Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-04-30T12:47:58.075Z Has data issue: false hasContentIssue false

3 - Role of the Plasma Membrane in Endothelial Cell Mechanosensation of Shear Stress

Published online by Cambridge University Press:  05 July 2014

Peter J. Butler
Affiliation:
The Pennsylvania State University
Shu Chien
Affiliation:
University of California, San Diego
Mohammad R. K. Mofrad
Affiliation:
University of California, Berkeley
Roger D. Kamm
Affiliation:
Massachusetts Institute of Technology
Get access

Summary

Introduction

Mechanotransduction, which is the process by which cells convert mechanical stimuli to biochemical signaling cascades, is involved in the homeostasis of numerous tissues (reviewed in [21] and [56]). The mechanotransduction of hemodynamic shear stress by endothelial cells (ECs) has garnered special attention because of its role in regulating vascular health and disease. In particular, there is intense interest in identifying the primary molecular mechanisms of the EC sensing of shear stress because its (or their) discovery may lead to clinical interventions in atherosclerosis and other diseases related to mechanobiology.

In this chapter, we address the hypothesis that the plasma membrane lipid bilayer is one endothelial cell mechanosensor. Here we define “mechanosensor” as a cellular structure that responds to mechanical stress and initiates mechanotransduction in response to shear stress without involving chemical second messengers. Mechanotransduction, then, is the process by which cells convert this sensory stimulus into changes in biochemical signaling. We define mechanobiology as the study of the entire process of sensation, transduction, and attendant changes in cell phenotype. Because mechanical linkages from the cell surface to lateral, internal, and basal parts of the cell redistribute forces imposed on the cell surface, many structures could serve as mechanosensors. Furthermore, mechanotransduction can involve direct force effects on molecules, diffusion- or convection-mediated transport of molecular second messengers, and the active transport of signaling molecules by molecular motors. Other chapters in this text will address other candidate mechanosensors (e.g., focal adhesions and their integrins). With respect to the membrane, in the context of these definitions, if shear stress induces a perturbation of the membrane constituents, and this perturbation is necessary for mechanotransduction, then the lipid bilayer is considered a mechanosensor. Similarly, if the shear stress acting on the apical portion of the cell leads to the perturbation of the membrane on the basal portion as a result of mechanical linkage, and this membrane perturbation is necessary for subsequent downstream signaling, then we consider the basal membrane also as a mechanosensor.

Type
Chapter
Information
Cellular Mechanotransduction
Diverse Perspectives from Molecules to Tissues
, pp. 61 - 88
Publisher: Cambridge University Press
Print publication year: 2009

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Andersen, O. S., et al. Ion channels as tools to monitor lipid bilayer-membrane protein interactions: Gramicidin channels as molecular force transducers. Methods Enzymol. 294, 208–224 (1999).CrossRefGoogle Scholar
Ando, J., Ohtsuka, A., Korenaga, R., Kawamura, T., & Kamiya, A. Wall shear stress rather than shear rate regulates cytoplasmic Ca++ responses to flow in vascular endothelial cells. Biochem. Biophys. Res. Commun. 190, 716–723 (1993).CrossRefGoogle ScholarPubMed
Arisaka, T., et al. Effects of shear stress on glycosaminoglycan synthesis in vascular endothelial cells. Ann. N.Y. Acad. Sci. 748, 543–554 (1995).CrossRefGoogle ScholarPubMed
Axelrod, D. Lateral motion of membrane proteins and biological function. J. Membr. Biol. 75, 1–10 (1983).CrossRefGoogle ScholarPubMed
Bao, X., Clark, C. B., & Frangos, J. A. Temporal gradient in shear-induced signaling pathway: Involvement of MAP kinase, c-fos, and connexin43. Am. J. Physiol. Heart Circ. Physiol. 278, H1598–H1605 (2000).CrossRefGoogle ScholarPubMed
Bevan, J. A. & Siegel, G. Blood vessel wall matrix flow sensor: Evidence and speculation. Blood Vessels 28, 552–556 (1991).Google ScholarPubMed
Boesze-Battaglia, K. Isolation of membrane rafts and signaling complexes. Methods Mol. Biol. 332, 169–179 (2006).Google ScholarPubMed
Broday, D. M. Diffusion of clusters of transmembrane proteins as a model of focal adhesion remodeling. Bull. Math. Biol. 62, 891–924 (2000).CrossRefGoogle ScholarPubMed
Brooks, A. R., Lelkes, P. I., & Rubanyi, G. M. Gene expression profiling of human aortic endothelial cells exposed to disturbed flow and steady laminar flow. Physiolog. Genom. 9, 27–41 (2002).CrossRefGoogle ScholarPubMed
Butler, P. J., Norwich, G., Weinbaum, S., & Chien, S. Shear stress induces a time- and position-dependent increase in endothelial cell membrane fluidity. Am. J. Physiol. Cell Physiol. 280, C962–C969 (2001).CrossRefGoogle ScholarPubMed
Butler, P. J., Tsou, T. C., Li, J. Y., Usami, S., & Chien, S. Rate sensitivity of shear-induced changes in the lateral diffusion of endothelial cell membrane lipids: A role for membrane perturbation in shear-induced MAPK activation 1. FASEB J. 16, 216–218 (2002).CrossRefGoogle Scholar
Butler, P. J., Weinbaum, S., Chien, S., & Lemons, D. E. Endothelium-dependent, shear-induced vasodilation is rate-sensitive. Microcirculation 7, 53–65 (2000).CrossRefGoogle ScholarPubMed
Cantor, R. S. Lipid composition and the lateral pressure profile in bilayers. Biophys. J. 76, 2625–2639 (1999).CrossRefGoogle ScholarPubMed
Casey, P. J. Protein lipidation in cell signaling. Science 268, 221–225 (1995).CrossRefGoogle ScholarPubMed
Charras, G. T. & Horton, M. A. Determination of cellular strains by combined atomic force microscopy and finite element modeling. Biophys. J. 83, 858–879 (2002).CrossRefGoogle ScholarPubMed
Chen, B. P., et al. DNA microarray analysis of gene expression in endothelial cells in response to 24-h shear stress. Physiol. Genomics 7, 55–63 (2001).CrossRefGoogle ScholarPubMed
Chien, S. Molecular basis of rheological modulation of endothelial functions: importance of stress direction. Biorheology 43, 95–116 (2006).Google ScholarPubMed
Chien, S., Li, S., & Shyy, Y. J. Effects of mechanical forces on signal transduction and gene expression in endothelial cells. Hypertension 31, 162–169 (1998).CrossRefGoogle ScholarPubMed
Czarny, M. & Schnitzer, J. E. Neutral sphingomyelinase inhibitor scyphostatin prevents and ceramide mimics mechanotransduction in vascular endothelium. Am. J. Physiol. Heart Circ. Physiol. 287, H1344–H1352 (2004).CrossRefGoogle ScholarPubMed
Dan, N. & Safran, S. A. Effect of lipid characteristics on the structure of transmembrane proteins. Biophys. J. 75, 1410–1414 (1998).CrossRefGoogle ScholarPubMed
Davies, P. F. Flow-mediated endothelial mechanotransduction. Physiolog. Rev. 75, 519–560 (1995).CrossRefGoogle ScholarPubMed
Davies, P. F., et al. Spatial relationships in early signaling events of flow-mediated endothelial mechanotransduction. [Review] [48 refs]. Annual Review of Physiology 59, 527–549 (1997).CrossRefGoogle ScholarPubMed
Davies, P. F., Polacek, D. C., Handen, J. S., Helmke, B. P., & DePaola, N. A spatial approach to transcriptional profiling: Mechanotransduction and the focal origin of atherosclerosis. Trends in Biotechnology 17, 347–351 (1999).CrossRefGoogle Scholar
Davies, P. F., Robotewskyj, A., & Griem, M. L. Quantitative studies of endothelial cell adhesion. Directional remodeling of focal adhesion sites in response to flow forces. J. Clin. Invest. 93, 2031–2038 (1994).Google ScholarPubMed
DePaola, N., Gimbrone, M. A., Davies, P. F., & Dewey, C. F.. Vascular endothelium responds to fluid shear stress gradients. Arterioscler. Thromb. 12, 1254–1257 (1992).CrossRefGoogle ScholarPubMed
Dietrich, C., Yang, B., Fujiwara, T., Kusumi, A., & Jacobson, K. Relationship of lipid rafts to transient confinement zones detected by single particle tracking. Biophys. J. 82, 274–284 (2002).CrossRefGoogle ScholarPubMed
Elson, E. & Magde, D. Fluorescence correlation spectroscopy. I. Conceptual basis and theory. Biopolymers 13, 1–27 (1974).CrossRefGoogle Scholar
Ferko, M. C., Bhatnagar, A., Garcia, M. B., & Butler, P. J. Finite-element stress analysis of a multicomponent model of sheared and focally-adhered endothelial cells. Annals of Biomedical Engineering 35 (2), 208–223 (2007).Google ScholarPubMed
Ferko, M. C., Patterson, B. W., & Butler, P. J. High-resolution solid modeling of biological samples imaged with 3D fluorescence microscopy. Microsc. Res. Tech. 69, 648–655 (2006).CrossRefGoogle ScholarPubMed
Florian, J. A., et al. Heparan sulfate proteoglycan is a mechanosensor on endothelial cells. Circ. Res. 93, e136–e142 (2003).CrossRefGoogle Scholar
Frangos, J. A., Eskin, S. G., McIntire, L. V., & Ives, C. L. Flow effects on prostacyclin production by cultured human endothelial cells. Science 227, 1477–1479 (1985).CrossRefGoogle ScholarPubMed
Frangos, J. A. & Gudi, S. Shear stress activates reconstituted G-proteins in the absence of protein receptors by modulating lipid bilayer fluidity. FASEB J. 11(3), A521 (1997).Google Scholar
Frangos, J. A., Huang, T. Y., & Clark, C. B. Steady shear and step changes in shear stimulate endothelium via independent mechanisms – superposition of transient and sustained nitric oxide production. Biochem. Biophys. Res. Commun. 224, 660–665 (1996).CrossRefGoogle ScholarPubMed
Fujiwara, K., Masuda, M., Osawa, M., Kano, Y., & Katoh, K. Is PECAM-1 a mechanoresponsive molecule?Cell Struct. Funct. 26, 11–17 (2001).CrossRefGoogle ScholarPubMed
Galbraith, C. G., Skalak, R., & Chien, S. Shear stress induces spatial reorganization of the endothelial cell cytoskeleton. Cell Motil. Cytoskeleton 40, 317–330 (1998).3.0.CO;2-8>CrossRefGoogle ScholarPubMed
Garcia-Cardena, G., Comander, J. I., Blackman, B. R., Anderson, K. R., & Gimbrone, M. A. Mechanosensitive endothelial gene expression profiles: scripts for the role of hemodynamics in atherogenesis?Ann. N.Y. Acad. Sci. 947, 1–6 (2001).CrossRefGoogle ScholarPubMed
Garcia-Cardena, G., Oh, P., Liu, J., Schnitzer, J. E., & Sessa, W. C. Targeting of nitric oxide synthase to endothelial cell caveolae via palmitoylation: Implications for nitric oxide signaling. Proc. Natl. Acad. Sci. U.S.A. 93, 6448–6453 (1996).CrossRefGoogle ScholarPubMed
Gaus, K., Le, L. S., Balasubramanian, N., & Schwartz, M. A. Integrin-mediated adhesion regulates membrane order. J. Cell Biol. 174, 725–734 (2006).CrossRefGoogle Scholar
Gautam, M., Shen, Y., Thirkill, T. L., Douglas, G. C., & Barakat, A. I. Flow-activated chloride channels in vascular endothelium: Shear stress sensitivity, desensitization dynamics, and physiological implications. J. Biol. Chem. 281 (48), 36492–36500 (2006).CrossRefGoogle ScholarPubMed
Geiger, R. V., Berk, B. C., Alexander, R. W., & Nerem, R. M. Flow-induced calcium transients in single endothelial cells: Spatial and temporal analysis. Am. J. Physiol. 262, C1411–C1417 (1992).CrossRefGoogle ScholarPubMed
Girard, P. R. & Nerem, R. M. Endothelial cell signaling and cytoskeletal changes in response to shear stress. Front. Med. Biol. Eng. 5, 31–36 (1993).Google ScholarPubMed
Girard, P. R. & Nerem, R. M. Shear stress modulates endothelial cell morphology and F-actin organization through the regulation of focal adhesion-associated proteins. J. Cell Physiol. 163, 179–193 (1995).CrossRefGoogle ScholarPubMed
Gojova, A. & Barakat, A. I. Vascular endothelial wound closure under shear stress: Role of membrane fluidity and flow-sensitive ion channels. J. Appl. Physiol. 98, 2355–2362 (2005).CrossRefGoogle ScholarPubMed
Gopalakrishna, P., Chaubey, S. K., Manogaran, P. S., & Pande, G. Modulation of alpha5beta1 integrin functions by the phospholipid and cholesterol contents of cell membranes. J. Cell Biochem. 77, 517–528 (2000).3.0.CO;2-6>CrossRefGoogle ScholarPubMed
Goulian, M., et al. Gramicidin channel kinetics under tension. Biophys. J. 74, 328–337 (1998).CrossRefGoogle Scholar
Gov, N. S. Diffusion in curved fluid membranes. Phys. Rev. E. Stat. Nonlin. Soft. Matter Phys. 73, 041918 (2006).CrossRefGoogle Scholar
Grabowski, E. F., Jaffe, E. A., & Weksler, B. B. Prostacyclin production by cultured endothelial cell monolayers exposed to step increases in shear stress. Lab. Clin. Med. 105, 36–43 (1985).Google ScholarPubMed
Green, J. M., et al. Role of cholesterol in formation and function of a signaling complex involving alphavbeta3, integrin-associated protein (CD47), and heterotrimeric G proteins. J. Cell Biol. 146, 673–682 (1999).CrossRefGoogle Scholar
Gudi, S., Nolan, J. P., & Frangos, J. A. Modulation of GTPase activity of G proteins by fluid shear stress and phospholipid composition. Proc. Natl. Acad. Sci. U.S.A. 95, 2515–2519 (1998).CrossRefGoogle ScholarPubMed
Gudi, S. R., Clark, C. B., & Frangos, J. A. Fluid flow rapidly activates G proteins in human endothelial cells. Involvement of G proteins in mechanochemical signal transduction. Circ. Res. 79, 834–839 (1996).CrossRefGoogle ScholarPubMed
Gullapalli, R. R., Tabouillot, T., Mathura, R., Dangaria, J., & Butler, P. J. Integrated multimodal microscopy, time resolved fluorescence, and optical-trap rheometry: Toward single molecule mechanobiology. J. Biomed. Opt. 12(1), 014012 (2007).CrossRefGoogle ScholarPubMed
Haidekker, M. A., L’Heureux, N., & Frangos, J. A. Fluid shear stress increases membrane fluidity in endothelial cells: A study with DCVJ fluorescence. Am. J. Physiol. Heart Circ. Physiol. 278, H1401–H1406 (2000).CrossRefGoogle ScholarPubMed
Hamill, O. P. & Martinac, B. Molecular basis of mechanotransduction in living cells. Physiol. Rev. 81, 685–740 (2001).CrossRefGoogle ScholarPubMed
Helmke, B. P., Goldman, R. D., & Davies, P. F. Rapid displacement of vimentin intermediate filaments in living endothelial cells exposed to flow. Circ. Res. 86, 745–752 (2000).CrossRefGoogle Scholar
Honda, H. M., et al. A complex flow pattern of low shear stress and flow reversal promotes monocyte binding to endothelial cells. Atherosclerosis 158, 385–390 (2001).CrossRefGoogle ScholarPubMed
Ingber, D. E. Mechanobiology and diseases of mechanotransduction. Ann. Med. 35, 564–577 (2003).CrossRefGoogle ScholarPubMed
Ishida, T., Takahashi, M., Corson, M. A., & Berk, B. C. Fluid shear stress-mediated signal transduction: How do endothelial cells transduce mechanical force into biological responses?Ann. N.Y. Acad. Sci. 811, 12–23 (1997).Google ScholarPubMed
Jacobs, E. R., et al. Shear activated channels in cell-attached patches of cultured bovine aortic endothelial cells. Pflugers Arch. 431, 129–131 (1995).CrossRefGoogle ScholarPubMed
Jacobson, K., Sheets, E. D., & Simson, R. Revisiting the fluid mosaic model of membranes. Science 268, 1441–1442 (1995).CrossRefGoogle ScholarPubMed
Jalali, S., et al. Integrin-mediated mechanotransduction requires its dynamic interaction with specific extracellular matrix (ECM) ligands. Proc. Natl. Acad. Sci. U.S.A. 98, 1042–1046 (2001).CrossRefGoogle ScholarPubMed
Jo, H., et al. Differential effect of shear stress on extracellular signal-regulated kinase and N-terminal Jun kinase in endothelial cells. Gi2- and Gbeta/gamma-dependent signaling pathways. J. Biol. Chem. 272, 1395–1401 (1997).CrossRefGoogle Scholar
Kaley, G., Koller, A., Messina, E. J., & Wolin, M. S. Role of endothelium-derived vasoactive factors in the control of the microcirculation, in Cardiovascular Significance of Endothelium-Derived Vasoactive Factors (ed. Rubanyi, G. M.), 179–195 (Futura Publishing Co., Mount Kisco, NY, 1991).Google Scholar
Kim, D. W., Langille, B. L., Wong, M. K., & Gotlieb, A. I. Patterns of endothelial microfilament distribution in the rabbit aorta in situ. Circ. Res. 64, 21–31 (1989).CrossRefGoogle ScholarPubMed
Koller, A. & Bagi, Z. On the role of mechanosensitive mechanisms eliciting reactive hyperemia. Am. J. Physiol. Heart Circ. Physiol. 283, H2250–H2259 (2002).CrossRefGoogle Scholar
Koller, A., Sun, D., & Kaley, G. Role of shear stress and endothelial prostaglandins in flow- and viscosity-induced dilation of arterioles in vitro. Circ. Res. 72, 1276–1284 (1993).CrossRefGoogle ScholarPubMed
Korenaga, R., et al. Laminar flow stimulates ATP- and shear stress-dependent nitric oxide production in cultured bovine endothelial cells. Biochem. Biophys. Res. Commun. 198, 213–219 (1994).CrossRefGoogle ScholarPubMed
Ku, D. N., Giddens, D. P., Zarins, C. K., & Glagov, S. Pulsatile flow and atherosclerosis in the human carotid bifurcation. Positive correlation between plaque location and low oscillating shear stress. Arteriosclerosis 5, 293–302 (1985).CrossRefGoogle ScholarPubMed
Kuchan, M. J., Jo, H., & Frangos, J. A. Role of G proteins in shear stress-mediated nitric oxide production by endothelial cells. Am. J. Physiol. 267, C753–C758 (1994).CrossRefGoogle ScholarPubMed
Kuo, L., Davis, M. J., & Chilian, W. M. Endothelium-dependent, flow-induced dilation of isolated coronary arterioles. Am. J. Physiol. 259, H1063–H1070 (1990).Google ScholarPubMed
Ladbrooke, B. D. & Chapman, D. Thermal analysis of lipids, proteins and biological membranes. A review and summary of some recent studies. Chem. Phys. Lipids 3, 304–356 (1969).CrossRefGoogle ScholarPubMed
LaDisa, J. F., Jr., et al. Three-dimensional computational fluid dynamics modeling of alterations in coronary wall shear stress produced by stent implantation. Ann. Biomed. Eng. 31, 972–980 (2003).CrossRefGoogle ScholarPubMed
Lakowicz, J. R. Principles of Fluorescence Spectroscopy (Springer, New York, 1999).CrossRefGoogle Scholar
Lange, Y. The rate of transmembrane movement of cholesterol in the human erythrocyte. J. Biol. Chem. 256, 5321–5323 (1981).Google ScholarPubMed
Lee, A. G. How lipids affect the activities of integral membrane proteins. Biochim. Biophys. Acta 1666, 62–87 (2004).CrossRefGoogle ScholarPubMed
Li, S., et al. The role of the dynamics of focal adhesion kinase in the mechanotaxis of endothelial cells. Proc. Natl. Acad. Sci. U.S.A. 99, 3546–3551 (2002).CrossRefGoogle ScholarPubMed
Li, S., et al. Fluid shear stress activation of focal adhesion kinase. Linking to mitogen-activated protein kinases. J. Biol. Chem. 272, 30455–30462 (1997).CrossRefGoogle ScholarPubMed
Macdonald, A. G. The homeoviscous theory of adaptation applied to excitable membranes: A critical evaluation. Biochim. Biophys. Acta Rev. Biomembr. 1031, 291–310 (1990).CrossRefGoogle ScholarPubMed
Magde, D., Elson, E., & Webb, W. W. Fluorescence correlation spectroscopy. II. An experimental realization. Biopolymers 13, 29–61 (1974).CrossRefGoogle ScholarPubMed
Makino, A., et al. G protein-coupled receptors serve as mechanosensors for fluid shear stress in neutrophils. Am. J. Physiol. Cell Physiol. 290, C1633–C1639 (2006).Google ScholarPubMed
Mateo, C. R. & Douhal, A. A coupled proton-transfer and twisting-motion fluorescence probe for lipid bilayers. Proc. Natl. Acad. Sci. U.S.A. 95, 7245–7250 (1998).CrossRefGoogle ScholarPubMed
McCormick, S. M., et al. DNA microarray reveals changes in gene expression of shear stressed human umbilical vein endothelial cells. Proc. Natl. Acad. Sci. U.S.A. 98, 8955–8960 (2001).CrossRefGoogle ScholarPubMed
Mitchell, D. C., Lawrence, J. T., & Litman, B. J. Primary alcohols modulate the activation of the G protein-coupled receptor rhodopsin by a lipid-mediated mechanism. J. Biol. Chem. 271, 19033–19036 (1996).CrossRefGoogle Scholar
Mochizuki, S., et al. Role of hyaluronic acid glycosaminoglycans in shear-induced endothelium-derived nitric oxide release. Am. J. Physiol. Heart Circ. Physiol. 285, H722–H726 (2003).CrossRefGoogle ScholarPubMed
Mouritsen, O. G. & Bloom, M. Mattress model of lipid-protein interactions in membranes. Biophys. 46, 141–153 (1984).Google ScholarPubMed
Mouritsen, O. G. & Bloom, M. Models of lipid-protein interactions in membranes. Ann. Rev. Biophys. & Biomolec. Struct. 22, 145–171 (1993).CrossRefGoogle ScholarPubMed
Muller, J. M., Chilian, W. M., & Davis, M. J. Integrin signaling transduces shear stress–dependent vasodilation of coronary arterioles. Circ. Res. 80, 320–326 (1997).CrossRefGoogle ScholarPubMed
Nagel, T., Resnick, N., Dewey, C. F., & Gimbrone, M. A.. Vascular endothelial cells respond to spatial gradients in fluid shear stress by enhanced activation of transcription factors. Arterioscler. Thromb. Vasc. Biol. 19, 1825–1834 (1999).CrossRefGoogle ScholarPubMed
Naruse, K. & Sokabe, M. Involvement of stretch-activated ion channels in Ca2+ mobilization to mechanical stretch in endothelial cells. Am. J. Physiol. 264, C1037–C1044 (1993).CrossRefGoogle ScholarPubMed
Nerem, R. M. Role of mechanics in vascular tissue engineering. Biorheology 40, 281–287 (2003).Google ScholarPubMed
Nerem, R. M., Levesque, M. J., & Cornhill, J. F. Vascular endothelial morphology as an indicator of the pattern of blood flow. J. Biomech. Eng. 103, 172–176 (1981).CrossRefGoogle ScholarPubMed
Nielsen, C., Goulian, M., & Andersen, O. S. Energetics of inclusion-induced bilayer deformations. Biophys. J. 74, 1966–1983 (1998).CrossRefGoogle ScholarPubMed
Oh, P. & Schnitzer, J. E. Segregation of heterotrimeric G proteins in cell surface microdomains. G(q) binds caveolin to concentrate in caveolae, whereas G(i) and G(s) target lipid rafts by default. Mol. Biol. Cell 12, 685–698 (2001).CrossRefGoogle Scholar
Ohno, M., Gibbons, G. H., Dzau, V. J., & Cooke, J. P. Shear stress elevates endothelial cGMP. Role of a potassium channel and G protein coupling. Circulation 88, 193–197 (1993).Google ScholarPubMed
Olesen, S. P., Clapham, D. E., & Davies, P. F. Haemodynamic shear stress activates a K+ current in vascular endothelial cells. Nature 331, 168–170 (1988).Google Scholar
Packard, B. S. & Wolf, D. E. Fluorescence lifetimes of carbocyanine lipid analogues in phospholipid bilayers. Biochemistry 24, 5176–5181 (1985).CrossRefGoogle ScholarPubMed
Pries, A. R., Reglin, B., & Secomb, T. W. Structural response of microcirculatory networks to changes in demand: Information transfer by shear stress. Am. J. Physiol. Heart Circ. Physiol. 284, H2204–H2212 (2003).CrossRefGoogle ScholarPubMed
Rizzo, V., Sung, A., Oh, P., & Schnitzer, J. E. Rapid mechanotransduction in situ at the luminal cell surface of vascular endothelium and its caveolae. J. Biol. Chem. 273, 26323–26329 (1998).CrossRefGoogle ScholarPubMed
Saffman, P. G. & Delbruck, M. Brownian motion in biological membranes. Proc. Natl. Acad. Sci. U.S.A. 72, 3111–3113 (1975).CrossRefGoogle ScholarPubMed
Sato, M., Nagayama, K., Kataoka, N., Sasaki, M., & Hane, K. Local mechanical properties measured by atomic force microscopy for cultured bovine endothelial cells exposed to shear stress. J. Biomech. 33, 127–135 (2000).CrossRefGoogle ScholarPubMed
Sato, M., Theret, D. P., Wheeler, L. T., Ohshima, N., & Nerem, R. M. Application of the micropipette technique to the measurement of cultured porcine aortic endothelial cell viscoelastic properties. J. Biomechan. Eng. 112, 263–268 (1990).Google ScholarPubMed
Satoh, A., Toida, T., Yoshida, K., Kojima, K., & Matsumoto, I. New role of glycosaminoglycans on the plasma membrane proposed by their interaction with phosphatidylcholine. FEBS Lett. 477, 249–252 (2000).CrossRefGoogle ScholarPubMed
Secomb, T. W., Hsu, R., & Pries, A. R. Effect of the endothelial surface layer on transmission of fluid shear stress to endothelial cells. Biorheology 38, 143–150 (2001).Google ScholarPubMed
Secomb, T. W. & Pries, A. R. Information transfer in microvascular networks. Microcirculation 9, 377–387 (2002).CrossRefGoogle Scholar
Shinitzky, M. The lipid regulation of receptor functions. Biomembranes and Receptor Mechanisms 7, 135–141 (1987).Google Scholar
Shyy, Y. J., Hsieh, H. J., Usami, S., & Chien, S. Fluid shear stress induces a biphasic response of human monocyte chemotactic protein 1 gene expression in vascular endothelium. Proc. Natl. Acad. Sci. U.S.A. 91, 4678–4682 (1994).CrossRefGoogle ScholarPubMed
Siegel, G., Malmsten, M., & Lindman, B. Flow sensing at the endothelium-blood interface. Colloids and Surfaces A—Physicochemical and Engineering Aspects 138, 345–351 (1998).Google Scholar
Sigurdson, W. J., Sachs, F., & Diamond, S. L. Mechanical perturbation of cultured human endothelial cells causes rapid increases of intracellular calcium. Am. J. Physiol. Heart Circ. Physiol. 264, H1745–H1752 (1993).CrossRefGoogle ScholarPubMed
Simionescu, M., Simionescu, N., & Palade, G. E. Segmental differentiations of cell junctions in the vascular endothelium: The microvasculature. J. Cell Biol. 67, 863–885 (1975).CrossRefGoogle ScholarPubMed
Singer, S. J. & Nicolson, G. L. The fluid mosaic model of the structure of cell membranes. Science 175, 720–731 (1972).CrossRefGoogle ScholarPubMed
Soubias, O., Jolibois, F., Reat, V., & Milon, A. Understanding sterol-membrane interactions, Part II: Complete 1H and 13C assignments by solid-state NMR spectroscopy and determination of the hydrogen-bonding partners of cholesterol in a lipid bilayer. Chemistry 10, 6005–6014 (2004).CrossRefGoogle Scholar
Sowa, G., Pypaert, M., & Sessa, W. C. Distinction between signaling mechanisms in lipid rafts vs. caveolae. Proc. Natl. Acad. Sci. U.S.A. 98, 14072–14077 (2001).CrossRefGoogle ScholarPubMed
Squire, J. M., et al. Quasi-periodic substructure in the microvessel endothelial glycocalyx: A possible explanation for molecular filtering?J. Struct. Biol. 136, 239–255 (2001).CrossRefGoogle ScholarPubMed
Stone, P. H., et al. Effect of endothelial shear stress on the progression of coronary artery disease, vascular remodeling, and in-stent restenosis in humans: In vivo 6-month follow-up study. Circulation 108, 438–444 (2003).CrossRefGoogle ScholarPubMed
Tardy, Y., Resnick, N., Nagel, T., Gimbrone, M. A., & Dewey, C. F.. Shear stress gradients remodel endothelial monolayers in vitro via a cell proliferation-migration-loss cycle. Arterioscler. Thromb. Vasc. Biol. 17, 3102–3106 (1997).CrossRefGoogle Scholar
Trudell, J. R. Role of membrane fluidity in anesthetic action in, Drug and Anesthetic Effects on Membrane Structure and Function (eds. Aloia, R. C., Curtain, C. C., & Gordon, L. M.), 1–14 (Wiley-Liss, Inc., New York, 1991).Google Scholar
Tzima, E., et al. A mechanosensory complex that mediates the endothelial cell response to fluid shear stress. Nature 437, 426–431 (2005).CrossRefGoogle ScholarPubMed
Wang, Y., et al. Visualizing the mechanical activation of Src 112. Nature 434, 1040–1045 (2005).CrossRefGoogle Scholar
Weinbaum, S. & Chien, S. Lipid transport aspects of atherogenesis. J. Biomech. Eng. 115, 602–610 (1993).CrossRefGoogle ScholarPubMed
Weinbaum, S., Zhang, X., Han, Y., Vink, H., & Cowin, S. C. Mechanotransduction and flow across the endothelial glycocalyx. Proc. Natl. Acad. Sci. U.S.A. 100, 7988–7995 (2003).CrossRefGoogle ScholarPubMed
Wentzel, J. J., et al. Shear stress, vascular remodeling and neointimal formation. J. Biomech. 36, 681–688 (2003).CrossRefGoogle ScholarPubMed
Wolf, A. A., Fujinaga, Y., & Lencer, W. I. Uncoupling of the cholera toxin-G(M1) ganglioside receptor complex from endocytosis, retrograde Golgi trafficking, and downstream signal transduction by depletion of membrane cholesterol. J. Biol. Chem. 277, 16249–16256 (2002).CrossRefGoogle ScholarPubMed
Zaidel-Bar, R., Kam, Z., & Geiger, B. Polarized downregulation of the paxillin-p130CAS-Rac1 pathway induced by shear flow. J. Cell Sci. 118, 3997–4007 (2005).CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×