Journal articles on the topic 'Cell-to-cell spreading'

To see the other types of publications on this topic, follow the link: Cell-to-cell spreading.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Cell-to-cell spreading.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Mothes, Walther, Nathan M. Sherer, Jing Jin, and Peng Zhong. "Virus Cell-to-Cell Transmission." Journal of Virology 84, no. 17 (April 7, 2010): 8360–68. http://dx.doi.org/10.1128/jvi.00443-10.

Full text
Abstract:
ABSTRACT Viral infections spread based on the ability of viruses to overcome multiple barriers and move from cell to cell, tissue to tissue, and person to person and even across species. While there are fundamental differences between these types of transmissions, it has emerged that the ability of viruses to utilize and manipulate cell-cell contact contributes to the success of viral infections. Central to the excitement in the field of virus cell-to-cell transmission is the idea that cell-to-cell spread is more than the sum of the processes of virus release and entry. This implies that virus release and entry are efficiently coordinated to sites of cell-cell contact, resulting in a process that is distinct from its individual components. In this review, we will present support for this model, illustrate the ability of viruses to utilize and manipulate cell adhesion molecules, and discuss the mechanism and driving forces of directional spreading. An understanding of viral cell-to-cell spreading will enhance our ability to intervene in the efficient spreading of viral infections.
APA, Harvard, Vancouver, ISO, and other styles
2

Salsmann, Alexandre, Elisabeth Schaffner-Reckinger, and Nelly Kieffer. "RGD, the Rho’d to cell spreading." European Journal of Cell Biology 85, no. 3-4 (April 2006): 249–54. http://dx.doi.org/10.1016/j.ejcb.2005.08.003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

McGrath, James L. "Cell Spreading: The Power to Simplify." Current Biology 17, no. 10 (May 2007): R357—R358. http://dx.doi.org/10.1016/j.cub.2007.03.057.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Vilette, Didier, Josquin Courte, Jean Michel Peyrin, Laurent Coudert, Laurent Schaeffer, Olivier Andréoletti, and Pascal Leblanc. "Cellular mechanisms responsible for cell-to-cell spreading of prions." Cellular and Molecular Life Sciences 75, no. 14 (May 14, 2018): 2557–74. http://dx.doi.org/10.1007/s00018-018-2823-y.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Vilette, Didier, Josquin Courte, Jean Michel Peyrin, Laurent Coudert, Laurent Schaeffer, Olivier Andréoletti, and Pascal Leblanc. "Correction to: Cellular mechanisms responsible for cell-to-cell spreading of prions." Cellular and Molecular Life Sciences 75, no. 14 (June 11, 2018): 2575. http://dx.doi.org/10.1007/s00018-018-2853-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Colin, Morvane, Meryem Tardivel, Séverine Bégard, Luc Bousset, Simon Dujardin, Kevin Richetin, Nicole Deglon, Ronald Melki, and Luc Buee. "TAU SPREADING: HOW ARE TAU ASSEMBLIES TRANSFERRED FROM CELL TO CELL?" Alzheimer's & Dementia 13, no. 7 (July 2017): P906. http://dx.doi.org/10.1016/j.jalz.2017.07.320.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

McInnes, C., P. Knox, and D. J. Winterbourne. "Cell spreading on serum is not identical to spreading on fibronectin." Journal of Cell Science 88, no. 5 (December 1, 1987): 623–29. http://dx.doi.org/10.1242/jcs.88.5.623.

Full text
Abstract:
Adhesion and spreading of cell lines on dishes coated with serum-derived proteins were studied after removal of cell-surface proteoglycans. A mixture of glycosaminoglycans lyases from heparin-induced Flavobacterium heparinum removed 80% of the [35S]sulphate-labelled glycosaminoglycans from the surface of attached cells within 30 min, but this had little effect on cell morphology. The rate of cell attachment to dishes coated with serum was unaffected by prior treatment of cells with this mixture of glycosaminoglycan lyases. While a heparan sulphate lyase preparation abolished cell spreading in response to fibronectin there was no effect of the enzyme on the spreading mediated by vitronectin. These results suggest that, although heparan sulphate is required for spreading on purified fibronectin, the spreading stimulated by serum under routine culture conditions requires neither cellular heparan sulphate nor serum-derived fibronectin.
APA, Harvard, Vancouver, ISO, and other styles
8

Mohamed, Nguyen-Vi, Thibaut Herrou, Vanessa Plouffe, Nicolas Piperno, and Nicole Leclerc. "Spreading of tau pathology in Alzheimer's disease by cell-to-cell transmission." European Journal of Neuroscience 37, no. 12 (June 2013): 1939–48. http://dx.doi.org/10.1111/ejn.12229.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Kerkeni, W., Y. Ayari, L. Charfi, A. Bouzouita, H. Ayed, M. Cherif, M. R. Ben Slama, K. Mrad, A. Derouiche, and M. Chebil. "Transitional Bladder Cell Carcinoma Spreading to the Skin." Urology Case Reports 11 (February 2017): 17–18. http://dx.doi.org/10.1016/j.eucr.2016.11.028.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Nisenholz, Noam, Aishwarya Paknikar, Sarah Köster, and Assaf Zemel. "Contribution of myosin II activity to cell spreading dynamics." Soft Matter 12, no. 2 (2016): 500–507. http://dx.doi.org/10.1039/c5sm01733e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

English, C. S., C. K. Puk, F. A. Petrigliano, B. M. Wu, and D. R. McAllister. "Ligament Engineering: The Contribution of Stretch to Cell Spreading." Journal of Investigative Medicine 54, no. 1_suppl (January 2006): 116–17. http://dx.doi.org/10.1177/108155890605401s97.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Doherty, Gary J., Monika K. Åhlund, Mark T. Howes, Björn Morén, Robert G. Parton, Harvey T. McMahon, and Richard Lundmark. "The endocytic protein GRAF1 is directed to cell-matrix adhesion sites and regulates cell spreading." Molecular Biology of the Cell 22, no. 22 (November 15, 2011): 4380–89. http://dx.doi.org/10.1091/mbc.e10-12-0936.

Full text
Abstract:
The rho GTPase-activating protein GTPase regulator associated with focal adhesion kinase-1 (GRAF1) remodels membranes into tubulovesicular clathrin-independent carriers (CLICs) mediating lipid-anchored receptor endocytosis. However, the cell biological functions of this highly prevalent endocytic pathway are unclear. In this article, we present biochemical and cell biological evidence that GRAF1 interacted with a network of endocytic and adhesion proteins and was found enriched at podosome-like adhesions and src-induced podosomes. We further demonstrate that these sites comprise microdomains of highly ordered lipid enriched in GRAF1 endocytic cargo. GRAF1 activity was upregulated in spreading cells and uptake via CLICs was concentrated at the leading edge of migrating cells. Depletion of GRAF1, which inhibits CLIC generation, resulted in profound defects in cell spreading and migration. We propose that GRAF1 remodels membrane microdomains at adhesion sites into endocytic carriers, facilitating membrane turnover during cell morphological changes.
APA, Harvard, Vancouver, ISO, and other styles
13

Woodring, Pamela J., Jill Meisenhelder, Sam A. Johnson, Guo-Lei Zhou, Jeffrey Field, Kavita Shah, Friedhelm Bladt, et al. "c-Abl phosphorylates Dok1 to promote filopodia during cell spreading." Journal of Cell Biology 165, no. 4 (May 17, 2004): 493–503. http://dx.doi.org/10.1083/jcb.200312171.

Full text
Abstract:
Filopodia are dynamic F-actin structures that cells use to explore their environment. c-Abl tyrosine kinase promotes filopodia during cell spreading through an unknown mechanism that does not require Cdc42 activity. Using an unbiased approach, we identified Dok1 as a specific c-Abl substrate in spreading fibroblasts. When activated by cell adhesion, c-Abl phosphorylates Y361 of Dok1, promoting its association with the Src homology 2 domain (SH2)/SH3 adaptor protein Nck. Each signaling component was critical for filopodia formation during cell spreading, as evidenced by the finding that mouse fibroblasts lacking c-Abl, Dok1, or Nck had fewer filopodia than cells reexpressing the product of the disrupted gene. Dok1 and c-Abl stimulated filopodia in a mutually interdependent manner, indicating that they function in the same signaling pathway. Dok1 and c-Abl were both detected in filopodia of spreading cells, and therefore may act locally to modulate actin. Our data suggest a novel pathway by which c-Abl transduces signals to the actin cytoskeleton through phosphorylating Dok1 Y361 and recruiting Nck.
APA, Harvard, Vancouver, ISO, and other styles
14

Runyan, R. B., J. Versalovic, and B. D. Shur. "Functionally distinct laminin receptors mediate cell adhesion and spreading: the requirement for surface galactosyltransferase in cell spreading." Journal of Cell Biology 107, no. 5 (November 1, 1988): 1863–71. http://dx.doi.org/10.1083/jcb.107.5.1863.

Full text
Abstract:
The molecular mechanisms underlying cell attachment and subsequent cell spreading on laminin are shown to be distinct form one another. Cell spreading is dependent upon the binding of cell surface galactosyltransferase (GalTase) to laminin oligosaccharides, while initial cell attachment to laminin occurs independent of GalTase activity. Anti-GalTase IgG, as well as the GalTase modifier protein, alpha-lactalbumin, both block GalTase activity and inhibited B16-F10 melanoma cell spreading on laminin, but not initial attachment. On the other hand, the addition of UDP galactose, which increases the catalytic turnover of GalTase, slightly increased cell spreading. None of these reagents had any effect on cell spreading on fibronectin. When GalTase substrates within laminin were either blocked by affinity-purified GalTase or eliminated by prior galactosylation, cell attachment appeared normal, but subsequent cell spreading was totally inhibited. The laminin substrate for GalTase was identified as N-linked oligosaccharides primarily on the A chain, and to a lesser extent on B chains. That N-linked oligosaccharides are necessary for cell spreading was shown by the inability of cells to spread on laminin surfaces pretreated with N-glycanase, even though cell attachment was normal. Cell surface GalTase was distinguished from other reported laminin binding proteins, most notably the 68-kD receptor, since they were differentially eluted from laminin affinity columns. These data show that surface GalTase does not participate during initial cell adhesion to laminin, but mediates subsequent cell spreading by binding to its appropriate N-linked oligosaccharide substrate. These results also emphasize that some of laminin's biological properties can be attributed to its oligosaccharide residues.
APA, Harvard, Vancouver, ISO, and other styles
15

Pinon, Perrine, Jenita Pärssinen, Patricia Vazquez, Michael Bachmann, Rolle Rahikainen, Marie-Claude Jacquier, Latifeh Azizi, et al. "Talin-bound NPLY motif recruits integrin-signaling adapters to regulate cell spreading and mechanosensing." Journal of Cell Biology 205, no. 2 (April 28, 2014): 265–81. http://dx.doi.org/10.1083/jcb.201308136.

Full text
Abstract:
Integrin-dependent cell adhesion and spreading are critical for morphogenesis, tissue regeneration, and immune defense but also tumor growth. However, the mechanisms that induce integrin-mediated cell spreading and provide mechanosensing on different extracellular matrix conditions are not fully understood. By expressing β3-GFP-integrins with enhanced talin-binding affinity, we experimentally uncoupled integrin activation, clustering, and substrate binding from its function in cell spreading. Mutational analysis revealed Tyr747, located in the first cytoplasmic NPLY747 motif, to induce spreading and paxillin adapter recruitment to substrate- and talin-bound integrins. In addition, integrin-mediated spreading, but not focal adhesion localization, was affected by mutating adjacent sequence motifs known to be involved in kindlin binding. On soft, spreading-repellent fibronectin substrates, high-affinity talin-binding integrins formed adhesions, but normal spreading was only possible with integrins competent to recruit the signaling adapter protein paxillin. This proposes that integrin-dependent cell–matrix adhesion and cell spreading are independently controlled, offering new therapeutic strategies to modify cell behavior in normal and pathological conditions.
APA, Harvard, Vancouver, ISO, and other styles
16

Kespichayawattana, W., S. Rattanachetkul, T. Wanun, P. Utaisincharoen, and S. Sirisinha. "Burkholderia pseudomallei Induces Cell Fusion and Actin-Associated Membrane Protrusion: a Possible Mechanism for Cell-to-Cell Spreading." Infection and Immunity 68, no. 9 (September 1, 2000): 5377–84. http://dx.doi.org/10.1128/iai.68.9.5377-5384.2000.

Full text
Abstract:
ABSTRACT Burkholderia pseudomallei, a facultative intracellular bacterium, is the causative agent of a broad spectrum of diseases collectively known as melioidosis. Its ability to survive inside phagocytic and nonphagocytic cells and to induce multinucleated giant cell (MNGC) formation has been demonstrated. This study was designed to assess a possible mechanism(s) leading to this cellular change, using virulent and nonvirulent strains of B. pseudomallei to infect both phagocytic and nonphagocytic cell lines. We demonstrated that when the cells were labeled with two different cell markers (CMFDA or CMTMR), mixed, and then infected with B. pseudomallei, direct cell-to-cell fusion could be observed, leading to MNGC formation. Staining of the infected cells with rhodamine-conjugated phalloidin indicated that immediately after the infection, actin rearrangement into a comet tail appearance occurred, similar to that described earlier for other bacteria. The latter rearrangement led to the formation of bacterium-containing, actin-associated membrane protrusions which could lead to a direct cell-to-cell spreading ofB. pseudomallei in the infected hosts. Results from 4′,6′-diamidine-2-phenylindole dihydrochloride (DAPI) nuclear staining, poly-ADP ribose polymerase cleavage, staining of infected cells for phosphatidylserine exposure with annexin V, and electrophoresis of the DNA extracted from these infected cells showed that B. pseudomallei could kill the host cells by inducing apoptosis in both phagocytic and nonphagocytic cells.
APA, Harvard, Vancouver, ISO, and other styles
17

Blondeau, C., D. Marc, K. Courvoisier, J. F. Vautherot, and C. Denesvre. "Functional Homologies between Avian and Human Alphaherpesvirus VP22 Proteins in Cell-to-Cell Spreading as Revealed by a New cis-Complementation Assay." Journal of Virology 82, no. 18 (July 16, 2008): 9278–82. http://dx.doi.org/10.1128/jvi.00598-08.

Full text
Abstract:
ABSTRACT VP22, encoded by the UL49 gene of Marek's disease virus (MDV), is indispensable for virus cell-to-cell spreading. We show herein that MDV UL49 can be functionally replaced with avian and human viral orthologs. Replacement of MDV VP22 with that of avian gallid herpesvirus 3 or herpesvirus of turkey, whose residue identity with MDV is close to 60%, resulted in 73 and 131% changes in viral spreading, respectively. In contrast, VP22 replacement with human herpes simplex virus type 1 resulted in 14% plaque formation. Therefore, heterologous avian and human VP22 proteins share sufficient structural homology to support MDV cell-to-cell spreading, albeit with different efficiencies.
APA, Harvard, Vancouver, ISO, and other styles
18

Berrier, Allison L., Anthony M. Mastrangelo, Julian Downward, Mark Ginsberg, and Susan E. LaFlamme. "Activated R-Ras, Rac1, Pi 3-Kinase and Pkcε Can Each Restore Cell Spreading Inhibited by Isolated Integrin β1 Cytoplasmic Domains." Journal of Cell Biology 151, no. 7 (December 25, 2000): 1549–60. http://dx.doi.org/10.1083/jcb.151.7.1549.

Full text
Abstract:
Attachment of many cell types to extracellular matrix proteins triggers cell spreading, a process that strengthens cell adhesion and is a prerequisite for many adhesion-dependent processes including cell migration, survival, and proliferation. Cell spreading requires integrins with intact β cytoplasmic domains, presumably to connect integrins with the actin cytoskeleton and to activate signaling pathways that promote cell spreading. Several signaling proteins are known to regulate cell spreading, including R-Ras, PI 3-kinase, PKCε and Rac1; however, it is not known whether they do so through a mechanism involving integrin β cytoplasmic domains. To study the mechanisms whereby cell spreading is regulated by integrin β cytoplasmic domains, we inhibited cell spreading on collagen I or fibrinogen by expressing tac-β1, a dominant-negative inhibitor of integrin function, and examined whether cell spreading could be restored by the coexpression of either V38R-Ras, p110α-CAAX, myr-PKCε, or L61Rac1. Each of these activated signaling proteins was able to restore cell spreading as assayed by an increase in the area of cells expressing tac-β1. R-Ras and Rac1 rescued cell spreading in a GTP-dependent manner, whereas PKCε required an intact kinase domain. Importantly, each of these signaling proteins required intact β cytoplasmic domains on the integrins mediating adhesion in order to restore cell spreading. In addition, the rescue of cell spreading by V38R-Ras was inhibited by LY294002, suggesting that PI 3-kinase activity is required for V38R-Ras to restore cell spreading. In contrast, L61Rac1 and myr-PKCε each increased cell spreading independent of PI 3-kinase activity. Additionally, the dominant-negative mutant of Rac1, N17Rac1, abrogated cell spreading and inhibited the ability of p110α-CAAX and myr-PKCε to increase cell spreading. These studies suggest that R-Ras, PI 3-kinase, Rac1 and PKCε require the function of integrin β cytoplasmic domains to regulate cell spreading and that Rac1 is downstream of PI 3-kinase and PKCε in a pathway involving integrin β cytoplasmic domain function in cell spreading.
APA, Harvard, Vancouver, ISO, and other styles
19

Cramer, L. P., and T. J. Mitchison. "Myosin is involved in postmitotic cell spreading." Journal of Cell Biology 131, no. 1 (October 1, 1995): 179–89. http://dx.doi.org/10.1083/jcb.131.1.179.

Full text
Abstract:
We have investigated a role for myosin in postmitotic Potoroo tridactylis kidney (PtK2) cell spreading by inhibitor studies, time-lapse video microscopy, and immunofluorescence. We have also determined the spatial organization and polarity of actin filaments in postmitotic spreading cells. We show that butanedione monoxime (BDM), a known inhibitor of muscle myosin II, inhibits nonmuscle myosin II and myosin V adenosine triphosphatases. BDM reversibly inhibits PtK2 postmitotic cell spreading. Listeria motility is not affected by this drug. Electron microscopy studies show that some actin filaments in spreading edges are part of actin bundles that are also found in long, thin, structures that are connected to spreading edges and substrate (retraction fibers), and that 90% of this actin is oriented with barbed ends in the direction of spreading. The remaining actin in spreading edges has a more random orientation and spatial arrangement. Myosin II is associated with actin polymer in spreading cell edges, but not retraction fibers. Myosin II is excluded from lamellipodia that protrude from the cell edge at the end of spreading. We suggest that spreading involves myosin, possibly myosin II.
APA, Harvard, Vancouver, ISO, and other styles
20

Dunn, G. A., and D. Zicha. "Dynamics of fibroblast spreading." Journal of Cell Science 108, no. 3 (March 1, 1995): 1239–49. http://dx.doi.org/10.1242/jcs.108.3.1239.

Full text
Abstract:
A new technique of microinterferometry permits cellular growth and motile dynamics to be studied simultaneously in living cells. In isolated chick heart fibroblasts, we have found that the non-aqueous mass of each cell tends to increase steadily, with minor fluctuations, throughout the cell cycle. The spread area of each cell also tends to increase during interphase but fluctuates between wide limits. These limits are dependent on the cell's mass and the upper limit is particularly sharp and directly proportional to mass. From a dynamical point of view, the spread area of a cell is determined by the balance between the rates of two antagonistic processes: protrusion of cellular material into new territory and retraction of material from previously occupied territory. The spatial asymmetry of these processes determines the translocation of the cell. We have found with the chick fibroblasts that the rates of the two processes are generally closely matched to each other and appear to be dependent on the cell's area of spreading. Both continue incessantly in well spread cells, even when there is no net translocation of the cell, and the lower limit of each activity is directly proportional to spread area. The two processes show different behaviour, however, during changes in the spread area of the cell. Both increases and decreases in area appear to be brought about by changes in the rate of retraction, the rate of protrusion remaining relatively constant. A simple stochastic model based on a limited supply of adhesion molecules can simulate all our observations including the mass-limited spreading, the strong correlation between protrusion and retraction and the retraction-dominated changes in area. We conclude that the spread area of the cell is actively regulated, possibly by a simple automatic mechanism that adjusts the area of spreading in relation to the mass of the cell and controls the rate of protrusion to compensate rapidly for spontaneous fluctuations in retraction.
APA, Harvard, Vancouver, ISO, and other styles
21

Stockton, Rebecca A., and Bruce S. Jacobson. "Modulation of Cell-Substrate Adhesion by Arachidonic Acid: Lipoxygenase Regulates Cell Spreading and ERK1/2-inducible Cyclooxygenase Regulates Cell Migration in NIH-3T3 Fibroblasts." Molecular Biology of the Cell 12, no. 7 (July 2001): 1937–56. http://dx.doi.org/10.1091/mbc.12.7.1937.

Full text
Abstract:
Adhesion of cells to an extracellular matrix is characterized by several discrete morphological and functional stages beginning with cell-substrate attachment, followed by cell spreading, migration, and immobilization. We find that although arachidonic acid release is rate-limiting in the overall process of adhesion, its oxidation by lipoxygenase and cyclooxygenases regulates, respectively, the cell spreading and cell migration stages. During the adhesion of NIH-3T3 cells to fibronectin, two functionally and kinetically distinct phases of arachidonic acid release take place. An initial transient arachidonate release occurs during cell attachment to fibronectin, and is sufficient to signal the cell spreading stage after its oxidation by 5-lipoxygenase to leukotrienes. A later sustained arachidonate release occurs during and after spreading, and signals the subsequent migration stage through its oxidation to prostaglandins by newly synthesized cyclooxygenase-2. In signaling migration, constitutively expressed cyclooxygenase-1 appears to contribute ∼25% of prostaglandins synthesized compared with the inducible cyclooxygenase-2. Both the second sustained arachidonate release, and cyclooxygenase-2 protein induction and synthesis, appear to be regulated by the mitogen-activated protein kinase extracellular signal-regulated kinase (ERK)1/2. The initial cell attachment-induced transient arachidonic acid release that signals spreading through lipoxygenase oxidation is not sensitive to ERK1/2 inhibition by PD98059, whereas PD98059 produces both a reduction in the larger second arachidonate release and a blockade of induced cyclooxygenase-2 protein expression with concomitant reduction of prostaglandin synthesis. The second arachidonate release, and cyclooxygenase-2 expression and activity, both appear to be required for cell migration but not for the preceding stages of attachment and spreading. These data suggest a bifurcation in the arachidonic acid adhesion-signaling pathway, wherein lipoxygenase oxidation generates leukotriene metabolites regulating the spreading stage of cell adhesion, whereas ERK 1/2-induced cyclooxygenase synthesis results in oxidation of a later release, generating prostaglandin metabolites regulating the later migration stage.
APA, Harvard, Vancouver, ISO, and other styles
22

Pasquali, Livia, Paola Lenzi, Francesca Biagioni, Gabriele Siciliano, and Francesco Fornai. "Cell to Cell Spreading of Misfolded Proteins as a Therapeutic Target in Motor Neuron Disease." Current Medicinal Chemistry 21, no. 31 (June 1, 2014): 3508–34. http://dx.doi.org/10.2174/0929867321666140601161534.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Nussbaum-Krammer, Carmen I., Kyung-Won Park, Liming Li, Ronald Melki, and Richard I. Morimoto. "Spreading of a Prion Domain from Cell-to-Cell by Vesicular Transport in Caenorhabditis elegans." PLoS Genetics 9, no. 3 (March 28, 2013): e1003351. http://dx.doi.org/10.1371/journal.pgen.1003351.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Sun, Jing, Dan Wei, Yuda Zhu, Meiling Zhong, Yicong Zuo, Hongsong Fan, and Xingdong Zhang. "A spatial patternable macroporous hydrogel with cell-affinity domains to enhance cell spreading and differentiation." Biomaterials 35, no. 17 (June 2014): 4759–68. http://dx.doi.org/10.1016/j.biomaterials.2014.02.041.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Kaul, Anirudh, and Parmod Kalsotra. "Extranodal B Cell Lymphoma Spreading from Skin to Oral Cavity." International Journal of Head and Neck Surgery 4, no. 3 (2013): 126–28. http://dx.doi.org/10.5005/jp-journals-10001-1159.

Full text
Abstract:
ABSTRACT Non-Hodgkin's lymphoma (NHL) belongs to a group of lymphoid neoplasms that is diverse in manner of presentation, response to therapy and prognosis. Usually oral manifestations of NHL are secondary to a more widespread involvement throughout th e bo dy. Tho ugh primary in traoral lesio ns in NHL are uncommon, it is important to be aware of them, since intraoral manifestations are the presenting symptoms in these patients.2 A 58-year-old man presented to our Department with swelling in the vestibule of mouth. CT scan revealed a mass on upper gingival without any bony involvement. Incisional biopsy of the lesion showed diffuse large B cell lymphoma. The patient was treated successfully with chemotherapy. One year after complete remission, patient developed recurrence of primary tumor for which patient was again given chemotherapy treatment but had no significant benefit and succumb to the disease. How to cite this article Kaul A, Kalsotra P. Extranodal B Cell Lymphoma Spreading from Skin to Oral Cavity. Int J Head Neck Surg 2013;4(3):126-128.
APA, Harvard, Vancouver, ISO, and other styles
26

English, C. S., C. K. Puk, F. A. Petrigliano, B. M. Wu, and D. R. McAllister. "214 LIGAMENT ENGINEERING: THE CONTRIBUTION OF STRETCH TO CELL SPREADING." Journal of Investigative Medicine 54, no. 1 (January 1, 2006): S116.6—S117. http://dx.doi.org/10.2310/6650.2005.x0004.213.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

LaLonde, David P., Michael C. Brown, Brian P. Bouverat, and Christopher E. Turner. "Actopaxin Interacts with TESK1 to Regulate Cell Spreading on Fibronectin." Journal of Biological Chemistry 280, no. 22 (April 6, 2005): 21680–88. http://dx.doi.org/10.1074/jbc.m500752200.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Cornaby, Caleb, Lauren Gibbons, Vera Mayhew, Chad S. Sloan, Andrew Welling, and Brian D. Poole. "B cell epitope spreading: Mechanisms and contribution to autoimmune diseases." Immunology Letters 163, no. 1 (January 2015): 56–68. http://dx.doi.org/10.1016/j.imlet.2014.11.001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Tsygankova, Oxana M., Changqing Ma, Waixing Tang, Christopher Korch, Michael D. Feldman, Yu Lv, Marcia S. Brose, and Judy L. Meinkoth. "Downregulation of Rap1GAP in Human Tumor Cells Alters Cell/Matrix and Cell/Cell Adhesion." Molecular and Cellular Biology 30, no. 13 (May 3, 2010): 3262–74. http://dx.doi.org/10.1128/mcb.01345-09.

Full text
Abstract:
ABSTRACT Rap1GAP expression is decreased in human tumors. The significance of its downregulation is unknown. We show that Rap1GAP expression is decreased in primary colorectal carcinomas. To elucidate the advantages conferred on tumor cells by loss of Rap1GAP, Rap1GAP expression was silenced in human colon carcinoma cells. Suppressing Rap1GAP induced profound alterations in cell adhesion. Rap1GAP-depleted cells exhibited defects in cell/cell adhesion that included an aberrant distribution of adherens junction proteins. Depletion of Rap1GAP enhanced adhesion and spreading on collagen. Silencing of Rap expression normalized spreading and restored E-cadherin, β-catenin, and p120-catenin to cell/cell contacts, indicating that unrestrained Rap activity underlies the alterations in cell adhesion. The defects in adherens junction protein distribution required integrin signaling as E-cadherin and p120-catenin were restored at cell/cell contacts when cells were plated on poly-l-lysine. Unexpectedly, Src activity was increased in Rap1GAP-depleted cells. Inhibition of Src impaired spreading and restored E-cadherin at cell/cell contacts. These findings provide the first evidence that Rap1GAP contributes to cell/cell adhesion and highlight a role for Rap1GAP in regulating cell/matrix and cell/cell adhesion. The frequent downregulation of Rap1GAP in epithelial tumors where alterations in cell/cell and cell/matrix adhesion are early steps in tumor dissemination supports a role for Rap1GAP depletion in tumor progression.
APA, Harvard, Vancouver, ISO, and other styles
30

Morrison, R. F., and E. R. Seidel. "Cell spreading and the regulation of ornithine decarboxylase." Journal of Cell Science 108, no. 12 (December 1, 1995): 3787–94. http://dx.doi.org/10.1242/jcs.108.12.3787.

Full text
Abstract:
The aim of this study was to investigate the effect of cell spreading on the induction of ornithine decarboxylase and the rate of putrescine uptake in anchorage-dependent and anchorage-independent cells. Plating non-transformed IEC-6 epithelial cells at high versus low cell density restricted cell spreading from 900 microns 2 to approximately 140 microns 2, blunted the transient induction of ornithine decarboxylase activity from 202 to 32 pmol 14CO2/mg protein per hour and reduced the rate of [14C] putrescine uptake from 46 to 23 pmol/10(5) cells per hour. The mean spreading area of the cell population was controlled by coating tissue culture dishes with the nonadhesive polymer, polyHEMA. Ornithine decarboxylase activity and putrescine uptake correlated with cell spreading with minimal spreading (263 microns 2) corresponding to an 83% decrease in ornithine decarboxylase activity and 51% decrease in the rate of putrescine uptake. Adding the RGD peptide, Gly-Arg-Gly-Glu-Ser-Pro to the medium of sparsely plated cells resulted in rapid reductions in cell spreading concomitant with dose-dependent decreases in ornithine decarboxylase activity and putrescine uptake. Finally, minimizing cell spreading by depriving cells of substratum contact completely abolished serum-induced increases in ornithine decarboxylase and reduced the rate of putrescine uptake by 47%. In contrast to IEC-6 cells, ornithine decarboxylase of neoplastic HTC-116 cells was constitutively expressed with basal and stimulated activity (193 and 982 pmol 14CO2/mg protein per hour, respectively) completely independent of cell adhesion. Putrescine uptake, however, was abolished in the absence of cell adhesion. These data suggest that the induction of ornithine decarboxylase activity and the rate of putrescine uptake correlate with spreading of anchorage-dependent IEC-6 cells and that ornithine decarboxylase activity but not putrescine uptake, appears to be independent of spreading of neoplastic HTC-116 cells.
APA, Harvard, Vancouver, ISO, and other styles
31

Young, Bradford A., Yasuyuki Taooka, Shouchun Liu, Karen J. Askins, Yasuyuki Yokosaki, Sheila M. Thomas, and Dean Sheppard. "The Cytoplasmic Domain of the Integrin α9 Subunit Requires the Adaptor Protein Paxillin to Inhibit Cell Spreading but Promotes Cell Migration in a Paxillin-independent Manner." Molecular Biology of the Cell 12, no. 10 (October 2001): 3214–25. http://dx.doi.org/10.1091/mbc.12.10.3214.

Full text
Abstract:
The integrin α9 subunit forms a single heterodimer, α9β1. The α9 subunit is most closely related to the α4 subunit, and like α4 integrins, α9β1 plays an important role in leukocyte migration. The α4 cytoplasmic domain preferentially enhances cell migration and inhibits cell spreading, effects that depend on interaction with the adaptor protein, paxillin. To determine whether the α9 cytoplasmic domain has similar effects, a series of chimeric and deleted α9 constructs were expressed in Chinese hamster ovary cells and tested for their effects on migration and spreading on an α9β1-specific ligand. Like α4, the α9 cytoplasmic domain enhanced cell migration and inhibited cell spreading. Paxillin also specifically bound the α9 cytoplasmic domain and to a similar level as α4. In paxillin −/− cells, α9 failed to inhibit cell spreading as expected but surprisingly still enhanced cell migration. Further, mutations that abolished the α9-paxillin interaction prevented α9 from inhibiting cell spreading but had no effect on α9-dependent cell migration. These findings suggest that the mechanisms by which the cytoplasmic domains of integrin α subunits enhance migration and inhibit cell spreading are distinct and that the α9 and α4 cytoplasmic domains, despite sequence and functional similarities, enhance cell migration by different intracellular signaling pathways.
APA, Harvard, Vancouver, ISO, and other styles
32

Chun, J. S., and B. S. Jacobson. "Requirement for diacylglycerol and protein kinase C in HeLa cell-substratum adhesion and their feedback amplification of arachidonic acid production for optimum cell spreading." Molecular Biology of the Cell 4, no. 3 (March 1993): 271–81. http://dx.doi.org/10.1091/mbc.4.3.271.

Full text
Abstract:
Release of arachidonic acid (AA) and subsequent formation of a lipoxygenase (LOX) metabolite(s) is an obligatory signal to induce spreading of HeLa cells on a gelatin substratum (Chun and Jacobson, 1992). This study characterizes signaling pathways that follow the LOX metabolite(s) formation. Levels of diacylglycerol (DG) increase upon attachment and before cell spreading on a gelatin substratum. DG production and cell spreading are insignificant when phospholipase A2 (PLA2) or LOX is blocked. In contrast, when cells in suspension where PLA2 activity is not stimulated are treated with exogenous AA, DG production is turned on, and inhibition of LOX turns it off. This indicates that the formation of a LOX metabolite(s) from AA released during cell attachment induces the production of DG. Consistent with the DG production is the activation of protein kinase C (PKC) which, as with AA and DG, occurs upon attachment and before cell spreading. Inhibition of AA release and subsequent DG production blocks both PKC activation and cell spreading. Cell spreading is also blocked by the inhibition of PKC with calphostin C or sphingosine. The inhibition of cell spreading induced by blocking AA release is reversed by the direct activation of PKC with phorbol ester. However, the inhibition of cell spreading induced by PKC inhibition is not reversed by exogenously applied AA. In addition, inhibition of PKC does not block AA release and DG production. The data indicate that there is a sequence of events triggered by HeLa cell attachment to a gelatin substratum that leads to the initiation of cell spreading: AA release, a LOX metabolite(s) formation, DG production, and PKC activation. The data also provide evidence indicating that HeLa cell spreading is a cyclic feedback amplification process centered on the production of AA, which is the first messenger produced in the sequence of messengers initiating cell spreading. Both DG and PKC activity that are increased during HeLa cell attachment to a gelatin substratum appear to be involved. DG not only activates PKC, which is essential for cell spreading, but is also hydrolyzed to AA. PKC, which is initially activated as consequence of AA production, also increases more AA production by activating PLA2.
APA, Harvard, Vancouver, ISO, and other styles
33

Flevaris, Panagiotis, Aleksandra Stojanovic, Haixia Gong, Athar Chishti, Emily Welch, and Xiaoping Du. "A molecular switch that controls cell spreading and retraction." Journal of Cell Biology 179, no. 3 (October 29, 2007): 553–65. http://dx.doi.org/10.1083/jcb.200703185.

Full text
Abstract:
Integrin-dependent cell spreading and retraction are required for cell adhesion, migration, and proliferation, and thus are important in thrombosis, wound repair, immunity, and cancer development. It remains unknown how integrin outside-in signaling induces and controls these two opposite processes. This study reveals that calpain cleavage of integrin β3 at Tyr759 switches the functional outcome of integrin signaling from cell spreading to retraction. Expression of a calpain cleavage–resistant β3 mutant in Chinese hamster ovary cells causes defective clot retraction and RhoA-mediated retraction signaling but enhances cell spreading. Conversely, a calpain-cleaved form of β3 fails to mediate cell spreading, but inhibition of the RhoA signaling pathway corrects this defect. Importantly, the calpain-cleaved β3 fails to bind c-Src, which is required for integrin-induced cell spreading, and this requirement of β3-associated c-Src results from its inhibition of RhoA-dependent contractile signals. Thus, calpain cleavage of β3 at Tyr759 relieves c-Src–mediated RhoA inhibition, activating the RhoA pathway that confines cell spreading and causes cell retraction.
APA, Harvard, Vancouver, ISO, and other styles
34

Potter, David A., Jennifer S. Tirnauer, Richard Janssen, Dorothy E. Croall, Christina N. Hughes, Kerry A. Fiacco, James W. Mier, Masatoshi Maki, and Ira M. Herman. "Calpain Regulates Actin Remodeling during Cell Spreading." Journal of Cell Biology 141, no. 3 (May 4, 1998): 647–62. http://dx.doi.org/10.1083/jcb.141.3.647.

Full text
Abstract:
Previous studies suggest that the Ca2+-dependent proteases, calpains, participate in remodeling of the actin cytoskeleton during wound healing and are active during cell migration. To directly test the role that calpains play in cell spreading, several NIH-3T3– derived clonal cell lines were isolated that overexpress the biological inhibitor of calpains, calpastatin. These cells stably overexpress calpastatin two- to eightfold relative to controls and differ from both parental and control cell lines in morphology, spreading, cytoskeletal structure, and biochemical characteristics. Morphologic characteristics of the mutant cells include failure to extend lamellipodia, as well as abnormal filopodia, extensions, and retractions. Whereas wild-type cells extend lamellae within 30 min after plating, all of the calpastatin-overexpressing cell lines fail to spread and assemble actin-rich processes. The cells genetically altered to overexpress calpastatin display decreased calpain activity as measured in situ or in vitro. The ERM protein ezrin, but not radixin or moesin, is markedly increased due to calpain inhibition. To confirm that inhibition of calpain activity is related to the defect in spreading, pharmacological inhibitors of calpain were also analyzed. The cell permeant inhibitors calpeptin and MDL 28, 170 cause immediate inhibition of spreading. Failure of the intimately related processes of filopodia formation and lamellar extension indicate that calpain is intimately involved in actin remodeling and cell spreading.
APA, Harvard, Vancouver, ISO, and other styles
35

Arthur, William T., Lawrence A. Quilliam, and Jonathan A. Cooper. "Rap1 promotes cell spreading by localizing Rac guanine nucleotide exchange factors." Journal of Cell Biology 167, no. 1 (October 11, 2004): 111–22. http://dx.doi.org/10.1083/jcb.200404068.

Full text
Abstract:
The Ras-related GTPase Rap1 stimulates integrin-mediated adhesion and spreading in various mammalian cell types. Here, we demonstrate that Rap1 regulates cell spreading by localizing guanine nucleotide exchange factors (GEFs) that act via the Rho family GTPase Rac1. Rap1a activates Rac1 and requires Rac1 to enhance spreading, whereas Rac1 induces spreading independently of Rap1. Active Rap1a binds to a subset of Rac GEFs, including VAV2 and Tiam1 but not others such as SWAP-70 or COOL-1. Overexpressed VAV2 and Tiam1 specifically require Rap1 to promote spreading, even though Rac1 is activated independently of Rap1. Rap1 is necessary for the accumulation of VAV2 in membrane protrusions at the cell periphery. In addition, if VAV2 is artificially localized to the cell edge with the subcellular targeting domain of Rap1a, it increases cell spreading independently of Rap1. These results lead us to propose that Rap1 promotes cell spreading by localizing a subset of Rac GEFs to sites of active lamellipodia extension.
APA, Harvard, Vancouver, ISO, and other styles
36

Gehler, Scott, Frances V. Compere, and Alex M. Miller. "Semaphorin 3A Increases FAK Phosphorylation at Focal Adhesions to Modulate MDA-MB-231 Cell Migration and Spreading on Different Substratum Concentrations." International Journal of Breast Cancer 2017 (2017): 1–16. http://dx.doi.org/10.1155/2017/9619734.

Full text
Abstract:
Interactions between integrin-mediated adhesions and the extracellular matrix (ECM) are important regulators of cell migration and spreading. However, mechanisms by which extracellular ligands regulate cell migration and spreading in response to changes in substratum concentration are not well understood. Semaphorin 3A (Sema3A) has been shown to inhibit cell motility and alter integrin signaling in various cell types. We propose that Sema3A alters focal adhesions to modulate breast carcinoma cell migration and spreading on substrata coated with different concentrations of ECM. We demonstrate that Sema3A inhibits MDA-MB-231 cell migration and spreading on substrata coated with high concentrations of collagen and fibronectin but enhances migration and spreading at lower concentrations of collagen and fibronectin. Sema3A increases focal adhesion kinase phosphorylation at tyrosine 397 (pFAK397) at focal adhesions on all substratum concentrations of collagen and fibronectin but decreased pFAK397 levels on laminin. Rho-associated protein kinase (ROCK) inhibition blocks the Sema3A-mediated effects on cell migration, spreading, and pFAK397 at focal adhesions when cultured on all concentrations of collagen. These results suggest that Sema3A shifts the optimal level of cell-matrix adhesions to a nonoptimal ECM coating concentration, in particular collagen, to yield maximal cell migration and spreading that may be mediated through a ROCK-dependent mechanism.
APA, Harvard, Vancouver, ISO, and other styles
37

Jarrell, B. E., S. K. Williams, D. Rose, D. Garibaldi, C. Talbot, and B. Kapelan. "Optimization of Human Endothelial Cell Attachment to Vascular Graft Polymers." Journal of Biomechanical Engineering 113, no. 2 (May 1, 1991): 120–22. http://dx.doi.org/10.1115/1.2891225.

Full text
Abstract:
Endothelial cells (EC) covering the blood-contacting surface of a prosthetic material could potentially enhance the subsequent nonthrombogenicity of the surface. In order to create such a surface, the EC must become attached to the surface, spread and ultimately form a monolayer. In this study we examined several factors that influence these processes. On ePTFE surfaces, surface pretreatment with human serum for 30 minutes at a concentration of 1.4 gm percent protein resulted in significantly more attached EC when compared to other concentrations or when compared to fetal calf serum or human serum albumin. The rate of EC spreading was strongly influenced by temperature, with a maximum occurring at 37°C. During real-time video microscopy, it was noted that the rate of EC attachment and spreading was primarily dependent on arrival of the EC to the surface rather than attachment and spreading. Thus as a method of increasing EC delivery, the concept of filtering EC onto the graft lumenal surface was tested by pressurizing the graft lumen to speed EC delivery to the surface. This technique produced a 2 to 5-fold increase in EC attachment when compared to gravity forced cell deposition. We conclude that an ePTFE graft can be rapidly endothelialized using these simple measures.
APA, Harvard, Vancouver, ISO, and other styles
38

Wu, Chih-Hang, Shu-Chuan Lee, and Chao-Wen Wang. "Viral protein targeting to the cortical endoplasmic reticulum is required for cell–cell spreading in plants." Journal of Cell Biology 193, no. 3 (April 25, 2011): 521–35. http://dx.doi.org/10.1083/jcb.201006023.

Full text
Abstract:
Many plant RNA viruses use their nonstructural proteins to target and move through the cortical endoplasmic reticulum (ER) tubules within the plant intercellular junction for cell-to-cell spreading. Most of these proteins, including the triple-gene-block 3 protein (TGBp3) of Potexvirus, are ER membrane proteins. We previously showed that TGBp3 of the Bamboo mosaic potexvirus partitions into tubular subdomains of the ER in both yeast and plants, but the mechanism and physiological significance of this localization is unclear. Here, we demonstrate that a sorting signal present in TGBp3 is necessary and sufficient for its oligomerization and for targeting integral membrane proteins into puncta within curved ER tubules. Mutations in the TGBp3 sorting signal impair viral spread, and plants infected with viruses harboring these mutants were either asymptomatic or had reduced symptoms. Thus, we propose that Potexvirus use the sorting signal in TGBp3 to target infectious viral derivatives to cortical ER tubules for transmission through the intercellular junctions in plants.
APA, Harvard, Vancouver, ISO, and other styles
39

Wei, Shuangshi, Xiaomei Liu, Bo Ma, Yihan Wu, Yan Liu, Mingchun Gao, Peifen Fu, and Junwei Wang. "The US2 protein is involved in the penetration and cell-to-cell spreading of DEVin vitro." Journal of Basic Microbiology 54, no. 9 (July 4, 2013): 1005–11. http://dx.doi.org/10.1002/jobm.201300068.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Jones, R. A., B. Nicholas, S. Mian, and M. Griffin. "REDUCED EXPRESSION OF TISSUE TRANSGLUTAMINASE IN A HUMAN ENDOTHELIAL CELL LINE LEADS TO CHANGES IN CELL SPREADING AND CELL ADHESION." Biochemical Society Transactions 24, no. 4 (November 1, 1996): 550S. http://dx.doi.org/10.1042/bst024550sc.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Francis, Emmet A., and Volkmar Heinrich. "Integrative experimental/computational approach establishes active cellular protrusion as the primary driving force of phagocytic spreading by immune cells." PLOS Computational Biology 18, no. 8 (August 26, 2022): e1009937. http://dx.doi.org/10.1371/journal.pcbi.1009937.

Full text
Abstract:
The dynamic interplay between cell adhesion and protrusion is a critical determinant of many forms of cell motility. When modeling cell spreading on adhesive surfaces, traditional mathematical treatments often consider passive cell adhesion as the primary, if not exclusive, mechanistic driving force of this cellular motion. To better assess the contribution of active cytoskeletal protrusion to immune-cell spreading during phagocytosis, we here develop a computational framework that allows us to optionally investigate both purely adhesive spreading (“Brownian zipper hypothesis”) as well as protrusion-dominated spreading (“protrusive zipper hypothesis”). We model the cell as an axisymmetric body of highly viscous fluid surrounded by a cortex with uniform surface tension and incorporate as potential driving forces of cell spreading an attractive stress due to receptor-ligand binding and an outward normal stress representing cytoskeletal protrusion, both acting on the cell boundary. We leverage various model predictions against the results of a directly related experimental companion study of human neutrophil phagocytic spreading on substrates coated with different densities of antibodies. We find that the concept of adhesion-driven spreading is incompatible with experimental results such as the independence of the cell-spreading speed on the density of immobilized antibodies. In contrast, the protrusive zipper model agrees well with experimental findings and, when adapted to simulate cell spreading on discrete adhesion sites, it also reproduces the observed positive correlation between antibody density and maximum cell-substrate contact area. Together, our integrative experimental/computational approach shows that phagocytic spreading is driven by cellular protrusion, and that the extent of spreading is limited by the density of adhesion sites.
APA, Harvard, Vancouver, ISO, and other styles
42

Roll, Richard L., Eve Marie Bauman, Joel S. Bennett, and Charles S. Abrams. "Phosphorylated Pleckstrin Induces Cell Spreading via an Integrin-Dependent Pathway." Journal of Cell Biology 150, no. 6 (September 18, 2000): 1461–66. http://dx.doi.org/10.1083/jcb.150.6.1461.

Full text
Abstract:
Pleckstrin is a 40-kD phosphoprotein containing NH2- and COOH-terminal pleckstrin homology (PH) domains separated by a disheveled-egl 10-pleckstrin (DEP) domain. After platelet activation, pleckstrin is rapidly phosphorylated by protein kinase C. We reported previously that expressed phosphorylated pleckstrin induces cytoskeletal reorganization and localizes in microvilli along with glycoproteins, such as integrins. Given the role of integrins in cytoskeletal organization and cell spreading, we investigated whether signaling from pleckstrin cooperated with signaling pathways involving the platelet integrin, αIIbβ3. Pleckstrin induced cell spreading in both transformed (COS-1 & CHO) and nontransformed (REF52) cell lines, and this spreading was regulated by pleckstrin phosphorylation. In REF52 cells, pleckstrin-induced spreading was matrix dependent, as evidenced by spreading of these cells on fibrinogen but not on fibronectin. Coexpression with αIIbβ3 did not enhance pleckstrin-mediated cell spreading in either REF52 or CHO cells. However, coexpression of the inactive variant αIIbβ3 Ser753Pro, or β3 Ser753Pro alone, completely blocked pleckstrin-induced spreading. This implies that αIIbβ3 Ser753Pro functions as a competitive inhibitor by blocking the effects of an endogenous receptor that is used in the signaling pathway involved in pleckstrin-induced cell spreading. Expression of a chimeric protein composed of the extracellular and transmembrane portion of Tac fused to the cytoplasmic tail of β3 completely blocked pleckstrin-mediated spreading, whereas chimeras containing the cytoplasmic tail of β3 Ser753Pro or αIIb had no effect. This suggests that the association of an unknown signaling protein with the cytoplasmic tail of an endogenous integrin β-chain is also required for pleckstrin-induced spreading. Thus, expressed phosphorylated pleckstrin promotes cell spreading that is both matrix and integrin dependent. To our knowledge, this is the first example of a mutated integrin functioning as a dominant negative inhibitor.
APA, Harvard, Vancouver, ISO, and other styles
43

Clarke, Dominic M., Michael C. Brown, David P. LaLonde, and Christopher E. Turner. "Phosphorylation of actopaxin regulates cell spreading and migration." Journal of Cell Biology 166, no. 6 (September 7, 2004): 901–12. http://dx.doi.org/10.1083/jcb.200404024.

Full text
Abstract:
Actopaxin is an actin and paxillin binding protein that localizes to focal adhesions. It regulates cell spreading and is phosphorylated during mitosis. Herein, we identify a role for actopaxin phosphorylation in cell spreading and migration. Stable clones of U2OS cells expressing actopaxin wild-type (WT), nonphosphorylatable, and phosphomimetic mutants were developed to evaluate actopaxin function. All proteins targeted to focal adhesions, however the nonphosphorylatable mutant inhibited spreading whereas the phosphomimetic mutant cells spread more efficiently than WT cells. Endogenous and WT actopaxin, but not the nonphosphorylatable mutant, were phosphorylated in vivo during cell adhesion/spreading. Expression of the nonphosphorylatable actopaxin mutant significantly reduced cell migration, whereas expression of the phosphomimetic increased cell migration in scrape wound and Boyden chamber migration assays. In vitro kinase assays demonstrate that extracellular signal-regulated protein kinase phosphorylates actopaxin, and treatment of U2OS cells with the MEK1 inhibitor UO126 inhibited adhesion-induced phosphorylation of actopaxin and also inhibited cell migration.
APA, Harvard, Vancouver, ISO, and other styles
44

Brancolini, Claudio, Paolo Edomi, Stefania Marzinotto, and Claudio Schneider. "Exposure at the Cell Surface Is Required for Gas3/PMP22 To Regulate Both Cell Death and Cell Spreading: Implication for the Charcot–Marie–Tooth Type 1A and Dejerine–Sottas Diseases." Molecular Biology of the Cell 11, no. 9 (September 2000): 2901–14. http://dx.doi.org/10.1091/mbc.11.9.2901.

Full text
Abstract:
Gas3/PMP22 is a tetraspan membrane protein highly expressed in myelinating Schwann cells. Point mutations in thegas3/PMP22 gene account for the dominant inherited peripheral neuropathies Charcot–Marie–Tooth type 1A disease (CMT1A) and Dejerine–Sottas syndrome (DSS). Gas3/PMP22 can regulate apoptosis and cell spreading in cultured cells.Gas3/PMP22 point mutations, which are responsible for these diseases, are defective in this respect. In this report, we demonstrate that Gas3/PMP22-WT is exposed at the cell surface, while its point-mutated derivatives are intracellularly retained, colocalizing mainly with the endoplasmic reticulum (ER). The putative retrieval motif present in the carboxyl terminus of Gas3/PMP22 is not sufficient for the intracellular sequestration of its point-mutated forms. On the contrary, the introduction of a retrieval signal at the carboxyl terminus of Gas3/PMP22-WT leads to its intracellular accumulation, which is accompanied by a failure to trigger cell death as well as by changes in cell spreading. In addition, by substituting the Asn at position 41 required for N-glycosylation, we provide evidence that N-glycosylation is required for the full effect on cell spreading, but it is not necessary for triggering cell death. In conclusion, we suggest that the DSS and the CMT1A neuropathies derived from point mutations ofGas3/PMP22 might arise, at the molecular level, from a reduced exposure of Gas3/PMP22 at the cell surface, which is required to exert its biological functions.
APA, Harvard, Vancouver, ISO, and other styles
45

Jones, R. A., B. Nicholas, S. Mian, P. J. Davies, and M. Griffin. "Reduced expression of tissue transglutaminase in a human endothelial cell line leads to changes in cell spreading, cell adhesion and reduced polymerisation of fibronectin." Journal of Cell Science 110, no. 19 (October 1, 1997): 2461–72. http://dx.doi.org/10.1242/jcs.110.19.2461.

Full text
Abstract:
Tissue transglutaminase (tTgase, type II) is a Ca2+-dependent GTP binding protein which crosslinks proteins via (epsilon)((gamma)-glutamyl)lysine bridges. Although essentially a cytosolic enzyme there is increasing evidence to suggest the enzyme is externalised where it may play a role in extracellular matrix organisation. To investigate the function of this enzyme in a human umbilical endothelial cell line ECV304 tTgase expression was reduced in these cells by up to 90% by stable transfection with a 1.1. kb antisense construct in the plasmid vector pSG5. Two clones showing a reduction in expression of tTgase activity of 70 and 90% have been isolated and characterised. These clones show a number of phenotypic differences when compared to the parent cell line and the transfected controls which include reduced cell spreading and a decreased adhesion of cells on different substrata as measured by their susceptibility to removal by trypsin. Reduced cell spreading in the antisense transfected clones was accompanied by a decrease in the crosslinking of fibronectin into polymeric multimers which could be correlated to the amount of tTgase externalised by cells. A novel assay was developed to measure externalised tTgase activity which is cell mediated, inhibited by preincubation of cells with anti-tTgase antibody and relies on the incorporation of biotinylated cadaverine into fibronectin. The results of these experiments suggest that externalised tTgase may play a key role in a number of cell behavioural patterns which might be related to the enzymes ability to bind and crosslink fibronectin.
APA, Harvard, Vancouver, ISO, and other styles
46

Lesjak, Michaela S., Rosemarie Marchan, Joanna D. Stewart, Eugen Rempel, Jörg Rahnenführer, and Jan G. Hengstler. "EDI3 links choline metabolism to integrin expression, cell adhesion and spreading." Cell Adhesion & Migration 8, no. 5 (September 3, 2014): 499–508. http://dx.doi.org/10.4161/cam.29284.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

THORNE, LUCY, WENDY DAVIES, SUSAN C. BROWN, CATHERINE A. RICE-EVANS, and JACK A. LUCY. "Cell spreading on collagen that has been exposed to reactive aldehydes." Biochemical Society Transactions 20, no. 4 (November 1, 1992): 369S. http://dx.doi.org/10.1042/bst020369s.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Lehmann, Paul V., Thomas Forsthuber, Alexander Miller, and Eli E. Sercarz. "Spreading of T-cell autoimmunity to cryptic determinants of an autoantigen." Nature 358, no. 6382 (July 1992): 155–57. http://dx.doi.org/10.1038/358155a0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Liu, Shouchun, Marina Slepak, and Mark H. Ginsberg. "Binding of Paxillin to the α9Integrin Cytoplasmic Domain Inhibits Cell Spreading." Journal of Biological Chemistry 276, no. 40 (July 27, 2001): 37086–92. http://dx.doi.org/10.1074/jbc.m105114200.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Morone, Simona, Stefania Augeri, Massimiliano Cuccioloni, Matteo Mozzicafreddo, Mauro Angeletti, Nicola Lo Buono, Alice Giacomino, Erika Ortolan, and Ada Funaro. "Binding of CD157 Protein to Fibronectin Regulates Cell Adhesion and Spreading." Journal of Biological Chemistry 289, no. 22 (April 21, 2014): 15588–601. http://dx.doi.org/10.1074/jbc.m113.535070.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography