Skip to main content Accessibility help
×
Hostname: page-component-76dd75c94c-nbtfq Total loading time: 0 Render date: 2024-04-30T09:22:15.621Z Has data issue: false hasContentIssue false

2 - Endothelial Mechanotransduction

Published online by Cambridge University Press:  05 July 2014

Peter F. Davies
Affiliation:
University of Pennsylvania
Brian P. Helmke
Affiliation:
University of Virginia
Mohammad R. K. Mofrad
Affiliation:
University of California, Berkeley
Roger D. Kamm
Affiliation:
Massachusetts Institute of Technology
Get access

Summary

Introduction

The endothelium is one of the most intensively studied tissues in cellular mechanotransduction. It is the interface between blood (or lymph) and the underlying vessel walls in arteries, microcirculatory beds, veins, and lymphatics. Mechanotransduction is of particular significance in high-pressure, high-flow arteries where considerable blood flow forces act on the endothelium lining the inner boundaries of the vessel walls. Consequently, endothelial mechanotransduction is studied principally as a flow-mediated mechanism.

Efficient vascular transport systems are central to the evolutionary success of all higher organisms. Throughout phylogeny there is a consistent pattern of structural relationships in branching vessels. For example, much of the mammalian arterial circulation obeys mathematical relationships of vessel geometry that ensure a continuum of flow characteristics (volumetric, velocity, flow profile, and shear relationships), where the major distributing arteries repeatedly branch to provide blood to the complex volume of widely dispersed tissues and organs throughout the body; Murray’s Law [1, 3] and Zamir’s Law [2] are examples. Similar relationships are found in fluid transport systems of primitive marine animals. General principles such as these reflect the interdependence of flow with vessel structure and function throughout evolution that ensures the efficient and successful distribution of fluid in primitive life forms and blood circulation in mammals.

Type
Chapter
Information
Cellular Mechanotransduction
Diverse Perspectives from Molecules to Tissues
, pp. 20 - 60
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

LaBarbera, M. Principles of design of fluid transport systems in zoology. Science 1990; 249:992–1000.CrossRefGoogle ScholarPubMed
Cheng, C, Helderman, F, Tempel, D, Segers, D, Hierck, B, Poelmann, R, et al. Large variations in absolute wall shear stress levels within one species and between species. Atherosclerosis 2007; 195:225–35.CrossRefGoogle ScholarPubMed
Davies, PF. Flow-mediated endothelial mechanotransduction. Physiol. Rev. 1995; 75:519–60.CrossRefGoogle ScholarPubMed
Nerem, RM. Hemodynamics and the vascular endothelium. J. Biomech. Eng. 1993; 115:510–14.CrossRefGoogle ScholarPubMed
Resnick, N, Gimbrone, MA. Hemodynamic forces are complex regulators of endothelial gene expression. FASEB J. 1995; 9:874–82.CrossRefGoogle ScholarPubMed
Davies, PF, Polacek, DC, Handen, JS, Helmke, BP, DePaola, N. A spatial approach to transcriptional profiling: mechanotransduction and the focal origin of atherosclerosis. Trends Biotechnol. 1999; 17:347–51.CrossRefGoogle ScholarPubMed
Shyy, JYJ, Chien, S. Role of integrins in endothelial mechanosensing of shear stress. Circ. Res. 2002; 91:769–75.CrossRefGoogle ScholarPubMed
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. 1993; 72:1276–84.CrossRefGoogle ScholarPubMed
Olesen, S-P, Clapham, DE, Davies, PF. Hemodynamic shear stress activates a K+ current in vascular endothelial cells. Nature 1988; 331:168–70.CrossRefGoogle Scholar
Gudi, SR, Nolan, JP, Frangos, JA. Modulation of GTPase activity of G proteins by fluid shear stress and phospholipid composition. Proc. Natl. Acad. Sci. USA 1998; 95:2515–19.CrossRefGoogle ScholarPubMed
Butler, PJ, Norwich, G, Weinbaum, S, Chien, S. Shear stress induces a time- and position-dependent increase in endothelial cell membrane fluidity. Am. J. Physiol. 2001; 280:C962–9.CrossRefGoogle ScholarPubMed
Aird, WC, ed. Endothelial Biomedicine. New York: Cambridge University Press, 2007.CrossRef
Pohl, U, Holtz, J, Busse, R, Bassenge, E. Crucial role of endothelium in the vasodilator response to increased flow in vivo. Hypertension 1986; 8:37–44.CrossRefGoogle ScholarPubMed
Moncada, S. Adventures in vascular biology: A tale of two mediators. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2006; 361:735–59.CrossRefGoogle ScholarPubMed
Corson, MA, James, NL, Latta, SE, Nerem, RM, Berk, BC, Harrison, DG. Phosphorylation of endothelial nitric oxide synthase in response to fluid shear stress. Circ. Res. 1996; 79:984–91.CrossRefGoogle ScholarPubMed
Griffith, TM. Endothelial control of vascular tone by nitric oxide and gap junctions: a haemodynamic perspective. Biorheology 2002; 39:307–18.Google ScholarPubMed
Langille, BL, O’Donnell, F. Reductions in arterial diameter produced by chronic decreases in blood flow are endothelium-dependent. Science 1986; 231:405–7.CrossRefGoogle ScholarPubMed
Lee, JS, Yu, Q, Shin, JT, Sebzda, E, Bertozzi, C, Chen, M, et al. Klf2 is an essential regulator of vascular hemodynamic forces in vivo. Dev. Cell 2006; 11:845–57.CrossRefGoogle ScholarPubMed
Glagov, S, Zarins, CK, Giddens, DP, Ku, DN. Hemodynamics and atherosclerosis. Insights and perspectives gained from studies of human arteries. Arch. Pathol. Lab. Med. 1988; 112:1018–31.Google ScholarPubMed
Ku, DN, Giddens, DP. Pulsatile flow in a model carotid bifurcation. Arteriosclerosis 1983; 3:31–9.CrossRefGoogle Scholar
Steinberg, D. Atherogenesis in perspective: Hypercholesterolemia and inflammation as partners in crime. Nat. Med. 2002; 8:1211–17.CrossRefGoogle ScholarPubMed
Helmke, BP, Rosen, AB, Davies, PF. Mapping mechanical strain of an endogenous cytoskeletal network in living endothelial cells. Biophys. J. 2003; 84:2691–9.CrossRefGoogle ScholarPubMed
Dull, RO, Davies, PF. Flow modulation of agonist (ATP)-response (Ca2+) coupling in vascular endothelial cells. Am. J. Physiol. 1991; 261:H149–H54.Google ScholarPubMed
Choi, H, Ferrara, K, Barakat, A. Modulation of ATP/ADP concentration at the endothelial surface by shear stress: Effect of flow recirculation. Ann. Biomed. Eng. 2007; 35:505–16.CrossRefGoogle ScholarPubMed
Dull, RO, Tarbell, JM, Davies, PF. Mechanisms of flow-mediated signal transduction in endothelial cells: Kinetics of ATP surface concentrations. J. Vasc. Res. 1992; 29:410–19.CrossRefGoogle ScholarPubMed
Davies, PF, Reidy, MA, Goode, TB, Bowyer, DE. Scanning electron microscopy in the evaluation of endothelial integrity of the fatty lesion in atherosclerosis. Atherosclerosis 1976; 25:125–30.CrossRefGoogle ScholarPubMed
Dewey, CF, Bussolari, SR, Gimbrone, MA, Davies, PF. The dynamic response of vascular endothelial cells to fluid shear stress. J. Biomech. Eng. 1981; 103:177–88.CrossRefGoogle ScholarPubMed
Davies, PF, Dewey, CF, Bussolari, SR, Gordon, EJ, Gimbrone, MA. Influence of hemodynamic forces on vascular endothelial function. In vitro studies of shear stress and pinocytosis in bovine aortic cells. J. Clin. Invest. 1984; 73:1121–9.CrossRefGoogle ScholarPubMed
Davies, PF, Remuzzi, A, Gordon, EJ, Dewey, CF, Gimbrone, MA. Turbulent fluid shear stress induces vascular endothelial cell turnover in vitro. Proc. Natl. Acad. Sci. USA 1986; 83:2114–17.CrossRefGoogle ScholarPubMed
Sprague, EA, Steinbach, BL, Nerem, RM, Schwartz, CJ. Influence of a laminar steady-state fluid-imposed wall shear stress on the binding, internalization, and degradation of low-density lipoproteins by cultured arterial endothelium. Circulation 1987; 76:648–56.CrossRefGoogle ScholarPubMed
Ingber, D. Integrins as mechanochemical transducers. Curr. Opin. Cell Biol. 1991; 3:841–18.CrossRefGoogle ScholarPubMed
Sims, JR, Karp, S, Ingber, DE. Altering the cellular mechanical force balance results in integrated changes in cell, cytoskeletal and nuclear shape. J. Cell Sci. 1992; 103:1215–22.Google ScholarPubMed
Davies, PF, Robotewskyj, A, Griem, ML. Endothelial cell adhesion in real time. Measurements in vitro by tandem scanning confocal image analysis. J. Clin. Invest. 1993; 91:2640–52.CrossRefGoogle ScholarPubMed
Davies, PF, Robotewskyj, A, Griem, ML. Quantitative studies of endothelial cell adhesion: Directional remodeling of focal adhesion sites in response to flow forces. J. Clin. Invest. 1994; 93:2031–8.CrossRefGoogle ScholarPubMed
Liu, Y, Chen, BPC, Lu, M, Zhu, Y, Stemerman, MB, Chien, S, et al. Shear stress activation of SREBP1 in endothelial cells is mediated by integrins. Arterioscler. Thromb. Vasc. Biol. 2002; 22:76–81.CrossRefGoogle ScholarPubMed
Fujiwara, K, Masuda, M, Osawa, M, Kano, Y, Katoh, K. Is PECAM-1 a mechanoresponsive molecule?Cell Struct. Funct. 2001; 26:11–17.CrossRefGoogle ScholarPubMed
Tzima, E, Irani-Tehrani, M, Kiosses, WB, Dejana, E, Schultz, DA, Engelhardt, B, et al. A mechanosensory complex that mediates the endothelial cell response to fluid shear stress. Nature 2005; 437:426–31.CrossRefGoogle ScholarPubMed
Helmke, BP, Goldman, RD, Davies, PF. Rapid displacement of vimentin intermediate filaments in living endothelial cells exposed to flow. Circ. Res. 2000; 86:745–52.CrossRefGoogle Scholar
Helmke, BP, Thakker, DB, Goldman, RD, Davies, PF. Spatiotemporal analysis of flow-induced intermediate filament displacement in living endothelial cells. Biophys. J. 2001; 80:184–94.CrossRefGoogle ScholarPubMed
Mott, RE, Helmke, BP. Mapping the dynamics of shear stress-induced structural changes in endothelial cells. Am. J. Physiol. 2007; 293:C1616–26.CrossRefGoogle ScholarPubMed
Davies, PF, Tripathi, SC. Mechanical stress mechanisms and the cell: An endothelial paradigm. Circ. Res. 1993; 72:239–45.CrossRefGoogle ScholarPubMed
Helmke, BP, Davies, PF. The cytoskeleton under external fluid mechanical forces: Hemodynamic forces acting on the endothelium. Ann. Biomed. Eng. 2002; 30:284–96.CrossRefGoogle ScholarPubMed
Jacobs, ER, Cheliakine, C, Gebremedhin, D, Davies, PF, Harder, DR. Shear activated channels in cell attached patches of vascular endothelial cells. Pflügers Arch. 1995; 431:129–31.CrossRefGoogle Scholar
Barbee, KA, Davies, PF, Lal, R. Shear stress-induced reorganization of the surface topography of living endothelial cells imaged by atomic force microscopy. Circ. Res. 1994; 74:163–71.CrossRefGoogle ScholarPubMed
Barbee, KA, Davies, PF, Lal, R. Subcellular distribution of shear stress at the surface of flow aligned and non-aligned endothelial monolayers. Am. J. Physiol. 1995; 268:H1765–72.Google Scholar
Weinbaum, S, Tarbell, JM, Damiano, ER. The structure and function of the endothelial glycocalyx layer. Annu. Rev. Biomed. Eng. 2007; 9:121–67.CrossRefGoogle ScholarPubMed
Malek, AM, Izumo, S. Mechanism of endothelial cell shape change and cytoskeletal remodeling in response to fluid shear stress. J. Cell Sci. 1996; 109:713–26.Google ScholarPubMed
Helmke, BP. Molecular control of cytoskeletal mechanics by hemodynamic forces. Physiology 2005; 20:43–53.CrossRefGoogle ScholarPubMed
Hu, S, Chen, J, Fabry, B, Numaguchi, Y, Gouldstone, A, Ingber, DE, et al. Intracellular stress tomography reveals stress focusing and structural anisotropy in cytoskeleton of living cells. Am. J. Physiol. 2003; 285:C1082–90.CrossRefGoogle ScholarPubMed
Hu, S, Wang, N. Control of stress propagation in the cytoplasm by prestress and loading frequency. Mol. Cell. Biomech. 2006; 3:49–60.Google ScholarPubMed
Janmey, PA, Euteneuer, U, Traub, P, Schliwa, M. Viscoelastic properties of vimentin compared with other filamentous biopolymer networks. J. Cell Biol. 1991; 113:155–60.CrossRefGoogle ScholarPubMed
Janmey, PA, Euteneuer, U, Traub, P, Schliwa, M. Viscoelasticity of intermediate filament networks. Sub-Cellular Biochem. 1998; 31:381–97.Google ScholarPubMed
Schmid-Schönbein, GW, Sung, K-LP, Tözeren, H, Skalak, R, Chien, S. Passive mechanical properties of human leukocytes. Biophys. J. 1981; 36:243–56.CrossRefGoogle ScholarPubMed
Theret, DP, Levesque, MJ, Sato, M, Nerem, RM, Wheeler, LT. The application of a homogeneous half-space model in the analysis of endothelial cell micropipette measurements. J. Biomech. Eng. 1988; 110:190–9.CrossRefGoogle ScholarPubMed
Sato, M, Theret, DP, Wheeler, LT, Ohshima, N, Nerem, RM. Application of the micropipette technique to the measurement of cultured porcine aortic endothelial cell viscoelastic properties. J. Biomech. Eng. 1990; 112:263–8.Google ScholarPubMed
Colangelo, S, Langille, BL, Steiner, G, Gotlieb, AI. Alterations in endothelial F-actin microfilaments in rabbit aorta in hypercholesterolemia. Arterioscler. Thromb. Vasc. Biol. 1998; 18:52–6.CrossRefGoogle ScholarPubMed
Galbraith, CG, Skalak, R, Chien, S. Shear stress induces spatial reorganization of the endothelial cell cytoskeleton. Cell Motil. Cytoskel. 1998; 40:317–30.3.0.CO;2-8>CrossRefGoogle ScholarPubMed
Sato, M, Ohshima, N, Nerem, RM. Viscoelastic properties of cultured porcine aortic endothelial cells exposed to shear stress. J. Biomech. 1996; 29:461–7.CrossRefGoogle ScholarPubMed
Satcher, RL, Dewey, CF. Theoretical estimates of mechanical properties of the endothelial cell cytoskeleton. Biophys. J. 1996; 71:109–18.CrossRefGoogle ScholarPubMed
Ferko, MC, Bhatnagar, A, Garcia, MB, Butler, PJ. Finite-element stress analysis of a multicomponent model of sheared and focally-adhered endothelial cells. Ann. Biomed. Eng. 2007; 35:208–23.CrossRefGoogle ScholarPubMed
Fuller, RB. Tensegrity. Portfolio Art News Annu. 1961; 4:112–27.Google Scholar
Wang, N, Butler, JP, Ingber, DE. Mechanotransduction across the cell surface and through the cytoskeleton. Science 1993; 260:1124–7.CrossRefGoogle ScholarPubMed
Fabry, B, Maksym, GN, Butler, JP, Glogauer, M, Navajas, D, Fredberg, JJ. Scaling the microrheology of living cells. Phys. Rev. Lett. 2001; 87:148102.CrossRefGoogle ScholarPubMed
Fabry, B, Maksym, GN, Butler, JP, Glogauer, M, Navajas, D, Taback, NA, et al. Time scale and other invariants of integrative mechanical behavior in living cells. Phys. Rev. E 2003; 68(4 pt. 1):041914.CrossRefGoogle ScholarPubMed
Bursac, P, Lenormand, G, Fabry, B, Oliver, M, Weitz, DA, Viasnoff, V, et al. Cytoskeletal remodeling and slow dynamics in the living cell. Nat. Mater. 2005; 4:557–61.CrossRefGoogle Scholar
Wootton, DM, Ku, DN. Fluid mechanics of vascular systems, diseases, and thrombosis. Annu. Rev. Biomed. Eng. 1999; 1:299–329.CrossRefGoogle ScholarPubMed
Friedman, MH, Giddens, DP. Blood flow in major blood vessels–Modeling and experiments. Ann. Biomed. Eng. 2005; 33:1710–13.CrossRefGoogle ScholarPubMed
Fung, YC, Liu, SQ. Elementary mechanics of the endothelium of blood vessels. J. Biomech. Eng. 1993; 115:1–12.CrossRefGoogle ScholarPubMed
Haidekker, MA, L’Heureux, N, Frangos, JA. Fluid shear stress increases membrane fluidity in endothelial cells: A study with DCVJ fluorescence. Am. J. Physiol. 2000; 278:H1401–6.Google ScholarPubMed
Fang, Y, Schram, G, Romanenko, VG, Shi, C, Conti, L, Vandenberg, CA, et al. Functional expression of Kir2.x in human aortic endothelial cells: The dominant role of Kir2.2. Am. J. Physiol. 2005; 289:C1134–44.CrossRefGoogle ScholarPubMed
Sukharev, SI, Sigurdson, WJ, Kung, C, Sachs, F. Energetic and spatial parameters for gating of the bacterial large conductance mechanosensitive channel, MscL. J. Gen. Physiol. 1999; 113:525–40.CrossRefGoogle ScholarPubMed
Hamill, OP, Martinac, B. Molecular basis of mechanotransduction in living cells. Physiol. Rev. 2001; 81:685–740.CrossRefGoogle ScholarPubMed
Martinac, B, Hamill, OP. Gramicidin A channels switch between stretch activation and stretch inactivation depending on bilayer thickness. Proc. Natl. Acad. Sci. USA 2002; 99:4308–12.CrossRefGoogle ScholarPubMed
Chachisvilis, M, Zhang, Y-L, Frangos, JA. G protein-coupled receptors sense fluid shear stress in endothelial cells. Proc. Natl. Acad. Sci. USA 2006; 103:15463–8.CrossRefGoogle ScholarPubMed
Romanenko, VG, Davies, PF, Levitan, I. Dual effect of fluid shear stress on volume-regulated anion current in bovine aortic endothelial cells. Am. J. Physiol. 2002; 282:C708–18.CrossRefGoogle ScholarPubMed
Balaban, NQ, Schwarz, US, Riveline, D, Goichberg, P, Tzur, G, Sabanay, I, et al. Force and focal adhesion assembly: A close relationship studied using elastic micropatterned substrates. Nat. Cell Biol. 2001; 3:466–72.CrossRefGoogle ScholarPubMed
Reitsma, S, Slaaf, D, Vink, H, van Zandvoort, M, oude Egbrink, M. The endothelial glycocalyx: Composition, functions, and visualization. Pflügers Arch. 2007; 454:345–59.CrossRefGoogle ScholarPubMed
Smith, ML, Long, DS, Damiano, ER, Ley, K. Near-wall μ-PIV reveals a hydrodynamically relevant endothelial surface layer in venules in vivo. Biophys. J. 2003; 85:637–45.CrossRefGoogle Scholar
Weinbaum, S, Zhang, X, Han, Y, Vink, H, Cowin, SC. Mechanotransduction and flow across the endothelial glycocalyx. Proc. Natl. Acad. Sci. USA 2003; 100:7988–95.CrossRefGoogle ScholarPubMed
Thi, MM, Tarbell, JM, Weinbaum, S, Spray, DC. The role of the glycocalyx in reorganization of the actin cytoskeleton under fluid shear stress: A “bumper-car” model. Proc. Natl. Acad. Sci. USA 2004; 101:16483–8.CrossRefGoogle ScholarPubMed
van den Berg, BM, Spaan, JAE, Rolf, TM, Vink, H. Atherogenic region and diet diminish glycocalyx dimension and increase intima-to-media ratios at murine carotid artery bifurcation. Am. J. Physiol. 2006; 290:H915–20.Google ScholarPubMed
Vink, H, Constantinescu, AA, Spaan, JA. Oxidized lipoproteins degrade the endothelial surface layer: Implications for platelet-endothelial cell adhesion. Circulation 2000; 101:1500–2.CrossRefGoogle ScholarPubMed
Pohl, U, Herlan, K, Huang, A, Bassenge, E. EDRF-mediated shear-induced dilation opposes myogenic vasoconstriction in small rabbit arteries. Am. J. Physiol. 1991; 261:H2016–23.Google ScholarPubMed
Hecker, M, Mulsch, A, Bassenge, E, Busse, R. Vasoconstriction and increased flow: Two principal mechanisms of shear stress-dependent endothelial autacoid release. Am. J. Physiol. 1993; 265:H828–33.Google ScholarPubMed
Pahakis, MY, Kosky, JR, Dull, RO, Tarbell, JM. The role of endothelial glycocalyx components in mechanotransduction of fluid shear stress. Biochem. Biophys. Res. Comm. 2007; 355:228–33.CrossRefGoogle ScholarPubMed
Florian, JA, Kosky, JR, Ainslie, K, Pang, Z, Dull, RO, Tarbell, JM. Heparan sulfate proteoglycan is a mechanosensor on endothelial cells. Circ. Res. 2003; 93:14.CrossRefGoogle ScholarPubMed
Gouverneur, M, Spaan, JAE, Pannekoek, H, Fontijn, RD, Vink, H. Fluid shear stress stimulates incorporation of hyaluronan into endothelial cell glycocalyx. Am. J. Physiol. 2006; 290:H458–2.Google ScholarPubMed
Yao, Y, Rabodzey, A, Dewey, CF. Glycocalyx modulates the motility and proliferative response of vascular endothelium to fluid shear stress. Am. J. Physiol. 2007; 293:H1023–30.Google ScholarPubMed
Kojimahara, M. Endothelial cilia in rat mesenteric arteries and intramyocardial capillaries. Z. Mikrosk. Anat. Forsch. 1990; 104:412–6.Google ScholarPubMed
Iomini, C, Tejada, K, Mo, W, Vaananen, H, Piperno, G. Primary cilia of human endothelial cells disassemble under laminar shear stress. J. Cell Biol. 2004; 164:811–17.CrossRefGoogle ScholarPubMed
van der Heiden, K, Groenendijk, BCW, Hierck, BP, Hogers, B, Koerten, HK, Mommaas, AM, et al. Monocilia on chicken embryonic endocardium in low shear stress areas. Developmental Dynamics 2006; 235:19–28.CrossRefGoogle ScholarPubMed
van der Heiden, K, Hierck, BP, Krams, R, de Crom, R, Cheng, C, Baiker, M, et al. Endothelial primary cilia in areas of disturbed flow are at the base of atherosclerosis. Atherosclerosis 2007; DOI:10.1016/j.atherosclerosis. 2007.05.030.Google ScholarPubMed
Ando, J, Komatsuda, T, Kamiya, A. Cytoplasmic calcium response to fluid shear stress in cultured vascular endothelial cells. In Vitro Cell. Dev. Biol. 1988; 24:871–7.CrossRefGoogle ScholarPubMed
Isshiki, M, Ando, J, Korenaga, R, Kogo, H, Fujimoto, T, Fujita, T, et al. Endothelial Ca2+ waves preferentially originate at specific loci in caveolin-rich cell edges. Proc. Natl. Acad. Sci. USA 1998; 95:5009–14.CrossRefGoogle ScholarPubMed
Isshiki, M, Ando, J, Yamamoto, K, Fujita, T, Ying, Y, Anderson, RGW. Sites of Ca2+ wave initiation move with caveolae to the trailing edge of migrating cells. J. Cell Sci. 2002; 115:475–84.Google Scholar
Wang, Y, Botvinick, EL, Zhao, Y, Berns, MW, Usami, S, Tsien, RY, et al. Visualizing the mechanical activation of Src. Nature 2005; 434:1040–5.CrossRefGoogle ScholarPubMed
Tzima, E, del Pozo, MA, Kiosses, WB, Mohamed, SA, Li, S, Chien, S, et al. Activation of Rac1 by shear stress in endothelial cells mediates both cytoskeletal reorganization and effects on gene expression. EMBO J. 2002; 21:6791–800.CrossRefGoogle ScholarPubMed
Yamada, S, Wirtz, D, Kuo, SC. Mechanics of living cells measured by laser tracking microrheology. Biophys. J. 2000; 78:1736–47.CrossRefGoogle ScholarPubMed
Crocker, JC, Hoffman, BD. Multiple-particle tracking and two-point microrheology in cells. Meth. Cell Biol. 2007; 83:141–78.CrossRefGoogle ScholarPubMed
Panorchan, P, Lee, JSH, Daniels, BR, Kole, TP, Tseng, Y, Wirtz, D. Probing cellular mechanical responses to stimuli using ballistic intracellular nanorheology. Meth. Cell Biol. 2007; 83:115–40.Google ScholarPubMed
Stamenovic, D, Mijailovich, SM, Tolic-Norrelykke, IM, Chen, J, Wang, N. Cell prestress. II. Contribution of microtubules. Am. J. Physiol. 2002; 282:C617–C24.CrossRefGoogle ScholarPubMed
Wang, N, Stamenovic, D. Contribution of intermediate filaments to cell stiffness, stiffening, and growth. Am. J. Physiol. 2000; 279:C188–94.CrossRefGoogle Scholar
Giancotti, FG, Ruoslahti, E. Integrin signaling. Science 1999; 285:1028–33.CrossRefGoogle ScholarPubMed
Davies, PF, Barbee, KA, Volin, MV, Robotewskyj, A, Chen, J, Joseph, L, et al. Spatial relationships in early signaling events of flow-mediated endothelial mechanotransduction. Annu. Rev. Physiol. 1997; 59:527–49.CrossRefGoogle ScholarPubMed
Li, S, Kim, M, Hu, YL, Jalali, S, Schlaepfer, DD, Hunter, T, et al. Fluid shear stress activation of focal adhesion kinase. Linking to mitogen-activated protein kinases. J. Biol. Chem. 1997; 272:30455–62.CrossRefGoogle ScholarPubMed
Chen, K-D, Li, Y-S, Kim, M, Li, S, Yuan, S, Chien, S, et al. Mechanotransduction in response to shear stress: Roles of receptor tyrosine kinases, integrins, and Shc. J. Biol. Chem. 1999; 274:18393–400.CrossRefGoogle ScholarPubMed
Jalali, S, del Pozo, MA, Chen, K-D, Miao, H, Li, Y-S, Schwartz, MA, et al. Integrin-mediated mechanotransduction requires its dynamic interaction with specific extracellular matrix (ECM) ligands. Proc. Natl. Acad. Sci. USA 2001; 98:1042–6.CrossRefGoogle ScholarPubMed
Hu, Y-L, Chien, S. Dynamic motion of paxillin on actin filaments in living endothelial cells. Biochem. Biophys. Res. Comm. 2007; 357:871–6.CrossRefGoogle ScholarPubMed
Tai, L-K, Okuda, M, Abe, J-I, Yan, C, Berk, BC. Fluid shear stress activates proline-rich tyrosine kinase via reactive oxygen species-dependent pathway. Arterioscler. Thromb. Vasc. Biol. 2002; 22:1790–6.CrossRefGoogle ScholarPubMed
Rizzo, V, Morton, C, DePaola, N, Schnitzer, JE, Davies, PF. Recruitment of endothelial caveolae into mechanotransduction pathways by flow conditioning in vitro. Am. J. Physiol. 2003; 285:H1720–H9.Google ScholarPubMed
Radel, C, Rizzo, V. Integrin mechanotransduction stimulates caveolin-1 phosphorylation and recruitment of Csk to mediate actin reorganization. Am. J. Physiol. 2005; 288:H936–H45.Google ScholarPubMed
Tzima, E, del Pozo, MA, Shattil, SJ, Chien, S, Schwartz, MA. Activation of integrins in endothelial cells by fluid shear stress mediates Rho-dependent cytoskeletal alignment. EMBO J. 2001; 20:4639–47.CrossRefGoogle ScholarPubMed
Ren, XD, Kiosses, WB, Sieg, DJ, Otey, CA, Schlaepfer, DD, Schwartz, MA. Focal adhesion kinase suppresses Rho activity to promote focal adhesion turnover. J. Cell Sci. 2000; 113:3673–8.Google ScholarPubMed
Arthur, WT, Burridge, K. RhoA inactivation by p190RhoGAP regulates cell spreading and migration by promoting membrane protrusion and polarity. Mol. Biol. Cell 2001; 12:2711–20.CrossRefGoogle ScholarPubMed
Maekawa, M, Ishizaki, T, Boku, S, Watanabe, N, Fujita, A, Iwamatsu, A, et al. Signaling from Rho to the actin cytoskeleton through protein kinases ROCK and LIM-kinase. Science 1999; 285:895–8.CrossRefGoogle ScholarPubMed
Orr, AW, Sanders, JM, Bevard, M, Coleman, E, Sarembock, IJ, Schwartz, MA. The subendothelial extracellular matrix modulates NF-κB activation by flow: A potential role in atherosclerosis. J. Cell Biol. 2005; 169:191–202.CrossRefGoogle ScholarPubMed
Orr, AW, Ginsberg, MH, Shattil, SJ, Deckmyn, H, Schwartz, MA. Matrix-specific suppression of integrin activation in shear stress signaling. Mol. Biol. Cell 2006; 17:4686–97.CrossRefGoogle ScholarPubMed
Orr, AW, Stockton, R, Simmers, MB, Sanders, JM, Sarembock, IJ, Blackman, BR, et al. Matrix-specific p21-activated kinase activation regulates vascular permeability in atherogenesis. J. Cell Biol. 2007; 176:719–27.CrossRefGoogle ScholarPubMed
Kano, Y, Katoh, K, Masuda, M, Fujiwara, K. Macromolecular composition of stress fiber-plasma membrane attachment sites in endothelial cells in situ. Circ. Res. 1996; 79:1000–6.CrossRefGoogle ScholarPubMed
Osawa, M, Masuda, M, Kusano K-i, Fujiwara K. Evidence for a role of platelet endothelial cell adhesion molecule-1 in endothelial cell mechanosignal transduction: Is it a mechanoresponsive molecule?J. Cell Biol. 2002; 158:773–85.CrossRefGoogle ScholarPubMed
Tai, L-k, Zheng, Q, Pan, S, Jin, Z-G, Berk, BC. Flow activates ERK1/2 and endothelial nitric oxide synthase via a pathway involving PECAM1, SHP2, and Tie2. J. Biol. Chem. 2005; 280:29620–4.CrossRefGoogle Scholar
Shay-Salit, A, Shushy, M, Wolfovitz, E, Yahav, H, Breviario, F, Dejana, E, et al. VEGF receptor 2 and the adherens junction as a mechanical transducer in vascular endothelial cells. Proc. Natl. Acad. Sci. USA 2002; 99:9462–7.CrossRefGoogle ScholarPubMed
Cornhill, JF, Roach, MR. A quantitative study of the localization of atherosclerotic lesions in the rabbit aorta. Atherosclerosis 1976; 23:489–501.CrossRefGoogle ScholarPubMed
Steinman, DA. Image-based computational fluid dynamics modeling in realistic arterial geometries. Ann. Biomed. Eng. 2002; 30:483–97.CrossRefGoogle ScholarPubMed
Caro, CG, Fitz-Gerald, JM, Schroter, RC. Arterial wall shear and distribution of early atheroma in man. Nature 1969; 223:1159–61.CrossRefGoogle ScholarPubMed
Lutz, RJ, Cannon, JN, Bischoff, KB, Dedrick, RL, Stiles, RK, Fry, DL. Wall shear stress distribution in a model canine artery during steady flow. Circ. Res. 1977; 41:391–9.CrossRefGoogle Scholar
Suo, J, Ferrara, DE, Sorescu, D, Guldberg, RE, Taylor, WR, Giddens, DP. Hemodynamic shear stresses in mouse aortas: Implications for atherogenesis. Arterioscler. Thromb. Vasc. Biol. 2007; 27:346–51.CrossRefGoogle ScholarPubMed
Passerini, AG, Polacek, DC, Shi, C, Francesco, NM, Manduchi, E, Grant, GR, et al. Coexisting proinflammatory and antioxidative endothelial transcription profiles in a disturbed flow region of the adult porcine aorta. Proc. Natl. Acad. Sci. USA 2004; 101:2482–7.CrossRefGoogle Scholar
Volger, OL, Fledderus, JO, Kisters, N, Fontijn, RD, Moerland, PD, Kuiper, J, et al. Distinctive expression of chemokines and transforming growth factor-β signaling in human arterial endothelium during atherosclerosis. Am. J. Pathol. 2007; 171:326–37.CrossRefGoogle ScholarPubMed
Simmons, CA, Grant, GR, Manduchi, E, Davies, PF. Spatial heterogeneity of endothelial phenotypes correlates with side-specific vulnerability to calcification in normal porcine aortic valves. Circ. Res. 2005; 96:792–9.CrossRefGoogle ScholarPubMed
Magid, R, Davies, PF. Endothelial protein kinase C isoform identity and differential activity of PKCζ in an athero-susceptible region of porcine aorta. Circ. Res. 2005; 97:443–9.CrossRefGoogle Scholar
Boon, RA, Fledderus, JO, Volger, OL, van Wanrooij, EJA, Pardali, E, Weesie, F, et al. KLF2 suppresses TGF-β signaling in endothelium through induction of Smad7 and inhibition of AP-1. Arterioscler. Thromb. Vasc. Biol. 2007; 27:532–9.CrossRefGoogle ScholarPubMed
Dai, G, Vaughn, S, Zhang, Y, Wang, ET, Garcia-Cardena, G, Gimbrone, MA. Biomechanical forces in atherosclerosis-resistant vascular regions regulate endothelial redox balance via phosphoinositol 3-kinase/Akt-dependent activation of Nrf2. Circ. Res. 2007; 101:723–33.CrossRefGoogle ScholarPubMed
Hajra, L, Evans, AI, Chen, M, Hyduk, SJ, Collins, T, Cybulsky, MI. The NF-κB signal transduction pathway in aortic endothelial cells is primed for activation in regions predisposed to atherosclerotic lesion formation. Proc. Natl. Acad. Sci. USA 2000; 97:9052–7.CrossRefGoogle ScholarPubMed
Dekker, RJ, van Soest, S, Fontijn, RD, Salamanca, S, de Groot, PG, VanBavel, E, et al. Prolonged fluid shear stress induces a distinct set of endothelial cell genes, most specifically lung Kruppel-like factor (KLF2). Blood 2002; 100:1689–98.CrossRefGoogle Scholar
Dekker, RJ, Boon, RA, Rondaij, MG, Kragt, A, Volger, OL, Elderkamp, YW, et al. KLF2 provokes a gene expression pattern that establishes functional quiescent differentiation of the endothelium. Blood 2006; 107: 4354–63.CrossRefGoogle ScholarPubMed
Parmar, KM, Larman, HB, Dai, G, Zhang, Y, Wang, ET, Moorthy, SN, et al. Integration of flow-dependent endothelial phenotypes by Kruppel-like factor 2. J. Clin. Invest. 2006; 116:49–58.CrossRefGoogle ScholarPubMed
Cheng, C, van Haperen, R, de Waard, M, van Damme, LCA, Tempel, D, Hanemaaijer, L, et al. Shear stress affects the intracellular distribution of eNOS: Direct demonstration by a novel in vivo technique. Blood 2005; 106:3691–8.CrossRefGoogle ScholarPubMed
Schoen, FJ, Levy, RJ. Tissue heart valves: Current challenges and future research perspectives. J. Biomed. Mater. Res. 1999; 47:439–65.3.0.CO;2-O>CrossRefGoogle ScholarPubMed
Poggianti, E, Venneri, L, Chubuchny, V, Jambrik, Z, Baroncini, LA, Picano, E. Aortic valve sclerosis is associated with systemic endothelial dysfunction. J. Amer. Coll. Cardiol. 2003; 41:136–41.CrossRefGoogle ScholarPubMed
Sacks, MS, Yoganathan, AP. Heart valve function: A biomechanical perspective. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2007; 362:1369–91.CrossRefGoogle ScholarPubMed
Davies, PF, Mundel, T, Barbee, KA. A mechanism for heterogeneous endothelial responses to flow in vivo and in vitro. J. Biomech. 1995; 28:1553–60.CrossRefGoogle ScholarPubMed
Mott, RE, Helmke, BP. Control of endothelial cell adhesion by mechanotransmission from cytoskeleton to substrate. In: King, MR, ed. Principles of Cellular Engineering: Understanding the Biomolecular Interface. Burlington, MA: Elsevier; 2006:25–50.CrossRefGoogle Scholar
Karino, T, Goldsmith, HL. Particle flow behavior in models of branching vessels. II. Effects of branching angle and diameter ratio on flow patterns. Biorheology 1985; 22:87–104.CrossRefGoogle ScholarPubMed
Steinman, D, Taylor, C. Flow imaging and computing: Large artery hemodynamics. Ann. Biomed. Eng. 2005; 33:1704–9.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
×