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.
Cdk5 regulates adhesion and migration in a variety of cell types. We previously showed that Cdk5 is strongly activated during stress fiber formation and contraction in spreading cells. Here we determine the mechanism linking Cdk5 to stress fiber contractility and its relevance to cell migration. Immunofluorescence showed that Cdk5 colocalized with phosphorylated myosin regulatory light chain (pMRLC) on contracting stress fibers. Inhibiting Cdk5 activity by various means significantly reduced pMRLC level and cytoskeletal contraction, with loss of central stress fibers. Blocking Cdk5 activity also reduced Rho-Rho kinase (ROCK) signaling, which is the principal pathway of myosin phosphorylation under these conditions. Next, we examined the effect of Cdk5 activity on Src, a known regulator of Rho. Inhibiting Cdk5 activity increased Src activation and phosphorylation of its substrate, p190RhoGAP, an upstream inhibitor of Rho. Inhibiting both Cdk5 and Src activity completely reversed the effect of Cdk5 inhibition on Rho and prevented the loss of central stress fibers, demonstrating that Cdk5 exerts its effects on Rho-ROCK signaling by suppressing Src activity. Moreover, inhibiting either Cdk5 or ROCK activity increased cell migration to an equal extent, while inhibiting both kinases produced no additional effect, demonstrating that Cdk5-dependent regulation of ROCK activity is a physiological determinant of migration rate.Cell migration is essential for morphogenesis during embryonic development and for epithelial homeostasis and wound healing throughout life. As myosin II is involved in all aspects of cell migration, from cell polarization and adhesion to protrusion and tail retraction (34, 48), the signaling pathways regulating myosin-dependent cytoskeletal contraction are of particular interest. Myosin contraction is regulated by phosphorylation of myosin regulatory light chain (MRLC) at Thr18/Ser19. Although a number of kinases have been identified which phosphorylate these sites, the principal kinases in most cells are myosin light chain kinase (MLCK), a calcium/calmodulin-regulated enzyme, and Rho kinase (ROCK), a downstream effector of the Rho family GTPase RhoA. To provide the stringent control of cytoskeletal contraction needed for migration, RhoA is subject to both positive regulation by guanine nucleotide exchange factors (GEFs), such as GEF-H1 (4, 21), and negative regulation by GTPase-activating proteins (GAPs), such as the Src-regulated protein p190RhoGAP (1, 3, 10, 13). An additional level of regulation is provided by guanine nucleotide dissociation inhibitors, which bind to inactive RhoA and other Rho family GTPases, sequestering them in the cytosol (3). Two major downstream effectors of RhoA with regard to the cytoskeleton are the mammalian homologue of diaphanous, involved in actin polymerization (43), and ROCK, which phosphorylates MRLC and myosin phosphatase (20).Cdk5, a serine/threonine kinase, is an atypical member of the well-known family of cyclin-dependent kinases (Cdks). Unlike the other Cdks, it has no known function in cell cycle regulation and is activated by one of two noncyclin proteins, p35 or p39 (16, 41). Phosphorylation of Cdk5 at Y15 increases its activity severalfold (36, 49). Although Cdk5 is most abundant in neuronal cells, where it regulates migration, cytoskeletal dynamics, and membrane trafficking (37, 38, 45), a growing body of evidence indicates that Cdk5 has similar functions in nonneuronal cells (35). In particular, Cdk5 has been shown to strengthen cell-to-matrix adhesion and regulate migration in lens epithelial cells (28), corneal epithelial cells (11, 12, 40), keratinocytes (27), and CHO-K1 cells (15). The effects of Cdk5 on adhesion and migration have been linked, at least in part, to Cdk5-dependent phosphorylation of talin, which strengthens adhesion by slowing the rate of focal adhesion turnover (15). However, we have observed that Cdk5 not only binds to focal adhesions, where talin is located, but also to stress fibers (33). Moreover, in spreading cells, Cdk5 exerts its greatest effect on adhesion 1 to 2 h after plating (28), when stress fiber contraction is pronounced and Cdk5 activity is maximum (33). Therefore, we hypothesized that Cdk5 might regulate the MRLC phosphorylation necessary for stress fiber contraction and stability. To test this possibility, we examined the relationship of Cdk5 activity to MRLC phosphorylation and cytoskeletal contraction in spreading human lens epithelial cells.  相似文献   

2.
The balance of transition between distinct adhesion types contributes to the regulation of mesenchymal cell migration, and the characteristic association of adhesions with actin filaments led us to question the role of actin filament-associating proteins in the transition between adhesive states. Tropomyosin isoform association with actin filaments imparts distinct filament structures, and we have thus investigated the role for tropomyosins in determining the formation of distinct adhesion structures. Using combinations of overexpression, knockdown, and knockout approaches, we establish that Tm5NM1 preferentially stabilizes focal adhesions and drives the transition to fibrillar adhesions via stabilization of actin filaments. Moreover, our data suggest that the expression of Tm5NM1 is a critical determinant of paxillin phosphorylation, a signaling event that is necessary for focal adhesion disassembly. Thus, we propose that Tm5NM1 can regulate the feedback loop between focal adhesion disassembly and focal complex formation at the leading edge that is required for productive and directed cell movement.Among the different modes of migration that cells adopt, mesenchymal cell migration is dependent on integrin-based adhesion to the extracellular matrix (14), and the cellular mechanisms regulating integrin adhesion formation and turnover (adhesion dynamics) are integral to this process. The fate of integrin adhesions is intimately linked with filaments of polymerized actin (4). At the molecular level, actin filaments are highly dynamic, and this aspect of actin polymer biology provides an important control mechanism by which cells can organize filaments into structures with distinct properties. Tropomyosins are a multi-isoform family of actin-associating proteins that confer isoform-specific regulation of diverse actin filaments (3, 16, 34, 35). The interdependence of integrin adhesions and actin filaments suggests that expression of actin-associated proteins such as the tropomyosins may represent a mechanism for the regulation of adhesion dynamics that determine cell migration.In migrating cells small integrin-based focal complexes form at the periphery of lamellipodial extensions (32). These complexes are characterized by their subcellular distribution, dot-like shape, dependence on Rac activity, phosphorylated paxillin, and association with the network of short, branched actin filaments at the leading edge. The focal complexes are short lived (43) but provide strong traction forces at the leading edge (2) and most likely regulate directional migration (19). Subsets of focal complexes mature into focal adhesions, structures characterized by: Rho GTPase and Rho kinase dependence, dash-like shape, high levels of paxillin and phosphorylated paxillin, and low levels of the actin-binding molecule tensin (43, 44). The focal adhesions play an important role in anchoring bundles of polymerized actin stress fibers, providing the contractile force necessary for the translocation of the cell body during migration. There are at least three distinct classes of stress fibers observed in migrating cells (20, 27). Dorsal stress fibers are inserted into focal adhesions at the ventral surface of the cell. The distal end of the dorsal fibers can associate with a second type of actin fiber, the transverse arcs that run parallel to the leading edge and are not directly connected to focal adhesions. Ventral stress fibers have focal adhesions at either end and can be established following the contraction of two dorsal stress fibers and the associated transverse arc to form one actin bundle (20).Increased ventral stress fibers and focal adhesions are characteristic of nonmotile cells, in contrast, cell migration depends on focal adhesion turnover at the leading edge, allowing the formation of newly protruding regions of membrane and focal complex formation (28, 39). While the precise mechanism of focal adhesion turnover is incompletely understood, activation and phosphorylation of Src kinase, p130Cas, and paxillin (13, 39, 45) have all been implicated in focal adhesion turnover. A biphasic relationship between cell adhesion and cell speed suggests that conditions that alter the turnover rate of focal adhesions (either too much or too little) can reduce cell speed (18, 22).In cells with a fibroblastic phenotype, increased levels of acto-myosin contractility promote focal adhesion transition to fibrillar adhesions (also known as ECM contacts) (6, 7): elongated, thin, central arrays of dots or elongated fibrils that characteristically contain tensin but low levels of phosphorylated paxillin (29, 44, 45) and bind fibrils of fibronectin parallel to actin bundles (23, 29). These adhesions are formed by ligand-occupied fibronectin integrin receptor translocation from focal adhesions along bundles of actin filaments toward the cell center, and the process is dependent on an intact actin cytoskeleton and myosin activity (29). Receptor translocation stimulates matrix reorganization by transmitting cytoskeleton-generated tension through the integrin receptors onto the surrounding matrix (25, 29). The rate of receptor translocation is apparently independent from the rate of cell migration (29). However, the cytoskeletal tension that causes the fibrillar adhesion formation is also reported to decrease paxillin phosphorylation (45). Since phosphorylated paxillin is required for the generation of new focal complexes (45), conditions which switch the balance of adhesion in favor of fibrillar adhesion should presumably result in significantly reduced paxillin phosphorylation, leading to reduced focal adhesion turnover and correspondingly decreased cell migration.The cytoskeletal tropomyosin Tm5NM1 is a broadly distributed isoform (37) that alters cell shape (34), localizes to and promotes stress fibers that are resistant to actin depolymerizing drugs (9), enhances myosin IIA activation and recruitment to stress fibers, and inhibits cell migration (3). Therefore, we hypothesized that Tm5NM1 expression might determine cell migration by coordinating actin-dependent transition toward a predominance of focal adhesions and fibrillar adhesions. Using overexpression, knockdown, and genetic knockout models, we demonstrate that Tm5NM1 inhibits cell migration by promoting selective stabilization of focal adhesions and transition to fibrillar adhesions via the regulation of paxillin phosphorylation.  相似文献   

3.
KSR1 is a mitogen-activated protein (MAP) kinase scaffold that enhances the activation of the MAP kinase extracellular signal-regulated kinase (ERK). The function of KSR1 in NK cell function is not known. Here we show that KSR1 is required for efficient NK-mediated cytolysis and polarization of cytolytic granules. Single-cell analysis showed that ERK is activated in an all-or-none fashion in both wild-type and KSR1-deficient cells. In the absence of KSR1, however, the efficiency of ERK activation is attenuated. Imaging studies showed that KSR1 is recruited to the immunological synapse during T-cell activation and that membrane recruitment of KSR1 is required for recruitment of active ERK to the synapse.Kinase suppressor of Ras was originally identified in Drosophila melanogaster (53) and Caenorhabditis elegans (19, 32, 52) as a positive regulator of the extracellular signal-regulated kinase (ERK) mitogen-activated protein (MAP) kinase signaling pathway. It is thought to function as a MAP kinase scaffold because it can bind to Raf, MEK, and ERK (18, 19, 27, 28, 44, 59). While the exact function of KSR is unknown, preassembling the three components of the ERK MAP kinase cascade could function to enhance the efficiency of ERK activation, potentially regulate the subcellular location of ERK activation, and promote access to specific subcellular substrates (16, 45, 46).While only one isoform of KSR is expressed in Drosophila (53), two KSR isoforms have been identified in C. elegans (19, 32, 52) and most higher organisms. They are referred to as KSR1 and KSR2 (32, 43). While KSR1 mRNA and protein are detectable in a wide variety of cells and tissues, including brain, thymus, and muscle (10, 11, 29), little is known about the expression pattern of KSR2.We previously reported the phenotype of KSR1-deficient mice (30). These mice are born at Mendelian ratios and develop without any obvious defects. Using gel filtration, we showed that KSR1 promotes the formation of large signaling complexes containing KSR1, Raf, MEK, and ERK (30). Using both primary T cells stimulated with antibodies to the T-cell receptor as well as fibroblasts stimulated with growth factors, we showed that KSR1-deficient cells exhibit an attenuation of ERK activation with defects in cell proliferation.Here we explored the role of KSR1 in NK cell-mediated cytolysis. The killing of a target cell by a cytolytic T cell or NK cell is a complicated process that involves cell polarization with microtubule-dependent movement of cytolytic granules to an area that is proximal to the contact surface or immunological synapse (7, 33, 34, 48-50, 54). A variety of different signaling molecules are also involved, including calcium (23), phosphatidylinositol-3,4,5-triphosphate (13, 17), and activation of the ERK MAP kinase (6, 42, 56). Recently, the recruitment of activated ERK to the immunological synapse (IS) has been shown to be a feature of successful killing of a target by cytotoxic T lymphocytes (58).How active ERK is recruited to the synapse is not known. Since KSR1 is known to be recruited to the plasma membrane by Ras activation (24), and since the immunological synapse is one of the major sites of Ras activation (26, 41), it seemed plausible to test the hypothesis that KSR1 recruitment to the plasma membrane functions to recruit ERK to the immunological synapse and facilitate its activation. We found that KSR1 was recruited to the immunological synapse and that KSR1 appeared to be required for the localization of active ERK at the contact site. As KSR1-deficient cells exhibit a defect in killing, this suggests that KSR1 recruitment to the synapse may be important in the cytolytic killing of target cells.  相似文献   

4.
5.
Protein tyrosine kinase 6 (PTK6) is a nonmyristoylated Src-related intracellular tyrosine kinase. Although not expressed in the normal mammary gland, PTK6 is expressed in a majority of human breast tumors examined, and it has been linked to ErbB receptor signaling and AKT activation. Here we demonstrate that AKT is a direct substrate of PTK6 and that AKT tyrosine residues 315 and 326 are phosphorylated by PTK6. Association of PTK6 with AKT occurs through the SH3 domain of PTK6 and is enhanced through SH2 domain-mediated interactions following tyrosine phosphorylation of AKT. Using Src, Yes, and Fyn null mouse embryonic fibroblasts (SYF cells), we show that PTK6 phosphorylates AKT in a Src family kinase-independent manner. Introduction of PTK6 into SYF cells sensitized these cells to physiological levels of epidermal growth factor (EGF) and increased AKT activation. Stable introduction of active PTK6 into SYF cells also resulted in increased proliferation. Knockdown of PTK6 in the BPH-1 human prostate epithelial cell line led to decreased AKT activation in response to EGF. Our data indicate that in addition to promoting growth factor receptor-mediated activation of AKT, PTK6 can directly activate AKT to promote oncogenic signaling.Protein tyrosine kinase 6 (PTK6; also known as the breast tumor kinase BRK) is an intracellular Src-related tyrosine kinase (9, 48). Human PTK6 was identified in cultured human melanocytes (32) and breast tumor cells (39), while its mouse orthologue was cloned from normal small intestinal epithelial cell RNA (50). Although PTK6 shares overall structural similarity with Src family tyrosine kinases, it lacks an N-terminal myristoylation consensus sequence for membrane targeting (39, 51). As a consequence, PTK6 is localized to different cellular compartments, including the nucleus (14, 15). PTK6 is expressed in normal differentiated epithelial cells of the gastrointestinal tract (34, 42, 51), prostate (14), and skin (51-53). Expression of PTK6 is upregulated in different types of cancers, including breast carcinomas (6, 39, 54), colon cancer (34), ovarian cancer (47), head and neck cancers (33), and metastatic melanoma cells (16). The significance of apparent opposing signaling roles for PTK6 in normal differentiation and cancer is still poorly understood.In human breast tumor cells, PTK6 enhances signaling from members of the ErbB receptor family (10, 29, 30, 36, 40, 49, 54). In the HB4a immortalized human mammary gland luminal epithelial cell line, PTK6 promoted epidermal growth factor (EGF)-induced ErbB3 tyrosine phosphorylation and AKT activation (29). In response to EGF stimulation, PTK6 promoted phosphorylation of the focal adhesion protein paxillin and Rac1-mediated cell migration (10). PTK6 can be activated by the ErbB3 ligand heregulin and promotes activation of extracellular signal-regulated kinase 5 (ERK5) and p38 mitogen-activated protein kinase (MAPK) in breast cancer cells (40). PTK6 can also phosphorylate p190RhoGAP-A and stimulate its activity, leading to RhoA inactivation and Ras activation and thereby promoting EGF-dependent breast cancer cell migration and proliferation (49). Expression of PTK6 has been correlated with ErbB2 expression in human breast cancers (4, 5, 54).AKT (also called protein kinase B) is a serine-threonine kinase that is activated downstream of growth factor receptors (38). It is a key player in signaling pathways that regulate energy metabolism, proliferation, and cell survival (7, 45). Aberrant activation of AKT through diverse mechanisms has been discovered in different cancers (2). AKT activation requires phosphorylation of AKT on threonine residue 308 and serine residue 473. The significance of phosphorylation of AKT on tyrosine residues is less well understood. Src has been shown to phosphorylate AKT on conserved tyrosine residues 315 and 326 near the activation loop (11). Substitution of these two tyrosine residues with phenylalanine abolished AKT kinase activity stimulated by EGF (11). Use of the Src family inhibitor PP2 impaired AKT activation following IGF-1 stimulation of oligodendrocytes (13). The RET/PTC receptor tyrosine kinase that responds to glial cell-line-derived neurotrophic factor also phosphorylated AKT tyrosine residue 315 promoting activation of AKT (28). AKT tyrosine residue 474 was phosphorylated when cells were treated with the tyrosine phosphatase inhibitor pervanadate, and phosphorylation of tyrosine 474 contributed to full activation of AKT (12). Recently, the nonreceptor tyrosine kinase Ack1 was shown to regulate AKT tyrosine phosphorylation and activation (37).Here we show that AKT is a cytoplasmic substrate of the intracellular tyrosine kinase PTK6. We identify the tyrosine residues on AKT that are targeted by PTK6, and we demonstrate that tyrosine phosphorylation plays a role in regulating association between PTK6 and AKT. In addition, we show that PTK6 promotes AKT activation and cell proliferation in a Src-independent manner.  相似文献   

6.
In many human cancers, p27 downregulation correlates with a worse prognosis, suggesting that p27 levels could represent an important determinant in cell transformation and cancer development. Using a mouse model system based on v-src-induced transformation, we show here that p27 absence is always linked to a more aggressive phenotype. When cultured in three-dimensional contexts, v-src-transformed p27-null fibroblasts undergo a morphological switch from an elongated to a rounded cell shape, accompanied by amoeboid-like morphology and motility. Importantly, the acquisition of the amoeboid motility is associated with a greater ability to move and colonize distant sites in vivo. The reintroduction of different p27 mutants in v-src-transformed p27-null cells demonstrates that the control of cell proliferation and motility represents two distinct functions of p27, both necessary for it to fully act as a tumor suppressor. Thus, we highlight here a new p27 function in driving cell plasticity that is associated with its C-terminal portion and does not depend on the control of cyclin-dependent kinase activity.Dissemination of tumor cells is strictly linked to their ability to attach to and move within the extracellular matrix (ECM) in a three-dimensional (3D) environment. The use of 3D experimental model systems revealed that a higher complexity in cell migration and adaptation responses exists in the 3D model than in the classical 2D model (10, 16, 41, 49). A striking example is given by the fact that only in 3D could individually migrating cells use different mechanisms such as mesenchymal and amoeboid motility (16, 17). The relative slow mesenchymal migration is characterized by a fibroblast-like spindle shape and is dependent on integrin-mediated adhesion and on protease function (16). The amoeboid motility can in some cases represent a less adhesive, integrin-independent type of movement. Cells use a propulsive mechanism and are highly deformable, and rather than degrade the matrix, they are able to squeeze through it (16). As a result, the cells that use the amoeboid motility can potentially move faster than cells that use a mesenchymal strategy. Mesenchymal and amoeboid movements are also characterized by a different involvement of small GTPases of the Rho family. A high RhoA activity is associated mainly with the amoeboid motility, while the mesenchymal migration needs a high Rac activity at the leading edge to promote the extension of cellular protrusions (41, 48). Under certain circumstances, cancer cells can undergo conversion from a mesenchymal toward an amoeboid motility, an event referred as mesenchymal-amoeboid transition (MAT) (50). MAT represents a putative escape mechanism in tumor cell dissemination that could be induced by inhibition of pericellular proteolysis (50) or by increased membrane-associated RhoA activity (18, 40).Key mediators of cell motility through ECM substrates are the members of the Src family kinases. The prototype of Src family kinases, c-Src (14), is activated following cell-ECM adhesion and contributes to regulate the focal adhesion turnover and the cytoskeletal modifications necessary for normal cell adhesion and motility (52). The c-Src gene is the proto-oncogene of the transforming gene v-src of Rous sarcoma virus, and its elevated protein level and activity have been found in many human tumors (20, 28, 27, 34). Despite the accumulation of information and new molecular understanding of how Src is controlled, there is still an incomplete picture about its role in the generation of the malignant phenotype. v-Src shows higher levels of the kinase activity and transforming ability than c-Src (14, 15, 52). It induces normal cells to acquire a variety of transformed features, including alteration of morphology and increase of invasion ability due to its role in focal adhesion remodeling (7, 9, 13).Many data suggest that there is a close relationship between cell-ECM interaction and the proliferation and movements in both normal and tumor cells (5, 38, 43). Accordingly, Src activation may influence not only cell motility but also cell cycle progression by targeting the cell cycle inhibitor p27kip1 to proteasomal degradation (22, 39). Recent evidences indicated that p27kip1 (hereafter called p27) can also regulate cell migration, even though its role still remains controversial since it has been reported to either block or stimulate cell movements (1, 4, 11, 19, 21, 23, 29, 45).Based on these notions, we tested the possible contribution of p27 to the growth and motility phenotypes induced by v-src transformation, with special regard to those cellular invasive features that can be observed in 3D environments. By studying in vitro and in vivo the behavior of wild-type (WT) and p27-null fibroblasts transformed with v-src, we highlight a new role for p27 in the regulation of cellular plasticity that can ultimately drive tumor cell shape, motility, and invasion.  相似文献   

7.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

8.
9.
10.
11.
Proteins of the 14-3-3 and Rho-GTPase families are functionally conserved eukaryotic proteins that participate in many important cellular processes such as signal transduction, cell cycle regulation, malignant transformation, stress response, and apoptosis. However, the exact role(s) of these proteins in these processes is not entirely understood. Using the fungal maize pathogen, Ustilago maydis, we were able to demonstrate a functional connection between Pdc1 and Rho1, the U. maydis homologues of 14-3-3ɛ and Rho1, respectively. Our experiments suggest that Pdc1 regulates viability, cytokinesis, chromosome condensation, and vacuole formation. Similarly, U. maydis Rho1 is also involved in these three essential processes and exerts an additional function during mating and filamentation. Intriguingly, yeast two-hybrid and epistasis experiments suggest that both Pdc1 and Rho1 could be constituents of the same regulatory cascade(s) controlling cell growth and filamentation in U. maydis. Overexpression of rho1 ameliorated the defects of cells depleted for Pdc1. Furthermore, we found that another small G protein, Rac1, was a suppressor of lethality for both Pdc1 and Rho1. In addition, deletion of cla4, encoding a Rac1 effector kinase, could also rescue cells with Pdc1 depleted. Inferring from these data, we propose a model for Rho1 and Pdc1 functions in U. maydis.Morphological switching is a unique attribute of all dimorphic fungi, which alternate between budding and filamentous growth. In some cases, as with mating, this is a prerequisite for genetic diversity for this subfamily of fungi. In addition, many dimorphic fungal pathogens rely on this ability in order to effectively invade their host. In general, the transition between these alternate life forms means a complete turnover of cellular and proteomic components, which often involves cell cycle arrest and/or cytoskeletal rearrangement. Although the cellular proteomes associated with these two processes share many components, there are both temporal and spatial regulations that are manifested during the transitional phase (4).Temporal-spatial regulation of the proteome during the dimorphic transition requires cooperation and synchronized communication among different regulatory pathways. Two highly intricate, yet well-established, signaling cascades that regulate fungal morphogenesis are the mitogen-activated protein kinase (MAPK) (34, 46) and protein kinase A pathways (11). These signaling cascades detect and perpetuate extracellular stimuli, e.g., pheromones and nutrients, which lead to phase transitions in dimorphic fungi. Although the mechanisms are not as fully understood, members of two highly conserved families of proteins, Rho/Rac GTPases and 14-3-3 proteins, have also been shown to control filamentation. Constituents of the Rho/Rac protein family have been shown to regulate actin organization (26, 35, 36), cytokinesis (3, 49, 52), cell integrity (42, 56), pathogenicity (29), signal transduction (22, 44, 56), and cell migration (8). Their activity is dependent upon the reversible binding of guanine nucleotides catalyzed by guanine nucleotide exchange factors (GEFs) and GTPase activating proteins (GAPs) (15, 22, 23, 25, 44). Upon activation, Rho-GTPases stimulate downstream effector proteins such as p21-activated kinases (PAKs) (29, 51) or Rho kinase (ROCK) (36). Based on in silico analysis of the genome sequence, the fungal pathogen of maize, Ustilago maydis, contains six different Rho/Rac encoding genes: cdc42, rac1, and four additional genes predicted to encode Rho-like proteins (20). Of these, only the roles of Cdc42 and Rac1 have been examined in depth. Cdc42 was shown to regulate cytokinesis, while Rac1 regulates hyphal development in U. maydis (26). We examine here the place of a Rho1 homologue, Rho1, in U. maydis cell morphology, polarity, and development.Similarly, the highly conserved, ubiquitously expressed 14-3-3 proteins that are found in most eukaryotes have also been shown to contribute to cellular differentiation and cytoskeletal organization. Like Rho-GTPases, 14-3-3 proteins play multiple roles, in cytoskeletal function, cell cycle regulation, apoptosis, and the regulation of a variety of signaling pathways (17, 19, 33, 50). These acidic proteins have been found in each cellular compartment and most organisms examined possess multiple isoforms: seven isoforms are found in mammals, and as many as fifteen isoforms have been identified in plants (31). Interestingly, the yeast Saccharomyces cerevisiae, the fruit fly, Drosophila melanogaster, and the nematode, Caenorhabditis elegans, each possess only two 14-3-3 isoforms (50), while Candida albicans contains a single isoform (37). They function typically by binding their particular ligands at phosphoserine or phosphothreonine residues (50). It is not clear what 14-3-3 proteins do in the processes mentioned above, whether they act as scaffolds or effectors. Inspection of the U. maydis genome sequence revealed that this organism could be ideal for the study of 14-3-3 proteins because, unlike most other organisms, the U. maydis genome contains only a single 14-3-3 homologue. Due to its predicted binding of phosphorylated proteins, we named this homologue Pdc1 (for phosphorylation domain coupling protein [10]). Recently, the protein (also designated Bmh1 [32]) was also shown to be involved in cell cycle regulation.Despite the functional differences between Rho-GTPases and 14-3-3 proteins, we provide evidence that members of these two families participate in the same regulatory cascade(s) that control morphogenesis in the dimorphic fungus U. maydis. We are able to demonstrate that both Pdc1 and Rho1 are essential for cell viability. In addition, overexpression of Rho1 led to the reduction of filamentation. Overexpression of Rac1 triggers filamentation in U. maydis (13, 29). We show here that deleting Rac1 eliminates the lethal effect imposed by either Rho1 or Pdc1 depletion. Our results have led us to predict that both Rho1 and Pdc1 are negative regulators of Rac1 in U. maydis and that they play important roles in polarized growth and cytokinesis.  相似文献   

12.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

13.
Endothelial cell (EC) migration, cell-cell adhesion, and the formation of branching point structures are considered hallmarks of angiogenesis; however, the underlying mechanisms of these processes are not well understood. Lipid phosphate phosphatase 3 (LPP3) is a recently described p120-catenin-associated integrin ligand localized in adherens junctions (AJs) of ECs. Here, we tested the hypothesis that LPP3 stimulates β-catenin/lymphoid enhancer binding factor 1 (β-catenin/LEF-1) to induce EC migration and formation of branching point structures. In subconfluent ECs, LPP3 induced expression of fibronectin via β-catenin/LEF-1 signaling in a phosphatase and tensin homologue (PTEN)-dependent manner. In confluent ECs, depletion of p120-catenin restored LPP3-mediated β-catenin/LEF-1 signaling. Depletion of LPP3 resulted in destabilization of β-catenin, which in turn reduced fibronectin synthesis and deposition, which resulted in inhibition of EC migration. Accordingly, reexpression of β-catenin but not p120-catenin in LPP3-depleted ECs restored de novo synthesis of fibronectin, which mediated EC migration and formation of branching point structures. In confluent ECs, however, a fraction of p120-catenin associated and colocalized with LPP3 at the plasma membrane, via the C-terminal cytoplasmic domain, thereby limiting the ability of LPP3 to stimulate β-catenin/LEF-1 signaling. Thus, our study identified a key role for LPP3 in orchestrating PTEN-mediated β-catenin/LEF-1 signaling in EC migration, cell-cell adhesion, and formation of branching point structures.Angiogenesis, the formation of new blood vessels, involves several well-coordinated cellular processes, including endothelial cell (EC) migration, synthesis and deposition of extracellular matrix proteins, such as fibronectin, cell-cell adhesion, and formation of branching point structures (1-3, 19, 33); however, less is known about the underlying mechanisms of these processes (6, 8, 12, 14, 16, 17). For example, adherens junctions (AJs), which mediate cell-cell adhesion between ECs, may be involved in limiting the extent of cell migration (2, 14, 38, 40). VE-cadherin, a protein found in AJs, is a single-pass transmembrane polypeptide responsible for calcium-dependent homophilic interactions through its extracellular domains (2, 38, 40). The VE-cadherin cytoplasmic domain interacts with the Armadillo domain-containing proteins, β-catenin, γ-catenin (plakoglobin), and p120-catenin (p120ctn) (2, 15, 38, 40, 43). Genetic and biochemical evidence documents a crucial role of β-catenin in regulating cell adhesion as well as proliferation secondary to the central position of β-catenin in the Wnt signaling pathway (13, 16, 25, 31, 44). In addition, the juxtamembrane protein p120ctn regulates AJ stability via binding to VE-cadherin (2, 7, 9, 15, 21, 28, 32, 43). The absence of regulation or inappropriate regulation of β-catenin and VE-cadherin functions is linked to cardiovascular disease and tumor progression (2, 6).We previously identified lipid phosphate phosphatase 3 (LPP3), also known as phosphatidic acid phosphatase 2b (PAP2b), in a functional assay of angiogenesis (18, 19, 41, 42). LPP3 not only exhibits lipid phosphatase activity but also functions as a cell-associated integrin ligand (18, 19, 35, 41, 42). The known LPPs (LPP1, LPP2, and LPP3) (20-23) are six transmembrane domain-containing plasma membrane-bound enzymes that dephosphorylate sphingosine-1-phosphate (S1P) and its structural homologues, and thus, these phosphatases generate lipid mediators (4, 5, 23, 35, 39). All LPPs, which contain a single N-glycosylation site and a putative lipid phosphatase motif, are situated such that their N and C termini are within the cell (4, 5, 22, 23, 35, 39). Only the LPP3 isoform contains an Arg-Gly-Asp (RGD) sequence in the second extracellular loop, and this RGD sequence enables LPP3 to bind integrins (18, 19, 22). Transfection experiments with green fluorescent protein (GFP)-tagged LPP1 and LPP3 showed that LPP1 is apically sorted, whereas LPP3 colocalized with E-cadherin at cell-cell contact sites with other Madin-Darby canine kidney (MDCK) cells (22). Mutagenesis and domain swapping experiments established that LPP1 contains an apical targeting signal sequence (FDKTRL) in its N-terminal segment. In contrast, LPP3 contains a dityrosine (109Y/110Y) basolateral sorting motif (22). Interestingly, conventional deletion of Lpp3 is embryonic lethal, since the Lpp3 gene plays a critical role in extraembryonic vasculogenesis independent of its lipid phosphatase activity (11). In addition, an LPP3-neutralizing antibody was shown to prevent cell-cell interactions (19, 42) and angiogenesis (42). Here, we addressed the hypothesis that LPP3 plays a key role in EC migration, cell-cell adhesion, and formation of branching point structures by stimulating β-catenin/lymphoid enhancer binding factor 1 (β-catenin/LEF-1) signaling.  相似文献   

14.
15.
16.
17.
18.
19.
20.
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