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.
2.
3.
Assembly of appropriately oriented actin cables nucleated by formin proteins is necessary for many biological processes in diverse eukaryotes. However, compared with knowledge of how nucleation of dendritic actin filament arrays by the actin-related protein-2/3 complex is regulated, the in vivo regulatory mechanisms for actin cable formation are less clear. To gain insights into mechanisms for regulating actin cable assembly, we reconstituted the assembly process in vitro by introducing microspheres functionalized with the C terminus of the budding yeast formin Bni1 into extracts prepared from yeast cells at different cell-cycle stages. EM studies showed that unbranched actin filament bundles were reconstituted successfully in the yeast extracts. Only extracts enriched in the mitotic cyclin Clb2 were competent for actin cable assembly, and cyclin-dependent kinase 1 activity was indispensible. Cyclin-dependent kinase 1 activity also was found to regulate cable assembly in vivo. Here we present evidence that formin cell-cycle regulation is conserved in vertebrates. The use of the cable-reconstitution system to test roles for the key actin-binding proteins tropomyosin, capping protein, and cofilin provided important insights into assembly regulation. Furthermore, using mass spectrometry, we identified components of the actin cables formed in yeast extracts, providing the basis for comprehensive understanding of cable assembly and regulation.Eukaryotic cells contain populations of actin structures with distinct architectures and protein compositions, which mediate varied cellular processes (1). Understanding how F-actin polymerization is regulated in time and space is critical to understanding how actin structures provide mechanical forces for corresponding biological processes. Branched actin filament arrays, which concentrate at sites of clathrin-mediated endocytosis (2, 3) and at the leading edge of motile cells (4), are nucleated by the actin-related protein-2/3 (Arp2/3) complex. In contrast, bundles of unbranched actin filaments, which sometimes mediate vesicle trafficking or form myosin-containing contractile bundles, often are nucleated by formin proteins (514).Much has been learned about how branched actin filaments are polymerized by the Arp2/3 complex and how these filaments function in processes such as endocytosis (2, 15). In contrast, relatively little is known about how actin cables are assembled under physiological conditions. In previous studies, branched actin filaments derived from the Arp2/3 complex have been reconstituted using purified proteins (1619) or cellular extracts (2025). When microbeads were coated with nucleation-promoting factors for the Arp2/3 complex and then were incubated in cell extracts, actin comet tails were formed by sequential actin nucleation, symmetry breaking, and tail elongation. Importantly, the motility behavior of F-actin assembled by the Arp2/3 complex using defined, purified proteins differs from that of F-actin assembled by the Arp2/3 complex in the full complexity of cytoplasmic extracts (19, 2628).Formin-based actin filament assembly using purified proteins also has been reported (29, 30). However, reconstitution of formin-derived actin cables under the more physiological conditions represented by cell extracts has not yet been reported.The actin nucleation activity of formin proteins is regulated by an inhibitory interaction between the N- and C-terminal domains, which can be released when GTP-bound Rho protein binds to the formin N-terminal domain, allowing access of the C terminus (FH1-COOH) to actin filament barbed ends (3140). In yeast, the formin Bni1 N terminus also has an inhibitory effect on actin nucleation through binding to the C terminus (41).Interestingly, several recent reports provided evidence for cell-cycle regulation of F-actin dynamics in oocytes and early embryos (4245). However, which specific types of actin structures are regulated by the cell cycle and what kind of nucleation factors and actin interacting-proteins are involved remain to be determined.Here, we report a reconstitution of actin cables in yeast extracts from microbeads derivatized with Bni1 FH1-COOH, identifying the proteins involved, increasing the inventory of the proteins that regulate actin cable dynamics and establishing that the actin cable reconstitution in cytoplasmic extracts is cell-cycle regulated.  相似文献   

4.
The plant hormone indole-acetic acid (auxin) is essential for many aspects of plant development. Auxin-mediated growth regulation typically involves the establishment of an auxin concentration gradient mediated by polarly localized auxin transporters. The localization of auxin carriers and their amount at the plasma membrane are controlled by membrane trafficking processes such as secretion, endocytosis, and recycling. In contrast to endocytosis or recycling, how the secretory pathway mediates the localization of auxin carriers is not well understood. In this study we have used the differential cell elongation process during apical hook development to elucidate the mechanisms underlying the post-Golgi trafficking of auxin carriers in Arabidopsis. We show that differential cell elongation during apical hook development is defective in Arabidopsis mutant echidna (ech). ECH protein is required for the trans-Golgi network (TGN)–mediated trafficking of the auxin influx carrier AUX1 to the plasma membrane. In contrast, ech mutation only marginally perturbs the trafficking of the highly related auxin influx carrier LIKE-AUX1-3 or the auxin efflux carrier PIN-FORMED-3, both also involved in hook development. Electron tomography reveals that the trafficking defects in ech mutant are associated with the perturbation of secretory vesicle genesis from the TGN. Our results identify differential mechanisms for the post-Golgi trafficking of de novo-synthesized auxin carriers to plasma membrane from the TGN and reveal how trafficking of auxin influx carriers mediates the control of differential cell elongation in apical hook development.Polar auxin transport (PAT) plays a key role in plant development (15). PAT is mediated by plasma membrane localized auxin influx and efflux carriers of the auxin-resistant (AUX)/like-AUX (LAX), pin-formed (PIN), and ABCB families (612). Highly regulated tissue, cellular localization, and amount of auxin carriers at the plasma membrane (PM) provide directionality to the auxin transport and underlies the creation of auxin concentration gradient that is essential for controlling several aspects of plant development (1318). One of the developmental programs in which auxin concentration gradient plays a central role is the formation of apical hook, a bending in the embryonic stem during early seedling germination (19). Hook formation involves differential elongation of cells on the two opposite sides of the hypocotyl. This process is mediated by the formation of an auxin maximum at the concave side of the hook, leading to the inhibition of cell elongation (2025). A model based on mutational analysis shows that auxin carriers including polarly localized auxin efflux and influx facilitators PIN3 and AUX1/LAX3, respectively, are important for hook development (23, 24). The amount of auxin carriers at the PM is important for the regulation of auxin concentration, and this depends on the balance between secretion, endocytosis, and recycling. The analysis of PIN efflux carriers has revealed how cell wall anchoring, endocytosis, targeted degradation, and also posttranslational modifications strongly influence the location and amount of these carriers at the PM (15, 17, 2629). In contrast, little is known about the mechanisms and molecular components underlying the deposition of auxin carriers at the PM. Post-Golgi secretion to the PM occurs via the trans-Golgi network (TGN), a post-Golgi compartment (30). The TGN is a complex tubulo-vesicular membrane network maturing from the trans-most cisternae of the Golgi apparatus to become a highly dynamic independent structure from which secretory vesicles (SVs) and CLATHRIN-coated vesicles (CCVs) originate (3134). Although auxin carriers traffic via TGN, components and mechanisms specifically involved in trafficking to the PM of de novo-synthesized auxin carriers remain largely undefined (35, 36). Importantly, it is not known whether auxin carriers traffic through SV or CCV sites of the TGN on their way to the PM. We have used apical hook development as a model system to investigate the mechanisms that link post-Golgi trafficking of auxin carriers to the PM with control of differential cell elongation. We previously identified the transmembrane TGN-localized protein ECHIDNA (ECH) that is required for cell elongation (37). We discovered that the ech mutant is defective in hook development and is insensitive to ethylene like the aux1 mutant. These data prompted us to investigate the role of ECH and the TGN in post-Golgi trafficking of auxin carriers during hook development. Using genetic, pharmacological, and cell biological approaches, we show that distinct mechanisms/components underlie post-Golgi trafficking of influx and efflux carriers. We show that post-Golgi trafficking of de novo-synthesized AUX1 occurs via an ECH-dependent SV-based pathway, whereas that of PIN3 and LAX3 are largely independent of ECH at the TGN. Thus, these results reveal the complexity of trafficking from the TGN to PM as shown by the differential trafficking of influx carriers AUX1 versus LAX3 and the efflux carrier PIN3. Hence, our results reveal an additional layer of regulatory control to auxin transport.  相似文献   

5.
Invariant natural killer T (iNKT) cells recognize endogenous and exogenous lipid antigens presented in the context of CD1d molecules. The ability of iNKT cells to recognize endogenous antigens represents a distinct immune recognition strategy, which underscores the constitutive memory phenotype of iNKT cells and their activation during inflammatory conditions. However, the mechanisms regulating such “tonic” activation of iNKT cells remain unclear. Here, we show that the spatiotemporal distribution of CD1d molecules on the surface of antigen-presenting cells (APCs) modulates activation of iNKT cells. By using superresolution microscopy, we show that CD1d molecules form nanoclusters at the cell surface of APCs, and their size and density are constrained by the actin cytoskeleton. Dual-color single-particle tracking revealed that diffusing CD1d nanoclusters are actively arrested by the actin cytoskeleton, preventing their further coalescence. Formation of larger nanoclusters occurs in the absence of interactions between CD1d cytosolic tail and the actin cytoskeleton and correlates with enhanced iNKT cell activation. Importantly and consistently with iNKT cell activation during inflammatory conditions, exposure of APCs to the Toll-like receptor 7/8 agonist R848 increases nanocluster density and iNKT cell activation. Overall, these results define a previously unidentified mechanism that modulates iNKT cell autoreactivity based on the tight control by the APC cytoskeleton of the sizes and densities of endogenous antigen-loaded CD1d nanoclusters.It is well-established that different populations of T lymphocytes can recognize not only peptides in the context of major histocompatibility complex (MHC) class I (MHCI) and MHCII molecules but also, foreign and self-lipids in association with CD1 proteins (1), antigen-presenting molecules that share structural similarities with MHCI molecules. Of five CD1 isoforms, CD1d restricts the activity of a family of cells known as invariant natural killer T (iNKT) cells because of their semiinvariant T-cell receptor (TCR) use (1). To date, the exogenous glycolipid α-GalactosylCeramide (α-GalCer) represents the best characterized CD1d-restricted agonist for iNKT cells (2). Unlike conventional peptide-specific T cells, iNKT cells react against CD1d+ antigen-presenting cells (APCs) in the absence of exogenous antigens, a feature defined as autoreactivity (3). iNKT cell autoreactivity underpins the constitutive memory phenotype of iNKT cells and their ability to be activated during a variety of immune responses from infections to cancer and autoimmunity (1). Some of the endogenous antigens known to elicit iNKT cell autoreactivity belong to glycosphingolipid families, with a mix of α- and β-anomeric configurations (47). How iNKT cell autoreactivity is fine-tuned to prevent autoimmunity is subject of much investigation. Previous results have shown that exposure of APCs to Toll-like receptor (TLR) agonists enhances iNKT cell autoreactivity (8, 9), consistent with the proposed mechanism by which ligand availability is regulated by lysosomal glycosidases (4, 6).The recent application of advanced optical techniques (1013) in combination with substrate patterning and functionalization (14, 15) is providing detailed information on how the lateral organization of a variety of molecules located on both sides of the immunological synapse contributes to controlling T-cell activation. Specifically, single-molecule dynamic approaches and superresolution optical nanoscopy experiments have provided indisputable proof that many receptors on the cell membrane organize in small nanoclusters before ligand activation (16). Membrane nanodomains enriched in cholesterol and sphingolipids (17), protein–protein interactions (18), and interactions between transmembrane proteins and the cytoskeleton (19, 20) have been all implicated in regulating receptor dynamics and nanoclustering. An emerging concept attributes the actin cytoskeleton the ability of imposing barriers or fences on the cell membrane, restricting the lateral mobility of transmembrane proteins (1921). This transient restriction would, in turn, increase the local concentration of transmembrane proteins, leading to protein nanoclusters. For instance, it has been shown that the actin cytoskeleton promotes the dimerization rate of EGF receptors and facilitates ligand binding and signaling activation (18, 22). Confinement of CD36 has also been observed as a result of its diffusion along linear channels dependent on the integrity of the cortical cytoskeleton (23). This constrained diffusion promotes CD36 clustering, influencing CD36-mediated signaling and internalization. A similar mechanism has been proposed for the maintenance of MHCI clusters on the cell membrane by the actin cytoskeleton, with loss of MHCI clustering resulting in a decreased CD8 T-cell activation (24, 25).Recent confocal microscopy studies have revealed that the association between agonist-loaded CD1d molecules and lipid rafts might contribute to the regulation of iNKT cell activation (26). This elegant study for the first time, to our knowledge, linked the spatial organization of CD1d molecules on the cell membrane of APCs with the activation profile of iNKT cells. However, it remains unclear whether the results of these experiments obtained using mouse cells can be extended to human cells and whether additional insights can be obtained by using higher-resolution microscopy. Indeed, it is not yet known whether surface-expressed CD1d molecules exist as monomers or nanoclusters and whether the actin cytoskeleton might regulate CD1d lateral organization and iNKT cell activation. Interestingly, it has been recently reported that the actin cytoskeleton impairs antigen presentation by CD1d and that disruption of F actin or inhibition of the ρ-associated protein kinase enhances CD1d-mediated antigen presentation (27). These results suggest that the actin cytoskeleton might regulate, in a not yet known manner, antigen presentation by CD1d molecules.Here, we combined dual-color single-molecule dynamic approaches with superresolution optical nanoscopy to characterize for the first time, to our knowledge, the spatiotemporal behavior of CD1d on living human myeloid cells. We find that α-GalCer–loaded human CD1d (hCD1d) molecules are organized in nanoclusters on the cell membrane of APCs. We report that the actin cytoskeleton prevents enhanced hCD1d nanoclustering by hindering physical encountering between hCD1d diffusing nanoclusters, thus reducing basal iNKT cell activation. Furthermore, we observed an increase in nanocluster density on activation of APCs with inflammatory stimuli, such as TLR stimulation, mirroring the increased iNKT cell stimulation. Notably, even during inflammation, the actin cytoskeleton retains an important role to limit hCD1d cluster size and iNKT cell activation. Overall, our results suggest that regulation of CD1d nanoclustering through the actin cytoskeleton represents a previously unidentified mechanism to fine-tune peripheral iNKT cell autoreactivity.  相似文献   

6.
Cells can interact with their surroundings via filopodia, which are membrane protrusions that extend beyond the cell body. Filopodia are essential during dynamic cellular processes like motility, invasion, and cell–cell communication. Filopodia contain cross-linked actin filaments, attached to the surrounding cell membrane via protein linkers such as integrins. These actin filaments are thought to play a pivotal role in force transduction, bending, and rotation. We investigated whether, and how, actin within filopodia is responsible for filopodia dynamics by conducting simultaneous force spectroscopy and confocal imaging of F-actin in membrane protrusions. The actin shaft was observed to periodically undergo helical coiling and rotational motion, which occurred simultaneously with retrograde movement of actin inside the filopodium. The cells were found to retract beads attached to the filopodial tip, and retraction was found to correlate with rotation and coiling of the actin shaft. These results suggest a previously unidentified mechanism by which a cell can use rotation of the filopodial actin shaft to induce coiling and hence axial shortening of the filopodial actin bundle.Tubular membrane remodeling driven by the actin cytoskeleton plays a major role in both pathogenesis and in a healthy immune response. For instance, invadopodia, podosomes, and filopodia are crucial for invasion and migration of cells (1). Filopodia are thin (100 to 300 nm), tube-like, actin-rich structures that function as “antennae” or “tentacles” that cells use to probe and interact with their microenvironment (24). Such structures have been studied in vitro using model systems where the point-like 3D contacts with the extracellular matrix (ECM) have been mimicked by using optically trapped dielectric particles, functionalized with relevant ligands. These model systems allow for mechanical and visual control over filopodial dynamics and have significantly advanced the understanding of cellular mechanosensing (5).Extraction of membrane tubes, using optical trapping, has been used to investigate mechanical properties of the membrane–cytoskeleton system (610), or membrane cholesterol content (11), and has revealed important insight into the mechanism that peripheral proteins use to shape membranes (12). Motivated by the pivotal role of F-actin in the mechanical behavior of filopodia and other cellular protrusions, special focus has been on revealing the presence of F-actin within extracted membrane tubes (1, 1316). However, apart from a single study (9), fluorescent visualization of the F-actin was achieved by staining and fixation of the cells, and literature contains conflicting results regarding the presence or absence of actin in membrane tubes pulled from living cells (8, 11, 17).Filopodia in living cells have the ability to rotate and bend by a so-far unknown mechanism (13, 1820). For instance, filopodia have been reported to exhibit sharp kinks in neuronal cells (2123) and macrophages (6). These filopodial kinks have been observed both for surface-attached filopodia (23) as well as for filopodia that were free to rotate in three dimensions (6, 19).Other filaments such as DNA and bacterial flagella are common examples of structures that shorten and bend in response to torsional twist. Filopodia have previously been shown to have the ability to rotate by a mechanism that exists at their base (18, 19) and could have the ability to twist in presence of a frictional force. Rotation of the actin shaft results in friction with the surrounding filopodial membrane, and hence torsional energy can be transferred to the actin shaft. The myosin-Vb motor has been found to localize at the base of filopodia in neurite cells and to be critical for rotational movement of the filopodia; however, inhibition of myosin-Vc did not affect the rotation (19).Here, we reveal how actin filaments can simultaneously rotate and helically bend within cellular membrane tubes obtained by elongation of preexisting filopodia by an optically trapped bead. Simultaneous force measurements and confocal visualization reveal how the actin transduces a force as it rotates and retracts. After ∼100 s, the force exerted by the filopodium starts to exhibit pulling events reflecting transient contact between the actin and the tip region of the filopodium, which is attached to the optically trapped bead. We show that helical bending and rotation of the actin shaft (defined as the visible part of the actin) occurs simultaneously with movement of actin coils inside filopodia, which can occur concomitantly with a traction force exerted at the tip. The velocity of the coils depends on their location along the tube, thus indicating that retrograde flow is not the only mechanism driving the motion. Our data, and accompanying calculations, show that the rotation of the actin shaft, and the resulting torsional twist energy accumulated in the actin shaft, can contribute to shortening and bending of the actin shaft in conjunction with a retrograde flow.  相似文献   

7.
Rap1 is a small GTPase regulating cell–cell adhesion, cell–matrix adhesion, and actin rearrangements, all processes dynamically coordinated during cell spreading and endothelial barrier function. Here, we identify the adaptor protein ras-interacting protein 1 (Rasip1) as a Rap1-effector involved in cell spreading and endothelial barrier function. Using Förster resonance energy transfer, we show that Rasip1 interacts with active Rap1 in a cellular context. Rasip1 mediates Rap1-induced cell spreading through its interaction partner Rho GTPase-activating protein 29 (ArhGAP29), a GTPase activating protein for Rho proteins. Accordingly, the Rap1–Rasip1 complex induces cell spreading by inhibiting Rho signaling. The Rasip1–ArhGAP29 pathway also functions in Rap1-mediated regulation of endothelial junctions, which controls endothelial barrier function. In this process, Rasip1 cooperates with its close relative ras-association and dilute domain-containing protein (Radil) to inhibit Rho-mediated stress fiber formation and induces junctional tightening. These results reveal an effector pathway for Rap1 in the modulation of Rho signaling and actin dynamics, through which Rap1 modulates endothelial barrier function.The small GTPase Rap1 regulates both integrin-mediated and cadherin-mediated adhesions. Rap1 can increase cell adhesion by inducing the allosteric activation and clustering of integrins, thereby increasing cell–extracellular matrix (ECM) adhesion (13). Upon cell–ECM engagement, Rap1 induces cell spreading, due to increased cell protrusion and decreased cell contraction, indicating changes in actin dynamics (4, 5). In addition, Rap1 regulates both epithelial and endothelial cell–cell adhesion (611). Particularly the role of Rap1 in controlling endothelial cell junctions is important, as weakening of the endothelial barrier can result in pathologies such as chronic inflammation, atherosclerosis, and vascular leakage (1214). Activation of Rap1 in endothelial cells results in stabilization of junctions and consequently increased barrier function through the recruitment of β-catenin, resulting in stabilization of vascular endothelial (VE)–cadherin at cell–cell junctions (1518) and rearrangements of the actin cytoskeleton (6, 7, 1921). These rearrangements of the actin cytoskeleton include the disruption of radial stress fibers and the induction of cortical actin bundles, and consequently a switch from discontinuous, motile junctions into linear, stable junctions (68, 20). Rap1 achieves this at least in part by regulating Rho-signaling (6, 7, 10, 19, 20). The molecular mechanism of how Rap1 regulates Rho, however, remains largely elusive, although the Rap1-effector Krev interaction trapped protein 1 (Krit-1)/cerebral cavernous malformations 1 protein (CCM1) has been proposed to be involved (15, 16, 22).In this study, we identified a Rap1-signaling cascade, comprising ras-interacting protein 1 (Rasip1), ras-association and dilute domain-containing protein (Radil), and Rho GTPase-activating protein 29 (ArhGAP29), affecting both cell spreading and endothelial barrier function by regulating the Rho-signaling cascade.  相似文献   

8.
The surface of a living cell provides a platform for receptor signaling, protein sorting, transport, and endocytosis, whose regulation requires the local control of membrane organization. Previous work has revealed a role for dynamic actomyosin in membrane protein and lipid organization, suggesting that the cell surface behaves as an active composite composed of a fluid bilayer and a thin film of active actomyosin. We reconstitute an analogous system in vitro that consists of a fluid lipid bilayer coupled via membrane-associated actin-binding proteins to dynamic actin filaments and myosin motors. Upon complete consumption of ATP, this system settles into distinct phases of actin organization, namely bundled filaments, linked apolar asters, and a lattice of polar asters. These depend on actin concentration, filament length, and actin/myosin ratio. During formation of the polar aster phase, advection of the self-organizing actomyosin network drives transient clustering of actin-associated membrane components. Regeneration of ATP supports a constitutively remodeling actomyosin state, which in turn drives active fluctuations of coupled membrane components, resembling those observed at the cell surface. In a multicomponent membrane bilayer, this remodeling actomyosin layer contributes to changes in the extent and dynamics of phase-segregating domains. These results show how local membrane composition can be driven by active processes arising from actomyosin, highlighting the fundamental basis of the active composite model of the cell surface, and indicate its relevance to the study of membrane organization.The cell surface mediates interactions between the cell and the outside world by serving as the site for signal transduction. It also facilitates the uptake and release of cargo and supports adhesion to substrates. These diverse roles require that the cell surface components involved in each function are spatially and temporally organized into domains spanning a few nanometers (nanoclusters) to several micrometers (microdomains). The cell surface itself may be considered as a fluid–lipid bilayer wherein proteins are embedded (1). In the living cell, this multicomponent system is supported by an actin cortex, composed of a branched network of actin and a collection of filaments (24).Current models of membrane organization fall into three categories: those invoking lipid–lipid and lipid–protein interactions in the plasma membrane [e.g., the fluid mosaic model (1, 5) and the lipid raft hypothesis (6)], or those that appeal to the membrane-associated actin cortex (e.g., the picket fence model) (7), or a combination of these (8, 9). Although these models based on thermodynamic equilibrium principles have successfully explained the organization and dynamics of a range of membrane components and molecules, there is a growing class of phenomena that appears inconsistent with chemical and thermal equilibrium, which might warrant a different explanation. These include aspects of the organization and dynamics of outer leaflet glycosyl-phosphatidylinositol-anchored proteins (GPI-anchored proteins) (1013), inner leaflet Ras proteins (14), and actin-binding transmembrane proteins (13, 15, 16).Recent experimental and theoretical work has shown that these features can be explained by taking into account that many cortical and membrane proteins are driven by ATP-consuming processes that drive the system out of equilibrium (13, 15, 17). The membrane models mentioned above have by-and-large neglected this active nature of the actin cortex where actin filaments are being continuously polymerized and depolymerized (1821), in addition to being persistently acted upon by a variety of myosin motors (2224) that consume ATP and exert contractile stresses on cortical actin filaments, continually remodeling the architecture of the cortex (4, 21, 25). These active processes in turn can generate tangential stresses and currents on the cell surface, which could drive the dynamics and local composition of membrane components at different scales (22, 2629).Actin polymerization is proposed to be driven at the membrane by two nucleators, the Arp2/3 complex, which creates a densely branched network, as well as formins that nucleate filaments (18, 21, 30). A number of myosin motors are also associated with the juxtamembranous actin cortex, of which nonmuscle myosin II is the major component in remodeling the cortex and creating actin flows (4, 23, 25, 26, 31, 32). Based on our observations that the clustering of cell surface components that couple directly or indirectly to cortical actin [e.g., GPI-anchored proteins, proteins of the Ezrin, Radaxin, or Moesin (ERM) family (13, 15)] depends on myosin activity, we proposed that this clustering arises from the coupling to contractile actomyosin platforms (called “actin asters”) produced at the cortex (15, 33).A coarse-grained theory describing this idea has been put forward and corroborated by the verification of its key predictions in live cells (15, 33), but a systematic identification of the underlying microscopic processes is lacking. Given the complexity of numerous processes acting at the membrane of a living cell, we use an in vitro approach to study the effect of an energy-consuming actomyosin network on the dynamics of membrane molecules that directly interact with filamentous actin.A series of in vitro studies have explored the organization of confined, dynamic filaments (both actin and microtubules) (3439) or the role of actin architecture on membrane organization (4046). Indeed, these studies have yielded insights into the nontrivial emergent configurations that mixtures of polar filaments and motors can adopt when fueled by ATP (3437), in particular constitutively remodeling steady states that display characteristics of active mechanics (38, 39, 47). However, the effect of linking these mechanics to the confining lipid bilayer and its organization has not been studied.The consequences of actin polymerization on membrane organization, in particular on giant unilamellar vesicles (GUVs), have been addressed in a number of studies on the propulsion of GUVs by an actin comet tail (40, 45, 46). In those experiments, the apparent advection of membrane bound ActA or WASP toward the site of actin polymerization is mainly due to the change in binding affinity of WASP to actin through Arp2/3 (44) and the spherical geometry resulting in the drag of actin to one pole of the vesicle after symmetry break of the actin shell. That this dynamic process changes the bulk properties of the bilayer, namely the critical temperature of a phase-separating lipid bilayer, was shown by Liu and Fletcher (40) when the actin nucleator N-WASP was connected to a lipid species (PIP2) that was capable of partitioning into one of the two phases.Besides these pioneering studies on the effects of active processes on membrane organization, little was done to directly test the effect of active lateral stresses as well as actomyosin remodeling at the membrane, particularly on the dynamics and organization of membrane-associated components.To this end, we build an active composite in vitro by stepwise addition of components: a supported lipid bilayer with an actin-binding component, actin filaments, and myosin motors. By systematically varying the concentrations of actin and myosin as well as the average actin filament length, we find distinct states of actomyosin organization at the membrane surface upon complete ATP consumption. More importantly, we find that the ATP-fueled contractile actomyosin currents induce the transient accumulation of actin-binding membrane components. As predicted, the active mechanics of actin and myosin at physiologically relevant ATP concentrations drives the system into a nonequilibrium steady state with anomalous density fluctuations and the transient clustering of actin-binding components of the lipid bilayer (15, 33). Finally, connection of this active layer of actomyosin to a phase-segregating bilayer, influences its phase behavior and coarsening dynamics.  相似文献   

9.
10.
Cofilin, a critical player of actin dynamics, is spatially and temporally regulated to control the direction and force of membrane extension required for cell locomotion. In carcinoma cells, although the signaling pathways regulating cofilin activity to control cell direction have been established, the molecular machinery required to generate the force of the protrusion remains unclear. We show that the cofilin phosphatase chronophin (CIN) spatiotemporally regulates cofilin activity at the cell edge to generate persistent membrane extension. We show that CIN translocates to the leading edge in a PI3-kinase–, Rac1-, and cofilin-dependent manner after EGF stimulation to activate cofilin, promotes actin free barbed end formation, accelerates actin turnover, and enhances membrane protrusion. In addition, we establish that CIN is crucial for the balance of protrusion/retraction events during cell migration. Thus, CIN coordinates the leading edge dynamics by controlling active cofilin levels to promote MTLn3 cell protrusion.Cofilin is one crucial mediator of actin cytoskeletal dynamics during cell motility (15). At the cell edge, cofilin severs F-actin filaments, generating substrates for Arp2/3-mediated branching activity and contributing to F-actin depolymerization by creating a new pointed end and F-actin assembly by increasing the pool of polymerization-competent actin monomers (G-actin) (6, 7). Because of its ability to sever actin filaments and thus, modulate actin dynamics, the precise spatial and temporal regulation of cofilin activity at the cell leading edge is crucial to cell protrusion, chemotaxis, and motility both in vitro and in vivo (2, 813). Misregulation of cofilin activity and/or expression is directly related to diseases, including tumor metastasis (1418) and Alzheimer’s disease (19).Several mechanisms regulate tightly the activation of cofilin in response to upstream stimuli, including interaction with phosphatidylinositol (4,5)-bisphosphate (2022), local pH changes (23, 24), and phosphorylation at a single regulatory serine (Ser3) (8, 25). The phosphorylation of cofilin, leading to its inactivation, is catalyzed by two kinase families: the LIM-kinases [LIMKs(Lin11, Isl-1, and Mec-3 domain)] and the testicular kinases (2527). Two primary families of ser/thr phosphatases dephosphorylate and reactivate the actin-depolymerizing and -severing functions of cofilin: slingshot (SSH) (28) and chronophin (CIN) (29).SSH was identified as a cofilin phosphatase through genetic studies in Drosophila (28). The most active and abundant SSH isoform, SSH-1L, has been implicated in such biological processes as cell division, growth cone motility/morphology, neurite extension, and actin dynamics during membrane protrusion (30). SSH dephosphorylates a number of actin regulatory proteins in addition to cofilin, including LIMK1 (31) and Coronin 1B (32). CIN is a haloacid dehydrogenase-type phosphatase, a family of enzymes with activity in mammalian cells that has been poorly characterized. CIN dephosphorylates a very limited number of substrates (33) and as opposed to SSH, has little phosphatase activity toward LIMK both in vitro and in vivo; thus, it seems to be the more specific activator of cofilin (29, 30). CIN exhibits several predicted interaction motifs potentially linking it to regulation by PI3-kinase and phospholipase Cγ (PLCγ), both of which have been implicated in signaling to cofilin activation in vivo in MTLn3 adenocarcinoma cells (10, 34). CIN has been involved in cell division (29), cofilin–actin rod formation in neurons (35), and chemotaxing leukocytes (36, 37). The molecular mechanisms that control the activity and localization of CIN in cells are still not well-understood. In neutrophils, CIN mediates cofilin dephosphorylation downstream of Rac2 (36), and stimulation of protease-activated receptor2 results in recruitment of CIN and cofilin at the cell edge by β-arrestins to promote localized generation of free actin barbed ends, membrane protrusion, and chemotaxis (37). Chemotaxis to EGF by breast tumor cells is directly correlated with cancer cell invasion and metastasis (38, 39). Although cofilin activity is required for tumor cell migration, the contribution(s) of CIN to the regulation of actin dynamics at the leading edge has not yet been investigated.The importance of cofilin in regulating tumor cell motility has been extensively studied using MTLn3 mammary carcinoma cells as a model system. The initial step of MTLn3 cell chemotaxis to EGF consists of a biphasic actin polymerization response resulting from two peaks of free actin barbed end formation (34, 40, 41). The first or early peak of actin polymerization occurs at 1 min after EGF stimulation and requires both cofilin and PLCγ activities (34), but it is not dependent on cofilin dephosphorylation (42). This first transient allows the cells to sense EGF gradients and initiate small-membrane protrusions (11). The second or late peak of actin polymerization occurs at 3 min and is dependent on both cofilin and PI3-kinase activities (43, 44). Cofilin activity in this late transient has been associated with full protrusion of lamellipodia (34). The mechanism by which cofilin becomes activated at the 3-min peak has not been identified, although it is likely to involve the phosphoregulation of Ser3 (42, 45).In this work, we determine the molecular mechanisms involved in the full protrusion of the leading edge upon EGF stimulation. We have identified CIN as a critical regulator of cofilin activation to coordinate leading edge dynamics. Our results yield insights into how CIN controls cell protrusion, a key step in the process of cell migration and metastasis.  相似文献   

11.
12.
13.
Auxin binding protein 1 (ABP1) has been studied for decades. It has been suggested that ABP1 functions as an auxin receptor and has an essential role in many developmental processes. Here we present our unexpected findings that ABP1 is neither required for auxin signaling nor necessary for plant development under normal growth conditions. We used our ribozyme-based CRISPR technology to generate an Arabidopsis abp1 mutant that contains a 5-bp deletion in the first exon of ABP1, which resulted in a frameshift and introduction of early stop codons. We also identified a T-DNA insertion abp1 allele that harbors a T-DNA insertion located 27 bp downstream of the ATG start codon in the first exon. We show that the two new abp1 mutants are null alleles. Surprisingly, our new abp1 mutant plants do not display any obvious developmental defects. In fact, the mutant plants are indistinguishable from wild-type plants at every developmental stage analyzed. Furthermore, the abp1 plants are not resistant to exogenous auxin. At the molecular level, we find that the induction of known auxin-regulated genes is similar in both wild-type and abp1 plants in response to auxin treatments. We conclude that ABP1 is not a key component in auxin signaling or Arabidopsis development.The auxin binding protein 1 (ABP1) was first isolated from maize plants based on its ability to bind auxin (1). The crystal structure of ABP1 demonstrated clearly that ABP1 has an auxin-binding pocket and, indeed, binds auxin (2). However, the elucidation of the physiological functions of ABP1 has been challenging because the first reported abp1 T-DNA insertion mutant in Arabidopsis was not viable (3). Nevertheless, ABP1 has been recognized as an essential gene for plant development and as a key component in auxin signaling (49). Because viable abp1 null mutants in Arabidopsis were previously unavailable, alternative approaches have been used to disrupt ABP1 function in Arabidopsis to determine the physiological roles of the protein. Cellular immunization approaches were used to generate ABP1 knockdown plants (10, 11). Inducible overexpression of the single chain fragment variable regions (scFv12) of the anti-ABP1 monoclonal antibody mAb12 both in cell lines and in Arabidopsis plants presumably neutralizes the endogenous ABP1 activities (10, 11). Two such antibody lines, SS12S and SS12K, have been widely used in many ABP1-related studies (4, 6, 911). The results obtained from the characterization of the antibody lines suggest that ABP1 regulates cell division, cell expansion, meristem activities, and root development (4, 6, 10, 12, 13). Transgenic plants that overexpress ABP1 antisense RNA were also used to elucidate the physiological functions of ABP1 (4, 10). Moreover, missense point mutation alleles of abp1 have also been generated through the Arabidopsis TILLING project. One such TILLING mutant, named abp1-5, harbors a mutation (His94 >Tyr) in the auxin-binding pocket and has been widely used in many ABP1-related studies (4, 8, 9). Previous studies based on the antisense lines, antibody lines, and Arabidopsis mutant alleles have led to the conclusion that ABP1 is essential for embryogenesis, root development, and many other developmental processes. However, the interpretation of results generated by using the ABP1 antisense and antibody lines are not straightforward and off-target effects have not been completely ruled out. We believe that characterization of abp1 null plants is urgently needed to unambiguously define the roles of ABP1 in auxin signaling and in plant development.In the past several years, studies of the presumed ABP1-mediated auxin signal transduction pathway were carried out in several laboratories. It has been hypothesized that ABP1 is an auxin receptor mediating fast, nongenomic effects of auxin (46, 8, 9), whereas the TIR1 family of F-box protein/auxin receptors are responsible for auxin-mediated gene regulation (14, 15). One of the proposed functions of ABP1 is to regulate subcellular distribution of PIN auxin efflux carriers (6, 9, 13). Furthermore, a recent report suggests that a cell surface complex consisting of ABP1 and transmembrane receptor-like kinases functions as an auxin receptor at the plasma membrane by activating the Rho-like guanosine triphosphatases (GTPases) (ROPs) in an auxin-dependent manner (8). ROPs have been reported to play a role in regulating cytoskeleton organization and PIN protein endocytosis (5, 6). However, it is important to unequivocally determine the biological processes that require ABP1 before extensive efforts are directed toward elucidating any ABP1-mediated signaling pathways.In this paper, we generate and characterize new abp1 null mutants in Arabidopsis. We are interested in elucidating the molecular mechanisms by which auxin regulates flower development because our previously identified auxin biosynthetic mutants display dramatic floral defects (1618). Because ABP1 was reported as an essential gene and ABP1 binds auxin (2, 3), we decided to determine whether ABP1 plays a role in flower development. We used our recently developed ribozyme-based CRISPR gene editing technology (19) to specifically inactivate ABP1 during flower development. Unexpectedly, we recovered a viable abp1 mutant (abp1-c1, c stands for alleles generated by using CRISPR) that contains a 5-bp deletion in the first exon of ABP1. We also isolated a T-DNA abp1 allele (abp1-TD1) that harbors a T-DNA insertion in the first exon of ABP1. We show that both abp1-c1 and abp1-TD1 are null mutants. Surprisingly, the mutants were indistinguishable from wild-type (WT) plants at all of the developmental stages we analyzed. Our data clearly demonstrate that ABP1 is not an essential gene and that ABP1 does not play a major role in auxin signaling and Arabidopsis development under normal growth conditions.  相似文献   

14.
The rapid reorganization and polarization of actin filaments (AFs) toward the pathogen penetration site is one of the earliest cellular responses, yet the regulatory mechanism of AF dynamics is poorly understood. Using live-cell imaging in Arabidopsis, we show that polarization coupled with AF bundling involves precise spatiotemporal control at the site of attempted penetration by the nonadapted barley powdery mildew fungus, Blumeria graminis f. sp. hordei (Bgh). We further show that the Bgh-triggered AF mobility and organelle aggregation are predominately driven by the myosin motor proteins. Inactivation of myosins by pharmacological inhibitors prevents bulk aggregation of organelles and blocks recruitment of lignin-like compounds to the penetration site and deposition of callose and defensive protein, PENETRATION 1 (PEN1) into the apoplastic papillae, resulting in attenuation of penetration resistance. Using gene knockout analysis, we demonstrate that highly expressed myosins XI, especially myosin XI-K, are the primary contributors to cell wall-mediated penetration resistance. Moreover, the quadruple myosin knockout mutant xi-1 xi-2 xi-i xi-k displays impaired trafficking pathway responsible for the accumulation of PEN1 at the cell periphery. Strikingly, this mutant shows not only increased penetration rate but also enhanced overall disease susceptibility to both adapted and nonadapted fungal pathogens. Our findings establish myosins XI as key regulators of plant antifungal immunity.In nature, plants are constantly exposed to a large number of pathogens including fungi, bacteria, and viruses. In response, plants have evolved multiple layers of defense mechanisms to resist the pathogen attack (1). The first line of plant defense against fungi is penetration resistance that is achieved by localized cell wall appositions (CWAs), also called papillae, on an inner surface of cell walls at the site of fungal penetration (2). CWAs consist primarily of callose (β-1,3-glucan), lignin, cell wall proteins, and reactive oxygen species (24). The focal deposition of these elements in papillae appears to be an early and essential factor in plant penetration resistance (5).Studies on the genetic basis of penetration resistance have revealed that entry control of A. thaliana against nonadapted powdery mildews largely depends on several PENETRATION (PEN) genes (PEN1, PEN2, and PEN3). All three PEN proteins are also recruited to attempted fungal penetration sites (611). Intriguing findings show that the focal accumulation of PEN1 and PEN3 occurs outside the plasma membrane and within papillae or haustorial encasements (3, 11, 12). Disruption of actin cytoskeleton by pharmacological inhibitors blocks PEN3–GFP accumulation at most penetration sites, but has a lesser effect on the recruitment of GFP–PEN1 to these sites (11), suggesting that transport pathways mediating PEN1 and PEN3 recruitment and export to the apoplastic papillae are distinct.Accumulation of dynamically moving cytoplasm near the pathogen penetration site is the most striking and microscopically visible early response in epidermal cells (2). The secretory vesicles and organelles, including peroxisomes, Golgi, mitochondria, and the nucleus, also move toward penetration sites (7, 13). In addition to the deposition of cell wall reinforcements and focal accumulation of penetration-related proteins such as PEN3, the accretion of cytoplasm and organelles at sites of attempted fungal penetration involves reorganization of actin cytoskeleton, which forms a radial array focused on penetration site (10, 11, 1418). Consistent with this finding, disruption of AFs hampers penetration resistance, leading to increased penetration frequency by various fungal and oomycete pathogens (1517, 19). However, the mechanisms that drive AF dynamics and active transport of cellular components toward sites of attempted pathogen penetration remain elusive. Myosins are molecular motors responsible for AF-based motility (20). Recently, plant class XI myosins were implicated in the organization of actin cytoskeleton, organelle and vesicle transport, cell expansion, and plant growth (2127). Although none of the individual myosin gene knockouts produces plant growth defects (22), progressive elimination of two to four highly expressed myosins results in concomitant reduction in cell and plant size (23, 24). However, relatively little is known about the functions of myosins in plant–pathogen interactions (28, 29).Using pharmacological and genetic approaches to disrupt myosin function in Arabidopsis, we show that transient assembly and polarization of actin filament (AF) bundles toward the fungal penetration site are regulated by myosin motors. Furthermore, we demonstrate that plant myosins contribute to focal aggregation of a battery of cellular defense activities at the infection site and papillary deposition of cell wall appositions of lignin-like compounds, callose and PEN1, and are required for plant penetration resistance.  相似文献   

15.
Actin filaments and integrin-based focal adhesions (FAs) form integrated systems that mediate dynamic cell interactions with their environment or other cells during migration, the immune response, and tissue morphogenesis. How adhesion-associated actin structures obtain their functional specificity is unclear. Here we show that the formin-family actin nucleator, inverted formin 2 (INF2), localizes specifically to FAs and dorsal stress fibers (SFs) in fibroblasts. High-resolution fluorescence microscopy and manipulation of INF2 levels in cells indicate that INF2 plays a critical role at the SF–FA junction by promoting actin polymerization via free barbed end generation and centripetal elongation of an FA-associated actin bundle to form dorsal SF. INF2 assembles into FAs during maturation rather than during their initial generation, and once there, acts to promote rapid FA elongation and maturation into tensin-containing fibrillar FAs in the cell center. We show that INF2 is required for fibroblasts to organize fibronectin into matrix fibers and ultimately 3D matrices. Collectively our results indicate an important role for the formin INF2 in specifying the function of fibrillar FAs through its ability to generate dorsal SFs. Thus, dorsal SFs and fibrillar FAs form a specific class of integrated adhesion-associated actin structure in fibroblasts that mediates generation and remodeling of ECM.The dynamic connection between the forces generated in the actomyosin cytoskeleton and integrin-mediated focal adhesions (FAs) to the extracellular matrix (ECM) is essential for many physiological processes including cell migration, vascular formation and function, the immune response, and tissue morphogenesis. These diverse functions are mediated by distinct cellular structures including protruding lamellipodia containing nascent FAs that mediate haptotaxis (1), ventral adhesive actin waves that mediate leukocyte transmigration through endothelia (2, 3), and stress fibers (SFs) and FAs that drive fibrillarization of ECM in developing embryos (4, 5). The coordination and interdependence of actin and integrin-based adhesion in these specialized cellular structures are rooted in their biochemical interdependence. Activation of integrins to their high-affinity ECM binding state requires the actin cytoskeleton (6). In turn, integrin engagement with ECM induces signaling that mediates actin polymerization and contractility downstream of Rho GTPases (6, 7). ECM-engaged integrins also affect cytoskeletal organization by physically linking the contractile actomyosin system to extracellular anchorage points (7). Thus, adhesion-associated actin structures are integrated systems that mediate cellular functions requiring coordination of intracellular cytoskeletal forces with ECM binding.Mesenchymal cells generally possess two main types of adhesion-associated actin structures: protruding lamellipodia containing nascent FAs at the cell edge and linear actin bundles in the cell body connected to FAs. Compared with architecturally invariant lamellipodia, adhesion-associated actin bundle structures, including filopodia, the perinuclear actin cap/transmembrane actin-associated nuclear lines, trailing edge bundles, and dorsal SFs, are more diverse in their morphology and less well understood in their architecture and function (810). The most-studied actin bundle structure is perhaps dorsal SFs, noncontractile bundles associated at one end with a ventral FA near the cell edge and that extend radially toward the cell center and join with dorsal actin arcs on their other end. How the functional specificity of dorsal SFs is generated apart from the many other distinct adhesion-associated actin bundle structures is not well understood.The functional specificity of adhesion-associated actin structures could be generated either on the adhesion side by compositional differences in FA proteins or on the actin side by differences in the nucleation mechanism and actin binding proteins. On the adhesion side, it is well known that different integrin family members bind distinct types of ECM (11, 12). However, cells adhered to different ECMs all form common structures including lamellipodia, filopodia, and multiple types of SFs. In addition to different integrins, FA function could be regulated by the process of “maturation” in which FAs undergo stereotypical dynamic changes in composition and morphology driven by actomyosin-mediated cellular tension (13, 14). Nascent FAs contain integrins, focal adhesion kinase (FAK), a-actinin, and paxillin (13, 15). When tension is applied, nascent FAs grow and recruit hundreds of proteins, including talin, vinculin, and zyxin (16). These mature FAs then either disassemble or further mature into tensin-containing fibrillar FAs that are responsible for fibronectin fibrillogenesis (17). Thus, the changes in FA size and protein content that accompany FA maturation could give rise to functional specialization of adhesion/actin systems.On the other hand, actin filaments in migrating cells are generated by two main classes of nucleators: the Arp2/3 complex and formins (18). Different nucleating proteins generate different actin organization and geometries, which could in turn dictate functional specificity of adhesions. Arp2/3 forms the branched network in lamellipodia and is thought to be linked to nascent FAs through interaction with FAK (1921) or vinculin (22). The formin family of actin nucleators, which generates linear actin bundles (23), is more diverse, although formins share a common actin assembly core domain (24), (25). Recent work has begun to ascribe the generation of particular actin structures to some of the 15 formins in mammalian cells, particularly members of the diaphanous family and FHOD1 (2629). Specifically regarding dorsal SFs, evidence points strongly to polymerization by a formin family member (23, 3032) but no formin has ever been localized to these SFs or their associated FAs in motile cells. Thus, although formins are clearly critical for forming distinct actin structures, whether they cooperate with FA proteins to specify the function of adhesion-associated actin structures in the cell is unclear.We hypothesized that inverted formin 2 (INF2), found in our recent FA proteome (33), may play a critical role in the formation and functional specificity of adhesion-associated actin structures. INF2 is expressed in cells in two isoforms, one containing a membrane-targeting CAAX-motif that plays a role in mitochondrial fission (34) and a non-CAAX isoform whose function is not well characterized. INF2 is an unusual formin insofar as it contains, in addition to the FH1–FH2 domains that polymerize actin, a WH2-like domain at the C terminus (35) that binds actin monomers to regulate autoinhibition, and also mediates filament severing (35, 36). INF2 also interacts with and inhibits members of the diaphanous family of formin proteins (37). INF2 therefore could have multiple possible roles at FAs in local modulation of actin.Here we explore the role of INF2 in mouse embryonic fibroblasts (MEFs). We find for the first time to our knowledge strong localization of an endogenous formin to FAs at the distal tips of dorsal SFs where it is required for actin polymerization at FAs to form dorsal SFs. We show that INF2 plays a role in controlling morphological, but not compositional maturation of FAs. Strikingly, INF2 is responsible for the formation of one specific class of FAs, the fibrillar FAs that organize the ECM; disruption of INF2 leads to defects in ECM fibrillogenesis. Thus, our study demonstrates that INF2 mediates the formation of dorsal SFs and fibrillar FAs, which together comprise a specific integrated adhesion-associated actin structure responsible for the fibrillogenesis of ECM by fibroblasts.  相似文献   

16.
An essential question of morphogenesis is how patterns arise without preexisting positional information, as inspired by Turing. In the past few years, cytoskeletal flows in the cell cortex have been identified as a key mechanism of molecular patterning at the subcellular level. Theoretical and in vitro studies have suggested that biological polymers such as actomyosin gels have the property to self-organize, but the applicability of this concept in an in vivo setting remains unclear. Here, we report that the regular spacing pattern of supracellular actin rings in the Drosophila tracheal tubule is governed by a self-organizing principle. We propose a simple biophysical model where pattern formation arises from the interplay of myosin contractility and actin turnover. We validate the hypotheses of the model using photobleaching experiments and report that the formation of actin rings is contractility dependent. Moreover, genetic and pharmacological perturbations of the physical properties of the actomyosin gel modify the spacing of the pattern, as the model predicted. In addition, our model posited a role of cortical friction in stabilizing the spacing pattern of actin rings. Consistently, genetic depletion of apical extracellular matrix caused strikingly dynamic movements of actin rings, mirroring our model prediction of a transition from steady to chaotic actin patterns at low cortical friction. Our results therefore demonstrate quantitatively that a hydrodynamical instability of the actin cortex can trigger regular pattern formation and drive morphogenesis in an in vivo setting.Self-organization is one of the principal mechanisms of biological pattern formation at the molecular, cellular, and tissue scale. Although the pioneering work of Turing (1) has suggested reaction–diffusion as a generic route toward pattern generation (2), a concrete biomolecular or mechanical understanding of how this might occur in vivo remains elusive, except in a few specific cases (35). For instance, Kondo and coworkers (6) demonstrated that pigment patterning on the skin of the Pomocanthus imperator can be understood quantitatively from the simple attraction–repulsion kinetics of two cell types.At the cellular level, active structures, such as the cytoskeleton, are generically expected to display a large variety of structures from a theoretical perspective (712), many of which have been reproduced in elegant in vitro studies (1315). In the case of actomyosin gels, the contractile stresses arising from molecular motors have been shown to create large actin flows that can reorganize the cortex (16, 17). Because actin filaments and motors are “self-advected,” or transported, by their own flow (18), there is a self-reinforcing loop in gel density, capable of creating patterns. Nevertheless, most theoretical studies do not consider the cross-effects of polymerization and diffusion, which resist pattern formation. Interestingly, in the past years, several groups have reported in vivo examples of actin patterns: mammalian axons (19), Caenorhabditis elegans embryo (20), and Drosophila trachea (21) are all cellular cylinders that display a regular array of concentric actin rings on their cortex.In this article, we study the example of ring formation in the Drosophila trachea and propose a generic mechanism for stable actin pattern formation, arising from the interplay of actin turnover and myosin activity. The model makes clear predictions, which we test through fly genetics and drug experiments.  相似文献   

17.
Filopodia are dynamic, finger-like plasma membrane protrusions that sense the mechanical and chemical surroundings of the cell. Here, we show in epithelial cells that the dynamics of filopodial extension and retraction are determined by the difference between the actin polymerization rate at the tip and the retrograde flow at the base of the filopodium. Adhesion of a bead to the filopodial tip locally reduces actin polymerization and leads to retraction via retrograde flow, reminiscent of a process used by pathogens to invade cells. Using optical tweezers, we show that filopodial retraction occurs at a constant speed against counteracting forces up to 50 pN. Our measurements point toward retrograde flow in the cortex together with frictional coupling between the filopodial and cortical actin networks as the main retraction-force generator for filopodia. The force exerted by filopodial retraction, however, is limited by the connection between filopodial actin filaments and the membrane at the tip. Upon mechanical rupture of the tip connection, filopodia exert a passive retraction force of 15 pN via their plasma membrane. Transient reconnection at the tip allows filopodia to continuously probe their surroundings in a load-and-fail manner within a well-defined force range.Filopodia are actin-rich cell membrane protrusions, involved in processes as diverse as cell migration, wound closure, and cell invasion by pathogens (13). During cell migration, filopodia can exert forces on the substrate (4, 5) and act as precursors of focal adhesions (68). Filopodia initiate contacts during wound closure and contribute to dorsal closure of the fruit fly embryo in a zipper-like fashion (912). Viruses can hijack filopodia and filopodia-like cell–cell bridges to surf toward the cell body (13, 14). Filopodia from macrophages and epithelial cells actively pull pathogens bound to their tips (1518). In all these examples filopodial retraction and retrograde force production are crucial. However, although filopodia formation and growth have been well studied (13), the mechanisms underlying their retraction are poorly understood.Filopodia show continuous rearward movement of their actin filaments in a process called “retrograde flow” (3, 19). In the lamellipodium, from which filopodia often emanate, the retrograde flow originates from actin treadmilling due to actin depolymerization at the rear and polymerization at the front of the lamellipodium. This retrograde flow is further amplified by the motor activity of myosins (2023). In neurons, the filopodial shaft is deeply anchored in the growth cone and filopodial dynamics depends on the balance between actin polymerization at the filopodial tip and its retrograde flow (19). In other cell types actin depolymerization at the tip has been associated with retracting filopodia (24).Different contributions to filopodial force production during retraction can be considered. A connection between the filopodial tip and retracting actin filaments through transmembrane receptors such as integrins could transduce cortical forces applied on the actin shaft. In macrophages, force measurements on retracting filopodia suggested a major role for cortical myosins pulling on filopodial actin bundles (16). These measurements showed that retraction could be slowed down for forces below 20 pN. Applied forces higher than 20 pN inverted filopodial retraction of macrophages (25).Filopodial force production can also be due to membrane mechanics (26). Forces exerted by actin-free tubes extruded from the cell plasma membrane typically range between 5 pN and 30 pN (27). Membrane tension could drive filopodial retraction by exerting inward forces against the actin filaments. Moreover, filopodial actin filaments have been found disconnected from the membrane at the tip (28, 29), underlining the importance of membrane properties in filopodial mechanics. The contributions of membrane- and actin-based forces, as well as the mechanical links controlling force production during filopodial retraction, are still unclear.Here, we studied the retraction dynamics and the forces exerted by a single filopodium that is contacting an optically trapped bead at its tip. We found that filopodia retracted in association with a reduced actin polymerization at their tip at rates below those needed to compensate for the retrograde flow. The speed of filopodial retraction was only marginally affected by counteracting forces up to 50 pN, suggesting that the driving forces for retraction were not limiting within this range. We argue that actin treadmilling in the cell cortex, that functions far from its stall regime, transduces inward forces to the filopodial actin shaft at the base via high friction. In addition we found that filopodia can exert passive inward forces of 15 pN by using cell membrane-based forces. External counterforces that are only 5 pN higher than the membrane force can lead to rupture of connections between the actin shaft and the membrane at the filopodial tip. These weak contacts at the tip define the maximal pulling force of filopodia and allow cytoskeletal inward forces to operate only for short time intervals (<25 s). We found that the mechanical disconnection between membrane and actin filaments is only transient as actin dynamics at the tip are altered after disconnection. A continuous load-and-fail behavior allows thus tip-bound filopodia to probe the mechanics of their environment.  相似文献   

18.
Grain size is one of the key factors determining grain yield. However, it remains largely unknown how grain size is regulated by developmental signals. Here, we report the identification and characterization of a dominant mutant big grain1 (Bg1-D) that shows an extra-large grain phenotype from our rice T-DNA insertion population. Overexpression of BG1 leads to significantly increased grain size, and the severe lines exhibit obviously perturbed gravitropism. In addition, the mutant has increased sensitivities to both auxin and N-1-naphthylphthalamic acid, an auxin transport inhibitor, whereas knockdown of BG1 results in decreased sensitivities and smaller grains. Moreover, BG1 is specifically induced by auxin treatment, preferentially expresses in the vascular tissue of culms and young panicles, and encodes a novel membrane-localized protein, strongly suggesting its role in regulating auxin transport. Consistent with this finding, the mutant has increased auxin basipetal transport and altered auxin distribution, whereas the knockdown plants have decreased auxin transport. Manipulation of BG1 in both rice and Arabidopsis can enhance plant biomass, seed weight, and yield. Taking these data together, we identify a novel positive regulator of auxin response and transport in a crop plant and demonstrate its role in regulating grain size, thus illuminating a new strategy to improve plant productivity.Because it is one of the most important staple food crops cultivated worldwide, improvement of grain yield is a major focus of rice-breeding programs (1). Grain size is one of the determining factors of grain yield (2, 3). A number of quantitative trait loci (QTLs) controlling rice grain size have been identified in recent years (411). However, functional mechanisms of these genes remain largely unknown. Because QTLs usually have important functions in determining grain size, many of them have been widely selected in breeding processes or existed in modern elite varieties, and a certain QTL could be only applicable in certain varieties (12). Thus, exploration of new grain size-associated genes and elucidation of their functional mechanisms have great significance for further improvement of rice yield (12).Seed size, as well as other organ size, is controlled by various plant hormones, such as auxin, brassinosteroid, and cytokinin (10, 13, 14). A number of studies have demonstrated that auxin plays a vital role in organ size determination by affecting cell division, cell expansion, and differentiation (1517). Auxin exists predominantly as indole-3-acetic acid (IAA) in plants, and genetic studies of its biosynthetic genes in Arabidopsis have demonstrated that IAA regulates many aspects of plant growth and development, including stem elongation, lateral branching, vascular development, and tropic growth responses (18, 19). Combined with biochemical studies, the tryptophan (Trp)-dependent IAA biosynthesis pathway has been clearly established involving the YUCCA family flavin monooxgenases (20). Importantly, the two-step pathway is highly conserved throughout the plant kingdom (21). Until very recently, the Trp-independent auxin biosynthetic pathway was elucidated as contributing to early embryogenesis in Arabidopsis (22). Primary auxin signaling is a rapid process initiated from the hormone perception by receptor TIR1, an F-box protein, followed by degradation of the negative regulator AUX/IAA proteins, and further release the downstream auxin response factors (ARFs) (2326). However, how the ARFs work in plants remains elusive. Auxin transport, generally referring to the cell-to-cell transportation of the hormone directed basipetally from shoots to roots in vascular tissues, plays a critical role in auxin response (18). The transport involves a number of membrane-associated proteins, such as PINs (protein inhibitor of nNOS), AUX1 (AUXIN TRANSPORTER PROTEIN 1), and ABCBs (ATP-BINDING CASSETTE, SUB-FAMILY B PROTEINS) as efflux or influx carriers (2730). Disruption of auxin transport induced by either gene mutations or chemical inhibitor treatment will lead to diverse development defects, such as decreased lateral organ initiation and defective tropic growth responses (27, 3134).In this study, we identify a rice mutant, named big grain1-D (Bg1-D) because it is a dominant mutant having extralarge grain size. BG1 encodes a novel plasma membrane-associated protein, and is specifically induced by auxin treatment. We show that BG1 is a new positive regulator of auxin response involved in auxin transport, and demonstrate that manipulation of BG1 expression can greatly improve grain size and plant productivity.  相似文献   

19.
ADP ribosylation factor (Arf) 6 anchors to the plasma membrane, where it coordinates membrane trafficking and cytoskeleton remodelling, but how it assembles actin filaments is unknown. By reconstituting membrane-associated actin assembly mediated by the WASP family veroprolin homolog (WAVE) regulatory complex (WRC), we recapitulated an Arf6-driven actin polymerization pathway. We show that Arf6 is divergent from other Arf members, as it was incapable of directly recruiting WRC. We demonstrate that Arf6 triggers actin assembly at the membrane indirectly by recruiting the Arf guanine nucleotide exchange factor (GEF) ARNO that activates Arf1 to enable WRC-dependent actin assembly. The pathogen Salmonella usurped Arf6 for host cell invasion by recruiting its canonical GEFs EFA6 and BRAG2. Arf6 and its GEFs facilitated membrane ruffling and pathogen invasion via ARNO, and triggered actin assembly by generating an Arf1–WRC signaling hub at the membrane in vitro and in cells. This study reconstitutes Arf6-dependent actin assembly to reveal a mechanism by which related Arf GTPases orchestrate distinct steps in the WRC cytoskeleton remodelling pathway.ADP ribosylation factor (Arf) GTPases are best known for their roles in vesicle and organelle trafficking (1). Class I and II Arfs (Arf1, Arf3, Arf4, and Arf5) are found predominantly in and around the Golgi apparatus. In contrast, the more divergent Class III Arf (Arf6) operates almost exclusively at the plasma membrane (2). Consistent with its localization, Arf6 has been heavily implicated in trafficking events at the cell surface, including the regulation of endocytosis and exocytosis (1). In particular, Arf6 and its guanine nucleotide exchange factors (GEFs) are believed to be pivotal to the recycling of endosomes and receptors to and from the plasma membrane (3, 4). Arf6 also has a clear role in cortical cytoskeleton rearrangement (5). This is strongly supported by evidence of Arf6 and Rac1 (Ras-related C3 botulinum toxin substrate) interplay (6), exemplified by Arf6 recruitment of the Rac1 GEF Kalirin, Arf6 promotion of Rac1 activation, and lamellipodia formation (79).Rac1 is required for generation of lamellipodia that lead to membrane ruffles and macropinocytosis (10). Rac1 is thought to achieve this by activating the WRC, which comprises WAVE (WASP family veroprolin homolog), Abi (abl-interactor 1), Cyfip (cytoplasmic FMR1 interacting protein), Nap1 (NCK-associated protein 1), HSPC300 (heat shock protein C300), or their homologs (1012). We recently established that Rac1 was not sufficient for WRC recruitment to the membrane and its activation, which instead requires direct binding by Rac1 and an Arf GTPase (13). This in vitro Arf activity could be supplied by multiple isoforms including Arf1 and Arf5, but only Arf1 facilitated WRC-dependent lamellipodia formation and macropinocytosis of the bacterial pathogen Salmonella into human host cells (1315). Intriguingly, Arf6 also promoted Salmonella invasion (14), and, given the capability of Arfs to modulate WRC, it seems likely that Arf6 also recruits and activates the WRC at the plasma membrane. However, despite mounting evidence that Arf6 remodels the cytoskeleton, a molecular mechanism by which Arf6 drives Arp2/3-dependent actin assembly has not been resolved.  相似文献   

20.
The actin cytoskeleton is a key element of cell structure and movement whose properties are determined by a host of accessory proteins. Actin cross-linking proteins create a connected network from individual actin filaments, and though the mechanical effects of cross-linker binding affinity on actin networks have been investigated in reconstituted systems, their impact on cellular forces is unknown. Here we show that the binding affinity of the actin cross-linker α-actinin 4 (ACTN4) in cells modulates cytoplasmic mobility, cellular movement, and traction forces. Using fluorescence recovery after photobleaching, we show that an ACTN4 mutation that causes human kidney disease roughly triples the wild-type binding affinity of ACTN4 to F-actin in cells, increasing the dissociation time from 29 ± 13 to 86 ± 29 s. This increased affinity creates a less dynamic cytoplasm, as demonstrated by reduced intracellular microsphere movement, and an approximate halving of cell speed. Surprisingly, these less motile cells generate larger forces. Using traction force microscopy, we show that increased binding affinity of ACTN4 increases the average contractile stress (from 1.8 ± 0.7 to 4.7 ± 0.5 kPa), and the average strain energy (0.4 ± 0.2 to 2.1 ± 0.4 pJ). We speculate that these changes may be explained by an increased solid-like nature of the cytoskeleton, where myosin activity is more partitioned into tension and less is dissipated through filament sliding. These findings demonstrate the impact of cross-linker point mutations on cell dynamics and forces, and suggest mechanisms by which such physical defects lead to human disease.Movement, morphology, and force production are essential aspects of animal life. At the cellular level, these mechanics are largely determined by the actin cytoskeleton—a network of actin filaments connected by cross-linkers to create a 3D biopolymer frame. These cross-linkers are not permanent, but bind transiently. Studies using reconstituted proteins show that when these cross-linkers are attached to and connect multiple filaments, they create a network that behaves like a weak elastic solid. When cross-links unbind, actin filaments are free to slide past one another, producing a network that behaves more like a viscous fluid (1, 2). This dynamic cross-linking makes the actin network a viscoelastic material that is solid-like on short timescales such as seconds, yet fluid-like on longer timescales such as minutes (3, 4).The timescale of the transition from solid to fluid-like behavior in reconstituted actin networks is set by the duration of cross-linking, which is in turn set by cross-linker dissociation rate, Koff (1, 5, 6). The ability for an actin network to move between solid and fluid-like states may be an essential mechanism to balance mechanical and structural integrity of an elastic solid with adaptability and movement (1, 79). Despite this clear physical role of actin cross-linking in reconstituted network mechanics, the impact of changes in cross-linker affinity on cell force generation has not, to our knowledge, been previously investigated.To examine the role of cross-linking dynamics on cellular forces, we study α-actinin, a 100-kDa actin cross-linking protein that exists as a dumbbell-shaped head-to-tail homodimer of ∼40 nm (10). α-Actinin binds to actin using two N-terminal calponin homology regions, which together create actin-binding domains (ABDs). These domains on opposite ends of the dimer allow α-actinin to act as a cross-linker, forming networks of loose bundles of actin filaments (11, 12). In humans, there are four genes that encode highly homologous forms of α-actinin. Point mutations in the ABD of ACTN4 cause a form of kidney damage known as focal segmental glomerulosclerosis (FSGS) (1315). With FSGS, the specialized podocyte cells that form part of the filtration barrier between the blood and urine lose their normal extended structure (13). This dysfunction leads to malfunction of the glomerular filter and decreased renal function that often progresses to kidney failure (15). In particular, one point mutation (K255E) appears to expose a cryptic actin-binding site, which increases the affinity of ACTN4 for actin. Previous research using purified proteins has shown that this K255E mutation creates a more elastic and solid-like actin network (6, 1618). Because the moduli of reconstituted actin networks are highly sensitive to cross-linker affinity and concentration (1, 2, 19), we hypothesized that changes in actin cross-linking will also significantly impact the total force and work exerted by an active cell on its environment. However, the effect of this mutation, or variable cross-linking in general, on the dynamics and forces of cells is largely unknown (20). Understanding how cross-linker binder affinity changes force generation will provide essential insight into how cells modulate their mechanics, and exert forces on their surrounding environment.In this article we quantify the affinity of variable ACTN4 cross-linking and its effects on cellular movement, intracellular transport, and traction forces. We measure that the K255E mutant form of ACTN4 has a threefold increase in affinity for actin in cells compared with wild type. This increased affinity creates a cell that is ∼50% less motile and displays 70% less cytoplasmic mobility. Surprisingly, though this mutation reduces both intra- and extracellular movement, it increases contractile stresses by 300%, and the total work done on the substrate by 500%. Thus, this point-mutation change in cross-linker binding has profound macroscopic effects on cell dynamics and force translation, providing a sensitive mechanism for the cell to modulate myosin work between cell contractility and movement.  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