Skip to main content
Log in

Fluid Dynamics of Heart Development

  • Review Paper
  • Published:
Cell Biochemistry and Biophysics Aims and scope Submit manuscript

Abstract

The morphology, muscle mechanics, fluid dynamics, conduction properties, and molecular biology of the developing embryonic heart have received much attention in recent years due to the importance of both fluid and elastic forces in shaping the heart as well as the striking relationship between the heart’s evolution and development. Although few studies have directly addressed the connection between fluid dynamics and heart development, a number of studies suggest that fluids may play a key role in morphogenic signaling. For example, fluid shear stress may trigger biochemical cascades within the endothelial cells of the developing heart that regulate chamber and valve morphogenesis. Myocardial activity generates forces on the intracardiac blood, creating pressure gradients across the cardiac wall. These pressures may also serve as epigenetic signals. In this article, the fluid dynamics of the early stages of heart development is reviewed. The relevant work in cardiac morphology, muscle mechanics, regulatory networks, and electrophysiology is also reviewed in the context of intracardial fluid dynamics.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

References

  1. Bartman, T., Walsh, E. C., Wen, K. K., McKane, M., Ren, J., Alexander, J., et al. (2004). Early myocardial function affects endocardial cushion development in zebrafish. PLoS Biology, 2, 673–681.

    Article  CAS  Google Scholar 

  2. Bennett, S. (1963). Morphological aspects of extracellular polysaccharides. Journal of Histochemistry and Cytochemistry, 11, 1423.

    Article  Google Scholar 

  3. Biben, C., Weber, R., Kesteven, S., Stanley, E., & McDonald, L. (2000). Cardiac septal and valvular dysmorphogenesis in mice heterozygous for mutations in the homeobox gene nkx2–5. Circulation Research, 87, 888–895.

    PubMed  CAS  Google Scholar 

  4. Biechler, S. V., Potts, J. D., Yost, M. J., Junor, L., Goodwin, R. L., & Weidner, J. W. (2010). Mathematical modeling of flow-generated forces in an in vitro system of cardiac valve development. Annals of Biomedical Engineering, 38(1), 109–117.

    Article  PubMed  Google Scholar 

  5. Boldt, T., Andersson, S., & Eronen, M. (2002). Outcome of structural heart disease diagnosed in utero. Scandinavian Cardiovascualr Journal, 36(2), 73–79.

    Article  Google Scholar 

  6. Bove, E. L., de Leval, M. R., Migliavacca, F., Guadagni, G., & Dubini, G. (2003). Computational fluid dynamics in the evaluation of hemodynamic performance of cavopulmonary connections after the norwood procedure for hypoplastic left heart syndrome. The Journal of Thoracic and Cardiovascular Surgery, 126(4), 1040–1047.

    Article  PubMed  Google Scholar 

  7. Bringley, T. T., Childress, S., Vandenberghe, N., & Zhang, J. (2008). An experimental investigation and a simple model of a valveless pump. Physics of Fluids, 20, 033602-1–033602-15.

    Article  CAS  Google Scholar 

  8. Brinkman, H. C. (1947). A calculation of the viscous force exerted by a flowing fluid on a dense swarm of particles. Applied Sciences Research Section A, 1, 27–34.

    Article  Google Scholar 

  9. Broboana, D., Muntean, T., & Balan, C. (2007). Experimental and numerical studies of weakly elastic viscous fluids in a Hele-Shaw geometry. Proceedings of the Romanian Academy, 8, 1–15.

    Google Scholar 

  10. Bruneau, B. G., Nemer, G., Schmitt, J. P., Charron, F., & Robitaille, L. (2001). A murine model of holtoram syndrome defines roles of the t-box transcription factor tbx5 in cardiogenesis and disease. Cell, 106, 709–721.

    Article  PubMed  CAS  Google Scholar 

  11. Burggren, W. W. (2004). What is the purpose of the embryonic heart beat? Or how facts can ultimately prevail over physiological dogma. Physiological and Biochemical Zoology, 77, 333–345.

    Article  PubMed  Google Scholar 

  12. Camenisch, T. D., Spicer, A. P., Brehm-Gibson, T., Biesterfeldt, J., & Augustine, M. L. (2000). Disruption of hyaluronan synthase-2 abrogates normal cardiac morphogenesis and hyaluronan-mediated transformation of epithelium to mesenchyme. Journal of Clinical Investigation, 106, 349–360.

    Article  PubMed  CAS  Google Scholar 

  13. Chapman, W. B. (1918). The effect of the heart-beat upon the development of the vascular system in the chick. The American Journal of Anatomy, 23, 175–203.

    Article  Google Scholar 

  14. Chien, S., Usami, S., Dellenback, R. J., & Gregersen, M. I. (1970). Shear-dependent deformation of erythrocytes in rheology of human blood. American Journal of Physiology, 219, 136–142.

    PubMed  CAS  Google Scholar 

  15. Cortez, R. (2001). The method of regularized Stokeslets. SIAM Journal of Scientific Computing, 23(4), 1204–1225.

    Google Scholar 

  16. Cokelet, G. R., Merrill, E. W., & Gilliland, E. R. (1963). The rheology of human blood: Measurement near and at zero shear rate. Transactions. Society of Rheology, 7, 303–317.

    Article  Google Scholar 

  17. Crowl, L. M., & Fogelson, A. L. (2009). Computational model of whole blood exhibiting lateral platelet motion induced by red blood cells. Communications in Numerical Methods in Engineering, 26, 471–487.

    Google Scholar 

  18. Damiano, E. R., Long, D. S., & Smith, M. L. (2004). Estimation of viscosity profiles using velocimetry data from parallel flows of linearly viscous: Application to microvascular hemodynamics. Journal of Fluid Mechanics, 512, 119.

    Article  Google Scholar 

  19. Davies, P. F. (1995). Flow-mediated endothelial mechanotransduction. Physiological Reviews, 75, 519–560.

    PubMed  CAS  Google Scholar 

  20. DeGroff, C. G., Thornburg, B. L., Pentecost, J. O., Thornburg, K. L., Gharib, M., Sahn, D. J., et al. (2003). Flow in the early embryonic human heart. Pediatric Cardiology, 24(4), 375–380.

    Article  PubMed  CAS  Google Scholar 

  21. Dekker, R. J., Van Soest, S., Fontijn, R. D., Salamanca, S., de Groot, P. G., VanBavel, E., et al. (2002). Prolonged fluid shear stress induces a distinct set of endothelial cell genes, most specifically lung Krüppel-like factor (KLF2). Blood, 100, 1689–1698.

    Article  PubMed  CAS  Google Scholar 

  22. Dewey, C. F., Jr., Bussolari, S. R., Gimbrone, M. A., Jr., & Davies, P. F. (1991). The dynamic response of vascular endothelial cells to fluid shear stress. Journal of Biomechanical Engineering, 103, 177–185.

    Article  Google Scholar 

  23. Fahraeus, R., & Lindqvist, T. (1931). The viscosity of the blood in narrow capillary tubes. American Journal of Physiology, 96, 562–568.

    CAS  Google Scholar 

  24. Fishman, M. C., & Chien, K. R. (1997). Fashioning the vertebrate heart: Earliest embryonic decisions. Development, 124, 2099–2117.

    PubMed  CAS  Google Scholar 

  25. Forouhar, A. S., Liebling, M., Hickerson, A., Moghaddam, A. N., Tsai, H. J., Hove, J. R., et al. (2006). The embryonic vertebrate heart tube is a dynamic suction pump. Science, 312, 751–753.

    Article  PubMed  CAS  Google Scholar 

  26. Fung, Y. C. (1996). Biomechanics: Circulation (2nd ed.). New York: Springer-Verlag.

    Google Scholar 

  27. Gerould, J. H. (1929). History of the discovery of periodic reversal of heartbeat in insects. Biological Bulletin, 56(3), 215–225.

    Article  Google Scholar 

  28. Gilbert, S. F. (2000). Developmental Biology. Sunderland, MA: Sinauer Associates, Inc.

    Google Scholar 

  29. Glickman, N. S., & Yelon, D. (2002). Cardiac development in zebrafish: Coordination of form and function. Seminars in Cell & Developmental Biology, 13, 507–513.

    Article  Google Scholar 

  30. Griffith, B., & Peskin, C. S. (2010). An immersed boundary formulation of the bidomain equations. Unpublished, details available at http://www.math.nyu.edu/griffith/.

  31. Groenendijk, B. C. W., Hierck, B. P., Vrolijk, J., Baiker, M., Pourquie, M. J. B. M., Gittenberger-de-Groot, A. C., et al. (2005). Changes in shear stress-related gene expression after experimentally altered venous return in the chicken embryo. Circulation Research, 96, 1291–1298.

    Article  PubMed  CAS  Google Scholar 

  32. Groenendijk, B. C. W., Van der Heiden, K., Hierck, B. P., & Poelmann, R. E. (2007). The role of shear stress on ET-1, KLF2, and NOS-3 expression in the developing cardiovascular system of chicken embryos in a venous ligation model. Physiology, 22, 380–389.

    Article  PubMed  CAS  Google Scholar 

  33. Gruber, J., & Epstein, J. A. (2004). Development gone awry–congenital heart disease. Circulation Research, 94, 273–283.

    Article  PubMed  CAS  Google Scholar 

  34. Hamburger, V., & Hamilton, H. (1951). A series of normal stages in the development of the chick embryo. Journal of Morphology, 88, 49–92.

    Article  Google Scholar 

  35. Henriquez, C. S. (1993). Simulating the electrical behavior of cardiac tissue using the bidomain model. Critical Reviews in Biomedical Engineering, 21, 1–77.

    PubMed  CAS  Google Scholar 

  36. Herrmann, C., Wray, J., Travers, F., & Bartman, T. (1992). Effect of 2,3-butanedione monoxime on myosin and myofibrillar atpases: An example of an uncompetitive inhibitor. Biochemistry, 31, 12227–12232.

    Article  PubMed  CAS  Google Scholar 

  37. Hickerson, A. I., & Gharib, M. (2006). On the resonance of a pliant tube as a mechanism for valveless pumping. Journal of Fluid Mechanics, 555(1), 141–148.

    Article  Google Scholar 

  38. Hickerson, A. I., Rinderknecht, D., & Gharib, M. (2005). Experimental study of the behavior of a valveless impedance pump. Experiments in Fluids, 38, 534–540.

    Article  Google Scholar 

  39. Hierck, B. B. P., Van der Heiden, K. K., Poelma, C., Westerweel, J., & Poelmann, R. E. (2008). Fluid shear stress and inner curvature remodeling of the embryonic heart. Choosing the right lane!. TheScientificWorldJournal, 8, 212–222.

    Article  PubMed  Google Scholar 

  40. Hogers, B., DeRuiter, M. C., Baasten, A. M., Gittenberger-de Groot, A. C., & Poelmann, R. E. (1995). Intracardiac blood flow patterns related to the yolk sac circulation of the chick embryo. Circulation Research, 76, 871–877.

    PubMed  CAS  Google Scholar 

  41. Hogers, B., DeRuiter, M. C., Gittenberger-de Groot, A. C., & Poelmann, R. E. (1997). Unilateral vitelline vein ligation alters intracardiac blood flow patterns and morphogenesis in the chick embryo. Circulation Research, 80, 473–481.

    PubMed  CAS  Google Scholar 

  42. Hove, J. R. (2004). In vivo biofluid imaging in the developing zebrafish. Birth Defects Research, 72, 277–289.

    Article  PubMed  CAS  Google Scholar 

  43. Hove, J. R. (2006). Quantifying cardiovascular flow dynamics during early development. Pediatric Research, 60, 6–13.

    Article  PubMed  Google Scholar 

  44. Hove, J. R., Köster, R. W., Forouhar, A. S., Bolton, G. A., Fraser, S. E., & Gharib, M. (2003). Intracardiac fluid forces are an essential epigenetic factor for embryonic cardiogenesis. Nature, 421, 172–177.

    Article  PubMed  CAS  Google Scholar 

  45. Hron, J., Malek, J., & Turek, S. (2000). A numerical investigation of flows of shear-thinning fluids with applications to blood rheology. International Journal for Numerical Methods in Fluids, 32, 863–879.

    Article  CAS  Google Scholar 

  46. Ichikawa, A., & Hoshino, Z. (1967). On the cell architecture of the heart of Ciona intestinalis. Zoological Magazine, 76, 148–153.

    Google Scholar 

  47. Iomini, C., Tejada, K., Mo, W., Vaananen, H., & Piperno, G. (2004). Primary cilia of human endothelial cells disassemble under laminar shear stress. Journal of Cell Biology, 164, 811–817.

    Article  PubMed  CAS  Google Scholar 

  48. Jones, E. A. V., Baron, M. H., Fraser, S. E., & Dickinson, M. E. (2004). Measuring hemodynamics during development. American Journal of Physiology. Heart and Circulatory Physiology, 287, H1561–H1569.

    Article  PubMed  CAS  Google Scholar 

  49. Jung, E., & Peskin, C. S. (2001). Two-dimensional simulations of valveless pumping using the immersed boundary method. SIAM Journal on Scientific Computing, 23, 19–45.

    Article  Google Scholar 

  50. Kim, G. B., & Lee, S. J. (2006). X-ray PIV measurements of blood flows without tracer particles. Experiments in Fluids, 41, 195–200.

    Article  Google Scholar 

  51. Lee, P., Griffith, B. E., & Peskin, C. S. (2010). The immersed boundary method for advection-electrodiffusion with implicit timestepping and local mesh refinement. Journal of Computational Physics, 229(13), 5208–5227.

    Google Scholar 

  52. Leiderman, K. M., Miller, L. A., & Fogelson, A. L. (2008). The effects of spatial inhomogeneities on flow through the endothelial surface layer. Journal of Theoretical Biology, 25, 313–325.

    Article  CAS  Google Scholar 

  53. Liao, E. C., Zapata, A., Kieran, M., Trede, N. S., Ransom, D., & Zon, L. I. (2002). Non-cell autonomous requirement for the bloodless gene in primitive hematopoiesis of zebrafish. Development, 129, 649–659.

    PubMed  CAS  Google Scholar 

  54. Liebau, G. (1954). Ber ein ventilloses pumpprinzip. Naturwissenschaften, 41(14), 327–328.

    Article  Google Scholar 

  55. Liebau, G. (1955). Die strmungsprinzipien des herzens. Zeitschrift für Kreislaufforschung, 44(17–18), 677–684.

    PubMed  CAS  Google Scholar 

  56. Liebau, G. (1956). Die bedeutung der trgheitskrfte fr die dynamik des blutkreislaufs. Zeitschrift für Kreislaufforschung, 45(13–14), 481–488.

    PubMed  CAS  Google Scholar 

  57. Liu, A., Rugonyi, S., Pentecost, J. O., & Thornburg, K. L. (2007). Finite element modeling of blood flow-induced mechanical forces in the outflow tract of chick embryonic hearts. Computers & Structures, 85(11), 11–14.

    Article  Google Scholar 

  58. Luft, J. H. (1966). Fine structures of capillary and endocapillary layer as revealed by ruthenium red. Federation Proceedings, 25, 1773–1783.

    PubMed  CAS  Google Scholar 

  59. Männer, J. (2000). Cardiac looping in the chick embryo: A morphological review with special reference to terminological and biomechanical aspects of the looping process. Anatomical Record, 259, 248–262.

    Article  PubMed  Google Scholar 

  60. Manopoulos, C. G., Mathioulakis, D. S., & Tsangaris, S. G. (2006). One-dimensional model of valveless pumping in a closed loop and a numerical solution. Physics of Fluids, 18(1), 017106–017116.

    Article  CAS  Google Scholar 

  61. Marshall, W. F., & Nonaka, S. (2006). Cilia: Tuning in to the cell’s antenna. Current Biology, 16, R604–R614.

    Article  PubMed  CAS  Google Scholar 

  62. McCann, F. V., & Sanger, J. W. (1969). Ultrastructure and function in an insect heart. Experientia. Supplementum, 15, 29–46.

    Google Scholar 

  63. McGrath, K. E., Koniski, A. D., Malik, J., & Palis, J. (2003). Circulation is established in a stepwise pattern in the mammalian embryo. Blood, 101, 1669–1676.

    Article  PubMed  CAS  Google Scholar 

  64. Merrill, E. W., Benis, A. M., Gilliland, E. R., Sherwood, T. K., & Salzman, E. W. (1965). Pressure flow relations of human blood in hollow fibers at low flow rates. Journal of Applied Physiology, 20, 954–967.

    PubMed  CAS  Google Scholar 

  65. Miller, L. A. (2011). Fluid dynamics of ventricular filling in the embryonic heart. Cell Biochemistry and Biophysics. doi:10.1007/s12013-011-9157-9.

  66. Mittal, R. (2005). Immersed boundary methods. Annual Review of Fluid Mechanics, 37, 239–261.

    Article  Google Scholar 

  67. Moorman, A. F. M., & Christoffels, V. M. (2003). Cardiac chamber formation: Development, genes, and evolution. Physiological Reviews, 83, 1223–1267.

    PubMed  CAS  Google Scholar 

  68. Moorman, A. F. M., Soufan, A. T., Hagoort, J., De Boer, P. A. J., & Christoffels, V. M. (2004). Development of the building plan of the heart. Annals of the New York Academy of Sciences, 1015, 171–181.

    Article  PubMed  Google Scholar 

  69. Moorman, A. F. M., Webb, S., Brown, N. A., Lamers, W., & Anderson, R. H. (2003). Development of the heart: (1) Formation of the cardiac chambers and arterial trunks. Heart, 89, 806–814.

    Article  PubMed  Google Scholar 

  70. Nauli, S. M., Alenghat, F. J., Luo, Y., Williams, E., Vassilev, P., Lil, X. G., et al. (2003). Polycystins 1 and 2 mediate mechanosensation in the primary cilium of kidney cells. Nature Genetics, 33, 129–137.

    Article  PubMed  CAS  Google Scholar 

  71. Nauli, S. M., Kawanabe, Y., Kaminski, J. J., Pearce, W. J., Ingber, D. E., & Zhou, J. (2008). Endothelial cilia are fluid-shear sensors that regulate calcium signaling and nitric oxide production through polycystin-1. Circulation, 117, 1161–1171.

    Article  PubMed  CAS  Google Scholar 

  72. Olesen, S. P., Clapham, D. E., & Davies, P. F. (1988). Hemodynamic shear-stress activates a K+ current in vascular endothelial cells. Nature, 331, 168–170.

    Article  PubMed  CAS  Google Scholar 

  73. Ottesen, J. T. (2003). Valveless pumping in a fluid-filled closed elastic tube system: One-dimensional theory with experimental validation. Journal of Mathematical Biology, 46(4), 309–332.

    Article  PubMed  CAS  Google Scholar 

  74. Patterson, C. (2005). Even flow: Shear cues vascular development. Arteriosclerosis, Thrombosis, and Vascular Biology, 25, 1761–1762.

    Article  PubMed  CAS  Google Scholar 

  75. Peskin, C. S. (2002). The immersed boundary method. Acta Numerica, 11, 479–517.

    Article  Google Scholar 

  76. Peskin, C. S., & McQueen, D. M. (1996). Fluid dynamics of the heart and its valves. In H. G. Othmer, F. R. Adler, M. A. Lewis, & J. C. Dallon (Eds.), Case studies in mathematical modeling: Ecology, physiology, and cell biology (2nd ed.). Upper Saddle River, NJ: Prentice-Hall.

    Google Scholar 

  77. Phoon, C. K. L., Aristizábal, O., & Turnbull, D. H. (2002). Spatial velocity profile in mouse embryonic aorta and Doppler-derived volumetric flow: A preliminary model. American Journal of Physiology. Heart and Circulatory Physiology, 283(3), H908–H916.

    PubMed  CAS  Google Scholar 

  78. Picart, C., Piau, J. M., & Galliard, H. (1998). Human blood shear yield stress and its hematocrit dependence. Journal of Rheology, 42, 1–12.

    Article  CAS  Google Scholar 

  79. Poelma, C., Van der Heiden, K., Hierck, B. P., Poelmann, R. E., & Westerweel, J. (2010). Measurements of the wall shear stress distribution in the outflow tract of an embryonic chicken heart. Journal of the Royal Society Interface, 7(42), 91–103.

    Article  CAS  Google Scholar 

  80. Praetorius, H. A., & Spring, K. R. (2001). Bending the MDCK cell primary cilium increases intracellular calcium. Journal of Membrane Biology, 184, 71–79.

    Article  PubMed  CAS  Google Scholar 

  81. Reckova, M., Rosengarten, C., de Almeida, A., Stanley, C. P., Wessels, A., Gourdie, R. G., et al. (2003). Hemodynamics is a key epigenetic factor in development of the cardiac conduction system. Circulation Research, 93, 77–85.

    Article  PubMed  CAS  Google Scholar 

  82. Reitsma, S., Slaaf, D. W., Vink, H., van Zandvoort, M. A. M. J., & oude Egbrink, M. G. A. (2007). The endothelial glycocalyx: Composition, functions, and visualization. European Journal of Physiology, 454, 345–359.

    Article  PubMed  CAS  Google Scholar 

  83. Rychter, Z., Kopecky, M., & Lemez, L. (1955). A micromethod for determination of the circulating blood volume in chick embryos. Nature, 175, 1126–1127.

    Article  PubMed  CAS  Google Scholar 

  84. Sadler, T. W. (1995). Langman’s medical embryology (7th ed.). Baltimore: Williams & Wilkins.

    Google Scholar 

  85. Santhanakrishnan, A., Nguyen, N., Cox, J. G., & Miller, L. A. (2009). Flow within models of the vertebrate embryonic heart. Journal of Theoretical Biology, 259, 449–464.

    Article  PubMed  Google Scholar 

  86. Santiago, J. G., Wereley, S. T., Meinhart, C. D., Beebe, D. J., & Adrian, R. J. (1998). A particle image velocimetry system for microfluidics. Experiments in Fluids, 25, 316–319.

    Article  CAS  Google Scholar 

  87. Savolainen, S. M., Foley, J. F., & Elmore, S. A. (2009). Histology atlas of the developing mouse heart with emphasis on E11.5 to E18.5. Toxicologic Pathology, 37, 395–414.

    Article  PubMed  Google Scholar 

  88. Schlichting, H., & Gersten, K. (2000). Boundary layer theory (8th ed.). New York: Springer-Verlag.

    Google Scholar 

  89. Secomb, T., Hsu, R., & Pries, A. (2001). Effect of the endothelial surface layer on transmission of fluid shear stress to endothelial cells. J. Biorheol., 38, 143–150.

    CAS  Google Scholar 

  90. Sehnert, A. J. (2002). Cardiac troponin t is essential in sarcomere assembly and cardiac contractility. Nature Genetics, 31, 106–110.

    Article  PubMed  CAS  Google Scholar 

  91. Selamet Tierney, E. S., Wald, R. M., McElhinney, D. B., Marshall, A. C., Benson, C. B., Colan, S. D., et al. (2007). Changes in left heart hemodynamics after technically successful in-utero aortic valvuloplasty. Ultrasound in Obstetrics and Gynecology, 30(5), 715–720.

    Article  PubMed  CAS  Google Scholar 

  92. Shelley, M., Fauci, L., & Teran, J. (2008). Large amplitude peristaltic pumping of an elastic fluid. Physics of Fluids, 20, 073101.

    Article  CAS  Google Scholar 

  93. Singla, V., & Reiter, J. F. (2006). The primary cilium as the cell’s antenna: Signaling at a sensory organelle. Science, 313, 629–633.

    Article  PubMed  CAS  Google Scholar 

  94. Skalak, R., & Özkaya, N. (1989). Biofluid mechanics. The Annual Review of Fluid Mechanics, 21, 167–204.

    Article  Google Scholar 

  95. Smith, M., Long, D., Damiano, E., & Ley, K. (2003). Near-wall micro-PIV reveals a hydrodynamically relevant endothelial surface layer in venules in vivo. Biophysical Journal, 85, 637–645.

    Article  PubMed  CAS  Google Scholar 

  96. Taber, L. A., Zhang, J., & Perucchio, R. (2007). Computational model for the transition From peristaltic to pulsatile flow in the embryonic heart tube. Journal of Biomechanical Engineering, 129(3), 441–450.

    Article  PubMed  Google Scholar 

  97. Thi, M., Tarbell, J., Weinbaum, S., & Spray, D. (2004). The role of the glycocalyx in reorganization of the actin cytoskeleton under fluid shear stress: A bumper car model. Proceedings of the National Academy of Sciences, 101(47), 16483–16488.

    Article  CAS  Google Scholar 

  98. Thisse, C., & Zon, L. I. (2002). Organogenesis-heart and blood formation from the zebrafish point of view. Science, 295(5554), 457–462.

    Article  PubMed  CAS  Google Scholar 

  99. Thomann, H. (1978). A simple pumping mechanism in a valveless tube. Zeitschrift fr Angewandte Mathematik und Physik (ZAMP), 29(2), 169–177.

    Article  Google Scholar 

  100. Topper, N., & Gimbrone, M. A., Jr. (1999). Blood flow and vascular gene expression: Fluid shear stress as a modulator of endothelial phenotype. Molecuar Medicine Today, 5, 40–46.

    Article  CAS  Google Scholar 

  101. Tretheway, D. C., & Meinhart, C. D. (2004). A generating mechanism for apparent fluid slip in hydrophobic microchannels. Physics of Fluids, 16, 1509–1515.

    Article  CAS  Google Scholar 

  102. Tucker, D. C., Snider, C., & Woods, W. T. (1988). Pacemaker development in embryonic rat heart cultured in oculo. Pediatric Research, 23, 637–642.

    Article  PubMed  CAS  Google Scholar 

  103. Ursem, N. T. C., Stekelenburg-de Vos, S., Wladimiroff, J. W., Poelmann, R. E., Gittenberger-de Groot, A. C., Hu, N., et al. (2004). Ventricular diastolic filling characteristics in stage-24 chick embryos after extra-embryonic venous obstruction. Journal of Experimental Biology, 207, 1487–1490.

    Article  PubMed  Google Scholar 

  104. Van der Heiden, K., Groenendijk, B. C. W., Hierck, B. P., Hogers, B., Koerten, H. K., Mommaas, A. M., et al. (2006). Monocilia on chicken embryonic endocardium in low shear stress areas. Developmental Dynamics, 235, 19–28.

    Article  PubMed  CAS  Google Scholar 

  105. Van der Heiden, K., Hierck, B. P., Krams, R., de Crom, R., Cheng, C., Baiker, M., et al. (2007). Endothelial primary cilia in areas of disturbed flow are at the base of atherosclerosis. Atherosclerosis, 196(2), 542–550.

    Article  PubMed  CAS  Google Scholar 

  106. Vennemann, P., Kiger, K. T., Lindken, R., Groenendijk, B. C. W., Stekelenburg-de Vos, S., ten Hagen, T. L. M., et al. (2006). In vivo micro particle image velocimetry measurements of blood-plasma in the embryonic avian heart. Journal of Biomechanics, 39(7), 1191–1200.

    Article  PubMed  Google Scholar 

  107. Vincent, P. E., Sherwin, S. J., & Weinberg, P. D. (2008). Viscous flow over outflow slits covered by an anisotropic Brinkman medium: A model of flow above interendothelial cell clefts. Physics of Fluids, 20(6), 063106–063111.

    Article  CAS  Google Scholar 

  108. Wang, Y., Dur, O., Patrick, M. J., Tinney, J. P., Tobita, K., Keller, B. B., et al. (2009). Aortic arch morphogenesis and flow modeling in the chick embryo. Annals of Biomedical Engineering, 37(6), 1069–1081.

    Article  PubMed  Google Scholar 

  109. Weinbaum, S., Zhang, X., Han, Y., Vink, H., & Cowin, S. C. (2003). Mechanotransduction and flow across the endothelial glycocalyx. Proceedings of the National Academy of Sciences of the United States of America, 100, 7988–7995.

    Article  PubMed  CAS  Google Scholar 

  110. Wenning, A., Cymbalyuk, G. S., & Calabrese, R. L. (2004). Heartbeat control in leeches. I. Constriction pattern and neural modulation of blood pressure in intact animals. Journal of Neurophysiology, 91, 382–396.

    Google Scholar 

  111. Wilkins-Haug, L. E., Benson, C. B., Tworetzky, W., Marshall, A. C., Jennings, R. W., & Lock, J. E. (2005). In-utero intervention for hypoplastic left heart syndrome—a perinatologist’s perspective. Ultrasound in Obstetrics and Gynecology, 26(5), 481–486.

    Article  PubMed  CAS  Google Scholar 

  112. Womersley, J. R. (1955). Method for the calculation of velocity, rate of flow and viscous drag in arteries when the pressure gradient is known. Journal of Physiology, 127, 553–563.

    PubMed  CAS  Google Scholar 

  113. Yao, Y., Rabodzey, A., & Dewey, C. F. (2007). Glycocalyx modulates the motility and proliferative response of vascular endothelium to fluid shear stress. American Journal of Physiology. Heart and Circulatory Physiology, 293, H1023–H1030.

    Article  PubMed  CAS  Google Scholar 

  114. Yoganathan, A. P., Rittgers, S. E., & Chandran, K. B. (2007). Biofluid mechanics: The human circulation (1st ed.). Boca Raton: Taylor and Francis.

    Google Scholar 

  115. Yost, H. J. (2003). Left-right asymmetry: Nodal cilia make and catch a wave. Current Biology, 13, R808–R809.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgments

We would like to thank the University of Utah Mathematical Biology Group and the UNC Fluids and Integrative & Mathematical Physiology Groups for their suggestions and insight. We would also like to thank Dr. Kathy K. Sulik for her excellent SEM images of the mouse embryonic heart used in this review. This work was funded by Miller’s Burroughs Wellcome Fund Career Award at the Scientific Interface.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Arvind Santhanakrishnan or Laura A. Miller.

Appendix

Appendix

Navier–Stokes Equations

In this section, the governing equations of incompressible, constant viscosity fluid flow is presented. Consider a fluid of constant density ρ (mass per unit volume) and dynamic viscosity μ (indicative of the effects of frictional forces/mixing in a fluid), flowing through a system with velocity components u, v, and w in x, y, and z coordinates, respectively. Let p represent the local fluid pressure. Equation 5 represents the conservation of mass, which requires that the mass flux (mass per unit time) of fluid entering a system must be equal to the mass flux of fluid leaving the system,

$$ {\frac{\partial u}{\partial x}} + {\frac{\partial v}{\partial y}} + {\frac{\partial w}{\partial z}} = 0 $$
(5)

Equations 68 represent the conservation of linear momentum in each coordinate direction, and this principle requires that the inertial force must be equal and opposite to the sum of the pressure force, viscous force, and body force (due to gravity in most cases). The inertial forces on the left hand side of the Eqs. 68 arise due to the fluid flow and include both unsteady (or transient) and convective (flow velocity and velocity gradient dependent) contributions.

$$ {\frac{\partial u}{\partial t}} + u{\frac{\partial u}{\partial x}} + v{\frac{\partial u}{\partial y}} + w{\frac{\partial u}{\partial z}} = - {\frac{ 1}{\rho }}\,{\frac{\partial p}{\partial x}} + \nu \left( {{\frac{{\partial^{ 2} u}}{{\partial x^{ 2} }}} + {\frac{{\partial^{ 2} u}}{{\partial y^{ 2} }}} + {\frac{{\partial^{ 2} u}}{{\partial z^{ 2} }}}} \right) + f_{{{\text{B,}}x}} $$
(6)
$$ {\frac{\partial v}{\partial t}} + u{\frac{\partial v}{\partial x}} + v{\frac{\partial v}{\partial y}} + w{\frac{\partial v}{\partial z}} = - {\frac{ 1}{\rho }}\,{\frac{\partial p}{\partial y}} + \nu \left( {{\frac{{\partial^{ 2} v}}{{\partial x^{ 2} }}} + {\frac{{\partial^{ 2} v}}{{\partial y^{ 2} }}} + {\frac{{\partial^{ 2} v}}{{\partial z^{ 2} }}}} \right) + f_{{{\text{B,}}y}} $$
(7)
$$ {\frac{\partial w}{\partial t}} + u{\frac{\partial w}{\partial x}} + v{\frac{\partial w}{\partial y}} + w{\frac{\partial w}{\partial z}} = - {\frac{ 1}{\rho }}\,{\frac{\partial p}{\partial z}} + \nu \left( {{\frac{{\partial^{ 2} w}}{{\partial x^{ 2} }}} + {\frac{{\partial^{ 2} w}}{{\partial y^{ 2} }}} + {\frac{{\partial^{ 2} w}}{{\partial z^{ 2} }}}} \right){ + }f_{{{\text{B,}}z}} $$
(8)

Note that ν is the kinematic viscosity of the fluid, which is the ratio of the coefficient of dynamic viscosity to the density of the fluid (ν = μ/ρ), while f B indicates the body force acting on the fluid flow. The collective set of governing conservation Eqs. 14 are also known as the Navier–Stokes equations. The nonlinear mathematical nature of these partial differential equations renders it difficult to solve, and only a few exact analytical solutions for specific problems are known. Further details on these equations may be obtained in the book by Schlichting and Gersten [88], for example.

Fluid Dynamic Scaling

To obtain a physical perspective, it is a useful exercise to non-dimensionalize the terms in the above governing conservation equations using equivalent scaling characteristics as shown below,

$$ x^{\prime } = \frac{x}{L} , \;y^{\prime } = \frac{y}{L} , \;z^{\prime } = \frac{z}{L} $$
(9)
$$ u^{\prime } = \frac{u}{U} , \;v^{\prime } = \frac{v}{U} , \;w^{\prime } = \frac{w}{U} $$
(10)
$$ t^{\prime } = \omega \, t ,\; p^{\prime } = {\frac{p}{{\rho U^{2} }}} $$
(11)

where L, U, and 1/ ω are characteristic flow length, velocity, and time scales respectively, As an example, in the case of blood flow through the arteries of the adult human circulatory system, the diameter of the vessel, the velocity along the centerline of the artery (after some distance devoid of any entrance effects), and the pumping rate of the heart are typically chosen to be the characteristic length, velocity, and time scales respectively. The application of these terms to Eqs. 14 results in the following set of equations after neglecting the body force contribution (as its importance in cardiovascular flows is insignificant),

$$ {\frac{{\partial u^{\prime } }}{{\partial x^{\prime } }}} + {\frac{{\partial v^{\prime } }}{{\partial y^{\prime } }}} + {\frac{{\partial w^{\prime } }}{{\partial z^{\prime } }}} = 0 $$
(12)
$$ \left( {{\frac{\omega L}{U}}} \right){\frac{{\partial u^{\prime } }}{{\partial t^{\prime } }}} + u^{\prime } {\frac{{\partial u^{\prime } }}{{\partial x^{\prime } }}} + v^{\prime } {\frac{{\partial u^{\prime } }}{{\partial y^{\prime } }}} + w^{\prime } {\frac{{\partial u^{\prime } }}{{\partial z^{\prime } }}} = - {\frac{{\partial p^{\prime } }}{{\partial x^{\prime } }}} + \left( {{\frac{UL}{\nu }}} \right)\left( {{\frac{{\partial^{ 2} u^{\prime } }}{{\partial x^{\prime 2} }}} + {\frac{{\partial^{ 2} u^{\prime } }}{{\partial y^{\prime 2} }}} + {\frac{{\partial^{ 2} u^{\prime } }}{{\partial z^{\prime 2} }}}} \right) $$
(13)
$$ \left( {{\frac{\omega L}{U}}} \right){\frac{{\partial v^{\prime } }}{{\partial t^{\prime } }}} + u^{\prime } {\frac{{\partial v^{\prime } }}{{\partial x^{\prime } }}} + v^{\prime } {\frac{{\partial v^{\prime } }}{{\partial y^{\prime } }}} + w^{\prime } {\frac{{\partial v^{\prime } }}{{\partial z^{\prime } }}} = - {\frac{{\partial p^{\prime } }}{{\partial y^{\prime } }}} + \left( {{\frac{UL}{\nu }}} \right)\left( {{\frac{{\partial^{ 2} v^{\prime } }}{{\partial x^{\prime 2} }}} + {\frac{{\partial^{ 2} v^{\prime } }}{{\partial y^{\prime 2} }}} + {\frac{{\partial^{ 2} v^{\prime } }}{{\partial z^{\prime 2} }}}} \right) $$
(14)
$$ \left( {{\frac{\omega L}{U}}} \right){\frac{{\partial w^{\prime } }}{{\partial t^{\prime } }}} + u^{\prime } {\frac{{\partial w^{\prime } }}{{\partial x^{\prime } }}} + v^{\prime } {\frac{{\partial w^{\prime } }}{{\partial y^{\prime } }}} + w^{\prime } {\frac{{\partial w^{\prime } }}{{\partial z^{\prime } }}} = - {\frac{{\partial p^{\prime } }}{{\partial z^{\prime } }}} + \left( {{\frac{UL}{\nu }}} \right)\left( {{\frac{{\partial^{ 2} w^{\prime } }}{{\partial x^{\prime 2} }}} + {\frac{{\partial^{ 2} w^{\prime } }}{{\partial y^{\prime 2} }}} + {\frac{{\partial^{ 2} w^{\prime } }}{{\partial z^{\prime 2} }}}} \right) $$
(15)

In the context of vertebrate embryonic heart development, it is useful to examine the limit of low Reynolds and Womersley numbers in the vector form of the Navier–Stokes equations 58 as given below:

$$ {\frac{{\partial \vec{u}}}{\partial t}} + \nabla \cdot \left( {\rho \vec{u}\vec{u}} \right) = - {\frac{ 1}{\rho }}\nabla p + \nu \nabla^{ 2} \vec{u} - g $$
(16)

As the Re is sufficiently small, the inertial terms in the momentum equations as well as the gravitational force can be ignored, and for small values of Wo the transient term can be neglected, resulting in the Stokes equations below:

$$ \nabla p = \mu \nabla^{ 2} \vec{u},\;\nabla \cdot \vec{u} = 0 $$
(17)

In this limit of very low Re and Wo, the flow is entirely driven through a balance between the pressure gradient and viscous diffusion.

Shear Stress on Blood Vessels

To analyze the flow within a blood vessel, it is useful to start with a simplified model of internal flow through a pipe. Consider the steady, incompressible, two-dimensional (radial r and axial x, see definitions in Fig. 16), incompressible, internal flow of a fluid of density ρ and uniform dynamic viscosity μ through a cylinder of radius R. The flow velocity is assumed to have no swirling component (u θ = 0), and the flow is considered to be axisymmetric about the central axis of the pipe, such that \( \partial /\partial \theta = 0 \).

$$ {\frac{{\partial u_{r} }}{\partial x}} + \frac{1}{r}{\frac{\partial }{\partial r}}\left( {ru_{r} } \right) = 0 $$
(18)
$$ u_{x} {\frac{{\partial u_{x} }}{\partial x}} \, + \, u_{r} {\frac{{\partial u_{x} }}{\partial r}} = - {\frac{ 1}{\rho }}\,{\frac{\partial p}{\partial x}} \, + \, \nu \left( {{\frac{{\partial^{ 2} u_{x} }}{{\partial x^{2} }}} \, + \, {\frac{{\partial^{ 2} u_{x} }}{{\partial r^{2} }}} + \frac{1}{r}{\frac{{\partial u_{x} }}{\partial r}}} \right) $$
(19)
$$ u_{x} {\frac{{\partial u_{r} }}{\partial x}} + u_{r} {\frac{{\partial u_{r} }}{\partial r}} = - {\frac{ 1}{\rho }}\,{\frac{\partial p}{\partial r}} + \nu \left( {{\frac{{\partial^{ 2} u_{r} }}{{\partial x^{2} }}} + {\frac{{\partial^{ 2} u_{r} }}{{\partial r^{2} }}} + \frac{1}{r}{\frac{{\partial u_{r} }}{\partial r}} - {\frac{{u_{r} }}{{r^{2} }}}} \right) $$
(20)
Fig. 16
figure 16

Internal flow of a fluid through a circular cylinder of radius R (also known as Hagen–Poiseuille flow). Under the simplifications of two-dimensional, time-invariant, axisymmetric (no fluid rotation, u θ = 0), incompressible flow of a fluid with uniform density and viscosity, the axial velocity u x remains unchanged along the longitudinal direction x after a certain length (typically 2–4 multiples of the radius) from the entrance, and this condition is also known as a fully developed flow. However, the radial variation of the axial velocity u x (r) is parabolic about the centerline axis as shown. The shear stress imposed by the flow τ in the radial direction varies linearly from the centerline, and the maximum value τ w occurs at the walls. Note that δ indicates the thickness of the boundary layer, which is the region near the solid boundaries where the deceleration of the flow on account of fluid viscosity is non-negligible. The coordinate system used for the analysis of this problem (r radial, x axial, θ rotational) is also shown

This problem can be solved analytically by considering a few simplifying assumptions. Typically, in the case of microcirculation, the flow reaches a fully developed state at a short distance from its entrance (small multiple of the vessel diameter) such that there is no variation in the flow velocity along the primary axial direction thereafter (\( \partial /\partial x = 0 \)). This reduces the above equation set 1820 to the following:

$$ \frac{1}{r}{\frac{\partial }{\partial r}}\left( {ru_{r} } \right) = 0 $$
(21)
$$ u_{r} {\frac{{\partial u_{x} }}{\partial r}} = - {\frac{ 1}{\rho }}\,{\frac{\partial p}{\partial x}} + \nu \left( {{\frac{{\partial^{ 2} u_{x} }}{{\partial r^{2} }}} + \frac{1}{r}{\frac{{\partial u_{x} }}{\partial r}}} \right) $$
(22)
$$ u_{r} {\frac{{\partial u_{r} }}{\partial r}} = - {\frac{ 1}{\rho }}\,{\frac{\partial p}{\partial r}} + \nu \left( {{\frac{{\partial^{ 2} u_{r} }}{{\partial r^{2} }}} + \frac{1}{r}{\frac{{\partial u_{r} }}{\partial r}} - {\frac{{u_{r} }}{{r^{2} }}}} \right) $$
(23)

The above equations of mass and momentum are subject to the following conditions at specific boundaries in the problem domain:

$$ u_{r} \left( {r = 0} \right) = 0 $$
(24)
$$ u_{r} \left( {r = R} \right) = 0 $$
(25)
$$ u_{x} \left( {r = R} \right) = 0 $$
(26)
$$ \left. {{\frac{{\partial u_{x} }}{\partial r}}} \right|_{{r = 0}} = 0 $$
(27)

The boundary conditions 24 and 27 are obtained by considering symmetry about the centerline. The boundary condition 25 ensures that there is no normal flow through the vessel wall, and condition 26 means that the layer of fluid that is in contact with the vessel wall remains at rest (“no slip” of fluid on the solid surface). From the continuity equation 21, \( ru_{r} \) must be constant. Applying the boundary conditions 24 and 25, it can be seen that there is no radial flow throughout the vessel, i.e., u r  = 0 everywhere. This simplifies the r-momentum equation 23 to the form given below:

$$ {\frac{1}{\rho }}\,{\frac{\partial p}{\partial r}} = 0 $$
(28)

As the flow is incompressible, this means that the dynamic pressure is invariant in the radial direction, and is only a function of the axial location, i.e., p = p(x). The axial momentum conservation equation 22 now becomes

$$ {\frac{1}{\rho }}\,{\frac{{{\text{d}}p}}{{{\text{d}}x}}} = \nu \left( {{\frac{{\partial^{ 2} u_{x} }}{{\partial r^{2} }}} + \frac{1}{r}{\frac{{\partial u_{x} }}{\partial r}}} \right) $$
(29)

which can be written as,

$$ {\frac{r}{\mu }}\,{\frac{{{\text{d}}p}}{{{\text{d}}x}}} = {\frac{\partial }{\partial r}}\left( {r{\frac{{\partial u_{x} }}{\partial r}}} \right) $$
(30)

Integrating both sides of the above equation in terms of r, we obtain

$$ {\frac{{r^{2} }}{2\mu }}\,{\frac{{{\text{d}}p}}{{{\text{d}}x}}} = r{\frac{{\partial u_{x} }}{\partial r}} + A $$
(31)

where A is the constant of integration, the value of which is determined by applying 27 to the above equation. The resultant equation can be integrated once again in terms of r to solve for the axial velocity profile u x (r),

$$ {\frac{{r^{2} }}{4\mu }}\,{\frac{{{\text{d}}p}}{{{\text{d}}x}}} = u_{x} + B $$
(32)

The constant of integration is determined by using boundary condition 26. The solution for the axial velocity is thus given by

$$ u_{x} \left( r \right) = - \left( {{\frac{{R^{2} }}{4\mu }}\,{\frac{{{\text{d}}p}}{{{\text{d}}x}}}} \right) \, \left( {1 - {\frac{{r^{2} }}{{R^{2} }}}} \right) = u_{\max } \left( {1 - {\frac{{r^{2} }}{{R^{2} }}}} \right) $$
(33)

The internal flow through a cylindrical vessel under the previously stated assumptions has a parabolic velocity profile with the peak located along the centerline, the magnitude of which depends on the pressure gradient at the particular axial location of interest, the dynamic viscosity of the fluid, and the vessel radius. The pressure gradient is referred to be adverse when dp/dx > 0 resulting in a decelerating flow, and is favorable when dp/dx < 0 and the flow accelerates.

The viscous fluid flow exerts a tangential shear stress, which can be determined as the gradient of the axial velocity as given below:

$$ \tau_{xr} = \tau_{rx} = \mu {\frac{{\partial u_{x} }}{\partial r}} = - {\frac{{2\mu u_{\max } r}}{{R^{2} }}} = {\frac{r}{2\mu }}\,{\frac{{{\text{d}}p}}{{{\text{d}}x}}} $$
(34)

Of special importance in developmental physiology is the shear stress imposed by blood flow on the walls of blood vessels, which is given by,

$$ \tau_{w} = \left. {\tau_{xr} } \right|_{{r{ = }R}} = - {\frac{{2\mu u_{\max } }}{R}} = {\frac{R}{2\mu }}\,{\frac{{{\text{d}}p}}{{{\text{d}}x}}} $$
(35)

The mean flow velocity through the vessel can be calculated by integrating the axial velocity profile over the cross section,

$$ \bar{u}_{x} = {\frac{1}{{\pi R^{2} }}}\int\limits_{0}^{2\pi } {\int\limits_{0}^{R} {u_{x} \left( r \right)r{\text{d}}r{\text{d}}\theta } } = {\frac{{u_{\max } }}{2}} $$
(36)

The shear stress can be redefined in terms of the volumetric flow rate Q based on mean flow velocity as

$$ \tau_{w} = - {\frac{{4\mu \bar{u}_{x} }}{R}} = - {\frac{4\mu Q}{{\pi R^{3} }}} = - {\frac{32\mu Q}{{\pi D^{3} }}} $$
(37)

Blood Rheology

One way to model the non-Newtonian properties of the blood is to consider it as a generalized Newtonian fluid, where the shear stress is a function of the shear rate at the particular time, and the fluid dynamics do not depend upon the history of deformation. This approach has been used previously to model the blood as a Cross fluid [9] and as a power law fluid [45, 114]. In both cases, the constitutive equations are the same as the traditional incompressible Navier–Stokes equations with the exception that the viscosity is no longer constant and depends upon the shear stress and/or the shear rate. For a power law fluid, the shear stress and effective viscosities are given by the equations:

$$ \tau = K\left( {{\frac{\partial u}{\partial y}}} \right)^{n} $$
(38)
$$ \mu_{\text{eff}} = K\left( {{\frac{\partial u}{\partial y}}} \right)^{n - 1} $$
(39)

where K is the flow consistency index, \( \partial u/\partial y \) is the velocity gradient perpendicular to the plane of shear, n is the flow behavior index, and μ eff is the effective viscosity. Note that for the Newtonian case n = 1. The disadvantage of this model is that it is only appropriate for shear rates over the range for which it was fitted. Notice that the effective viscosity goes to infinity as the shear rate approaches zero for shear thinning fluids (n < 1). A more reasonable generalized Newtonian model might be the Cross model. In this case, the effective viscosity is a function of the shear rate and is given by the equation:

$$ \mu_{eff} \left( {\dot{\gamma }} \right) = {\frac{{\mu_{0} }}{{1 + \left( {{\frac{{\mu_{ 0} \dot{\gamma }}}{{\tau^{*} }}}} \right)^{1 - n} }}} $$
(40)

where μ 0, τ*, and n are experimentally determined coefficients. The shear rate, \( \dot{\gamma } \), is set to the gradient of the velocity of the fluid. In this model, the fluid behaves as a Newtonian fluid at low shear rates (μ 0 \( \dot{\gamma } \) ≪ τ*) and as a power law fluid at high shear rates (μ 0 \( \dot{\gamma } \) ≫ τ*).

For modeling purposes, μ eff is usually fit in the biologically relevant range of the non-Newtonian characteristics of the blood. These models are typically used for intermediate sized blood vessels (diameter > 22 μm) where blood is treated as a homogenous fluid. If the cell diameter is comparable to the vessel diameter (or on the same order of magnitude), this continuum approximation is not appropriate. A Newtonian fluid approximation for blood viscosity is acceptable typically for larger vessels where the diameter of the vessel (typically >0.5 mm) is well above the diameter of the red blood cells (roughly 8 μm). Skalak and Özkaya [94] present a detailed review of blood rheology, and may be referred to for further information.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Santhanakrishnan, A., Miller, L.A. Fluid Dynamics of Heart Development. Cell Biochem Biophys 61, 1–22 (2011). https://doi.org/10.1007/s12013-011-9158-8

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12013-011-9158-8

Keywords

Navigation