Show simple item record

The steady propagation of a surfactant-laden liquid plug in a two-dimensional channel

dc.contributor.authorFujioka, Hidekien_US
dc.contributor.authorGrotberg, James B.en_US
dc.date.accessioned2011-11-15T15:58:11Z
dc.date.available2011-11-15T15:58:11Z
dc.date.issued2005-08en_US
dc.identifier.citationFujioka, Hideki; Grotberg, James B. (2005). "The steady propagation of a surfactant-laden liquid plug in a two-dimensional channel." Physics of Fluids 17(8): 082102-082102-17. <http://hdl.handle.net/2027.42/87308>en_US
dc.identifier.urihttps://hdl.handle.net/2027.42/87308
dc.description.abstractIn this study, we investigate the steady propagation of a liquid plug in a two-dimensional channel lined by a uniform, thin liquid film. The liquid contains soluble surfactant that can exist both in the bulk fluid and on the air-liquid interface. The Navier-Stokes equations with free-surface boundary conditions and the surfactant transport equations are solved using a finite volume numerical scheme. The adsorption/desorption process of the surfactant is modeled based on pulmonary surfactant properties. As the plug propagates, the front meniscus sweeps preexisting interfacial surfactant from the precursor film, and the surfactant accumulates on the front meniscus interface. As the front meniscus converges on the precursor film from the region where the interfacial surfactant concentration is maximized, the Marangoni stress opposes the flow. In this region, the Marangoni stress results in nearly zero surface velocity, which causes the precursor film thickness near the meniscus to be thicker than the leading film thickness. Since the peaks of wall pressure and wall shear stress occur due to narrowing of the film thickness, the observed increase of the minimum film thickness weakens these stresses. In the thicker film region, however, the drag forces increase due to an increase in the surfactant concentration. This causes the overall pressure drop across the plug to increase as a result of the increasing surfactant concentration. A recirculation flow forms inside the plug core and is skewed toward the rear meniscus as the Reynolds number increases. When no surfactant exists, the recirculation flow is in contact with both the front and the rear interfaces. As the surfactant concentration increases, the Marangoni stress begins to rigidify the front interface and forces the recirculation flow away from the front interface. Subsequently, the recirculation flow is directed away from the rear interface in a manner similar to that for the front interface. When the plug length is shorter, this change in recirculation pattern occurs at a smaller surfactant concentration.en_US
dc.publisherThe American Institute of Physicsen_US
dc.rights© The American Institute of Physicsen_US
dc.titleThe steady propagation of a surfactant-laden liquid plug in a two-dimensional channelen_US
dc.typeArticleen_US
dc.subject.hlbsecondlevelPhysicsen_US
dc.subject.hlbtoplevelScienceen_US
dc.description.peerreviewedPeer Revieweden_US
dc.contributor.affiliationumDepartment of Biomedical Engineering, The University of Michigan, Ann Arbor, Michigan 48109en_US
dc.description.bitstreamurlhttp://deepblue.lib.umich.edu/bitstream/2027.42/87308/2/082102_1.pdf
dc.identifier.doi10.1063/1.1948907en_US
dc.identifier.sourcePhysics of Fluidsen_US
dc.identifier.citedreferenceM. E. Avery and J. Mead, “Surface properties in relation to atelectasis and hyaline membrane disease,” Am. J. Dis. Child 97, 517 (1959).en_US
dc.identifier.citedreferenceJ. M. B. Hughes, D. Y. Rosenzweig, and P. B. Kivitz, “Site of airway closure in excised dog lungs: Histologic demonstration,” J. Appl. Physiol. 29, 340 (1970).en_US
dc.identifier.citedreferenceP. T. Macklem, D. F. Proctor, and J. C. Hogg, “The stability of peripheral airways,” Respir. Physiol. 8, 191 (1970).en_US
dc.identifier.citedreferenceR. D. Kamm and R. C. Schroter, “Is airway closure caused by a thin liquid instability?” Respir. Physiol. 75, 141 (1989).en_US
dc.identifier.citedreferenceD. Halpern and J. B. Grotberg, “Fluid-elastic instabilities of liquid-lined flexible tubes,” J. Fluid Mech. 244, 615 (1992).en_US
dc.identifier.citedreferenceA. M. Bilek, K. C. Dee, and D. P. Gaver, “Mechanisms of surface-tension-induced epithelial cell damage in a model of pulmonary airway reopening,” J. Appl. Physiol. 94, 770 (2003).en_US
dc.identifier.citedreferenceW. Long, T. Thompson, H. Sundell, and R. Schumacher, F. Volberg, and R. Guthrie, “Effects of two rescue doses of a synthetic surfactant on mortality rate and survival without broncho-pulmonary dysplasia in 700- to 1350-gram700-to1350-gram infants with respiratory distress syndrome. The American Exosurf Neonatal Study Group I,” J. Pediatr. (St. Louis) 118, 595 (1991).en_US
dc.identifier.citedreferenceB. Robertson, “Surfactant replacement therapy for severe neonatal respiratory distress syndrome: An international randomized clinical trial,” J. Pediatr. (St. Louis) 82, 683 (1988).en_US
dc.identifier.citedreferenceH. Yapicioglu, D. Yildizdas, I. Bayram, Y. Sertdemir, and H. L. Yilmaz, “The use of surfactant in children with acute respiratory distress syndrome: Efficacy in terms of oxygenation, ventilation and mortality,” Pulm. Pharmacol. Ther. 16, 327 (2003).en_US
dc.identifier.citedreferenceM. D. Salvia-Roiges, X. Carbonell-Estrany, J. Figueras-Aloy, and J. M. Rodriguez-Miguelez, “Efficacy of three treatment schedules in severe meconium aspiration syndrome,” Acta Pediatr. Esp. 93, 60 (2004).en_US
dc.identifier.citedreferenceE. M. Zola, J. H. Gunkel, R. K. Chan, M. O. Lim, I. Knox, B. H. Feldman, S. E. Denson, B. S. Stonestreet, B. R. Mitchell, M. M. Wyza, K. J. Bennett, and A. J. Gold, “Comparison of 3 dosing procedures for administration of bovine surfactant to neonates with respiratory-distress syndrome,” J. Pediatr. (St. Louis) 122, 453 (1993).en_US
dc.identifier.citedreferenceR. B. Hirschl, R. Tooley, A. C. Parent, K. Johnson, and R. H. Bartlett, “Improvement of gas exchange, pulmonary function, and lung injury with partial liquid ventilation. A study model in a setting of severe respiratory failure,” Chest 108, 500 (1995).en_US
dc.identifier.citedreferenceT. H. Shaffer and M. R. Wolfson, “Liquid ventilation: An alternative ventilation strategy for management of neonatal respiratory distress,” Eur. J. Pediatr. 155, S30 (1996).en_US
dc.identifier.citedreferenceH. P. Baden, J. D. Mellema, S. L. Bratton, P. P. O'Rourke, and J. C. Jackson, “High-frequency oscillatory ventilation with partial liquid ventilation in a model of acute respiratory failure,” Crit. Care Med. 25, 299 (1997).en_US
dc.identifier.citedreferenceC. L. Leach, J. S. Greenspan, S. D. Rubenstein, T. H. Shaffer, M. R. Wolfson, J. C. Jackson, R. DeLemos, and B. P. Fuhrman, “Partial liquid ventilation with perflubron in premature infants with severe respiratory distress syndrome. The LiquiVent Study Group,” N. Engl. J. Med. 335, 761 (1996).en_US
dc.identifier.citedreferenceK. Mikawa, K. Nishina, Y. Takao, and H. Obara, “Efficacy of partial liquid ventilation in improving acute lung injury induced by intratracheal acidified infant formula: Determination of optimal dose and positive end-expiratory pressure level,” Crit. Care Med. 32, 209 (2004).en_US
dc.identifier.citedreferenceC. W. Choi, J. H. Hwang, Y. S. Chang, and W. S. Park, “Combined effect of low-dose nitric oxide gas inhalation with partial liquid ventilation on hemodynamics, pulmonary function, and gas exchange in acute lung injury of newborn piglets,” J. Korean Med. Sci. 18, 813 (2003).en_US
dc.identifier.citedreferenceP. N. Cox, H. Frndova, O. Karlsson, S. Holowka, and C. A. , “Fluorocarbons facilitate lung recruitment,” Intensive Care Med. 29, 2297 (2003).en_US
dc.identifier.citedreferenceD. J. Weiss, L. Bonneau, and D. Liggitt, “Use of perfluorochemical liquid allows earlier detection of gene expression and use of less vector in normal lung and enhances gene expression in acutely injured lung,” Mol. Ther. 3, 734, Part 1 (2001).en_US
dc.identifier.citedreferenceK. Nakazawa, K. Yokoyama, Y. Matsuzawa, K. Makita, and K. Amaha, “Pulmonary administration of prostacyclin (PGI(2)) during partial liquid ventilation in an oleic acid-induced lung injury: inhalation of aerosol or intratracheal instillation?” Intensive Care Med. 27, 243 (2001).en_US
dc.identifier.citedreferenceE. W. Dickson, S. O. Heard, T. E. Tarara, J. G. Weers, A. B. Brueggemann, and G. V. Doern, “Liquid ventilation with perflubron in the treatment of rats with pneumococcal pneumonia,” Crit. Care Med. 30, 393 (2002).en_US
dc.identifier.citedreferenceJ. Yu and Y. W. Chien, “Pulmonary drug delivery: Physiologic and mechanistic aspects,” Crit. Rev. Ther. Drug Carrier Syst. 14, 395 (1997).en_US
dc.identifier.citedreferenceE. Raczka, J. E. Kukowska-Latallo, M. Rymaszewski, C. Chen, and J. R. Baker, Jr., “The effect of synthetic surfactant Exosurf on gene transfer in mouse lung in vivo,” Gene Ther. 5, 1333 (1998).en_US
dc.identifier.citedreferenceA. H. Jobe, T. Ueda, J. A. Whitsett, B. C. Trapnell, and M. Ikegami, “Surfactant enhances adenovirus-mediated gene expression in rabbit lungs,” Gene Ther. 3, 775 (1996).en_US
dc.identifier.citedreferenceO. E. Jensen, D. Halpern, and J. B. Grotberg, “Transport of a passive solute by surfactant-driven flows,” Chem. Eng. Sci. 49, 1107 (1994).en_US
dc.identifier.citedreferenceY. L. Zhang, O. K. Matar, and R. V. Craster, “A theoretical study of chemical delivery within the lung using exogenous surfactant,” Med. Eng. Phys. 25, 115 (2003).en_US
dc.identifier.citedreferenceS. Iqbal, S. Ritson, I. Prince, J. Denyer, and M. L. Everard, “Drug delivery and adherence in young children,” Pediatr. Pulmonol 37, 311 (2004).en_US
dc.identifier.citedreferenceP. B. Myrdal, K. L. Karlage, S. W. Stein, B. A. Brown, and A. Haynes, “Optimized dose delivery of the peptide cyclosporine using hydrofluoroalkane-based metered dose inhalers,” J. Pharm. Sci. 93, 1054 (2004).en_US
dc.identifier.citedreferenceK. J. Cassidy, J. L. Bull, M. R. Glucksberg, C. A. Dawson, S. T. Haworth, R. B. Hirschl, N. Gavriely, and J. B. Grotberg, “A rat lung model of instilled liquid transport in the pulmonary airways,” J. Appl. Physiol. 90, 1955 (2001).en_US
dc.identifier.citedreferenceS. L. Waters and J. B. Grotberg, “The propagation of a surfactant laden liquid plug in a capillary tube,” Phys. Fluids 14, 471 (2002).en_US
dc.identifier.citedreferenceP. D. Howell, S. L. Waters, and J. B. Grotberg, “The propagation of a liquid bolus along a liquid-lined flexible tube,” J. Fluid Mech. 406, 309 (2000).en_US
dc.identifier.citedreferenceH. Fujioka and J. B. Grotberg, “Steady propagation of a liquid plug in a 2-dimensional channel,” ASME J. Biomech. Eng. 126, 567 (2004).en_US
dc.identifier.citedreferenceG. M. Ginley and C. J. Radke, “Influence of soluble surfactants on the flow of long bubbles through a cylindrical capillary,” ACS Symp. Ser. 396, 480 (1989).en_US
dc.identifier.citedreferenceJ. Ratulowski and H.-C. Chang, “Marangoni effects of trace impurities on the motion of long gas bubbles in capillaries,” J. Fluid Mech. 210, 303 (1990).en_US
dc.identifier.citedreferenceF. Wassmuth, W. G. Laidlaw, and D. A. Coombe, “Calculation of interfacial flows and surfactant redistribution as a gas-liquid interface moves between 2 parallel plates,” Phys. Fluids A 5, 1533 (1993).en_US
dc.identifier.citedreferenceS. N. Ghadiali and D. P. Gaver, “The influence of non-equilibrium surfactant dynamics on the flow of a semi-infinite bubble in a rigid cylindrical capillary tube,” J. Fluid Mech. 478, 165 (2003).en_US
dc.identifier.citedreferenceM. Severino, M. D. Giavedoni, and F. A. Saita, “A gas phase displacing a liquid with soluble surfactants out of a small conduit: The plane case,” Phys. Fluids 15, 2961 (2003).en_US
dc.identifier.citedreferenceM. A. Krueger and D. P. Gaver, “A theoretical model of pulmonary surfactant multilayer collapse under oscillating area conditions,” J. Colloid Interface Sci. 229, 353 (2000).en_US
dc.identifier.citedreferenceD. R. Otis, Jr., E. P. Ingenito, R. D. Kamm, and M. Johnson, “Dynamic surface tension of surfactant TA: Experiments and theory,” J. Appl. Physiol. 77, 2681 (1994).en_US
dc.identifier.citedreferenceC. H. Chang and E. I. Franses, “Adsorption dynamics of surfactants at the air/water interface—A critical-review of mathematical-models, data, and mechanisms,” Colloids Surf., A 100, 1 (1995).en_US
dc.identifier.citedreferenceS.V. Patankar, Numerical Heat Transfer and Fluid Flow (Hemisphere, London, 1980).en_US
dc.identifier.citedreferenceJ. F. Thompson and Z. U. Warsi, “Boundary-fitted coordinate system for numerical solution of partial differential equations—A review,” J. Comput. Phys. 47, 1 (1982).en_US
dc.identifier.citedreferenceS. Muzaferija and M. Peric, “Computation of free-surface flows using the finite-volume method and moving grids,” Numer. Heat Transfer, Part B 32, 369 (1997).en_US
dc.identifier.citedreferenceJ.F. Thompson, B.K. Soni, and N.P. Weatherill, Handbook of Grid Generation (CRC, Boca Raton, FL, 1999), p. 4.en_US
dc.identifier.citedreferenceP. Wesseling, Principles of Computational Fluid Dynamics (Springer, New York, 2000), p. 538.en_US
dc.identifier.citedreferenceM. D. Giavedoni and F. A. Saita, “The axisymmetric and plane cases of a gas phase steadily displacing a Newtonian liquid—A simultaneous solution of the governing equations,” Phys. Fluids 9, 2420 (1997).en_US
dc.identifier.citedreferenceD. Halpern and D. P. Gaver, “Boundary-element analysis of the time-dependent motion of a semiinfinite bubble in a channel,” J. Comput. Phys. 115, 366 (1994).en_US
dc.identifier.citedreferenceF. P. Bretherton, “The motion of long bubbles in tubes,” J. Fluid Mech. 10, 166 (1961).en_US
dc.identifier.citedreferenceM. Heil, “Finite Reynolds number effects in the Bretherton problem,” Phys. Fluids 13, 2517 (2001).en_US
dc.identifier.citedreferenceK. J. Cassidy, N. Gavriely, and J. B. Grotberg, “Liquid plug flow in straight and bifurcating tubes,” ASME J. Biomech. Eng. 123, 580 (2001).en_US
dc.identifier.citedreferenceM. A. Launoissurpas, T. Ivanova, I. Panaiotov, J. E. Proust, F. Puisieux, and G. Georgiev, “Behavior of pure and mixed Dppc liposomes spread or adsorbed at the air-water-interface,” Colloid Polym. Sci. 270, 901 (1992).en_US
dc.identifier.citedreferenceS. N. Ghadiali and D. P. Gaver, “An investigation of pulmonary surfactant physicochemical behavior under airway reopening conditions,” J. Appl. Physiol. 88, 493 (2000).en_US
dc.identifier.citedreferenceS. Schurch, H. Bachofen, J. Goerke, and F. Possmayer, “A captive bubble method reproduces the in situ behavior of lung surfactant monolayers,” J. Appl. Physiol. 67, 2389 (1989).en_US
dc.identifier.citedreferenceM. L. Agrawal and R. D. Neuman, “Surface-diffusion in monomolecular films. 2. Experiment and theory,” J. Colloid Interface Sci. 121, 366 (1988).en_US
dc.identifier.citedreferenceW. R. Schief, M. Antia, B. M. Discher, S. B. Hall, and V. Vogel, “Liquid-crystalline collapse of pulmonary surfactant monolayers,” Biophys. J. 84, 3792 (2003).en_US
dc.identifier.citedreferenceM. C. Phillips and D. Chapman, “Monolayer characteristics of saturated I,2-diacyl phosphatidylcholines (lecithins) and phosphatidylethanolamines at air-water interface,” Biochim. Biophys. Acta 163, 301 (1968).en_US
dc.identifier.citedreferenceM. M. Lipp, K. Y. C. Lee, D. Y. Takamoto, J. A. Zasadzinski, and A. J. Waring, “Coexistence of buckled and flat monolayers,” Phys. Rev. Lett. 81, 1650 (1998).en_US
dc.identifier.citedreferenceH. E. Ries and H. Swift, “Twisted double-layer ribbons and the mechanism for monolayer collapse,” Langmuir 3, 853 (1987).en_US
dc.identifier.citedreferenceH. E. Ries, “Stable ridges in a collapsing monolayer,” Nature (London) 281, 287 (1979).en_US
dc.identifier.citedreferenceJ. N. Hildebran, J. Goerke, and J. A. Clements, “Pulmonary surface-film stability and composition,” J. Appl. Physiol.: Respir., Environ. Exercise Physiol. 47, 604 (1979).en_US
dc.identifier.citedreferenceR. J. King and J. A. Clements, “Surface-active materials from dog lung. 2. Composition and physiological correlations,” Am. J. Physiol. 223, 715 (1972).en_US
dc.identifier.citedreferenceC.-W. Park, “In fluence of soluble surfactants on the motion of finite bubble in a capillary tube,” Phys. Fluids A 4, 2335 (1992).en_US
dc.identifier.citedreferenceM. D. Giavedoni and F. A. Saita, “The rear meniscus of a long bubble steadily displacing a Newtonian liquid in a capillary tube,” Phys. Fluids 11, 786 (1999).en_US
dc.identifier.citedreferenceM. S. Borgas and J. B. Grotberg, “Monolayer flow on a thin film,” J. Fluid Mech. 193, 151 (1988).en_US
dc.identifier.citedreferenceN. Antonova, R. Todorov, and D. Exerowa, “Rheological behavior and parameters of the in vitro model of lung surfactant systems: The role of the main phospholipid component,” Biorheology 40, 531 (2003).en_US
dc.owningcollnamePhysics, Department of


Files in this item

Show simple item record

Remediation of Harmful Language

The University of Michigan Library aims to describe library materials in a way that respects the people and communities who create, use, and are represented in our collections. Report harmful or offensive language in catalog records, finding aids, or elsewhere in our collections anonymously through our metadata feedback form. More information at Remediation of Harmful Language.

Accessibility

If you are unable to use this file in its current format, please select the Contact Us link and we can modify it to make it more accessible to you.