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.
Organization of multiprotein complexes at cell–cell junctions   总被引:1,自引:1,他引:0  
The formation of stable cell-cell contacts is required for the generation of barrier-forming sheets of epithelial and endothelial cells. During various physiological processes like tissue development, wound healing or tumorigenesis, cellular junctions are reorganized to allow the release or the incorporation of individual cells. Cell-cell contact formation is regulated by multiprotein complexes which are localized at specific structures along the lateral cell junctions like the tight junctions and adherens junctions and which are targeted to these site through their association with cell adhesion molecules. Recent evidence indicates that several major protein complexes exist which have distinct functions during junction formation. However, this evidence also indicates that their composition is dynamic and subject to changes depending on the state of junction maturation. Thus, cell-cell contact formation and integrity is regulated by a complex network of protein complexes. Imbalancing this network by oncogenic proteins or pathogens results in barrier breakdown and eventually in cancer. Here, I will review the molecular organization of the major multiprotein complexes at junctions of epithelial cells and discuss their function in cell-cell contact formation and maintenance.  相似文献   

2.
Epithelial sheets often present a “cobblestone” appearance, but the mechanisms underlying the dynamics of this arrangement are unclear. In this issue, Choi et al. (2016. J. Cell Biol. http://dx.doi.org/10.1083/jcb.201506115) show that afadin and ZO-1 regulate tension and maintain zonula adherens architecture in response to changes in contractility.The textbook view of epithelial cells is that once such cells adopt a close, hexagonal packing, their “honeycomb” or “cobblestone” arrangement is static. This fixed appearance is misleading, as these cells are more like players in a rugby scrum, locked in a tussle in which the forces exerted by each of the players on the others maintains their seemingly static arrangement, but by a very dynamic force balance. How such balance is maintained in epithelia is a subject of substantial interest. A crucial role is played by F-actin and nonmuscle myosin II isoforms, which are deployed in contractile networks that transiently attach to cell–cell junctions to generate tensile forces along cell–cell boundaries (Lecuit and Yap, 2015). Contractile arrays of actomyosin are regulated by the monomeric G protein Rho, its upstream regulators, including Rho guanine nucleotide exchange factors (Quiros and Nusrat, 2014), and its effectors ROCK/Rho kinase and Shroom3 (Nishimura and Takeichi, 2008), but also by tension-mediated feedback between the myosin network and the junction (Lecuit and Yap, 2015). Cell–cell adhesion, including cadherin-dependent adhesion, also plays a crucial role in this process. As cells engage with one another via interactions of the extracellular domains of their cadherin complexes, they transduce forces to the actomyosin cytoskeleton through catenins. β-Catenin binds to the cytoplasmic domain of classical cadherins and recruits α-catenin, which binds F-actin.Given the dynamic nature of epithelia, the attachment of contractile actomyosin networks to junctions are also subject to regulation. One aspect of epithelial architecture that has received relatively little attention is that a typical epithelial monolayer (Fig. 1 A) displays two main types of cell–cell interfaces: bilateral junctions (BCJs), in which two cells establish a relatively long stretch of contact, and cellular vertices, which represent a confluence of three or more cell edges to form tricellular junctions (TCJs) or multicellular junctions. TCJs are not well understood, but are known to contain unique molecular components (Furuse et al., 2014; Flores-Benitez and Knust, 2015). In this issue, Choi et al. show that the multivalent scaffolding proteins afadin and ZO-1/2 regulate the spacing of and tension along lateral contacts in cultured cells, thereby shedding light on how contractile arrays containing bilateral and tri- or multicellular contact points are regulated in epithelia.Open in a separate windowFigure 1.ZO proteins and afadin regulate junctional tension and organization in cultured cells. (A) Untreated MDCK cells have sinuous cell boundaries, whereas ZO KD cells show extremely straight boundaries. When ZO proteins and afadin are knocked down, cells adopt contact zones of irregular length with other cells, sometimes clustering into foci (asterisks). Images courtesy of Mark Peifer (Choi et al., 2016). (B) A model for actomyosin organization at adherens junctions (adapted from Choi et al., 2016). Contractile actomyosin arrays run parallel to bicellular junctions and are anchored by side-on attachments (pink circles). At TCJs, end-on binding of actin, likely stabilized by afadin, anchors actomyosin filaments. In ZO KD cells, contractile elements and cadherin complexes collapse toward TCJs, and myosin minifilaments adopt a regularly spaced arrangement.Afadin and ZO-1/2 are far from new players at junctions. Afadin binds α-catenin, actin, and other cytoskeletal and junctional proteins and associates with the transmembrane protein nectin, which appears to form an alternative adhesion system at adherens junctions (Mandai et al., 2013). The zonula occludens proteins ZO-1 and ZO-2 are tight junction proteins that bind claudins and are required for tight junction formation (Itoh et al., 1999; Balda and Matter, 2008). In addition, ZO proteins also bind to α-catenin (Itoh et al., 1997), are involved in establishing the zonula adherens (ZA; Ikenouchi et al., 2007), and potentiate cadherin-dependent adhesion in Caenorhabditis elegans (Lockwood et al., 2008) and Drosophila melanogaster (Choi et al., 2011). Knockdown of ZO-1 and ZO-2 (ZO KD) in MDCK cells has previously been shown (Fanning et al., 2012) to lead to dramatic alterations of the ZA: F-actin and myosin IIs assemble into striking apical arrays at the ZA, spaced at regular intervals. In addition, the normally sinuous boundaries between cells give way to very straight borders (Fig. 1 A).Using superresolution microscopy, diffraction-limited junctional laser ablation, cell morphometry, kinetic analysis, and a whole-monolayer approach to contractility, Choi et al. (2016) now extend this story. To test whether contractility is increased after ZO KD, the authors first measured the recoil after laser ablation of ZO KD cells; an increase in recoil velocity indicated that the straight junctional boundaries between ZO-depleted cells are under tension. Imaging analysis of BCJs showed that the increase in contractility in ZO KD cells is associated with a strikingly dynamic behavior of the BCJs. Individual BCJs were found to undergo periods of shortening and elongation, whereas neighboring BCJs underwent compensatory, opposite changes in length. These changes in contractility have effects on the entire tissue sheet as well: whereas control cell sheets remained flat when detached from the substratum, ZO KD cells contracted into a cup-like shape. This constriction was blocked by the myosin inhibitor blebbistatin. Overall, these experiments indicated that ZO proteins regulate myosin assembly and contractility across the cellular sheet.To dissect the protein network mediating increased contractility in ZO KD cells, Choi et al. (2016) examined the role of ROCK and found that ROCK inhibitors abolished the straight BCJs, which became curvilinear. Additionally, Shroom3, which is known to recruit ROCK (Nishimura and Takeichi, 2008), was cytoplasmic in control cells but junctional in ZO KD cells. Transient Shroom3 overexpression led to ROCK recruitment to the ZA and drove formation of an actomyosin network similar to that in ZO KD cells. Conversely, Shroom3 knockdown resulted in loss of the actomyosin arrays in ZO KD cells. Collectively, these data indicated that Shroom3 is an effector of increased apical contractility in ZO KD cells.The researchers used ZO KD cells to test how tissue integrity is maintained despite elevated contractibility and how junctions are remodeled to maintain integrity when increased tension is present. Afadin is a good candidate: the Drosophila homologue of afadin, Canoe, plays roles in convergent extension and collective cell migration; in its absence, actomyosin networks at the apex of constricting epithelial cells in the embryo contract in a catastrophic, uncontrolled fashion (Sawyer et al., 2009), suggesting a potential role for afadin in the maintenance of tissue integrity during morphogenetic movements. Choi et al. (2016) therefore turned their attention to afadin. ZO KD cells have significantly more afadin at their adherens junctions and TCJs, a pattern reminiscent of the normal distribution of Canoe in Drosophila (Sawyer et al., 2009). Knocking down afadin by shRNA in ZO KD cells led to further defects in cell–cell boundary maintenance. In addition to the taut appearance of bicellular borders, cell boundary length became much more irregular, with occasional foci of highly contracted cells (Fig. 1 A). Velocimetry analysis and live-cell imaging indicated that loss of both ZO proteins and afadin led to large-scale cell movements within the monolayer not seen after ZO KD alone.New imaging techniques used by Choi et al. (2016) revealed further details about the changes in actomyosin arrays in ZO KD cells. Superresolution imaging of myosin light chain kinase staining via structured illumination showed that myosin II assembles into arrays of myosin minifilaments spaced ∼415 nm apart along bicellular contacts. Superresolution and transmission electron microscopy also revealed reorganization of F-actin and E-cadherin at TCJs in ZO KD cells. Lateral F-actin bundles appeared to terminate end-on at TCJs at sites where E-cadherin was present. ZO KD therefore induces assembly of a remarkably ordered actomyosin array along BCJs, and these arrays appear to be separate contractile units that anchor end-on at the ZA. Moreover, based on staining for vinculin and a specific epitope in αE-catenin that serve as markers for regions under high tension (Yonemura et al., 2010), the end-on attachments of actin cables to the ZA at TCJs experience significant tensile stress. Strikingly, although vinculin and αE-catenin accumulation at TCJs was relatively uniform after ZO KD, their distribution was more heterogeneous after ZO/afadin KD. Differences in staining paralleled differences in cell border length and correlated with the level of tension measured at BCJs after laser cutting, suggesting that afadin contributes to the ability of cells to distribute forces at TCJ/multicellular junctions throughout the monolayer. Lastly, the researchers investigated whether internal cues downstream of ZO KD are sufficient for myosin recruitment or whether such recruitment depends on mechanical cues exerted by neighboring cells. They designed an assay mixing small islands of wild-type cells surrounded by ZO KD cells (or vice versa) and found that the development of the contractile array at the ZA depends on the contractility of neighboring cells; however, afadin recruitment to the ZA was less dependent on the sustained contractility of neighboring cells. Taking these data together, Choi et al. (2016) propose that cells respond to elevated contractility by increasing junctional afadin; because combined ZO/afadin knockdown dramatically alters cell shape and barrier function in response to elevated contractility, afadin acts as a robust scaffold that maintains ZA architecture most crucially at TCJs.Although many aspects of the model proposed by Choi et al. (2016) remain to be tested, their data suggest new features regarding the detailed assembly of actomyosin contractile arrays in confluent cells (Fig. 1 B). In control cells, actomyosin arrays presumably extend parallel to individual BCJs. Choi et al. (2016) propose that these actomyosin bundles act as separate contractile units that terminate near TCJs, allowing the generation of tension along BCJs. In ZO KD cells, excessive assembly of actomyosin filaments, perhaps exacerbated by the tendency of F-actin/myosin minifilament arrays to self-assemble, somehow leads to regularly spaced actomyosin arrays, and perhaps collapse of cadherin complexes and other components toward TCJs. There is a precedent for such lateral collapse of cadherin-dependent attachments: it is a prominent feature of cadherin complexes at sites of high tension in the epidermis of the C. elegans embryo (Choi et al., 2015). If the new model of Choi et al. (2016) is correct, then the foci seen in ZO KD/afadin KD cells may be similar to what happens in a game of tug of war when one team stops pulling. If some end-on attachments (assisted by afadin) fail, filaments might be expected to collapse along BCJs as the other, still tethered end of a set of filaments contracts toward the remaining attachment at the opposite cell vertex.Several other interesting questions remain. First, what is the relationship of the striking, regularly spaced bipolar myosin II minifilaments that form in ZO KD cells to myosin arrays in normal cells? It is clear that untreated cells have junctional actomyosin networks, but not with this strict periodicity. One possibility is that this spacing is simply an epiphenomenon; when not appropriately anchored along junctions, actomyosin networks may self-organize as they are known to do in other systems, such as in the contractile ring and in migrating cells (Srivastava et al., 2015; Fenix et al., 2016). More optimistically, the spacing may represent an intensified version of processes that operate in normal cells at bicellular and multicellular contact sites. If so, components of the model of Choi et al. (2016) will require further investigation. For example, the organization of F-actin along BCJs remains unclear, as are the proteins that mediate the side-on binding envisioned in this model. It is also uncertain whether proteins assist bundling of filaments and what role dynamic growth and shrinkage of actin filaments plays in end-on binding. In some contexts, junctions are capable of seeding polymerization of F-actin (Brieher and Yap, 2013), and it may be that actin dynamics are important in the processes studied here.A second question has to do with the community events within monolayers that Choi et al. (2016) describe. The neighbor effects on ZA morphology that they document are intriguing, as are the long-range accelerated movements of cells lacking both ZO proteins and afadin. Collective properties of monolayers are only beginning to be explored; connecting these properties with subcellullar events is an exciting future challenge. Whatever the answers to these new questions, the work of Choi et al. (2016) refines our understanding of the roles of key scaffolding proteins in organizing and anchoring junctions in epithelia.  相似文献   

3.
4.
Summary Seminiferous tubules of 1-day-old rats were maintained in organ culture for up to 40 days. Five classes of intercellular junctions between Sertoli cells were observed by the freeze-fracture method as the tissue aged: (a) typical gap junctions; (b) focal tight junctions; (c) macular tight junctions; (d) meandering tight junctions; and (e) extensive tight junctions. The relative proportions of these types of Sertoli cell junctions were quantitated as the organ cultures progressed. The junctional structures observed and classified in organ culture were identical to those seen in vivo, but the timing of their appearance and/or disappearance, as well as their relative proportions, was different from that observed in the developing animal. Extensive tight junctions, with numerous parallel strands, were observed in the 40-day cultures; however, their oblique orientation with respect to the myoid layer was in contrast to the parallel orientation observed in vivo. This study was supported by Grant 801D185 (Dr. Posalaky) from the Medical Education and Research Foundation, St. Paul-Ramsey Hospital.  相似文献   

5.
Shen S  Kepp O  Kroemer G 《Autophagy》2012,8(1):1-3
In the mammalian system, cell death is often preceded or accompanied by autophagic vacuolization, a finding that initially led to the widespread belief that so-called "autophagic cell death" would be mediated by autophagy. Thanks to the availability of genetic tools to disable the autophagic machinery, it has become clear over recent years that autophagy usually constitutes a futile attempt of dying cells to adapt to lethal stress rather than a mechanism to execute a cell death program. Recently, we systematically addressed the question as to whether established or prospective anticancer agents may induce "autophagic cell death". Although a considerable portion among the 1,400 compounds that we evaluated induced autophagic puncta and actually increased autophagic flux, not a single one turned out to kill tumor cells through the induction of autophagy. Thus, knockdown of essential autophagy genes (such as ATG5 and ATG7) failed to prevent and rather accelerated chemotherapy-induced cell death, in spite of the fact that this manipulation efficiently inhibits autophagosome formation. Herein, we review these finding and--polemically--raise doubts as to the very existence of "autophagic cell death".  相似文献   

6.
Immunocytochemical, electron-, and immunoelectron-microscopical studies have revealed that, in addition to the four major “textbook categories” of cell-cell junctions (gap junctions, tight junctions, adherens junctions, and desmosomes), a broad range of other junctions exists, such as the tiny puncta adhaerentia minima, the taproot junctions (manubria adhaerentia), the plakophilin-2-containing adherens junctions of mesenchymal or mesenchymally derived cell types including malignantly transformed cells, the composite junctions (areae compositae) of the mature mammalian myocardium, the cortex adhaerens of the eye lens, the interdesmosomal “sandwich” or “stud” junctions in the subapical layers of stratified epithelia and the tumors derived therefrom, and the complexus adhaerentes of the endothelial and virgultar cells of the lymph node sinus. On the basis of their sizes and shapes, other morphological criteria, and their specific molecular ensembles, these junctions and the genes that encode them cannot be subsumed under one of the major categories mentioned above but represent special structures in their own right, appear to serve special functions, and can give rise to specific pathological disorders.  相似文献   

7.
The skeletal muscle satellite cell: stem cell or son of stem cell?   总被引:18,自引:0,他引:18  
The concept of the adult tissue stem cell is fundamental to models of persistent renewal in functionally post-mitotic tissues. Although relatively ignored by stem cell biology, skeletal muscle is a prime example of an adult tissue that can generate terminally differentiated cells uniquely specialized to carry out tissue-specific functions. This capacity is attributed to satellite cells, a population of undifferentiated, quiescent precursors that become activated to divide and differentiate in response to the demands of growth or damage. The aim of this review is to discuss the role of the satellite cell as an adult tissue-specific stem cell. We examine evidence for the presence of behaviourally and phenotypically distinct subpopulations of precursor within the satellite cell pool. Further, we speculate on the possible identity, origins and relevance of multipotent muscle stem cells, a population with both myogenic and hematopoietic potentials that has been isolated from whole muscle. Taken together, current evidence suggests the possibility that the regenerative compartment of adult skeletal muscle may conform to an archetypal stem cell-based hierarchy, maintained within a stem cell niche. It therefore remains to be seen whether all satellite cells are skeletal muscle-specific stem cells, or whether some or all are the progeny of an as yet unidentified muscle stem cell.  相似文献   

8.
We prepared encapsulated whole cell β-galactosidase using E. coli. The cell culture was divided into two steps for the cell accumulation inside the capsule and enzyme production in the cell. Growth and production media were used individually for this purpose. The dry cell weight of the free cell culture was increased 2.8 times by controlling the pH of the growth medium during cultivation. However, the weight of cells accumulated in the capsule reduced 40% with pH control. The dry cell weight increased with lactose concentration of the production medium for both cases of free and capsule cultures. The dry cell weights were 1.5?g/l for free culture and 100?g/l in the capsule when the lactose concentration of the production medium was 10?g/l. The dry cell weight increased about 60% for both cases as the lactose concentration increased from 10 to 50?g/l. The specific activity of whole cell enzyme decreased with lactose concentration from 5 to 1.4?unit/g dry cell for free culture and from 1.1 to 0.65?unit/g dry cell in the capsule. The value of Michaelis constant, Km, of whole cell enzyme increased 3 times because of the resistance of mass transfer through the capsule membrane. The constants of Michaelis-Menten equation for the whole cell enzyme in the capsule were Vm: 0.0479?mM/min and Km: 44.86?mM. These constants of the membrane-free cells were Vm: 0.0464?mM/min and Km: 15.64?mM. To increase the whole cell enzyme activity, we treated encapsulated cells with organic solvents. The activity of encapsulated whole cell enzyme was increased 3.5 times with the treatment of chloroform and ethanol. The activity of the encapsulated whole cell enzymes was reserved after repeating the process 30 times.  相似文献   

9.
Research in many fields of biology has been extremely successful in decomposing biological mechanisms to discover their parts and operations. It often remains a significant challenge for scientists to recompose these mechanisms to understand how they function as wholes and interact with the environments around them. This is true of the eukaryotic cell. Although initially identified in nineteenth-century cell theory as the fundamental unit of organisms, researchers soon learned how to decompose it into its organelles and chemical constituents and have been highly successful in understanding how these carry out many operations important to life. The emphasis on decomposition is particularly evident in modern cell biology, which for the most part has viewed the cell as merely a locus of the mechanisms responsible for vital phenomena. The cell, however, is also an integrated system and for some explanatory purposes it is essential to recompose it and understand it as an organized whole. I illustrate both the virtues of decomposition (treating the cell as a locus) and recomposition (treating the cell as an object) with recent work on circadian rhythms. Circadian researchers have both identified critical intracellular operations that maintain endogenous oscillations and have also addressed the integration of cells into multicellular systems in which cells constitute units.  相似文献   

10.
11.
Santangelo F 《Amino acids》2002,23(4):359-365
Summary.  The diet of industrialised countries is usually rich in amino acids, which are in part used as a source of calories. However, metabolic alterations are observed in diseased patients and a preferential retention of Sulphurated Amino Acids (SAA) occurs during the inflammatory response. Moreover, it has been demonstrated in a model of an acute sepsis phase of rats that the metabolism of Cysteine is modified. The liver converts Cysteine at a different ratio of Sulphate to Taurine (Tau) i.e. the sulphate production decreases while the Tau conversion increases. The Glutathione (GSH) concentration is greater in the liver, kidneys and other organs and the Cysteine incorporation into proteins is higher in the spleen, lungs and plasma (Acute Phase Proteins) while the Albumin level decreases. The pro-inflammatory cytokines such as Interleukin-1, Interleukin-6 and TNF-α are the main initiators that alter protein and amino acid metabolism. Another important phenomenon is the impairment of Methionine conversion to Cysteine during stress. For example, premature infants or AIDS patients are capable of synthesizing Cysteine from Methionine at a much lower rate. Thus, the metabolic flow through the trans-sulphuration path may be inadequate to meet the Cysteine demand under critical conditions. In this complex picture, an SAA supply may contribute to an immune system regulation. Received November 30, 2001 Accepted January 15, 2002 Published online August 30, 2002 Author's address: Francesco Santangelo, Zambon Group, via Lillo del Duca 10, I-20091 Bresso, Milan, Italy, Email: francesco.santangelo@zambongroup.com  相似文献   

12.
This study investigates how epithelial cells moving together function to coordinate their collective movement to repair a wound. Using a lens ex vivo mock cataract surgery model we show that region-specific reorganization of cell–cell junctions, cytoskeletal networks and myosin function along apical and basal domains of an epithelium mediates the process of collective migration. An apical junctional complex composed of N-cadherin/ZO-1/myosin II linked to a cortical actin cytoskeleton network maintains integrity of the tissue during the healing process. These cells’ basal domains often preceded their apical domains in the direction of movement, where an atypical N-cadherin/ZO-1 junction, linked to an actin stress fiber network rich in phosphomyosin, was prominent in cryptic lamellipodia. These junctions joined the protruding forward-moving lamellipodia to the back end of the cell moving directly in front of it. These were the only junctions detected in cryptic lamellipodia of lens epithelia migrating in response to wounding that could transmit the protrusive forces that drive collective movement. Both integrity of the epithelium and ability to effectively heal the wound was found to depend on myosin mechanical cues.  相似文献   

13.
Impaired endothelial barrier function results in a persistent increase in endothelial permeability and vascular leakage. Repair of a dysfunctional endothelial barrier requires controlled restoration of adherens junctions, comprising vascular endothelial (VE)-cadherin and associated β-, γ-, α-, and p120-catenins. Little is known about the mechanisms by which recovery of VE-cadherin–mediated cell–cell junctions is regulated. Using the inflammatory mediator thrombin, we demonstrate an important role for the Src homology 2-domain containing tyrosine phosphatase (SHP2) in mediating recovery of the VE-cadherin–controlled endothelial barrier. Using SHP2 substrate-trapping mutants and an in vitro phosphatase activity assay, we validate β-catenin as a bona fide SHP2 substrate. SHP2 silencing and SHP2 inhibition both result in delayed recovery of endothelial barrier function after thrombin stimulation. Moreover, on thrombin challenge, we find prolonged elevation in tyrosine phosphorylation levels of VE-cadherin–associated β-catenin in SHP2-depleted cells. No disassembly of the VE-cadherin complex is observed throughout the thrombin response. Using fluorescence recovery after photobleaching, we show that loss of SHP2 reduces the mobility of VE-cadherin at recovered cell–cell junctions. In conclusion, our data show that the SHP2 phosphatase plays an important role in the recovery of disrupted endothelial cell–cell junctions by dephosphorylating VE-cadherin–associated β-catenin and promoting the mobility of VE-cadherin at the plasma membrane.  相似文献   

14.
15.
Mechanical stimuli play important roles in proliferation and differentiation of connective tissue cells, and development and homeostatic maintenance of tissues. However, excessive mechanical loading to a tissue can injure cells and disrupt the matrix, as occurs in tendinopathy. Tendinopathy is a common clinical problem in athletes and in many occupational settings due to overuse of the tendon. Moreover, interleukin (IL)-1β is generally considered to be a "bad" cytokine, activating NF-κb and cell death and inducing matrix metalloproteinase (MMPs 1, 2, 3) expression and matrix destruction. However, activated NF-κB can also drive a cell survival pathway. We have reported that cyclic strain induced tenocyte death in three-dimensional (3D) cultures, and IL-1β could promote cell survival under strain. Therefore, it was hypothesized that 1) cyclic strain could induce cell death in tenocytes as observed in pathologic tendons in vivo; 2) a gene expression profile indicative of tendinopathy could be identified; and 3) low-dose IL-1β could protect cells from strain-induced, tendinopathy-like changes. Human tenocytes were cultured in 3D type I collagen hydrogels and subjected to 3.5% elongation at 1 Hz for 1 h/day for up to 5 days with or without IL-1β. Real-time RT-PCR data showed that cyclic strain regulated the expression of tendinopathy marker genes in a manner similar to that found in pathological tendons from patients and that addition of IL-1β reversed the gene expression changes to control levels. Results of further studies showed that IL-1β may modulate cell survival through upregulating the expression of connexin 43, which is involved in the modulation of cell death/survival in a variety of cells and tissues. The elucidation of the mechanisms underlying strain-induced cell death and recovery from strain injury will facilitate our understanding of the pathogenesis of tendinopathy and may lead to the discovery of new molecular targets for early diagnosis and treatment of tendinopathy.  相似文献   

16.
Cadherin-mediated cell–cell adhesion is required for epithelial tissue integrity in homeostasis, during development, and in tissue repair. E-cadherin stability depends on F-actin, but the mechanisms regulating actin polymerization at cell–cell junctions remain poorly understood. Here we investigated a role for formin-mediated actin polymerization at cell–cell junctions. We identify mDia1 and Fmnl3 as major factors enhancing actin polymerization and stabilizing E-cadherin at epithelial junctions. Fmnl3 localizes to adherens junctions downstream of Src and Cdc42 and its depletion leads to a reduction in F-actin and E-cadherin at junctions and a weakening of cell–cell adhesion. Of importance, Fmnl3 expression is up-regulated and junctional localization increases during collective cell migration. Depletion of Fmnl3 or mDia1 in migrating monolayers results in dissociation of leader cells and impaired wound repair. In summary, our results show that formin activity at epithelial cell–cell junctions is important for adhesion and the maintenance of epithelial cohesion during dynamic processes, such as wound repair.  相似文献   

17.
Intercellular junctions are crucial for mechanotransduction, but whether tight junctions contribute to the regulation of cell–cell tension and adherens junctions is unknown. Here, we demonstrate that the tight junction protein ZO-1 regulates tension acting on VE-cadherin–based adherens junctions, cell migration, and barrier formation of primary endothelial cells, as well as angiogenesis in vitro and in vivo. ZO-1 depletion led to tight junction disruption, redistribution of active myosin II from junctions to stress fibers, reduced tension on VE-cadherin and loss of junctional mechanotransducers such as vinculin and PAK2, and induced vinculin dissociation from the α-catenin–VE-cadherin complex. Claudin-5 depletion only mimicked ZO-1 effects on barrier formation, whereas the effects on mechanotransducers were rescued by inhibition of ROCK and phenocopied by JAM-A, JACOP, or p114RhoGEF down-regulation. ZO-1 was required for junctional recruitment of JACOP, which, in turn, recruited p114RhoGEF. ZO-1 is thus a central regulator of VE-cadherin–dependent endothelial junctions that orchestrates the spatial actomyosin organization, tuning cell–cell tension, migration, angiogenesis, and barrier formation.  相似文献   

18.
19.
Localisation of -amylase (EC 3.2.1.1.) in low-temperature-embedded isolated barley (Hordeum vulgare L.) aleurone has been achieved using rhodamine-labelled secondary antibodies and the protein A-gold technique. Treatment with gibberellic acid (GA3) resulted in an increase of immunofluorescence in the cytoplasm of aleurone cells and also its appearance in specific regions of the cell walls. Cytoplasmic label was neither perinuclear nor associated specifically with aleurone grains as had been found in earlier work, but was present throughout the cytoplasm of all cells. A relatively high level of labelling occurred in hydrolysed wall regions. Label was also associated with plasmodesmata in both hydrolysed and unhydrolysed wall regions. The pattern of labelling indicates that -amylase is released from aleurone via digested wall channels and that, except for the inner wall layer, unhydrolysed regions are impermeable to the enzyme. It is suggested that the resistant wall tubes around plasmodesmata may facilitate enzyme release by providing a pathway for transfer, especially of wall hydrolases, into the more impermeable parts of the wall.Abbreviations ER endoplasmic reticulum - GA3 gibberellic acid - RER rough endoplasmic reticulum  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号