overlap only within domain 6, and domains 1-5 do not pair with domains from the other chain. This information allows construction of a model of the gelation factor molecule and suggests how the chains in the related molecule filamin (ABP280) may interact.  相似文献   

13.
利用盘基网柄菌表达可溶性人Fas配体   总被引:1,自引:0,他引:1  
用PCR扩增从激活的人中性粒细胞中得到的编码可溶性Fas配体胞外区中第141个到第281个氨基酸的cDNA ,将其与hCG-β信号肽片段融合到质粒MB12neo中,随后导入到盘基网柄菌AX3细胞中,得到分泌性表达hFasL的重组菌AX3-H3。为提高shFasL的表达量,对质粒pMB12neo作了改造,得到衍生质粒pMB74。利用质粒pMB74克隆表达shFasL ,得到高通量表达shFasL的重组菌AX3_pLu8。在复杂培养基HL_5C中,重组菌的细胞密度可达(1.5~2 )×107 mL ,AX3-H3及AX3_pLu8分泌的shFasL浓度分别为23.5 μg/L及206μg/L。利用合成培养基SIH培养重组菌AX3-H3及AX3-pLu8,细胞密度均达到(4~5)×107m/L ,shFasL浓度则分别达到111μg/L和420μg/L。  相似文献   

14.
Mechanical stresses applied to the plasma membrane of an adherent cell induces strain hardening of the cytoskeleton, i.e. the elasticity of the cytoskeleton increases with its deformation. Strain hardening is thought to mediate the transduction of mechanical signals across the plasma membrane through the cytoskeleton. Here, we describe the strain dependence of a model system consisting of actin filaments (F-actin), a major component of the cytoskeleton, and the F-actin cross-linking protein alpha-actinin, which localizes along contractile stress fibers and at focal adhesions. We show that the amplitude and rate of shear deformations regulate the resilience of F-actin networks. At low temperatures, for which the lifetime of binding of alpha-actinin to F-actin is long, F-actin/alpha-actinin networks exhibit strong strain hardening at short time scales and soften at long time scales. For F-actin networks in the absence of alpha-actinin or for F-actin/alpha-actinin networks at high temperatures, strain hardening appears only at very short time scales. We propose a model of strain hardening for F-actin networks, based on both the intrinsic rigidity of F-actin and dynamic topological constraints formed by the cross-linkers located at filaments entanglements. This model offers an explanation for the origin of strain hardening observed when shear stresses are applied against the cellular membrane.  相似文献   

15.
16.
The focal adhesion kinase (FAK) is a key regulator of cell migration. Phosphorylation at Tyr-397 activates FAK and creates a binding site for Src family kinases. FAK phosphorylates the cytoskeletal protein alpha-actinin at Tyr-12. Here we report that protein-tyrosine phosphatase 1B (PTP 1B) is an alpha-actinin phosphatase. PTP 1B-dependent dephosphorylation of alpha-actinin was seen in COS-7 cells and PTP 1B-null fibroblasts reconstituted with PTP 1B. Furthermore, we show that coexpression of wild-type alpha-actinin and PTP 1B causes dephosphorylation at Tyr-397 in FAK. No dephosphorylation was observed in cells coexpressing the alpha-actinin phosphorylation mutant Y12F and PTP 1B. Furthermore, the phosphorylation at four other sites in FAK was not altered by PTP 1B. In addition, we found that phosphorylated alpha-actinin bound to Src and reduced the binding of FAK to Src. The dephosphorylation at Tyr-397 in FAK triggered by wild-type alpha-actinin and PTP 1B caused a significant increase in cell migration. We propose that phosphorylated alpha-actinin disrupts the FAK x Src complex exposing Tyr-397 in FAK to PTP 1B. These findings uncover a novel feedback loop involving phosphorylated alpha-actinin and PTP 1B that regulates FAK x Src interaction and cell migration.  相似文献   

17.
Dictyostelium discoideum alpha-actinin (D.d. alpha-actinin) is a calcium and pH-regulated actin-binding protein that can cross-link F-actin into a gel at a submicromolar free calcium concentration and a pH less than 7 [Fechheimer, et al., 1982]. We examined mixtures of actin and D.d. alpha-actinin at four pH and calcium concentrations that exhibited various degrees of gelation or solation. The macroscopic viscosities of these mixtures were measured by falling ball viscometry (FBV) and compared to the translational diffusion coefficients measured by gaussian spot and periodic-pattern fluorescence photobleaching recovery (FPR) of both the actin filaments and D.d. alpha-actinin. A homogeneous, macroscopic gel was not composed of a static actin network. Instead, the filament diffusion coefficient decreased to approximately 65% of the control value. If the D.d. alpha-actinin concentration was increased, the solution became inhomogeneous, consisting of domains of higher actin concentration. These domains were often composed of a static actin network. The mobility of D.d. alpha-actinin consisted of a major fraction that freely diffused and a minor fraction that appeared immobile under the conditions employed. This suggested that D.d. alpha-actinin binding to the actin filaments was static over the time course of measurement (approximately 5 sec). Under solation conditions, there was no apparent interaction of actin with D.d. alpha-actinin. These results demonstrate that 1) actin filaments need not be cross-linked into an immobile, static array in order to have macroscopic properties of a gel; 2) interpretation of the rheological properties of actin:alpha-actinin gels are complicated by spatial heterogeneity of the filament concentration and mobility; and 3) a fraction of D.d. alpha-actinin binds statically to actin in undisturbed gels. The implications of these results are discussed in relation to cytoplasmic structure and contractility.  相似文献   

18.
The Dictyostelium 30,000-dalton protein is a calcium-regulated actin filament-bundling protein which has been suggested to contribute to the structure and reorganization of filopodia and pseudopodia accompanying cell movements. cDNAs encoding this protein were isolated using antibody and oligonucleotide probes to screen cDNA libraries in phage lambda. The sequence of the cDNA predicts a protein of 295 amino acids with a molecular weight of 33,355. The sequence reveals two EF-hand calcium-binding regions that provide a structural explanation for calcium regulation of the activity of this protein. The putative calcium-binding region of the 30,000-dalton protein has similarity to sequences of other calcium-regulated actin-binding proteins such as alpha-actin and fimbrin. One region of the sequence with similarity to both Dictyostelium gelation factor (ABP 120) and fructose bisphosphate aldolase is a potential actin-binding sequence. A highly charged region of the protein is similar to a sequence in human cytovillin that is repeated eight times in chicken gizzard caldesmon. No strong homology to previously identified actin-binding sequences of other actin-binding proteins is apparent. Results from Southern blot experiments indicate that the 30,000-dalton protein is encoded by a single gene in the Dictyostelium genome.  相似文献   

19.
BackgroundConsidering the importance of cellular mechanics in the birth and evolution of cancer towards increasingly aggressive stages, we compared nano-mechanical properties of non-tumoral (WPMY-1) and highly aggressive metastatic (PC-3) prostate cell lines both on cell aggregates, single cells, and membrane lipids.MethodsCell aggregate rheological properties were analyzed during dynamic compression stress performed on a homemade rheometer. Single cell visco-elasticity measurements were performed by Atomic Force Microscopy using a cantilever with round tip on surface-attached cells. At a molecular level, the lateral diffusion coefficient of total extracted lipids deposited as a Langmuir monolayer on an air-water interface was measured by the FRAP technique.ResultsAt cellular pellet scale, and at single cell scale, PC-3 cells were less stiff, less viscous, and thus more prone to deformation than the WPMY-1 control. Interestingly, stress-relaxation curves indicated a two-step response, which we attributed to a differential response coming from two cell elements, successively stressed. Both responses are faster for PC-3 cells. At a molecular scale, the dynamics of the PC-3 lipid extracts are also faster than that of WPMY-1 lipid extracts.ConclusionsAs the evolution of cancer towards increasingly aggressive stages is accompanied by alterations both in membrane composition and in cytoskeleton dynamical properties, we attribute differences in viscoelasticity between PC-3 and WPMY-1 cells to modifications of both elements.General significanceA decrease in stiffness and a less viscous behavior may be one of the diverse mechanisms that cancer cells adopt to cope with the various physiological conditions that they encounter.  相似文献   

20.
《Biorheology》1998,35(3):171-191
Cheese viscoelasticity is commonly measured using steady uniaxial compression, steady uniaxial extension, shear and compressive creep, and stress relaxation in shear and compression. Viscoelastic properties for many cheeses have also been studied using small amplitude oscillatory shear (SAOS). However, there is little on the measurement of nonlinear viscoelastic properties. The large deformation test usually conducted on cheese to study nonlinear viscoelasticity is uniaxial compression, but this test hardly departs from linear behavior. In this work, the nonlinear viscoelasticity of four cheese varieties was studied using large amplitude oscillatory shear (LAOS). A sliding plate rheometer incorporating a shear stress transducer was used. The data were evaluated using spectral analysis, and results are presented mainly in the form of shear stress versus shear rate loops.  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