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.
Polarity is a fundamental cell property essential for differentiation, proliferation and morphogenesis in unicellular and multicellular organisms. We have recently demonstrated that phosphatidylinositol 3-kinase (PI3K) activity is required for the establishment of anterior-posterior axis, leading to asymmetrical localization of F-actin in migrating monospores of the red alga Porphyra yezoensis. We also showed that the formation of the apical-basal axis via adhesion of monospores to the substratum after the cessation of migration requires newly synthesized proteins and does not depend on PI3K activity. However, little is known about the mechanism and regulation of axis conversion during development of monospores. In this addendum, we report our investigation as to the role of the cell wall in axis conversion. Our results indicate that inhibition of cell wall synthesis prevented the development of germlings. Also, defects in the cell wall disrupted the asymmetrical distribution of F-actin and inhibited the adhesion to the substratum that is required for establishment of apical-basal axis. Hence, we conclude that the cell wall is critical for the maintenance of cell polarity in migrating cells, which is indirectly involved in axis conversion via enabling monospores to adhere to the substratum.Key words: BFA, cell polarity, cell wall, F-actin, monospores, PI3K, Porphyra yezoensisThe initial establishment of cell polarity, which is exhibited in asymmetrical cell division and directional migration, depends on asymmetrical cues that lead to reorganization of the cytoskeleton and polarized distribution of cortical proteins and membrane lipids.13 For directional migration of Dictyostelium cells and leukocytes, cells in the axialized form can rapidly change their body shape along with the formation of cell polarity in response to external impulses such as cAMP and cytokines, enabling them to migrate toward the external impulse with driving and contractile forces provided by asymmetrically distributed cytoskeletal elements.4,5 Evidence is growing that in both asymmetrical cell division and migration, intracellular compartmentalization of phosphatidylinositol 3-kinase (PI3K) and phosphatidylinositol polyphosphate phosphatases is responsible for the asymmetrical and reciprocal distributions of PI(3,4,5)P3 and PI(4,5)P2 on plasma membranes. This helps cells to define their polarity by organizing polarized localization of F-actin and myosin.68We used the monospores of the red alga Porphyra yezoensis to elucidate the molecular mechanisms involved in the establishment of cell polarity in plants. Migration and asymmetrical cell division are both observed during the early development of monospores released from monosporangia produced at the marginal region of the thallus.9,10 Thus, monospores are thought to be unique and useful materials for investigating polarity determination in plant cells. In the early development of monospores, there are two different cellular axes: the anterior-posterior axis during migration and the apical-basal axis in asymmetrical cell division and upward growth of a thallus. The use of LY294002, a PI3K inhibitor, prevented the migration of monospores because the anterior-posterior axis cannot be established. This is evidence that the PI3K activity is essential for the establishment of cell polarity and asymmetric distribution of F-actin for migration of monospores.10 Thus, the formation of the former axis requires PI3K activity and asymmetrical distribution of F-actin.10 These results are similar to those observed in Dictyostelium cells and leukocytes, suggesting that the role of PI3K-dependent F-actin asymmetry in the establishment of cell polarity might be evolutionarily conserved in migrating eukaryotic cells. However, it is still unclear whether PI(3,4,5)P3 corresponds to D3-phosphorylated phosphatidylinositol in P. yezoensis, since this phosphoinositide has not yet been detected in any plant cell.11,12In addition to D3-phosphorylated phosphatidylinositols, it is well known that the cell wall plays an essential role in the establishment of cell polarity, a phenomenon documented in the brown algae Fucus.1315 It has been demonstrated that the cell wall is required for the fixation, but not the formation, of the apical-basal axis.13,14 In this case, polarized secretion via Golgi apparatus is needed for synthesis of the cell wall.15 In fact, we observed that monospores whose migration was inhibited by treatment with PI3K and cytoskeleton inhibitors have no cell wall,10 suggesting the importance of the cell wall in formation and/or maintenance of the cell axis in P. yezoensis.To confirm this possibility, we used Brefeldin A (BFA), a specific inhibitor of polysaccharide biosynthesis required for cell wall formation via Golgi-derived vesicle trafficking. As shown in Figure 1A, part a, the cell wall was synthesized during migration of monospores. However, when freshly released monospores were treated with BFA for 3 h, there was no cell wall synthesis in monospores with a rounded shape or in migrating monospores with a tapered shape (Fig. 1A, part b). This evidence led us to conclude that Golgi-derived vesicle trafficking is responsible for cell wall formation in these monospores. These results also indicated that the anterior-posterior axis during migration can be established without cell wall synthesis. Indeed, F-actin accumulated at the leading edge in the migrating monospores in the presence of BFA (Fig. 1A, part d).Open in a separate windowFigure 1Effect of BFA on the cell wall synthesis, F-actin localization and development of monospores. (A) Freshly released monospores were treated with (parts b–d) or without (part a) BFA (MP Biomedicals) at 20 µM for 3 h incubation. The cell wall (parts a and b) and F-actin (parts c and d) were stained with 0.01% Fluorescent Brightener 28 (Sigma) and 5 U· mL−1 Alex Flour 488 phalloidin (Molecular probe), respectively. Migrating monospores are indicated by arrowheads. (Part c) Cell with a round shape, (Part d) cell with a tapered shape during migration. Upper and lower photographs in each panel show bright field and fluorescent images, respectively. Direction of migrating monospores is indicated by an arrow. Scale bars: (Parts a and b) 10 µm; (Parts c and d) 5 µm. (B) Freshly released monospores were treated with (parts b–d) or without (part a) BFA at 20 µM for 24 h incubation. The cell wall (parts a and b) and F-actin (parts c and d) were stained with 0.01% Fluorescent Brightener 28 and 5 U· mL−1 Alex Flour 488 phalloidin, respectively. Migrating monospores are indicated by arrowheads. (Part c) Cell with a round shape, (Part d) cell with a tapered shape during migration. Upper and lower photographs in each panel show bright field and fluorescent images, respectively. Direction of migrating monospores is indicated by an arrow. Scale bars: (Parts a and b) 10 µm; (Parts c and d) 5 µm. (C) Dose-dependent effect of BFA on early development of monospores. Freshly released monospores were treated with an increasing concentration (2–20 µM) of BFA for 24 h, and the number of germlings was counted. Data are presented as mean ± SD (n = 3).Next, we analyzed the relationship between cell wall synthesis and development of germlings to confirm the functional significance of cell wall synthesis in the establishment of apical-basal axis. In the control medium, the cell wall was observed in 2-celled germlings 24 h after monospores release (Fig. 1B, part a), while the BFA-treated monospores still retained the migrating form in which the cell wall was not synthesized (Fig. 1B, part b) and the adhesion of monospores to the substratum and development of germlings was prevented (Fig. 1B, parts b–d). Moreover, the rate of the development of germlings from monospores decreased in a dose-dependent manner after 24 h incubation with BFA (Fig. 1C). Notably, it is significant that the polarized localization of F-actin in migrating cells was destroyed during BFA treatment (Fig. 1B, part d), suggesting the involvement of the cell wall in the maintenance of the asymmetrical distribution of F-actin in migrating monospores. Thus, the cell wall is indispensable for maintenance of anterior-posterior axis in migrating cells. Moreover, since adhering is trigger of the formation of the apical-basal axis, synthesis of cell wall could enable cells to develop further into germlings. We therefore concluded that cell wall plays a role indirectly in axis conversion during the development of monospores. Future work should be focused on the nature of the cell wall factors involved in the maintenance of cell axis and the adhesion to the substratum and how the function and expression of these factors are regulated.In summary, the establishment and maintenance of cell polarity during development of monospores is under complex regulation. Dissecting the molecular mechanisms of this regulatory system could help in further understanding the interrelation among PI3K signaling, the actin-based system and cell wall formation, which can provide new insight into the machinery regulating the establishment and maintenance of cell polarity in plants.  相似文献   

2.
3.
Organelle movement in plants is dependent on actin filaments with most of the organelles being transported along the actin cables by class XI myosins. Although chloroplast movement is also actin filament-dependent, a potential role of myosin motors in this process is poorly understood. Interestingly, chloroplasts can move in any direction and change the direction within short time periods, suggesting that chloroplasts use the newly formed actin filaments rather than preexisting actin cables. Furthermore, the data on myosin gene knockouts and knockdowns in Arabidopsis and tobacco do not support myosins'' XI role in chloroplast movement. Our recent studies revealed that chloroplast movement and positioning are mediated by the short actin filaments localized at chloroplast periphery (cp-actin filaments) rather than cytoplasmic actin cables. The accumulation of cp-actin filaments depends on kinesin-like proteins, KAC1 and KAC2, as well as on a chloroplast outer membrane protein CHUP1. We propose that plants evolved a myosin XI-independent mechanism of the actin-based chloroplast movement that is distinct from the mechanism used by other organelles.Key words: actin, Arabidopsis, blue light, kinesin, myosin, organelle movement, phototropinOrganelle movement and positioning are pivotal aspects of the intracellular dynamics in most eukaryotes. Although plants are sessile organisms, their organelles are quickly repositioned in response to fluctuating environmental conditions and certain endogenous signals. By and large, plant organelle movements and positioning are dependent on actin filaments, although microtubules play certain accessory roles in organelle dynamics.1,2 Actin inhibitors effectively retard the movements of mitochondria,36 peroxisomes,5,711 Golgi stacks,12,13 endoplasmic reticulum (ER),14,15 and nuclei.1618 These organelles are co-aligned and associated with actin filaments.5,7,8,1012,15,18 Recent progress in this field started to reveal the molecular motility system responsible for the organelle transport in plants.19Chloroplast movement is among the most fascinating models of organelle movement in plants because it is precisely controlled by ambient light conditions.20,21 Weak light induces chloroplast accumulation response so that chloroplasts can capture photosynthetic light efficiently (Fig. 1A). Strong light induces chloroplast avoidance response to escape from photodamage (Fig. 1B).22 The blue light-induced chloroplast movement is mediated by the blue light receptor phototropin (phot). In some cryptogam plants, the red light-induced chloroplast movement is regulated by a chimeric phytochrome/phototropin photoreceptor neochrome.2325 In a model plant Arabidopsis, phot1 and phot2 function redundantly to regulate the accumulation response,26 whereas phot2 alone is essential for the avoidance response.27,28 Several additional factors regulating chloroplast movement were identified by analyses of Arabidopsis mutants deficient in chloroplast photorelocation.2932 In particular, identification of CHUP1 (chloroplast unusual positioning 1) revealed the connection between chloroplasts and actin filaments at the molecular level.29 CHUP1 is a chloroplast outer membrane protein capable of interacting with F-actin, G-actin and profilin in vitro.29,33,34 The chup1 mutant plants are defective in both the chloroplast movement and chloroplast anchorage to the plasma membrane,22,29,33 suggesting that CHUP1 plays an important role in linking chloroplasts to the plasma membrane through the actin filaments. However, how chloroplasts move using the actin filaments and whether chloroplast movement utilizes the actin-based motility system similar to other organelle movements remained to be determined.Open in a separate windowFigure 1Schematic distribution patterns of chloroplasts in a palisade cell under different light conditions, weak (A) and strong (B) lights. Shown as a side view of mid-part of the cell and a top view with three different levels (i.e., top, middle and bottom of the cell). The cell was irradiated from the leaf surface shown as arrows. Weak light induces chloroplast accumulation response (A) and strong light induces the avoidance response (B).Here, we review the recent findings pointing to existence of a novel actin-based mechanisms for chloroplast movement and discuss the differences between the mechanism responsible for movement of chloroplasts and other organelles.  相似文献   

4.
Cell surface receptors of the integrin family are pivotal to cell adhesion and migration. The activation state of heterodimeric αβ integrins is correlated to the association state of the single-pass α and β transmembrane domains. The association of integrin αIIbβ3 transmembrane domains, resulting in an inactive receptor, is characterized by the asymmetric arrangement of a straight (αIIb) and tilted (β3) helix relative to the membrane in congruence to the dissociated structures. This allows for a continuous association interface centered on helix-helix glycine-packing and an unusual αIIb(GFF) structural motif that packs the conserved Phe-Phe residues against the β3 transmembrane helix, enabling αIIb(D723)β3(R995) electrostatic interactions. The transmembrane complex is further stabilized by the inactive ectodomain, thereby coupling its association state to the ectodomain conformation. In combination with recently determined structures of an inactive integrin ectodomain and an activating talin/β complex that overlap with the αβ transmembrane complex, a comprehensive picture of integrin bi-directional transmembrane signaling has emerged.Key words: cell adhesion, membrane protein, integrin, platelet, transmembrane complex, transmembrane signalingThe communication of biological signals across the plasma membrane is fundamental to cellular function. The ubiquitous family of integrin adhesion receptors exhibits the unusual ability to convey signals bi-directionally (outside-in and inside-out signaling), thereby controlling cell adhesion, migration and differentiation.15 Integrins are Type I heterodimeric receptors that consist of large extracellular domains (>700 residues), single-pass transmembrane (TM) domains, and mostly short cytosolic tails (<70 residues). The activation state of heterodimeric integrins is correlated to the association state of the TM domains of their α and β subunits.610 TM dissociation initiated from the outside results in the transmittal of a signal into the cell, whereas dissociation originating on the inside results in activation of the integrin to bind ligands such as extracellular matrix proteins. The elucidation of the role of the TM domains in integrin-mediated adhesion and signaling has been the subject of extensive research efforts, perhaps commencing with the demonstration that the highly conserved GFFKR sequence motif of α subunits (Fig. 1), which closely follows the first charged residue on the intracellular face, αIIb(K989), constrains the receptor to a default low affinity state.11 Despite these efforts, an understanding of this sequence motif had not been reached until such time as the structure of the αIIb TM segment was determined.12 In combination with the structure of the β3 TM segment13 and available mutagenesis data,6,9,10,14,15 this has allowed the first correct prediction of the overall association of an integrin αβ TM complex.12 The predicted association was subsequently confirmed by the αIIbβ3 complex structure determined in phospholipid bicelles,16 as well as by the report of a similar structure based on molecular modeling using disulfide-based structural constraints.17 In addition to the structures of the dissociated and associated αβ TM domains, their membrane embedding was defined12,13,16,18,19 and it was experimentally recognized that, in the context of the native receptor, the TM complex is stabilized by the inactive, resting ectodomain.16 These advances in integrin membrane structural biology are complemented by the recent structures of a resting integrin ectodomain and an activating talin/β cytosolic tail complex that overlap with the αβ TM complex,20,21 allowing detailed insight into integrin bi-directional TM signaling.Open in a separate windowFigure 1Amino acid sequence of integrin αIIb and β3 transmembrane segments and flanking regions. Membrane-embedded residues12,13,16,18,19 are enclosed by a gray box. Residues 991–995 constitute the highly conserved GFFKR sequence motif of integrin α subunits.  相似文献   

5.
6.
Epithelial to mesenchymal transition (EMT) is a critical event in embryogenesis and plays a fundamental role in cancer progression and metastasis.1 Numb has been shown to play an important role in the proper functions of Par protein complex and in cell-cell junctions,2,3 both of which are associated with EMT.4,5 However, the role of Numb in EMT has not been fully elucidated. Recently, we showed that Numb is capable of binding to both Par3 and E-cadherin. Intriguingly, the interaction of Numb with E-cadherin or the Par protein complex is dynamically regulated by tyrosine phosphorylation induced by HGF or Src. Knockdown of Numb by shRNA in MDCK cells led to a lateral to apical translocation of E-cadherin and β-catenin, active F-actin polymerization, mis-localization of Par3 and aPKC, a decrease in cell-cell adhesion and an increase in cell migration and proliferation. These data suggest a diverse role for Numb in regulating cell-cell adhesion, polarity and migration during EMT.6Key words: Numb, E-cadherin, tyrosine kinase, cell polarity, adhesion, EMTNumb was originally identified as a gene required for cell fate determination during neuroblast division and sensory organogenesis.7 Recently, a number of proteins involved in cell polarity, cell-cell adhesion and tumorigenesis have been identified as binding partners for Numb. These include the Par3-Par6-aPKC polarity complex, E-cadherin, integrin, Notch, WNT and p53.810 Although these new data implicate Numb in multiple signaling pathways, questions remain as to how the various interactions are regulated and in which biological context they occur. Interestingly, most binding partners of Numb are involved in one way or another in the onset and/or progression of cancer. For instance, the Par complex is involved in regulating the formation and stability of tight junction whereas E-cadherin is a key component of the adherens junction in epithelial cells. Understandably, deregulation of the Par protein complex and/or E-cadherin is implicated in EMT. A variety of stimuli have been identified to induce EMT,1 including transforming growth factor-β (TGFβ), hepatocyte growth factor (HGF), fibroblast growth factor (FGF) and activation of tyrosine kinase Src. During the progression of EMT, non-motile epithelial cells gradually lose their apical-basal polarity and cell-cell junctions and become mesenchymal cells with an ability to migrate away from the primary site to surrounding tissues.1 Therefore, the study of Numb interactions with the Par protein complex and E-cadherin in the context of EMT provides a good point of entry to decode the complex signaling network mediated by Numb.Based on data obtained from Madin-Darby canine kidney (MDCK) cells,6 we propose a model in which Numb regulates epithelial polarity and cell-cell adhesion in EMT (Fig. 1). In epithelial cells, Numb binds to E-cadherin or the Par protein complex via Par3 under a normal physiological condition to stabilize adherens and tight junctions. Under the influence of an extracellular cue, such as HGF, the Met receptor recruits and activates one of its downstream components, c-Src. C-Src subsequently phosphorylates the DNVYYY motif on E-cadherin, resulting in Numb dissociation from phosphorylated E-cadherin. Numb then binds and sequesters phosphorylated aPKC and Par6, while phosphorylated Par3 is released from the Par complex. The Numb-aPKC-Par6 complex remains on the plasma membrane or in the cytoplasm, whereas Par3 is transported into the nucleus (by an unknown mechanism). Phosphorylated E-cadherin is relocated to an apico-lateral domain accompanied with active F-actin polymerization. Enhanced F-actin polymerization, together with reduced cell-cell adhesion and increased cell proliferation, promotes cell migration (Fig. 1).Open in a separate windowFigure 1A model depicting the role of Numb in epithelial-mesenchymal transition (EMT). In epithelial cells, Numb stabilizes both E-cadherin based adherens junctions through a PTB domain binding to NVYY motif on E-cadherin, and Par protein complex on tight junctions by interacting with Par3. During the early stage of EMT, c-Met recruits the tyrosine kinase c-Src upon its activation by HGF. C-Src phosphorylates E-cadherin on the NVYY motif, which leads to dissociation of Numb from E-cadherin and the translocation of E-cadherin to an apical domain. Numb forms a complex with phosphorylated aPKC and Par6, whereas phosphorylated Par3 is transported into the nucleus. Enhanced F-actin polymerization, together with reduced cell-cell adhesions, promotes the transition to mesenchymal cells.While our work in MDCK epithelial cells identifies a key role for Numb in regulating the sub-cellular localizationsand functions of E-cadherin and the Par protein complex, it raises a number of unaddressed questions. First, biochemical data obtained from temperature sensitive src mutant (ts-src) MDCK cell line suggest that tryosine phosphorylation plays a critical role in regulating the dynamic interactions of Numb with E-cadherin or the Par protein complex. Using a peptide array screen, we found that a conserved DNVYYY motif in E-cadherin is the binding site for phosphotyrosine binding (PTB) domain of Numb. Interestingly, peptide analogs in which a Tyr residue in the YYY triad is replaced by a pTyr were deficient in binding. This suggests that the Numb PTB domain may act in a phosphotyrosine-independent manner in EMT signal transduction events, and that the interaction of Numb with VIEW E-cadherin is negatively regulated by tyrosine kinase signaling. An intriguing question is why E-caderin harbors a conserved YYY triad6 when phosphorylation of one Tyr is sufficient to eliminate Numb binding. Furthermore, when and where are the YYY triad phosphorylated during EMT? It is likely that one or more tyrosine residues are phosphorylated by Src, depending on the duration of stimulus that a cell receives during EMT. Phosphorylated E-cadherin may be targeted to an alternative location in the cell by the E3 ligase Hakai which mediates the endocytosis and degradation of E-cadherin11 after its phosphorylation and dissociation from Numb. To complicate the issue, Par3 and aPKC themselves are also shown to be tyrosine phosphorylated by HGF treatment or Src activation in our studies, although the precise phosphorylation sites on either protein remain to be determined. In a previous report, a high-throughput phosphoproteomic screen has identified multiple tyrosine phosphorylation sites in the carboxyl terminus of Par-3.12 Phosphorylation of Y1127 in Par-3 reduced its interaction with LIMK2 and delayed tight junction assembly in mammalian epithelial cells with a constitutive Src activation or under EGF treatment.12 A recent study, which shows that the SH2 domain of C-terminal Src kinase (Csk) binds to the Y1127 site on Par3, adds another layer of complexity to the scenario.13 In the same vein, Src has been shown to associate with aPKC and phosphorylate the latter in PC12 cells.14 Par6 appears also to be subjected to Src regulation as recruitment of Src by Pals leads to Pals1 and Par6 to be sequestered away from JAM-C of desmosome adhesions in the blood-testis-barrier, causing a disruption in cell adhesion.15 It will be of interest to determine whether Par6 is directly phosphorylated by Src in this case.15 Together, these studies suggest a role for tyrosine phosphorylation, besides serine/threonine phosphorylation, in regulating the functions of various polarity proteins.Second, our biochemical and immunfluorescence data suggest that the interactions of Numb with E-cad/Par complex are spatially and temporally regulated in response to HGF treatment. However, the molecular mechanisms for some remarkable phenotypes remain poorly defined. For instance, shRNA-mediated Numb knockdown caused a dramatic apicolateral mis-localization of E-cadherin and β-catenin. Dynamic cellular localization may be regulated by endocytosis and post-translational modifications such as phosphorylation. In keeping with this, Numb can directly bind to several endocytic proteins, including AP2 and Eps15, through its conserved DPF and NPF motifs.16 Our data suggest that Numb is responsible for targeting E-cadherin to correct localization on the basolateral membrane, but we cannot rule out the possibility that other proteins may be involved in this process. One candidate that may facilitate E-cadherin localization is Rab11. Previous studies revealed that in newly polarized MDCKII cells, Rab11 mutation causes an apical mis-localization of E-cadherin and aberrant actin localization, while leaving ZO-1 localization unchanged. This mirrors the phenotype we observed in the Numb shRNA cells.17 In support of the notion that Rab11 and Numb may be functionally related, Rab11 and Numb are segregated selectively in the pIIb daughter cell during the division of a sensory organ precursor (SOP).18 It is likely that the mis-localization of E-cadherin is caused by an abnormal enodocytosis of E-cadherin in the absence of Numb.In contrast, the Par protein complex provides critical spatial information in the formation of tight junctions. Treatment of MDCK cells by HGF caused Par3 to dissociate from Par6-aPKC, suggesting that the Par3-Par6-aPKC complex19 is dynamically regulated. Additionally, both Par3 and aPKC are found to translocate to the nucleus following HGF treatment. A key question, therefore, is how the localizations of these polarity proteins themselves are regulated during EMT. Through a PB1-PB1 domain heterodimarization,20 Par-6 binds to aPKC and thereby inhibits its catalytic. Moreover, Par6 also helps recruit substrates to aPKC,20 one of which is Par3. Par6 binds to Par3, through a PDZ-PDZ interaction, but aPKC can also bind directly to Par3 through its kinase domain, and phosphorylates a Ser residue on Par3.21 Importantly, depending on the cellular context, Par3, Par6 and aPKC may not form a constitutive complex.20 For example, in Drosophila neuroblasts and embryonic epithelial cells, Par3 apical localization is independent of aPKC,22 while in mammalian epithelial cells, Par3 is not apical but is associated with tight junctions.23 Interactions of these polarity proteins are dynamically regulated by multiple protein kinases, small GTPases, or competition from other binding partners.19 The dynamic change of components of the Par complex leads to the different sub-cellular distribution of these individual polarity proteins, and thereby alters their functions.19 We showed that activation of Src kinase or HGF treatment reduces the association of Par3 with aPKC, but does not change the binding between aPKC and Par6. In support of this, a similar observation is made in MDCK cells with activation of tyrosine kinase ErbB2, although neither Par3 nor Par6 is a substrate of ErbB2.24 The dynamic assembly and dissolution of the Par3-Par6-aPKC complex in response to intracellular cues or extracellular stimuli may be a general mechanism used by metazoan to control cell polarity and movement.Lastly, active F-actin polymerization is one of most striking morphologies we observed in the Numb-shRNA cells. Given that the Numb-shRNA cells exhibit a faster rate of migration and wound healing than control cells, it is likely that the active F-actin polymerization is caused by aberrant activation of the Rho family GTPase activity. Small GTPases of the Rho family control organization of cytoskeleton, cell motility, cell growth, morphogenesis, cytokinesis and trafficking.25 The most common members of small Rho GTPases are RhoA, Rac1 and Cdc42. RhoA is responsible for the activation of stress fibers and cell contractility.25 Rac1 activation leads to polymerization of filamentous actin, which results in lamellipodium formation and membrane ruffling at the leading edge of migrating cells.25 Cdc42 activation causes the formation of filopodia, long finger-like protrusions at the edges of lamellipodia.25 A Rho family of GTPase exists in two states: a GDP-bound inactive state and a GTP-bound active state. The switch between the two states is regulated by a large group of guanine nucleotide exchange factors (GEFs) and GTPase activating proteins (GAPs).25 GEFs and GAPs typically contain interaction domains which direct the enzymes to specific sub-cellular locations and help recruit upstream/downstream partners to affect processes such as actin cytoskeleton, cell polarity, microtubule dynamics and membrane transport.25 It is often difficult to pinpoint exactly which member of the Rho family GTPases leads to a specific phenotype due to the overlapping functions of the Rho GTPase members. However, several observations have suggested that certain small Rho GTPases are essential for the establishment of the apico-basal polarity and likely interplay with Numb. Deletion of Cdc42 abolished normal localization of aPKC, Par6 and Numb in neuroepithelium.26 Par6 binds to GTP-bound Cdc42 through a Rho GTPase-bindingCdc42/Rac-interactive binding (CRIB) domain.23,27 Another line of evidence is that Numb binds to intersectin, a Cdc42 GEF, and enhances the GEF activity of intersectin, leading to activation of Cdc42 in vivo.28 It appears that Numb both regulates and is regulated by Cdc42 temporally and spatially. Par3 has also been reported to sequester Tiam1, a Rac GEF, to inhibit Rac activation in hippocampal neurons.29 Nevertheless, Par3 recruits Tiam1 to activate Rac in keratinocytes.30 Thus, the dynamic interplay between Rho GTPases and the Par protein complex may be cell type-dependent. It will also be of interest in the future to determine whether other GEFs/GAPs interact with the Par protein complex, and whether and how these interactions are modulated by Numb.A more detailed understanding of the role of Numb in EMT will undoubtedly provide valuable insights into the molecular basis of cancer metastasis and suggest novel therapeutic strategies for cancer.  相似文献   

7.
Cell motility is a highly coordinated multistep process. Uncovering the mechanism of myosin II (MYO2) activation responsible for the contractility underlying cell protrusion and retraction provides clues on how these complementary activities are coordinated. Several protein kinases have been shown to activate MYO2 by phosphorylating the associated myosin light chain (MLC). Recent work suggests that these MLC kinases are strategically localized to various cellular regions during cell migration in a polarized manner. This localization of the kinases together with their specificity in MLC phosphorylation, their distinct enzymatic properties and the distribution of the myosin isoforms generate the specific contractile activities that separately promote the cell protrusion or retraction essential for cell motility.Key words: myosin, MLCK, ROK, MRCK, phosphorylation, cell migrationCell movement is a fundamental activity underlying many important biological events ranging from embryological development to immunological responses in the adult. A typical cell movement cycle entails polarization, membrane protrusion, formation of new adhesions, cell body translocation and finally rear retraction.1 A precise temporal and spatial coordination of these separate steps that take place in different parts of the cell is important for rapid and efficient movement.2One major event during eukaryotic cell migration is the myosin II (MYO2)-mediated contraction that underlies cell protrusion, traction and retraction.1,3 An emerging theme from collective findings is that there are distinct myosin contractile modules responsible for the different functions which are separately regulated by local myosin regulatory light chain (MLC) kinases. These kinases contribute to contractile forces that connect adhesion, protrusion and actin organization.2 Unraveling the regulation of these contractile modules is therefore pivotal to a better understanding of the coordination mechanism.At the lamellipodium, the conventional calcium/calmodulin-dependent myosin light chain kinase (MLCK) has been shown to play an essential role in a Rac-dependent lamellipodial extension.4 Inhibition of calmodulin or MLCK activity by specific photoactivatable peptides in motile eosinophils effectively blocks lamellipodia extension and net movement.5 Furthermore, there is a strong correlation between activated MLCK and phosphorylated MLC within the lamellipodia of Ptk-2 cells as revealed by fluorescence resonance energy transfer (FRET) analysis.6 More recent studies showed MLCK to regulate the formation of focal complexes during lamellipodia extension.7,8 Functionally, MLCK is thought to play a critical role in the environment-sensing mechanism that serves to guide membrane protrusion. It mediates contraction that exerts tension on integrin-extracellular matrix (ECM) interaction, which, depending on the rigidity of the substratum, will lead to either stabilization of adhesion resulting in protrusion or destabilization of attachment seen as membrane ruffling on non-permissive surfaces.8,9As a Rho effector, Rho-associated kinase (ROK/ROCK/Rho-kinase) has been shown to regulate stress fibers and focal adhesion formation by activating myosin, an effect that can be blocked by the specific ROK inhibitor Y-27632.10,11 Myosin activation by ROK is the effect of two phosphorylation events: the direct phosphorylation on MLC and the inhibition of myosin phosphatase through phosphorylation of its associated myosin-binding subunit (MBS).11 Consistent with this notion of a localization-function relationship, ROK and MBS, which can interact simultaneously with activated RhoA,11 have been shown to colocalize on stress fibers.12,13 In migrating cells, Rho and ROK activities have been mostly associated with the regulation of tail retraction, as inhibition of their activities often results in trailing tails due to the loss of contractility specifically confined to the cell rear.14,15 Tail retraction requires high contractile forces to overcome the strong integrin-mediated adhesion established at the rear end, an event which coincides with the strategic accumulation of highly stable and contractile stress fibers that assemble at the posterior region of migrating cells.MRCK was previously shown to phosphorylate MLC and promote Cdc42-mediated cell protrusion.16 More recently, it was found to colocalize extensively with and regulate the dynamics of a specific actomyosin network located in the lamella and cell center, in a Cdc42-dependent manner but independent of MLCK and ROK.17 The lamellar actomyosin network physically overlaps with, but is biochemically distinct from the lamellipodial actin meshwork.9,18 The former network consists of an array of filaments assembled in an arrangement parallel to the leading edge, undergoing continuous retrograde flow across the lamella, with their disassembly occurring at the border of the cell body zone sitting in a deeper region.1719 Retrograde flow of the lamellar network plays a significant role in cell migration as it is responsible for generating contractile forces that support sustained membrane protrusion and cell body advancement.1719It is therefore conceivable that these three known MLC kinases are regulated by different signaling mechanisms at different locations and on different actomyosin contractile modules. The coordination of the various modules will ensure persistent directional migration (Figure 1). Phosphorylation of MLC by PAK and ZIP kinase has also been reported, but their exact roles in this event have yet to be determined.20,21 It is also noteworthy that individual kinases can work independently of each other, as amply shown by evidence from inhibitor treatments. This is particularly true for MRCK in the lamella, whose activity on lamellar actomyosin flow is not affected by ML7 and Y-27632, the inhibitors of MLCK and ROK respectively.17 These findings further indicate that although both ROK and MRCK have been shown to upregulate phosphorylated MLC levels by inhibiting the myosins phosphatases,11,22 they are likely to act as genuine MLC kinases themselves, without the need of MLCK as previously suggested.11Open in a separate windowFigure 1Upper panel depicts a model for the specific activation of the different MLC kinases at various locations in the cell. In response to upstream signals, MLC kinases MLCK, MRCK and ROK are activated and localized to different regions. In the case of MRCK and ROK, the interaction of the GTP-bound Rho GTPase binding domain will determine the specific action of the downstream kinase, resulting in actomyosin contractility at different locations. The coordination of these signalling events is crucial for directional cell migration. Lower panel shows a typical front-rear location for Myosin 2A and 2B in a migrating U2OS cell.In conjunction with their differences in localization, the three MLC kinases show apparent individual preferences and specificity towards the MYO2 isoforms that they associate with. The two major MYO2 isoforms MYO2A and 2B are known to have distinct intracellular distributions that are linked to their individual functions (Figure 1).23,24 In motile cells, MYO2A localization that is skewed towards the protruding cell front is consistent with it being the major myosin 2 component of the lamellar filaments regulated by MRCK as well as its regulation by MLCK in lamellipodial contraction.8,17,19 In contrast, the enrichment of MYO2B at retracting cell rear conforms well with the requirement of thick and stable stress fibers capable of causing tail contraction and prevention of protrusion under the control of Rho/ROK signaling.23,25 The selection for MYO2B filaments in the cell rear stems from their more contractile and stable nature compared with MYO2A, a consequence of their higher time-averaged association with actin.26,27 Conversely, the lower tension property of MYO2A filaments suggests that they are more dynamic in nature,26,27 a characteristic which fits well with the dynamic actomyosin activities at the leading edge and lamella that regulate protrusion.It deserves special mention that the three MLC kinases display subtle differences in their specificity towards MLC. While MLCK and MRCK phosphorylate only a single Ser19 site (monophosphorylation),18,28 ROK is able to act on both Thr18 and Ser19 residues causing diphosphorylation of MLC,29 MLCK only causes diphosphorylation when present at higher concentrations.30 By further increasing its actin-activated ATPase activity, diphosphorylation of MLC has been shown to induce a higher myosin activation and filament stability.3032 The use of specific antibodies that can differentiate between the two populations of phosphorylated MLC has been instrumental in revealing their localization and correlation with the activity of the MLC kinases. The emerging picture from these experiments is that mono and diphosphorylated MLC exhibit distinct distributions in migrating cells, with the monophosphorylated MLC localized more towards the protrusive region, while the diphosphorylated form is more enriched at the posterior end.21,33 Taking into account their biochemical properties, the polarized distributions of these differentially phosphorylated MLC coincide functionally with the segregation of the MYO2 isoforms and their corresponding regulators. These findings provide further support for the existence of segregated contractile modules in migrating cell and their distinctive regulation.The mechanisms that determine the specific segregation of the contractile modules and their regulation are unclear. However, some clues have emerged from recent studies. It has been shown that the C-terminal coiled-coil region of MYO2B is important for determining its localization in cell rear25 and which requires Rho/ROK activity as their inhibition resulted in the loss of this specific localization.23 Correspondingly, the inhibition of MRCK activity resulted in the loss of lamella-localized MYO2A.17 These findings suggest that activation of MYO2 filaments by their upstream regulators is important for their functional segregation and maintenance. It is noteworthy that both ROK and MRCK have distinct regulatory domains including the pleckstrin homology domains which have been shown to be essential for their localization, a process which may involve myosin interaction and lipid-dependent targeting as has been respectively shown for ROK and MRCK.11,13,16 Further, the specificity of MRCK for lamellar actomyosin is believed to be largely determined by the two proteins it forms a complex with: the adaptor LRAP35a, and the MYO2-related MYO18A. Activation of MYO18A by MRCK, a process bridged by LRAP35a, is a crucial step which facilitates MRCK regulation on lamellar MYO2A.17The mechanisms responsible for segregating the contractile modules and their regulators may also comprise a pathway that parallels the microtubule-modulatory Par6/aPKC/GSK3β signalling pathway which regulates cellular polarization. This notion is supported by both Cdc42 and Rho being common upstream regulators of these two pathways.34 GTPase activation may determine the localized activities of the separate contractile modules and create an actomyosin-based asymmetry across the cell body, which together with the microtubule-based activities, result in the formation of a front-back axis important for directional movement. The involvement of MRCK in MTOC reorientation and nuclear translocation events,35 and our unpublished observation that LRAP35a has a GSK3β-dependent microtubule stabilizing function are supportive of a possible cross-talk between these two pathways.In conclusion, the complex regulation of contractility in cell migration emphasizes the importance of the localization, specificity and enzymatic properties of the different MLC kinases and myosin isoforms involved. The initial excitement and confusion caused by the emergence of the different MLC kinases are fading, being now overtaken by the curiosity about how they cooperate and are coordinated while promoting cell motility.  相似文献   

8.
Directional cell migration is essential for almost all organisms during embryonic development, in adult life and contributes to pathological conditions. This is particularly critical during embryogenesis where it is essential that cells end up in their correct, precise locations in order to build a normal embryo. Many cells have solved this problem by following a gradient of a chemoattractant usually secreted by their target tissues. Our recent research has found an alternative, complimentary, mechanism where intracellular signals are able to generate cell polarity and directional migration in absence of any external chemoattactant. We used neural crest cells to study cell migration in vivo, by performing live imagining of the neural crest cell migrating during embryo development. We show that the Planar Cell Polarity (PCP) or non-canonical Wnt signaling pathway interacts with the proteoglycan syndecan-4 to control the direction in which cell protrusions are generated, and in consequence, the direction of migration. By analyzing the activity of the small GTPases using in vivo FRET imaging we showed that PCP signaling activates RhoA, while syndecan-4 inhibits Rac, both at the back of the neural crest cell. Here we discuss a model where these signals are integrated to generate directional migration in vivo.Key words: directional migration, cell migration, syndecan-4, PCP, non-canonical Wnt, neural crest, RhoA, RacThe ability of cells to move in a directed manner is a fundamental requirement for life. In multi-cellular organisms, this requirement begins in the embryo, where morphogenetic processes are dependent on the correct movement of large numbers of cells. In the adult too, cell migration plays a vital role in many systems including the immune system and wound healing. Cell migration defects can contribute to the pathology of many diseases including vascular diseases such as atherosclerosis, and chronic inflammatory diseases like asthma and multiple sclerosis. Likewise, metastasis in cancer is characterized by mis-regulation of the normal cell migration machinery and results in cells that are normally static becoming aggressively motile and invasive.Cell migration requires cell polarization and the formation of protrusions at one end of the cell. Polarization results in a different molecular ensemble at the front of the cell compared to that at the back. Cell protrusion formation at the front of the cell requires reorganization of the actin and microtubule cytoskeleton to produce a protrusion either in the form of a broad sheet-like lamellipodium or spiky filopodium. Small GTPases are well known modulators of these processes (reviewed in ref. 1).Several mechanism has been proposed as involved in directional migration during embryo development, such as chemotaxis (migration toward an soluble chemoattractant),2 haptotaxis (migration toward a substrate-bound chemoattractant),3 population pressure (migration from a region of high towards a region of low cell density)4 and contact inhibition of locomotion (change in the direction of migration as a consequence of cell-cell contact),5 being chemotaxis the most widely accepted and studied.The correct orientation of the cell and its protrusion is the keystone of directional migration and, in the case of chemotaxis, it is supposed to be controlled by the action of external chemical cues (chemoattractants) that are produced by or near to the target tissue.6 One of the best examples for chemoattraction in vivo is the migration of the progenitor germ cells, which are attracted by the chemokine SDF-1.2 It has been shown in vitro and in vivo, that upon receiving a chemotactic signal, the cell becomes polarized in the direction of migration. Nevertheless, it is known that cells cultured in vitro can became polarized and exhibit directional migration in absence of extrinsic chemoattractants.7 Pankov et al. showed that persistent directional migration in vitro can be achieved solely by modulating the activity of the small GTPase, Rac: high levels of Rac promotes the formation of peripheral lamella during random migration, while slightly lower levels of Rac suppress peripheral lamella and favour the formation of a polarized cell with lamella just at the leading edge.7 Is it possible that a similar mechanism of directional migration could occur in vivo?The migration of Neural Crest (NC) cells has been used as a model to study directional cell migration in vivo.810 The neural crest is an embryonic population of cells that are specified at the border between the neural plate and the epidermis.11 Upon induction neural crest cells undergo an epithelial to mesenchymal transition,12 detach from the neural tube and migrate following defined pathways that eventually allow them to colonize almost the entire embryo.13 Finally, after reaching their destination NC cells differentiate to form many different cell types including neurons, glia, cartilage, skeleton and pigment cells.14 The migration of the NC cells is critical for the proper differentiation of their derivatives and there are several human syndromes associated with failures in this process.The migration of NC cells is a highly ordered process; individual NC cells migrate with high persistence towards the direction of their targets,8 but until now it was not known how this directionality is controlled. A number of molecules have been identified as key players in neural crest migration, such as Ephrins, Semaphorins, Slit/Robo, etc. (reviewed in ref. 13). However most of these molecules work as inhibitory signals, which are required to restrict the migration of NC cells from prohibited areas. Although chemoattraction has been one of the proposed mechanisms to explain this directional migration, no chemoattractant has thus far been found in the NC.It has been known for many years that NC cells can migrate in vitro with a high directionality even in the absence of external signals.15 Therefore, our work has been focused on understanding how NC directionality is controlled. Recently, we have unveiled some of the molecules that control this directional migration in vitro. More importantly, we have been able to show that the same molecular machinery controls directional migration in vivo.9,10One of the key factors that controls directional migration of NC cells is the Planar Cell Polarity (PCP) or non-canonical Wnt signaling pathway.9,10,16 PCP signaling was first described in Drosophila, where a number of mutations were identified that disrupt the formation of bristles and hairs on the adult cuticle.17 In the Drosophila wing, epithelial cells are highly polarized, with a single hair outgrowth forming at the distal end of each cell. Mutations in PCP genes cause loss in cell polarity in this tissue with hairs forming in a disorganized pattern.18 In vertebrates, PCP signaling also regulates cell polarity during a number of different developmental processes including neural tube closure, cochlear hair orientation and ciliogenesis.19We have shown that the PCP pathway is essential for correct neural crest migration in Xenopus. Injection of dominant negative forms of the intracellular PCP component Dishevelled (Dsh), which inhibit the PCP pathway but not canonical Wnt signaling, block the migration of cranial neural crest cells in vivo.9 Recently this role has also been extended to zebrafish where directional migration of neural crest is severely disrupted in the PCP mutant trilobite (strabismus) and in embryos injected with a dominant negative form of Dsh or a morpholino against wnt5a,10 with no effect in neural crest cell motility.9,10 Two factors, pescadillo and syndecan-4 that have recently been proposed as modulators of the PCP signaling,20,21 are also required for NC migration.10,21 Taken together, these data point to an essential role for PCP signaling in neural crest migration.What is the cellular and molecular mechanism by which PCP signaling controls migration of NC cells? In order to investigate this question we analyzed the direction of neural crest migration and cell polarity in vitro and in vivo after interfering with two elements of the PCP signaling pathway: syndecan-4 and Dsh. One of the key finding of our work was that the inhibition of NC migration through syndecan-4 depletion does not affect the velocity of cell migration, but significantly reduces the directional migration of the cells in vivo (Fig. 1A and B). Consequently, when the orientation of cell protrusions was analyzed we found that syndecan-4 depletion does not affect the formation of cell protrusions, but the direction in which the cell protrusions are generated during migration. More precisely, normal cells extend their lamellipodia at the front of the cell (Fig. 1D), while cells where syndecan-4 is inhibited generate protrusion in all directions (Fig. 1E). A similar analysis was performed for embryos expressing a mutated form of Dsh that works as a dominant negative of PCP signaling and an equivalent effect on directional migration and the orientations of cell protrusions was observed (Fig. 1C and F).Open in a separate windowFigure 1Directional migration of neural crest cells. (A and B) Example of track of a single cell migrating in vivo. (A) Control cell showing persistent directional migration. (B) Cell in which the PCP signaling has been inhibited, showing absence of directional migration. (C) Cell in which syndecan-4 has been inhibited, showing no persistent migration. (D–F) Analysis of cell polarity and model of directional migration. Fn: fibronectin; Syn4: syndecan-4. (D) Control cell. Activation of Fn/Syn4 and PCP/RhoA lead to inhibition of Rac at the back of the cell, with the consequence polarization and directional migration. (E) Inhibition of PCP signaling leads to absence of RhoA activity, and in consequence an increase of Rac activity at the back of the cell. It seems that the inhibition of Rac activity by Syn4 is not sufficient to keep low levels of Rac at the back of the cells. High levels of Rac at the back produce a loss in cell polarity and in directional migration. (F) Inhibition of Syn4 generates high levels of Rac activity by a double mechanism: absence of direct inhibition of Rac and absence of RhoA which is dependent on PCP signaling. High levels of Rac at the back produce a loss of cell polarity and directional migration.As cell protrusions are known to be controlled by small GTPases and as PCP and syndecan-4 signaling regulates the activities of small GTPases,18,22 we analyzed the activity of cdc42, RhoA and Rac after interfering with Dsh and syndecan-4. We choose to perform FRET analysis of these molecules as it is a technique that allows the visualization of their localized activity. More interestingly we succeeded in performing FRET analysis in cells migrating in vivo for the first time. Our results show that syndecan-4 inhibits Rac activity, while Dsh signaling promotes RhoA activity. In addition, we show that RhoA inhibits Rac in neural crest cells.10 The regulation of Rac by syndecan-4 is similar to that seen in other cells types in vitro.23,24The model that emerges from these results to explain directional migration of NC cells in vivo is as follows (Fig. 1D). After delamination NC cells come into contact with fibronectin in the extracellular matrix, which is known to provide the main substrate for neural crest cells during their migration.25,26 The interaction of fibronectin with syndecan-4 leads to two major changes in the cell: activation of PCP signaling and inhibition of Rac activity. The activated PCP signaling becomes localized at the back of the cell. From here, PCP contributes to the inhibition of Rac at the back of the cell, through the activation of RhoA. The coordinated activities of syndecan-4 and PCP signaling lead to polarised Rac activity across the cell, with Rac enriched at the leading edge, where it promotes the polymerization of actin and formation of lamellipodia, resulting in directional migration (Fig. 1D). Inhibition of PCP signaling produces high levels of Rac all over the cell as Rac, an inhibitor of RhoA in many cell types including neural crest cells, is absent (Fig. 1E). This generates cell protrusions in all directions with the consequent loss of cell polarity. If syndecan-4 is absent, the levels of Rac activity are also high all over the cell as the inhibition of Rac by syndecan-4 is absent (Fig. 1F), which also leads to a loss of cell polarity.Although detailed study of the localized activity of small GTPases has not been performed for other migratory cells in vivo, it is likely that the machinery will be similar to the one described here for NC cells. For example, it is well established in Xenopus, zebrafish and chick embryos that the migration of mesodermal cells during gastrulation requires PCP signaling.2729 It has also been shown that gastrulation in Xenopus20 and in zebrafish (unpublished observations) requires the activity of syndecan-4. Thus, it is expected that cell polarity established during the migration of mesodermal cells will be dependent on small GTPases controlled by non-canonical Wnt signaling and syndecan-4.This novel integrated view of PCP, syndecan-4 and small GTPase activity during directional cell migration in vivo is an important advance in our knowledge of cell migration. Nevertheless, how the PCP signaling becomes activated only at the back of the cell, is a key question that needs to be answered. Future studies will be necessary to solve this and other crucial problems.  相似文献   

9.
Growth factors regulate a diverse array of cellular functions including proliferation, survival and movement, and the ability to do this often involves interactions with the extracellular matrix (ECM) and particularly heparan sulfate proteoglycans (HSPGs). HSPGs have been shown to sequester growth factors and to act as growth factor co-receptors or receptors themselves. Recent studies, however, have revealed a new role for HSPGs in mediating the interactions of growth factors with the ECM. Specifically, heparan sulfate has been shown to modulate fibronectin structure to reveal previously masked growth factor binding sites. In vivo, this mechanism appears to control the guidance of migrating cells during embryonic development as HSPG-modification of fibronectin enables direct platelet derived growth factor-fibronectin interactions necessary for this process. A model based on this observation is discussed here as well as the possibility that other growth factors/morphogens utilize similar mechanisms involving fibronectin or additional ECM proteins.Key words: heparin, heparan sulfate, fibronectin, cell movement, cell motility, mesoderm, mesendoderm, embryonic development, embryo, XenopusThe ability of cells to move is critical to normal processes, such as embryonic development and wound healing, and is also a feature of pathological conditions such as cancer metastasis. In normal circumstances, cell movement is tightly regulated and leads to precise cell reorganizations. This is particularly evident during embryonic development where cells need to be in the right place at the right time to give and receive signals, to form tissues and organs and ultimately, to become a functional organism. To achieve this, cells must be able to move in a directed way. This requires that cells sense and move in response to extracellular signals, while coordinating changes in shape with the creation of traction and force and balancing attachment and detachment to the extracellular matrix (ECM) and neighboring cells.Evidence from a number of systems suggests that growth factors play a significant role in the spatial and temporal control of cell movements. In particular, members of the platelet derived growth factor (PDGF)/vascular endothelial growth factor (VEGF) family act as guidance cues for cells in a variety of processes including the development of both invertebrate and vertebrate embryos (reviewed in ref. 1). How such cues are laid down is an area of intensive investigation; however, a new study from our group suggests that the microenvironment within the ECM, principally the composition of heparan sulfate, modulates direct interactions between PDGF and fibronectin necessary for directed migration of embryonic cells during Xenopus development.2During the development of all vertebrates, cell movements begin at gastrulation, when changes in the shape, adhesion and motility of cells drive the complex series of tissue rearrangements necessary to establish the basic body plan of the embryo. In Xenopus embryos, some of the first cells to move during gastrulation are the anterior mesendoderm cells. These cells migrate in a characteristic pattern that provides a tractable model to identify the molecular control of directed cell movement. Anterior mesendoderm cells move away from the blastopore lip and towards the animal pole across a fibrillar, fibronectin-rich ECM that covers the ectoderm (see Fig. 1A; reviewed in ref. 3). They migrate as a sheet and are polarized with lamelliform protrusions extending in the direction of movement and the trailing edges underlapping one another like shingles on a roof (reviewed in ref. 3). This mesendoderm movement is guided by PDGF-A signaling.4 Immediately prior to and during Xenopus gastrulation, PDGF-A is expressed in the ectoderm while its receptor, PDGFRα, is expressed in the migrating mesendoderm cells (reviewed in ref. 5). Disruption of PDGF-A signaling by interference with either the receptor or ligand in vivo or in an ex vivo assay disorients the cells and the direction of migration is randomized, although the cells are still able to physically move.4Open in a separate windowFigure 1Heparan sulfate modification of fibronectin enables PDGF-AA-fibronectin interactions necessary for directed cell migration during Xenopus gastrulation. (A) A schematic of a Xenopus embryo immediately after the onset of gastrulation (stage 10+). The prospective mesoderm (red), ectoderm (blue) and endoderm (yellow) are indicated. The black circle indicates migrating anterior mesendoderm cells and the region of tissue shown in (B and C). (B) In vivo the inner surface of the ectoderm (blue blocks) is lined with a fibronectin-rich ECM. This fibronectin is proposed to be in its open conformation (blue dumbbells) due to the presence of heparan sulfate proteoglycans on the surface of these cells (indicated in red). PDGF (green triangle) secreted by the ectoderm cells binds to the C-terminus of the fibronectin. This PDGF is then available for PDGFRα (green fork) binding and activation. The PDGFRα is expressed on the migrating mesendoderm cells. This ligand receptor interaction guides the directed movement of these cells. (C) Heparinase III treatment of these tissues during matrix deposition digests the heparan sulfate side chains leaving the proteoglycan core (red line), which does not alter fibronectin structure leaving it in the closed conformation. PDGF secreted by the ectoderm cells can then diffuse from the site of secretion leaving it available to bind PDGFRαs over the entire surface of the mesendoderm cells and thus, disrupting directed movement. (D and E) A schematic of a “higher magnification” view of one cell in (B and C). A key of the symbols used in (B–D) is shown.PDGF-A must be associated with the ECM for directed mesendoderm migration. In an ex vivo assay, overexpression of the long form of PDGF-A, which associates with the ECM via a positively charged carboxy-terminal retention motif (reviewed in ref. 6) in the ectoderm cells, disrupts the normal guidance cues causing randomization of mesendoderm movement.4 In contrast, overexpression of the short form of PDGF-A, which diffuses from its site of secretion,6 does not.4 Our recent work now suggests that this growth factor-ECM association represents a direct interaction between PDGF-A and fibronectin.2 Importantly, this interaction requires that the fibronectin is pre-exposed to heparan sulfate proteoglycan (HSPG) before PDGF-A can bind.2 A similar interaction between the related growth factor VEGF and fibronectin is also enhanced by pre-treatment of the fibronectin with heparin or heparan sulfate.79 Heparin acts by modifying the structure of fibronectin,8 mediating the transition from the globular form to a stable, more extended conformation,7,8,1012 which reveals growth factor binding sites.7,8 These data are significant because in an ex vivo directed migration assay, treatment of the native ECM with heparinase III, which degrades heparan sulfate chains, abolishes the guidance cues and results in randomized mesendoderm migration.2 Here we propose a model in which prior to and during Xenopus gastrulation, fibronectin is modified to an extended conformation by endogenous HSPGs as it is being laid down (Fig. 1B and D). PDGF-AA secreted from the ectoderm cells binds to this modified fibronectin and is thus, retained in the ECM (Fig. 1B and D). PDGF-AA then guides the mesendoderm cells by local activation of their cell surface PDGFRαs (Fig. 1B and D). In the absence of HSPG, the fibronectin remains in a more globular form and PDGF-AA binding sites are shielded (Fig. 1C and E). In this case, it is proposed that PDGF-AA diffuses away from the site of secretion resulting in a less localized activation of PDGFRαs and a loss of persistent cell migration (Fig. 1C and E). This idea is supported by data showing that flooding the native ECM with exogenous PDGF-AA prior to performing the ex vivo migration assay, also causes a randomization of directed movement.4 In this case, it seems likely that any gradient or uneven distribution of PDGF-AA is overwhelmed and the guidance cues consequently lost.In cultured cells, PDGF can also associate with HSPGs via its retention motif.13 However, in the Xenopus ex vivo directed migration assay, heparinase treatment of the ectoderm ECM after its deposition does not alter the direction of mesendoderm migration.2,14 This suggests that in this system, PDGF-AA interacts directly with fibronectin and not with heparan sulfate chains, since heparinase III would have degraded them and caused the release of HS-bound PDGF-AA.13The type of proteoglycan as well as the specific structural composition of their heparan sulfate chains likely contributes to the modulation of PDGF-fibronectin binding. Both chain length and the chemical composition of heparan sulfate play crucial roles in the binding of VEGF to fibronectin, in which long (>22 saccharides) heparin chains with sulfation on the 6-O and N positions of glucosamine units are required for activity.8 This suggests that a localized control of heparan sulfate biosynthesis through the regulation of the array of enzymes involved might be a critical component to directed cell migration. Interestingly, an analysis of heparan sulfate expression during mouse lung development noted dynamic and discrete localization within the mesenchyme at sites of FGF10-mediated epithelial budding,15 indicating that rapid alterations of heparan sulfate structure may provide a general mechanism for positional control of growth factor activity.Several HSPGs are expressed in tissue specific patterns during Xenopus gastrulation including Syndecans-1 and -2 (expressed in the ectoderm);16 and Syndecan-4, Glypican-4 and Biglycan (expressed in the ectoderm and mesoderm).1619 Although the role(s) of these HSPGs in mesendoderm migration are not known, they do affect embryological processes that involve cell rearrangements. For example, Syndecan-2 has recently been shown to regulate the migration of organ primordia and fibrillogenesis in zebrafish embryos.20 During gastrulation, knock-down of either Syndecan-1 or -2 disrupts the native ECM for mesendoderm migration and results in gaps in fibril deposition.21 In addition, Syndecan-4 and Glypican-4 are both essential for convergent extension, another important form of cell movement during gastrulation that involves the coordinated rearrangement of cells.17,18 Similarly, Syndecan-4 is necessary for the directional migration of neural crest cells.22 In these cases, Syndecan-4 and Glypican-4 regulate the planar cell polarity (PCP) pathway, which is also involved in polarized matrix deposition during convergent extension.23 Compellingly, embryos develop with anterior defects when either Syndecan-4 or Glypican-4 is knocked-down, indicating that mesendoderm cell migration is also likely to be affected.17,18The Slb/Wnt-11 mediated-PCP pathway has been shown in zebrafish embryos to regulate the polarity and directional movement of ingressing mesendoderm cells.24 In Slb/Wnt-11 mutants, however, although these processes are disrupted they are not completely abolished and evidence suggests that PDGF signaling through phosphatidyl inositol 3-kinase (PI3K) is responsible for establishing cell polarity and controlling the velocity of these cells.25 Rho GTPases act downstream in both PCP and PDGF signaling pathways. Recent evidence suggests that RhoA and Rac1 are important for the polarity and protrusive activity of migrating mesendoderm cells and play a role in the guidance of neural crest cells,22,26 but the mechanisms by which these pathways are integrated with signals from other factors that also modulate these processes are still to be determined.How retention of PDGF-AA in the ECM translates into directional migration still needs to be determined. For example, PDGF-AA may be present in a gradient such that receptors at the front of each cell are exposed to a different level of PDGF than those at the rear. Alternatively, control might involve selective expression of molecules that locally release PDGF from matrix binding sites. Either process might provide a means for uneven activation of PDGFRs and a polarized distribution of downstream signaling components at the leading and lagging edges. Such signaling mechanisms have been identified in other cell types and organisms including leukocytes, fibroblasts and Dictyostelium in which the initial response to a chemoattractant creates cell polarity that is further amplified by feedback (reviewed in ref. 27). For example, in Dictyostelium PI3K and PTEN (phosphatase and tensin homolog) are regulated reciprocally to control the spatial and temporal levels of phosphatidyl inositol-3,4,5-triphosphate (PI(3,4,5)P3). PI(3,4,5)P3 is localized to the leading edge of the cell in response to an extracellular signal, whereas PTEN is downregulated in these regions. Recruitment of proteins that bind preferentially to PI(3,4,5)P3, many of which affect remodeling of the actin cytoskeleton, further enhances this polarity in the direction of the chemoattractant.27 In Xenopus, PDGF signaling does appear to regulate cell polarity since overexpression of PDGF-A in the ectoderm disorients the protrusive activity and abolishes the normal shingle arrangement of the receptor expressing, migrating mesendoderm cells.4 Similarly, in the zebrafish mesendoderm, PDGF signaling induces cell protrusions and polarization.25 In this case, upon PDGF treatment, protein kinase B (PKB)/Akt become localized to the leading edge of the cells in a PI3K dependent manner.25 Such mechanisms require the activation of PDGFR at the leading edge of the cell. In mammalian cells, however, it has been demonstrated that activation of epidermal growth factor receptor (EGFR) can spread laterally in the cell membrane.28 If the same is true for PDGFR, even cells exposed to a gradient or localized source of PDGF-AA may lose spatial information. Recent evidence in Drosophila, however, suggests an alternate mechanism involving receptor endocytosis that ensures that a graded extracellular signal is maintained intracellularly.29The Drosophila member of the PDGF/VEGF family, PVF1, along with epidermal growth factor, directs the migration of border cells towards the oocyte during oogenesis.3032 The patterns of expression of ligand and receptor are similar to those during Xenopus gastrulation in that a stationary cell, the oocyte, expresses the ligand, PVF1, while the motile border cells express its receptor, PVR.30 PVR activation is concentrated at the leading edge, and specifically maintained in a localized region of the cell by endocytosis.29 PDGFRβ endocytosis also regulates PDGF-dependent migration of NIH3T3 cells in culture.33 This process involves recruitment of a ternary complex of DOCK4-Grb2-Dynamin2 to the leading edge of the migrating cell, which results in rapid receptor endocytosis, where it is maintained in an active phosphorylation state and leads to polarized signaling. Disruption of the complex, for example by overexpression of DOCK4 ΔC, which cannot bind Grb2, blocks chemotaxis in response to PDGF. Whether a similar mechanism creates spatial signaling during Xenopus gastrulation requires further investigation, but this tantalizing evidence suggests that it is a possibility.  相似文献   

10.
Lens development and differentiation are intricate and complex processes characterized by distinct molecular and morphological changes. The growth of a transparent lens involves proliferation of the epithelial cells and their subsequent differentiation into secondary fiber cells. Prior to differentiation, epithelial cells at the lens equator exit from the cell cycle and elongate into long, ribbon-like cells. Fiber cell elongation takes place bidirectionally as fiber tips migrate both anteriorly and posteriorly along the apical surface of the epithelium and inner surface of the capsule, respectively. The differentiating fiber cells move inward from the periphery to the center of the lens on a continuous basis as the lens grows throughout life. Finally, when fiber cells reach the center or suture line, their basal and apical tips detach from the epithelium and capsule, respectively, and interlock with cells from the opposite direction of the lens and form the suture line. Further, symmetric packing of fiber cells and degradation of most of the cellular organelle during fiber cell terminal differentiation are crucial for lens transparency. These sequential events are presumed to depend on cytoskeletal dynamics and cell adhesive interactions; however, our knowledge of regulation of lens fiber cell cytosketal reorganization, cell adhesive interactions and mechanotransduction, and their role in lens morphogenesis and function is limited at present. Recent biochemical and molecular studies have targeted cytoskeletal signaling proteins, including Rho GTPases, Abl kinase interacting proteins, cell adhesion molecules, myosin II, Src kinase and phosphoinositide 3-kinase in the developing chicken and mouse lens and characterized components of the fiber cell basal membrane complex. These studies have begun to unravel the vital role of cytoskeletal proteins and their regulatory pathways in control of lens morphogenesis, fiber cell elongation, migration, differentiation, survival and mechanical properties.Key words: lens, fiber cells, elongation, migration, adhesion, Rho GTPasesLens morphogenesis involves a complex network of regulatory genes and interplay between growth factor, mitogenic, cell adhesive and cytoskeletal signaling pathways. The lens originates from surface ectoderm near the optic vesicle and lens vesicle that is formed via invagination of lens placode differentiates into primary fibers (the posterior half ) and epithelial cells (the anterior half ). These changes in embryonic cells control the lens distinctive anterior-posterior polarity. Subsequently, the lens grows through the proliferation of epithelial cells and the differentiation of their progeny into secondary fiber cells.1,2 The continuous addition of new fiber cells at the lens periphery leads to a gradual inward movement of older cells to the center of the lens. The ectodermal basement membrane that surrounds the lens vesicle thickens to form the lens capsule and is composed of mainly proteins of extracellular matrix.2,3 Since the lens does not shed cells, they are retained throughout the lens''s life and are packed symmetrically within the lens4 (Fig. 1).Open in a separate windowFigure 1Diagram of organization of lens epithelial and differentiating fiber cells. The lens is enclosed by a thick capsule consisting of various extracellular matrix proteins. Lens epithelial cells at the equator divide and exit from the cell cycle, and as they exit from the cell cycle, they start to elongate bidirectionally by making apical (AMC) and basal (BMC) membrane complexes with epithelium and capsule, respectively. As fiber cells elongate, they are pushed down and migrate toward the center. As the fiber cells migrate toward the center, both the basal and apical membrane complexes are expected to undergo changes in a regulated manner to control fiber cell adhesive, protrusive and contractile activity. Finally, when the fiber cells reach the center or suture line, their basal and apical ends detach from the epithelium and capsule, respectively and interlock with cells from the opposite direction of the lens and form suture. During fiber cell elongation and differentiation, cell adhesive interactions are reorganized extensively, and terminally differentiated fiber cells exhibit loss of cellular organelle and extensive membrane remodeling with unique ball and socket interdigitations. Arrows indicate the direction of fiber cell movement. This schematic is a modified version of Figure 2 from Lovicu and McAvoy.1Lens fiber cell elongation and differentiation is associated with a remarkable change in cell morphology, with the length of fiber cells increasing on the order of several hundredfold. These morphological changes are associated with extensive membrane and cortical cytoskeletal remodeling, actomyosin reorganization and cell adhesion turnover.517 Additionally, the tips of the elongating fiber cells at both the anterior and posterior terminals slide along the lens epithelium and capsule, respectively, as these cells migrate inward, and finally detach at the suture, where they form contacts with their counterparts from the opposite side of the lens.4,12 These cell movements are fundamental for maintaining distinct lens fiber cell polarity and are temporally and spatially regulated as the lens grows continuously throughout life.1,2,12 Another unique feature of the lens is that during fiber cell terminal differentiation, all the cellular organelles, including nuclei, endoplasmic reticulum and mitochondria, are degraded in a programmed manner.18 It has been well documented that lens epithelial cell elongation and differentiation is associated with reorganization of actin cytoskeleton, increased ratio of G-actin to F-actin, integrin switching, formation of N-cadherin linked cell adhesions, and expression of actin capping protein tropomodulin.5,6,9,10,13,15,17,1921 Importantly, disruption of actin cytoskeletal organization has been shown to impair lens epithelial differentiation and induce cataract formation, indicating the significance of actin cytoskeleton in lens differentiation and maintenance of lens optical quality.14,22 Further, during accommodation, lens shape is changed in a reversible manner. Therefore, the tensional homeostasis between actomyosin inside the fiber cell and fiber cell adhesion on the inner side of the lens capsule is considered to be crucial for accommodation.12In the developing mouse and chicken lens, the tips of the fiber cells (both apical and basal) have been reported to cluster with different cytoskeletal proteins, including actin, myosin II, actin capping protein tropomodulin, and N-cadherins.10,19,21 Similarly, adhesion regulating signaling molecules including integrins, focal adhesion kinase, Cdk5, abl kinase interacting protein (Abi-2), and Rho GTPases have been shown to localize to the fiber cell apical and basal tips.20,2326 Moreover, isolation and characterization of the fiber cell basal membrane complexes (BMCs) had revealed a symmetric organization of N-cadherin, myosin II, actin in association with myosin light chain kinase, focal adhesion kinase, β1 integrin and caldesmon.12 The signaling activity, tensional property and dynamics of BMCs are thought to control the coordinated migration of fiber cells along the lens capsule, formation of lens suture line, and lens accommodation.12 Additionally, the BMCs have been shown to undergo a characteristic regional rearrangement (including size and shape) during lens elongation and migration along the lens capsule.27 Therefore, impaired fiber cell migration on the lens capsule is expected to induce cataractogenesis.27 Taken together, these different observations convincingly indicate the importance of cytoskeleton and cell adhesion regulatory mechanisms in lens fiber cell elongation and migration.Although important insights have emerged regarding external cues controlling lens epithelial cell proliferation, elongation and differentiation, little is known regarding the specific signaling pathways that drive the processes culminating in fiber cell formation, migration, packing and maturation.1,7,28 For example, growth factors are known to play key roles in influencing cell fates during development. Some of the major growth factor families, including FGFs and TGFβ/BMPs, have been shown to be involved in the regulation of lens developmental processes and primary fiber cell differentiation via ERK kinase activation.1,28,29 However, the identity and role of signaling pathways acting downstream to growth factors regulating lens secondary fiber cell elongation, migration, adhesion, membrane remodeling and survival are poorly understood.1,12,21,30 In particular, regulatory mechanisms involved in cytoskeletal reorganization, tensional force and cell adhesive interactions during these cellular processes have yet be identified and characterized.7,9,12,21,3032Our laboratory has been working on a broad hypothesis that the actin cytoskeletal and cell adhesive signaling mechanisms composed of Rho GTPases (Rho, Rac and Cdc42) and their effector molecules play a critical role in controlling lens growth and differentiation, and in maintaining lens integrity.7 The Rho family of small GTPases regulates morphogenesis, polarity, migration and cell adhesion.33 These proteins bind GTP, exhibit GTPase activity, and cycle between an inactive GDP-bound form and an active GTP-bound form. This cycling is regulated by three groups of proteins: guanine-nucleotide exchange factors, which facilitate the exchange of GDP for GTP, thus rendering Rho GTPases active; GTPase-activating proteins, which regulate the inactivation of Rho by accelerating intrinsic GTPase activity and converting Rho GTPases back to their GDP-bound form; and GDP dissociation inhibitors (GDIs), which inhibit the dissociation of GDP bound to Rho GTPases.33,34 The GTP-bound form of the Rho GTPases interact with downstream effectors, which include protein kinases (e.g., ROCK and PAK), regulators of actin polymerization (e.g., N-WASP/WAVE, PI3-kinase and mDia), and other proteins with adaptor functions.33 The selective interaction of the different Rho GTPases with a variety of effectors determines the final outcome of their activation.33 For example, during cell movement, Rac and Cdc42 stimulate formation of protrusions at the leading edges of cells, and RhoA induces retraction at the tail ends of cells. This coordinated cytoskeletal reorganization permits cells to move toward a target.35 PI3-kinase and PI (3, 4, 5) P3 have also been widely implicated in controlling cell migration and polarity in a Rac GTPase-dependent manner.35 Members of the Wiskott-Aldrich syndrome protein (WASP) and WASP-family verprolin homologous protein (WAVE) families serve to link Rho GTPases signals to the ARP2/3 complex, leading to actin polymerization that is crucial for the reorganization of the actin cytoskeleton at the leading edge for processes such as cell movement and protrusions.36 Importantly, all three Rho GTPases also regulate microtubule polymerization and assembly of adherens junctions to influence polarity and cell adhesion, respectively.33,37Likewise, a tensional balance between cell adhesion on the outside and myosin II-based contractility on the inside of the cells is regulated by Rho GTPases.38To explore the role of the Rho GTPases in lens morphogenesis and differentiation, we have targeted the lens Rho GTPases by overexpressing either the C3 exoenzyme (inactivator of RhoA and RhoB) or RhoGDIα (Rho GDP dissociation inhibitor) in a lens-specific manner in transgenic mice and followed their effects developmentally. These two transgenic mouse models exhibited ocular phenotype, including lens opacity (cataract) and microphthalmic eyes. Importantly, various histological, immunofluorescence and biochemical analyses performed in these developing transgenic mice have revealed defective lens morphogenesis, abnormal fiber cell migration, elongation, disrupted cytoskeletal organization and adhesive interactions, along with changes in proteins of the fiber cell gap junctions and water channels.32,39 These lenses have also shown decreased ERM (ezrin, radixin, moesin) protein phosphorylation,40 proteins that are involved in crosslinking of the plasma membrane with actin cytoskeleton,41 and increased apoptosis.32 Defective fiber cell migration has been found to be more notable in the Rho GDI overexpressing lenses than in the C3 exoenzyme expressing lenses (Fig. 2). The Rho GDI overexpressing lenses have shown a defective membrane localization of Rho, Rac and Cdc42 confirming their inactivation. These data, together with mechanistic studies performed using the lens epithelial cells and the noted effects on cell shape, actin polymerization, myosin phosphorylation and cell adhesive interactions, reveal the importance of Rho GTPase-dependent signaling pathways in processes underlying fiber cell migration, elongation, cytoskeletal and membrane organization and survival in the developing lens.7 Lens fiber cell BMC has been found to be localized intensely with Rac GTPase involved in cell migration (our unpublished work). Additionally, the Rho GDI transgenic lenses showed an impaired apical-apical cell-cell interactions between the fiber cells and epithelial cells.32 Moreover, the ruptured posterior capsule and disrupted suture lines in these lenses are indicative of defective BMC organization and activity.32Open in a separate windowFigure 2Abnormal lens phenotype in the neonatal Rho GDIα overexpressing transgenic mouse. Hematoxylin and eosin-stained sagittal sections of P1 RhoGDIα transgenic eyes reveal abnormal migration and morphology of the posterior lens fibers as compared with the symmetric organization of lens fibers and their migration toward the lens suture in the wild type mouse (reproduced with permission from Maddala et al.)32.Further support for involvement of Rho GTPases in lens fiber cell differentiation and survival has come from studies conducted with chick lens epithelial explants and cultured epithelial cells. Inactivation of Rho kinase or Rac activation by PI3 kinase in chick lens epithelial cells has been reported to induce fiber cell differentiation and survival in association with distinct cortical actin cytoskeletal reorganization, indicating the significance of Rho GTPases in lens fiber cell differentiation and survival.9,42 Additionally, lens fiber cell elongation and differentiation has been found to be associated with increased myosin light chain (MLC) phosphorylation, and inhibition of MLC phosphorylation regulated by MLC kinase and Rho kinase has induced lens opacity and disruption of cytoskeletal integrity, supporting the importance of myosin II activity in maintaining lens architecture and transparency.10 Importantly, various growth factors that regulate lens morphogenesis, fiber cell differentiation, and survival have been found to activate Rho and Rac GTPases and to induce MLC phosphorylation, actin cytoskeletal reorganization, and focal adhesion formation in lens epithelial cells.7,30 In addition to Rho GTPases, inhibition of Src kinase has been shown to induce fiber cell differentiation in association with actin cytoskeletal reorganization and cell adhesive interactions.43 Also, the expression and activation of focal adhesion kinase has been reported to increase in differentiating and migrating lens epithelial cells.44 Both these molecules are well recognized to regulate cell migration by participating in the disassembly of cell adhesions at the front of migrating cells.35Additional evidence for the participation of actin cytoskeletal organization and Rho GTPases in lens fiber cell migration and elongation has been derived from the studies of Abi-2 deficient mouse. Abl-interactor adaptor proteins Abi-1 and Abi-2 are linked to the Rac-WAVE-Arp2/3 signaling pathway and regulate actin polymerization and cell-cell adhesive interactions.45 Homozygous deletion of Abi-2 in mice has been shown to exhibit ocular phenotype including microphthalmia and lens opacity similar to the Rho GDI overexpressing transgenic mouse eyes noted in previous studies.23,32 In the absence of Abi-2, the secondary lens fiber orientation, migration and elongation were found to be defective, supporting the importance of Rac-WAVE-Arp2/3 signaling in lens fiber cell migration and cell adhesion.23 Abi-2 has been shown to localize intensely to the both basal and apical regions of the fiber cells and adherens junctions, and suppression of Abi-2 expression in epithelial cells resulted in impaired adherens junctions and downregulation of actin nucleation promoting factors.23 The significance of cytoskeletal signaling in lens has also been implicated in Lowe syndrome, a rare X-linked disorder characterized by congenital cataracts, results from mutations in the OCRL1 gene. The OCRL1 protein product (phosphatidylinositol 4, 5 bisphosphate 5-phosphatase) has been shown to participate in Rac GTPase regulated actin cytoskeletal organization, cell migration, and cell adhesion in various cell types.46 Finally, Wnt/PCP signaling via activation of Rho GTPases has been suggested to control lens morphogenesis, fiber cell migration and differentiation.26Importantly, given how the activity of the Rho GTPases is regulated by external cues and various effector proteins, a detailed understanding of the regulation of Rho GTPase signaling is necessary for a better appreciation of their role in lens morphogenesis, fiber cell elongation and differentiation, and tensional homeostasis. Further mechanistic studies are critical to unravel the specific role(s) of Rho GTPases and other cytoskeletal regulatory mechanisms involved in regulating the formation and disassembly of fiber cell basal and apical membrane complexes, fiber cell lateral membrane remodeling, and fiber cell-cell adhesive interactions during lens differentiation. Very little is known in terms of the assembly of different cell adhesive molecules at the apical-apical interface between the lens fibers and epithelial cells. We are only beginning to glimpse the regulatory networks involved in the regulation of fiber cell elongation, polarity, migration and adhesion. Many challenging questions remain: for example, how are the pathways regulating migration, basal and apical membrane complexes, and tensional homeostasis controlled by extracellular signals, and how are they integrated during fiber cell migration, suture formation, and packing? Novel insights into the molecular mechanisms regulating these cellular processes are expected to advance our understanding of lens morphogenesis, function and cataractogenesis.  相似文献   

11.
12.
13.
The prion hypothesis13 states that the prion and non-prion form of a protein differ only in their 3D conformation and that different strains of a prion differ by their 3D structure.4,5 Recent technical developments have enabled solid-state NMR to address the atomic-resolution structures of full-length prions, and a first comparative study of two of them, HET-s and Ure2p, in fibrillar form, has recently appeared as a pair of companion papers.6,7 Interestingly, the two structures are rather different: HET-s features an exceedingly well-ordered prion domain and a partially disordered globular domain. Ure2p in contrast features a very well ordered globular domain with a conserved fold, and—most probably—a partially ordered prion domain.6 For HET-s, the structure of the prion domain is characterized at atomic-resolution. For Ure2p, structure determination is under way, but the highly resolved spectra clearly show that information at atomic resolution should be achievable.Key words: prion, NMR, solid-state NMR, MAS, structure, Ure2p, HET-sDespite the large interest in the basic mechanisms of fibril formation and prion propagation, little is known about the molecular structure of prions at atomic resolution and the mechanism of propagation. Prions with related properties to the ones responsible for mammalian diseases were also discovered in yeast and funghi8,9 which provide convenient model system for their studies. Prion proteins described include the mammalian prion protein PrP, Ure2p,10 Rnq1p,11 Sup35,12 Swi1,13 and Cyc8,14 from bakers yeast (S. cervisiae) and HET-s from the filamentous fungus P. anserina. The soluble non-prion form of the proteins characterized in vitro is a globular protein with an unfolded, dynamically disordered N- or C-terminal tail.1518 In the prion form, the proteins form fibrillar aggregates, in which the tail adopts a different conformation and is thought to be the dominant structural element for fibril formation.Fibrills are difficult to structurally characterize at atomic resolution, as X-ray diffraction and liquid-state NMR cannot be applied because of the non-crystallinity and the mass of the fibrils. Solid-state NMR, in contrast, is nowadays well suited for this purpose. The size of the monomer, between 230 and 685 amino-acid residues for the prions of Figure 1, and therefore the number of resonances in the spectrum—that used to be large for structure determination—is now becoming tractable by this method.Open in a separate windowFigure 1Prions identified today and characterized as consisting of a prion domain (blue) and a globular domain (red).Prion proteins characterized so far were found to be usually constituted of two domains, namely the prion domain and the globular domain (see Fig. 1). This architecture suggests a divide-and-conquer approach to structure determination, in which the globular and prion domain are investigated separately. In isolation, the latter, or fragments thereof, were found to form β-sheet rich structures (e.g., Ure2p(1-89),6,19 Rnq1p(153-405)20 and HET-s(218-289)21). The same conclusion was reached by investigating Sup35(1-254).22 All these fragements have been characterized as amyloids, which we define in the sense that a significant part of the protein is involved in a cross-beta motif.23 An atomic resolution structure however is available presently only for the HET-s prion domain, and was obtained from solid-state NMR24 (vide infra). It contains mainly β-sheets, which form a triangular hydrophobic core. While this cross-beta structure can be classified as an amyloid, its triangular shape does deviate significantly from amyloid-like structures of smaller peptides.23Regarding the globular domains, structures have been determined by x-ray crystallography (Ure2p25,26 and HET-s27), as well as NMR (mammal prions15,2830). All reveal a protein fold rich in α-helices, and dimeric structures for the Ure2 and HET-s proteins. The Ure2p fold resembles that of the β-class glutathione S-transferases (GST), but lacks GST activity.25It is a central question for the structural biology of prions if the divide-and-conquer approach imposed by limitations in current structural approaches is valid. Or in other words: can the assembly of full-length prions simply be derived from the sum of the two folds observed for the isolated domains?  相似文献   

14.
The conserved eukaryotic protein SGT1 (suppressor of G2 allele of skp1) participates in diverse physiological processes such as cell cycle progression in yeast, plant immunity against pathogens and plant hormone signalling. Recent genetic and biochemical studies suggest that SGT1 functions as a novel co-chaperone for cytosolic/nuclear HSP90 and HSP70 molecular chaperones in the folding and maturation of substrate proteins. Since proteins containing the leucine-rich repeat (LRR) protein-protein interaction motif are overrepresented in SGT1-dependent phenomena, we consider whether LRR-containing proteins are preferential substrates of an SGT1/HSP70/HSP90 complex. Such a chaperone organisation is reminiscent of the HOP/HSP70/HSP90 machinery which controls maturation and activation of glucocorticoid receptors in animals. Drawing on this parallel, we discuss the possible contribution of an SGT1-chaperone complex in the folding and maturation of LRR-containing proteins and its evolutionary consequences for the emergence of novel LRR interaction surfaces.Key words: heat shock protein, SGT1, co-chaperone, HSP90, HSP70, leucine-rich repeat, LRR, resistance, SCF, ubiquitinThe proper folding and maturation of proteins is essential for cell viability during de novo protein synthesis, translocation, complex assembly or under denaturing stress conditions. A complex machinery composed of molecular chaperones (heat-shock proteins, HSPs) and their modulators known as co-chaperones, catalyzes these protein folding events.1,2 In animals, defects in the chaperone machinery is implicated in an increasing number of diseases such as cancers, susceptibility to viruses, neurodegenerative disease and cystic fibrosis, and thus it has become a major pharmacological target.3,4 In plants, molecular genetic studies have identified chaperones and co-chaperones as components of various physiological responses and are now starting to yield important information on how chaperones work. Notably, processes in plant innate immunity rely on the HSP70 and HSP9057 chaperones as well as two recently characterised co-chaperones, RAR1 (required for Mla12 resistance) and SGT1 (suppressor of G2 allele of skp1).811SGT1 is a highly conserved and essential co-chaperone in eukaryotes and is organized into three structural domains: a tetratricopeptide repeat (TPR), a CHORD/SGT1 (CS) and an SGT1-specific (SGS) domain (Fig. 1A). SGT1 is involved in a number of apparently unrelated physiological responses ranging from cell cycle progression and adenylyl cyclase activity in yeast to plant immunity against pathogens, heat shock tolerance and plant hormone (auxin and jasmonic acid) signalling.79,12,13 Because the SGT1 TPR domain is able to interact with Skp1, SGT1 was initially believed to be a component of SCF (Skp1/Cullin/F-box) E3 ubiquitin ligases that are important for auxin/JA signalling in plants and cell cycle progression in yeast.13,14 However, mutagenesis of SGT1 revealed that the TPR domain is dispensable for plant immunity and auxin signalling.15 Also, SGT1-Skp1 interaction was not observed in Arabidopsis.13 More relevant to SGT1 functions appear to be the CS and SGS domains.16 The former is necessary and sufficient for RAR1 and HSP90 binding. The latter is the most conserved of all SGT1 domains and the site of numerous disabling mutations.14,16,17Open in a separate windowFigure 1Model for SGT1/chaperone complex functions in the folding of LRR-containing proteins. (A) The structural domains of SGT1, their sites of action (above) and respective binding partners (below) are shown. N- and C-termini are indicated. TPR, tetratricopeptide repeat; CS, CHORD/SGT1; SGS, SGT1-specific. (B) Conceptual analogy between steroid receptor folding by the HOP/chaperone machinery and LRR protein folding by the SGT1/chaperone machinery. LRR motifs are overrepresented in processes requiring SGT1 such as plant immune receptor signalling, yeast adenylyl cyclase activity and plant or yeast SCF (Skp1/Cullin/F-box) E3 ubiquitin ligase activities. (C) Opposite forces drive LRR evolution. Structure of LRRs 16 to 18 of the F-box auxin receptor TIR1 is displayed as an illustration of the LRR folds.30 Leucine/isoleucine residues (side chain displayed in yellow) are under strong purifying selection and build the hydrophobic LRR backbone (Left). By contrast, solvent-exposed residues of the β-strands define a polymorphic and hydrophilic binding surface conferring substrate specificity to the LRR (Right) and are often under diversifying selection.We recently demonstrated that Arabidopsis SGT1 interacts stably through its SGS domain with cytosolic/nuclear HSP70 chaperones.7 The SGS domain was both necessary and sufficient for HSP70 binding and mutations affecting SGT1-HSP70 interaction compromised JA/auxin signalling and immune responses. An independent in vitro study also found interaction between human SGT1 and HSP70.18 The finding that SGT1 protein interacts directly with two chaperones (HSP90/70) and one co-chaperone (RAR1) reinforces the notion that SGT1 behaves as a co-chaperone, nucleating a larger chaperone complex that is essential for eukaryotic physiology. A future challenge will be to dissect the chaperone network at the molecular and subcellular levels. In plant cells, SGT1 localization appears to be highly dynamic with conditional nuclear localization7 and its association with HSP90 was recently shown to be modulated in vitro by RAR1.16A co-chaperone function suits SGT1 diverse physiological roles better than a specific contribution to SCF ubiquitin E3 ligases. Because SGT1 does not affect HSP90 ATPase activity, SGT1 was proposed rather as a scaffold protein.16,19 In the light of our findings and earlier studies,20 SGT1 is reminiscent of HOP (Hsp70/Hsp90 organizing protein) which links HSP90 and HSP70 activities and mediates optimal substrate channelling between the two chaperones (Fig. 1B).21 While the contribution of the HSP70/HOP/HSP90 to the maturation of glucocorticoid receptors is well established,21 direct substrates of an HSP70/SGT1/HSP90 complex remain elusive.It is interesting that SGT1 appears to share a functional link with leucine-rich repeat- (LRR) containing proteins although LRR domains are not so widespread in eukaryotes. For example, plant SGT1 affects the activities of the SCFTIR1 and SCFCOI1 E3 ligase complexes whose F-box proteins contain LRRs.13 Moreover, plant intracellular immune receptors comprise a large group of LRR proteins that recruit SGT1.8,9 LRRs are also found in yeast adenylyl cyclase Cyr1p and the F-box protein Grr1p which is required for SGT1-dependent cyclin destruction during G1/S transition.12,14 Yeast 2-hybrid interaction assays also revealed that yeast and plant SGT1 tend to associate directly or indirectly with LRR proteins.12,22,23 We speculate that SGT1 bridges the HSP90-HSC70 chaperone machinery with LRR proteins during complex maturation and/or activation. The only other structural motif linked to SGT1 are WD40 domains found in yeast Cdc4p F-box protein and SGT1 interactors identified in yeast two-hybrid screens.12What mechanisms underlie a preferential SGT1-LRR interaction? HSP70/SGT1/HSP90 may have co-evolved to assist specifically in folding and maturation of LRR proteins. Alternatively, LRR structures may have an intrinsically greater need for chaperoning activity to fold compared to other motifs. These two scenarios are not mutually exclusive. The LRR domain contains multiple 20 to 29 amino acid repeats, forming an α/β horseshoe fold.24 Each repeat is rich in hydrophobic leucine/isoleucine residues which are buried inside the structure and form the structural backbone of the motif (Fig. 1C, left). Such residues are under strong purifying selection to preserve structure. These hydrophobic residues would render the LRR a possible HSP70 substrate.25 By contrast, hydrophilic solvent- exposed residues of the β strands build a surface which confers ligand recognition specificity of the LRRs (Fig. 1C). In many plant immune receptors for instance, these residues are under diversifying selection that is likely to favour the emergence of novel pathogen recognition specificities in response to pathogen evolution.26 The LRR domain of such a protein has to survive such antagonist selection forces and yet remain functional. Under strong selection pressure, LRR proteins might need to accommodate less stable LRRs because their recognition specificities are advantageous. This could be the point at which LRRs benefit most from a chaperoning machinery such as the HSP90/SGT1/HSP70 complex. This picture is reminiscent of the genetic buffering that HSP90 exerts on many traits to mask mutations that would normally be deleterious to protein folding and/or function, as revealed in Drosophila and Arabidopsis.27 It will be interesting to test whether the HSP90/SGT1/HSP70 complex acts as a buffer for genetic variation, favouring the emergence of novel LRR recognition surfaces in, for example, highly co-evolved plant-pathogen interactions.28,29  相似文献   

15.
Cell migration is an integrated process that involves cell adhesion, protrusion and contraction. We recently used CAS (Crk-associated substrate, 130CAS)-deficient mouse embryo fibroblasts (MEFs) to examined contribution made to v-Crk to that process via its interaction with Rac1. v-Crk, the oncogene product of avian sarcoma virus CT10, directly affects membrane ruffle formation and is associated with Rac1 activation, even in the absence of CAS, a major substrate for Crk. In CAS-deficient MEFs, cell spreading and lamellipodium dynamics are delayed; moreover, Rac activation is significantly reduced and it is no longer targeted to the membrane. However, expression of v-Crk by CAS-deficient MEFs increased cell spreading and active lamellipodium protrusion and retraction. v-Crk expression appears to induce Rac1 activation and its targeting to the membrane, which directly affects membrane dynamics and, in turn, cell migration. It thus appears that v-Crk/Rac1 signaling contributes to the regulation of membrane dynamics and cell migration, and that v-Crk is an effector molecule for Rac1 activation that regulates cell motility.Key words: v-Crk, Rac, lamellipodia dynamics, cell migration, p130CASCell migration is a central event in a wide array of biological and pathological processes, including embryonic development, inflammatory responses, angiogenesis, tissue repair and regeneration, cancer invasion and metastasis, osteoporosis and immune responses.1,2 Although the molecular basis of cell migration has been studied extensively, the underlying mechanisms are still not fully understood. It is known that cell migration is an integrated process that involves formation of cell adhesions and/or cell polarization, membrane protrusion in the direction of migration (e.g., filopodium formation and lamellipodium extension), cell body contraction and tail detachment.13 Formation of cell adhesions, including focal adhesions, fibrillar adhesions and podosomes are the first step in cell migration. Cell adhesions are stabilized by attachment to the extracellular matrix (ECM) mediated by integrin transmembrane receptors, which are also linked to various cytoplasmic proteins and the actin cytoskeleton, which provide the mechanical force necessary for migration.2,4 The next steps in the process of cell migration are filopodium formation and lamellipodium extension. These are accompanied by actin polymerization and microtubule dynamics, which also contribute to the control of cell adhesion and migration.5Focal adhesions are highly dynamic structures that form at sites of membrane contact with the ECM and involve the activities of several cellular proteins, including vinculin, focal adhesion kinase (FAK), Src family kinase, paxillin, CAS (Crk-associated substrate, p130CAS) and Crk.6 A deficiency in focal adhesion protein is associated with the severe defects in cell motility and results in embryonic death. For example, FAK deficiency disrupts mesoderm development in mice and delays cell migration in vitro,7 which reflects impaired assembly and disassembly the focal adhesions.8 In addition, mouse embryonic fibroblasts (MEFs) lacking Src kinase showed a reduced rate of cell spreading that resulted in embryonic death.9 Taken together, these findings strongly support the idea that cell adhesion complexes play crucial roles in cell migration.CAS is a hyperphosphorylated protein known to be a major component of focal adhesion complexes and to be involved in the transformation of cells expressing v-Src or v-Crk.10 CAS-deficient mouse embryos die in utero and show marked systematic congestion and growth retardation,4 while MEFs lacking CAS show severely impaired formation and bundling of actin stress fibers and delayed cell motility.4,11,12 Conversely, transient expression of CAS in COS7 cells increases cell migration.11 Crk-null mice also exhibit lethal defects in embryonic development,13 which is consistent with the fact that CAS is a major substrate for v-Crk, and both CAS and v-Crk are necessary for induction of cell migration.14 v-Crk consists of a viral gag sequence fused to cellular Crk sequences, which contain Src homology 2 (SH2) and SH3 domains but no kinase domain, and both CAS and paxillin bind to SH2 domains.12,15,16 Despite the absence of a kinase domain, cell expressing v-Crk show upregulation of tyrosine phosphorylation of CAS, FAK and paxillin, which is consistent with v-Crk functioning as an adaptor protein.17 Moreover, this upregulation of tyrosine phosphorylation correlates well with the transforming activity of v-Crk.17 By contrast, tyrosine phosphorylation of FAK and CAS is diminished in Src kinase-deficient cells expressing v-Crk, and they are not targeted to the membrane, suggesting v-Crk signaling is Src kinase-dependent. After formation of the CAS/v-Crk complex, v-Crk likely transduces cellular signaling to Src kinase and FAK.12 Notably, tyrosine phosphorylation of FAK and cell migration and spreading are all enhanced when v-Crk is introduced into CAS-deficient MEFs.12 We therefore suggest that v-Crk activity, but not cellular Crk activity, during cell migration and spreading is CAS-independent.Membrane dynamics such as lamellipodium protrusion and membrane ruffling reportedly involve Rac1,18 α4β1 integrin,19 Arp2/3,6 and N-WASP,20 and are enhanced in v-Crk-expressing CAS-deficient MEFs.21 Moreover, expression in those cells of N17Rac1, a dominant defective Rac1 mutant, abolished membrane dynamics at early times and delayed cell migration.21 v-Crk-expressing, CAS-deficient MEFs transfected with N17Rac1 did not begin spreading until one hour after being plated on fibronectin, and blocking Rac activity suppressed both membrane dynamics and cell migration. We therefore suggest that v-Crk is involved in cell attachment and spreading, and that this process is mediated by Rac1 activation. In addition, v-Crk expression apparently restores lamellipodium formation and ruffle retraction in CAS-deficient MEFs. Thus v-Crk appears to participate in a variety cellular signaling pathways leading to cell spreading, Rac1 activation, membrane ruffling and cell migration, even in the absence of CAS, its major substrate protein.In fibroblasts, the Rho family of small GTP-binding proteins (e.g., Cdc42, Rac and Rho) functions to control actin cytoskeleton turnover, including filopodium extension, lamellipodium formation and generation of actin stress fibers and focal adhesions.22 These GTPases function in a cascade, such that activation of Cdc42 leads to activation of Rac1, which in turn activates Rho.22 Once activated, Rho controls cell migration. Cell adhesion to ECM leads to the translocation of Rac1 and Cdc42 from the cytosol to the plasma membrane,23 where they regulate actin polymerization at the leading edge.19,24 Dominant negative Rac and Cdc42 mutants inhibit the signaling to cell spreading initiated by the interaction of integrin with ECM.24 The fact that cellular levels of activated Rac are higher in cells adhering to ECM than in suspended cells further suggests that activation of Rac and Cdc42 is a critical step leading to membrane protrusion and ruffle formation. It is noteworthy in this regard that v-Crk is able to induce Rac activation and its translocation to plasma membrane.21Overall, the findings summarized in this article demonstrate that v-Crk participates in several steps leading to cell adhesion and spreading (Fig. 1), and the targeting of v-Crk to focal adhesion sites appears to be a prerequisite for regulation of cell migration and spreading via Rac activation. To fully understand its function, however, it will be necessary to clarify the role of v-Crk in Rac1 and Cdc42 activation initiated by integrin-ECM interactions.Open in a separate windowFigure 1Schematic diagram of v-Crk signaling in MEFs. Cell adhesion signaling initiated by the integrin-ECM interaction triggers v-Crk signaling mediated by Src kinase, after which focal adhesion proteins are tyrosine phosphorylated. These events lead to translocation of Rac from the cytosol to the membrane, where it promotes membrane protrusion and ruffle formation. Under basal conditions, Rac is bound with GDP and is inactive. Upon stimulation, Rac activation is mediated by guanine nucleotide exchange factors (GEFs) that stimulate the release of bound GDP and the binding of GTP. Activation of Rac is transient, however, as it is inactivated by GTPase activating protein (GAP).  相似文献   

16.
Junctional Adhesion Molecule A (JAM-A) is a member of the Ig superfamily of membrane proteins expressed in platelets, leukocytes, endothelial cells and epithelial cells. We have previously shown that in endothelial cells, JAM-A regulates basic fibroblast growth factor, (FGF-2)-induced angiogenesis via augmenting endothelial cell migration. Recently, we have revealed that in breast cancer cells, downregulation of JAM-A enhances cancer cell migration and invasion. Further, ectopic expression of JAM-A in highly metastatic MDA-MB-231 cells attenuates cell migration, and downregulation of JAM-A in low-metastatic T47D cells enhance migration. Interestingly, JAM-A expression is greatly diminished as breast cancer disease progresses. The molecular mechanism of this function of JAM-A is beyond its well-characterized barrier function at the tight junction. Our results point out that JAM-A differentially regulates migration of endothelial and cancer cells.Key words: JAM-A, integrin, αvβ3, FGF-2, breast cancer, cell migration and invasion, T47D, MDA-MB-231, siRNAEndothelial and epithelial cells exhibit cell polarity and have characteristic tight junctions (TJs) that separate apical and basal surfaces. TJs are composed of both transmembrane and cytoplasmic proteins. The three major families of transmembrane proteins include claudins, occludin and JAM family members.13 Additionally, interaction between the peripheral proteins such as PDS-95/Discs large/ZO family (PDZ) domain-containing proteins in TJs plays an important role in maintaining the junctional integrity.2,4,5JAMs are type I membrane proteins (Fig. 1) predominately expressed in endothelial and epithelial cell TJs, platelets and some leukocytes.68 The classical JAMs are JAM-A, JAM-B and JAM-C, which can all regulate leukocyte-endothelial cell interaction through their ability to undergo heterophilic binding with integrins αLβ2 or αvβ3, α4β1 and αMβ2 respectively. The cytoplasmic tail of JAMs contains a type II PDZ-domain-binding motif (Fig. 1) that can interact with the PDZ domain containing cytoplasmic molecules such as ZO-1, ASIP/PAR-3 or AF-6.9,10 Additionally, consistant with their junctional localization and their tendency to be involved in homophilic interactions, JAMs have been shown to modulate paracellular permeability and thus may play an important role in regulating the epithelial and endothelial barrier.11,12 In addition, ectopic expression of JAM-A in CHO cells promotes localization of ZO-1 and occludin at points of cell contacts, which suggests a role for JAM-A in TJ assembly.10,13,14 Recently, it has been shown that JAM-A regulates epithelial cell morphology by modulating the activity of small GTPase Rap1 suggesting a role for JAM-A in intracellular signaling.15Open in a separate windowFigure 1Schematic representation of the domain structure of JAM family proteins. V, variable Ig domain; C2, constant type 2 Ig domain; TM, transmembrane domain; T-II, Type II PDZ-domain binding motif.We have previously shown that JAM-A is a positive regulator of fibroblast growth factor-2 (FGF-2) induced angiogenesis.16 Evidence was provided to support the notion that JAM-A forms a complex with integrin αvβ3 at the cell-cell junction in quiescent human umbilical cord vein endothelial cells (HUVECs) and FGF-2 dissociates this complex.16 It was further established that inhibition of JAM-A using a function-blocking antibody also inhibits FGF-2 induced HUVECs migration in vitro and angiogenesis in vivo. Overexpression of JAM-A induced a change in HUVECs morphology similar to that observed when treated with FGF-2.17 Furthermore, overexpression of JAM-A, but not its cytoplasmic domain deletion mutant, augmented cell migration in the absence of FGF-2.17 In addition, downregulation of JAM-A in HUVECs using specific siRNA, resulted in reduced FGF-2-induced cell migration and inhibition of mitogen activated protein (MAP) kinase activation.18 These findings clearly suggested that JAM-A positively regulates FGF-2-induced endothelial cell migration. This was further confirmed in vivo by using JAM-A null mouse in which FGF-2 failed to support angiogenesis.19It is known that JAM-C, a JAM family member, is involved in the process of tumor cell metastasis.20 However, little is known about JAM-A''s role in cancer progression. We recently found that JAM-A is expressed in breast cancer tissues and cell lines.21 Based on our studies with endothelial cells it was felt that JAM-A expression in breast cancer cells may also enhance the migratory ability of these cells. Surprisingly, we found an inverse relation between the expression of JAM-A and the metastatic ability of breast cancer cells. T47D cells, which express high levels of JAM-A, are the least migratory; whereas MDA-MB-231 cells, which are highly migratory, are found to express the least amount of JAM-A.21 We also found that overexpression of JAM-A in MDA-MB-231 cells caused a change in cell morphology from spindle-like to rounded shape and formed cobblestone-like clusters.21 This is consistent with the previous report, that downregulation of JAM-A expression from epithelial cells using siRNA results in the change of epithelial cell morphology.15 This change in cell morphology by knockdown of JAM-A was attributed to the disruption of epithelial cell barrier function.15 It was further shown that knockdown of JAM-A affects epithelial cell morphology through reduction of β1integrin expression due to decreased Rap1 activity.15 Our observed effect of JAM-A downregulation in T47D cells, however, is not due to downregulation of β1integrin, since the level of this integrin was not affected in these cells. Interestingly, overexpression of JAM-A significantly affected both the cell migration and invasion of MDA-MB-231 cells. Furthermore, knockdown of JAM-A using siRNA enhanced invasiveness of MDA-MB-231 cells, as well as T47D cells.21 The ability of JAM-A to attenuate cell invasion was found to be due to the formation of functional tight junctions as observed by distinct accumulation of JAM-A and ZO-1 at the TJs and increased transepithelial resistance. These results identify, for the first time, a tight junctional cell adhesion protein as a key negative regulator of breast cancer cell migration and invasion.21JAM-A has been shown to be important in maintaining TJ integrity.15,2225 Disruption of TJs has been implicated to play a role in cancer cell metastasis by inducing epithelial mesenchymal transition.26 Several laboratories, including ours, have shown that cytokines and growth factors redistribute JAM-A from TJs.16,27,28 Consistent with this finding, it has been shown that hepatocyte growth factor (HGF) disrupts TJs in human breast cancer cells and downregulates expression of several TJ proteins.29 It is therefore conceivable that the loss of JAM-A in highly metastatic cells is a consequence of disruption of TJs. This was further supported by the findings that overexpression of JAM-A forms functional TJs in MDA-MB-231 cells and attenuates their migratory behavior. Our result is the first report correlating an inverse relationship of JAM-A expression in breast cancer cells to their invasive ability.21Using cDNA microarray technology, it has been revealed how genes involved in cell-cell adhesion, including those of the TJ, are under or overexpressed in different carcinomas.15,30 Cell-cell adhesion molecules have been well documented to regulate cancer cell motility and invasion. Of these, the cadherin family have been studied the most.31,32 It was proposed that a cadherin switch, that is, the loss of E-cadherin and subsequent expression of N-cadherin, may be responsible for breast cancer cell invasion.33,34 Although the role of cadherins is well-documented, it remains controversial since some breast cancer cell lines that do not express these proteins still posses highly invasive characteristics.33,34 However, the observed effect of overexpression of JAM-A does not appear to be simply due to the formation of TJs, since individual cells that express increased JAM-A show reduced migration.21 This is not surprising, considering the fact that JAM-A in addition to its function of regulating TJ integrity is also shown to participate in intracellular signaling. JAM-A is capable of interacting homotypically as well as heterotypically on the cell surface.35,36 It has also been shown that it interacts with several cytoplasmic proteins through its PDZ domain-binding motif and recruits signaling proteins at the TJs.37 Recent findings using site-directed mutagenesis suggest that cis-dimerization of JAM-A is necessary for it to carry out its biological functions.38 Our own observations suggest that a JAM-A function-blocking antibody inhibits focal adhesion formation in endothelial cells (unpublished data), whereas overexpresion of JAM-A in MDA-MB-231 cells show increased and stable focal adhesions.21 It is therefore conceivable that in quiescent endothelial/epithelial cells JAM-A associates with integrin to form an inactive complex at the TJ (Fig. 2). Growth factors such as FGF-2 signaling dissociates this complex thus allowing dimerization of JAM-A and activation of integrin augmenting cell migration (Fig. 2). On the contrary, in MDA-MB-231 cancer cells, which express low levels of JAM-A and do not form tight junctions, there may not be efficient inactive complex formation between JAM-A and integrin. Overexpression of JAM-A in these cells however, may promote such inactive complex formation leading to inhibition of integrin activation and JAM-A dimerization, both necessary events for cell migration. We are currently in the process of determining the specificity of interaction of JAM-A with integrins. Further experimentation is ongoing to determine the contribution of JAM-A dependent signaling in cell migration.Open in a separate windowFigure 2Schematic representation of JAM-A regulation of cell migration. JAM-A forms an inactive complex with the integrin and sequesters it at the TJs. Growth factor signaling dissociates this complex, promoting integrin activation and JAM-A dimerization leading to cell migration via MAP kinase activation. Ectopic expression of JAM-A in cancer cells may induce its association with integrin, forming an inactive complex and hence attenuation of migration.JAM-A differentially regulates cell migration in endothelial and cancer cells due to its ability to form inactive complex with integrin, making it a metastasis suppressor. The downregulation of JAM-A in carcinoma cells may be detrimental to the survival of breast cancer patients. It is therefore very important to determine the molecular determinants that are responsible for the downregulation of JAM-A during cancer progression. Thus, JAM-A, a molecule that dictates breast cancer cell invasion, could be used as a prognostic marker for metastatic breast cancer.  相似文献   

17.
18.
19.
Polar auxin transport (PAT), which is controlled precisely by both auxin efflux and influx facilitators and mediated by the cell trafficking system, modulates organogenesis, development and root gravitropism. ADP-ribosylation factor (ARF)-GTPase protein is catalyzed to switch to the GTP-bound type by a guanine nucleotide exchange factor (GEF) and promoted for hybridization to the GDP-bound type by a GTPase-activating protein (GAP). Previous studies showed that auxin efflux facilitators such as PIN1 are regulated by GNOM, an ARF-GEF, in Arabidopsis. In the November issue of The Plant Journal, we reported that the auxin influx facilitator AUX1 was regulated by ARF-GAP via the vesicle trafficking system.1 In this addendum, we report that overexpression of OsAGAP leads to enhanced root gravitropism and propose a new model of PAT regulation: a loop mechanism between ARF-GAP and GEF mediated by vesicle trafficking to regulate PAT at influx and efflux facilitators, thus controlling root development in plants.Key Words: ADP-ribosylation factor (ARF), ARF-GAP, ARF-GEF, auxin, GNOM, polar transport of auxinPolar auxin transport (PAT) is a unique process in plants. It results in alteration of auxin level, which controls organogenesis and development and a series of physiological processes, such as vascular differentiation, apical dominance, and tropic growth.2 Genetic and physiological studies identified that PAT depends on efflux facilitators such as PIN family proteins and influx facilitators such as AUX1 in Arabidopsis.Eight PIN family proteins, AtPIN1 to AtPIN8, exist in Arabidopsis. AtPIN1 is located at the basal side of the plasma membrane in vascular tissues but is weak in cortical tissues, which supports the hypothesis of chemical pervasion.3 AtPIN2 is localized at the apical side of epidermal cells and basally in cortical cells.1,4 GNOM, an ARF GEF, modulates the localization of PIN1 and vesicle trafficking and affects root development.5,6 The PIN auxin-efflux facilitator network controls root growth and patterning in Arabidopsis.4 As well, asymmetric localization of AUX1 occurs in the root cells of Arabidopsis plants,7 and overexpression of OsAGAP interferes with localization of AUX1.1 Our data support that ARF-GAP mediates auxin influx and auxin-dependent root growth and patterning, which involves vesicle trafficking.1 Here we show that OsAGAP overexpression leads to enhanced gravitropic response in transgenic rice plants. We propose a model whereby ARF GTPase is a molecular switch to control PAT and root growth and development.Overexpression of OsAGAP led to reduced growth in primary or adventitious roots of rice as compared with wild-type rice.1 Gravitropism assay revealed transgenic rice overxpressing OsAGAP with a faster response to gravity than the wild type during 24-h treatment. However, 1-naphthyl acetic acid (NAA) treatment promoted the gravitropic response of the wild type, with no difference in response between the OsAGAP transgenic plants and the wild type plants (Fig. 1). The phenotype of enhanced gravitropic response in the transgenic plants was similar to that in the mutants atmdr1-100 and atmdr1-100/atpgp1-100 related to Arabidopsis ABC (ATP-binding cassette) transporter and defective in PAT.8 The physiological data, as well as data on localization of auxin transport facilitators, support ARF-GAP modulating PAT via regulating the location of the auxin influx facilitator AUX1.1 So the alteration in gravitropic response in the OsAGAP transgenic plants was explained by a defect in PAT.Open in a separate windowFigure 1Gravitropism of OsAGAP overexpressing transgenic rice roots and response to 1-naphthyl acetic acid (NAA). (A) Gravitropism phenotype of wild type (WT) and OsAGAP overexpressing roots at 6 hr gravi-stimulation (top panel) and 0 hr as a treatment control (bottom panel). (B) Time course of gravitropic response in transgenic roots. (C and D) results correspond to those in (A and B), except for treatment with NAA (5 × 10−7 M).The polarity of auxin transport is controlled by the asymmetric distribution of auxin transport proteins, efflux facilitators and influx carriers. ARF GTPase is a key member in vesicle trafficking system and modulates cell polarity and PAT in plants. Thus, ARF-GDP or GTP bound with GEF or GAP determines the ARF function on auxin efflux facilitators (such as PIN1) or influx ones (such as AUX1).ARF1, targeting ROP2 and PIN2, affects epidermal cell polarity.9 GNOM is involved in the regulation of PIN1 asymmetric localization in cells and its related function in organogenesis and development.6 Although VAN3, an ARF-GAP in Arabidopsis, is located in a subpopulation of the trans-Golgi transport network (TGN), which is involved in leaf vascular network formation, it does not affect PAT.10 OsAGAP possesses an ARF GTPase-activating function in rice.11 Specifically, our evidence supports that ARF-GAP bound with ARF-GTP modulates PAT and gravitropism via AUX1, mediated by vesicle trafficking, including the Golgi stack.1Therefore, we propose a loop mechanism between ARF-GAP and GEF mediated by the vascular trafficking system in regulating PAT at influx and efflux facilitators, which controls root development and gravitropism in plants (Fig. 2). Here we emphasize that ARF-GEF catalyzes a conversion of ARF-bound GDP to GTP, which is necessary for the efficient delivery of the vesicle to the target membrane.12 An opposite process of ARF-bound GDP to GTP is promoted by ARF-GTPase-activating protein via binding. A loop status of ARF-GTP and ARF-GDP bound with their appurtenances controls different auxin facilitators and regulates root development and gravitropism.Open in a separate windowFigure 2Model for ARF GTPase as a molecular switch for the polar auxin transport mediated by the vesicle traffic system.  相似文献   

20.
A role for SR proteins in plant stress responses   总被引:1,自引:0,他引:1  
  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