Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-27T03:56:05.865Z Has data issue: false hasContentIssue false

Splitting of a two-dimensional liquid plug at an airway bifurcation

Published online by Cambridge University Press:  14 March 2016

Benjamin L. Vaughan Jr*
Affiliation:
Department of Mathematical Sciences, University of Cincinnati, 4199 French Hall West, Cincinnati, OH 45221-0025, USA Department of Biomedical Engineering, University of Michigan, 2200 Bonisteel Blvd., Ann Arbor, MI 48109, USA
James B. Grotberg
Affiliation:
Department of Biomedical Engineering, University of Michigan, 2200 Bonisteel Blvd., Ann Arbor, MI 48109, USA
*
Email address for correspondence: vaughabn@ucmail.uc.edu

Abstract

Certain medical treatments involve the introduction of exogenous liquids in the lungs. These liquids can form plugs within the airways. The plugs propagate throughout the branching network in the lungs being forced by airflow. They leave a deposited film on the airway walls and split at bifurcations. Understanding the resulting distribution of liquid throughout the lungs is important for the effective administration of the prescribed treatments. In this paper, we investigate numerically the splitting of a liquid plug by a two-dimensional pulmonary bifurcation under the influence of a transverse gravitational field. The splitting is characterized by the splitting ratio, which is the ratio of volume of the liquid plug in the daughter channels and depends on the capillary number and the orientation of the bifurcation plane with respect to a three-dimensional gravitational field. It is observed that gravity induces asymmetry in the splitting, causing the splitting ratio to be reduced. This effect is mitigated as the capillary number is increased. It is also observed that there exists a critical capillary number where the plug will not split and will instead propagate entirely into the gravitationally favoured daughter channel. We also compute the wall stresses on the bifurcation walls and observe the locations where stresses and their gradients are the highest in magnitude.

Type
Papers
Copyright
© 2016 Cambridge University Press 

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

Abadie, T., Aubin, J., Legendre, D. & Xuereb, C. 2012 Hydrodynamics of gas–liquid Taylor flow in rectangular microchannels. Microfluid. Nanofluid. 12 (1–4), 355369.Google Scholar
Anderson, J. C., Molthen, R. C., Dawson, C. A., Haworth, S. T., Bull, J. L., Glucksberg, M. R. & Grotberg, J. B. 2004 Effect of ventilation rate on instilled surfactant distribution in the pulmonary airways of rats. J. Appl. Physiol. 97 (1), 4556.CrossRefGoogle ScholarPubMed
Angeli, P. & Gavriilidis, A. 2008 Hydrodynamics of Taylor flow in small channels: a review. Proc. Inst. Mech. Engrs C 222 (5), 737751.Google Scholar
Aussillous, P. & Quéré, D. 2000 Quick deposition of a fluid on the wall of a tube. Phys. Fluids 12 (10), 23672371.Google Scholar
Bilek, A. M., Dee, K. C. & Gaver, D. P. III 2003 Mechanisms of surface-tension-induced epithelial cell damage in a model of pulmonary airway reopening. J. Appl. Physiol. 94 (2), 770783.CrossRefGoogle Scholar
Bretherton, F. P. 1961 The motion of long bubbles in tubes. J. Fluid Mech. 10 (02), 166188.Google Scholar
Bull, J. L., Tredici, S., Komori, E., Brant, D. O., Grotberg, J. B. & Hirschl, R. B. 2004 Distribution dynamics of perfluorocarbon delivery to the lungs: an intact rabbit model. J. Appl. Physiol. 96 (5), 16331642.CrossRefGoogle Scholar
Cassidy, K. J., Bull, J. L., Glucksberg, M. R., Dawson, C. A., Haworth, S. T., Hirschl, R., Gavriely, N. & Grotberg, J. B. 2001 A rat lung model of instilled liquid transport in the pulmonary airways. J. Appl. Physiol. 90 (5), 19551967.Google Scholar
Cassidy, K. J., Halpern, D., Ressler, B. G. & Grotberg, J. B. 1999 Surfactant effects in model airway closure experiments. J. Appl. Physiol. 87 (1), 415427.Google Scholar
Chen, X., Zielinski, R. & Ghadiali, S. N. 2014 Computational analysis of microbubble flows in bifurcating airways: role of gravity, inertia, and surface tension. Trans. ASME J. Biomech. Engng 136 (10), 101007.Google ScholarPubMed
Cox, P. N., Frndova, H., Karlsson, O., Holowka, S. & Bryan, C. A. 2003 Fluorocarbons facilitate lung recruitment. Intens. Care Med. 29 (12), 22972302.Google Scholar
Cubaud, T. & Ho, C.-M. 2004 Transport of bubbles in square microchannels. Phys. Fluids 16 (12), 45754585.CrossRefGoogle Scholar
Donn, S. M. & Dalton, J. 2009 Surfactant replacement therapy in the neonate: beyond respiratory distress syndrome. Respir. Care 54 (9), 12031208; 54th International Respiratory Congress of the American-Association-for-Respiratory-Care, Anaheim, CA, Dec 13–16, 2008.Google Scholar
Douville, N. J., Zamankhan, P., Tung, Y.-C., Li, R., Vaughan, B. L., Tai, C.-F., White, J., Christensen, P. J., Grotberg, J. B. & Takayama, S. 2011 Combination of fluid and solid mechanical stresses contribute to cell death and detachment in a microfluidic alveolar model. Lab on a Chip 11 (4), 609619.Google Scholar
Espinosa, F. F. & Kamm, R. D. 1998 Meniscus formation during tracheal instillation of surfactant. J. Appl. Physiol. 85 (1), 266272.CrossRefGoogle ScholarPubMed
Filoche, M., Tai, C.-F. & Grotberg, J. B. 2015 Three-dimensional model of surfactant replacement therapy. Proc. Natl Acad. Sci. USA 112 (30), 92879292.Google Scholar
Fujioka, H. & Grotberg, J. B. 2004 Steady propagation of a liquid plug in a two-dimensional channel. J. Biomed. Engng 126 (5), 567577.Google Scholar
Fujioka, H. & Grotberg, J. B. 2005 The steady propagation of a surfactant-laden liquid plug in a two-dimensional channel. Phys. Fluids 17, 082102.CrossRefGoogle Scholar
Gaver, D. P., Halpern, D., Jensen, O. E. & Grotberg, J. B. 1996 The steady motion of a semi-infinite bubble through a flexible-walled channel. J. Fluid Mech. 319, 2565.Google Scholar
Ghadiali, S. N. & Gaver, D. P. 2003 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, 165196.CrossRefGoogle Scholar
Ghadiali, S. N., Halpern, D. & Gaver, D. P. 2001 A dual-reciprocity boundary element method for evaluating bulk convective transport of surfactant in free-surface flows. J. Comput. Phys. 171 (2), 534559.CrossRefGoogle Scholar
Gilliard, N., Richman, P. M., Merritt, T. A. & Spragg, R. G. 1990 Effect of volume and dose on the pulmonary distribution of exogenous surfactant administered to normal rabbits or to rabbits with oleic-acid lung injury. Am. Rev. Respir. Dis. 141 (3), 743747.CrossRefGoogle ScholarPubMed
Gregory, T. J., Steinberg, K. P., Spragg, R., Gadek, J. E., Hyers, T. M., Longmore, W. J., Moxley, M. A., Cai, G.-Z., Hite, R. D., Smith, R. M. et al. 1997 Bovine surfactant therapy for patients with acute respiratory distress syndrome. Am. J. Respir. Crit. Care Med. 155 (4), 13091315.CrossRefGoogle ScholarPubMed
Guo, Z. L., Lu, G. P., Ren, T., Zheng, Y. H., Gong, J. Y., Yu, J. & Liang, Y. J. 2009 Partial liquid ventilation confers protection against acute lung injury induced by endotoxin in juvenile piglets. Respir. Physiol. Neurobiol. 167 (3), 221226.Google Scholar
Gupta, R., Fletcher, D. F. & Haynes, B. S. 2010 Taylor flow in microchannels: a review of experimental and computational work. J. Comput. Multiphase Flows 2 (1), 132.Google Scholar
Halpern, D. & Gaver, D. P. 1994 Boundary element analysis of the time-dependent motion of a semi-infinite bubble in a channel. J. Comput. Phys. 115 (2), 366375.CrossRefGoogle Scholar
Halpern, D., Jensen, O. E. & Grotberg, J. B. 1998 A theoretical study of surfactant and liquid delivery into the lung. J. Appl. Physiol. 85 (1), 333352.Google Scholar
Heil, M. 2000 Finite Reynolds number effects in the propagation of an air finger into a liquid-filled flexible-walled channel. J. Fluid Mech. 424, 2144.Google Scholar
Hirschl, R. B., Croce, M., Gore, D., Wiedemann, H., Davis, K., Zwischenberger, J. & Bartlett, R. H. 2002 Prospective, randomized, controlled pilot study of partial liquid ventilation in adult acute respiratory distress syndrome. Am. J. Respir. Crit. Care Med. 165 (6), 781787.Google Scholar
Huh, D., Fujioka, H., Tung, Y.-C., Futai, N., Paine, R., Grotberg, J. B. & Takayama, S. 2007 Acoustically detectable cellular-level lung injury induced by fluid mechanical stresses in microfluidic airway systems. Proc. Natl Acad. Sci. USA 104 (48), 1888618891.CrossRefGoogle ScholarPubMed
Iqbal, S., Ritson, S., Prince, I., Denyer, J. & Everard, M. L. 2004 Drug delivery and adherence in young children. Pediatr. Pulm. 37 (4), 311317.CrossRefGoogle ScholarPubMed
Jeng, M.-J., Soong, W.-J. & Lee, Y.-S. 2009 Effective lavage volume of diluted surfactant improves the outcome of meconium aspiration syndrome in newborn piglets. Pediatr. Res. 66 (1), 107112.Google Scholar
Jensen, M. H., Libchaber, A., Pelce, P. & Zocchi, G. 1987 Effect of gravity on the Saffman–Taylor meniscus – theory and experiment. Phys. Rev. A 35 (5), 22212227.CrossRefGoogle ScholarPubMed
Jensen, O. E., Halpern, D. & Grotberg, J. B. 1994 Transport of a passive solute by surfactant-driven flows. Chem. Engng Sci. 49 (8), 11071117.CrossRefGoogle Scholar
Klaseboer, E., Gupta, R. & Manica, R. 2014 An extended Bretherton model for long taylor bubbles at moderate capillary numbers. Phys. Fluids 26 (3), 032107.Google Scholar
Kolb, W. B. & Cerro, R. L. 1993 The motion of long bubbles in tubes of square cross section. Phys. Fluids A 5 (7), 15491557.Google Scholar
Long, W., Thompson, T., Sundell, H., Schumacher, R., Volberg, F. & Guthrie, R. 1991 Effects of 2 rescue doeses of a dynthetic surfactant on mortality-rate and survival without bronchopulmonary dysplasia in 700-gram to 1350-gram infants with respiratory-distress syndrome. J. Pediatr. 118 (4, Part 1), 595605.Google Scholar
Mikawa, K., Nishina, K., Takao, Y. & Obara, H. 2004 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 (1), 209216.Google Scholar
Myrdal, P. B., Karlage, K. L., Stein, S. W., Brown, B. A. & Haynes, A. 2004 Optimized dose delivery of the peptide cyclosporine using hydrofluoroalkane-based metered dose inhalers. J. Pharm. Sci. 93 (4), 10541061.CrossRefGoogle ScholarPubMed
Rasche, S., Friedrich, S., Bleyl, J. U., de Abreu, M. G., Koch, T. & Ragaller, M. 2010 Pilot study of vaporization of perfluorohexane during high-frequency oscillatory ventilation in experimental acute lung injury. Exp. Lung Res. 36 (9), 538547.CrossRefGoogle ScholarPubMed
Salvia-Roiges, M. D., Carbonell-Estrany, X., Figueras-Aloy, J. & Rodriguez-Miguelez, J. M. 2004 Efficacy of three treatment schedules in severe meconium aspiration syndrome. Acta Paediatr. 93 (1), 6065.Google Scholar
Smith, B. J. & Gaver, D. P. 2008 The pulsatile propagation of a finger of air within a fluid-occluded cylindrical tube. J. Fluid Mech. 601, 123.Google Scholar
Spragg, R. G., Lewis, J. F., Walmrath, H.-D., Johannigman, J., Bellingan, G., Laterre, P.-F., Witte, M. C., Richards, G. A., Rippin, G., Rathgeb, F. et al. 2004 Effect of recombinant surfactant protein c-based surfactant on the acute respiratory distress syndrome. New Engl. J. Med. 351 (9), 884892.CrossRefGoogle ScholarPubMed
Spragg, R. G., Taut, F. J. H., Lewis, J. F., Schenk, P., Ruppert, C., Dean, N., Krell, K., Karabinis, A. & Günther, A. 2011 Recombinant surfactant protein c-based surfactant for patients with severe direct lung injury. Am. J. Respir. Crit. Care Med. 183 (8), 10551061.Google Scholar
Suresh, V. & Grotberg, J. B. 2005 The effect of gravity on liquid plug propagation in a two-dimensional channel. Phys. Fluids 17 (3), 031507.Google Scholar
Taylor, G. I. 1961 Deposition of a viscous fluid on the wall of a tube. J. Fluid Mech. 10 (02), 161165.CrossRefGoogle Scholar
Ueda, T., Ikegami, M., Rider, E. D. & Jobe, A. H. 1994 Distribution of surfactant and ventilation in surfactant-treated preterm lambs. J. Appl. Physiol. 76 (1), 4555.Google Scholar
Yalcin, H. C., Perry, S. F. & Ghadiali, S. N. 2007 Influence of airway diameter and cell confluence on epithelial cell injury in an in vitro model of airway reopening. J. Appl. Physiol. 103 (5), 17961807.CrossRefGoogle Scholar
Yapicioglu, H., Yildizdas, D., Bayram, I., Sertdemir, Y. & Yilmaz, H. L. 2003 The use of surfactant in children with acute respiratory distress syndrome: efficacy in terms of oxygenation, ventilation and mortality. Pulm. Pharmacol. Ther. 16 (6), 327333.Google Scholar
Yu, J. W. & Chien, Y. W. 1997 Pulmonary drug delivery: physiologic and mechanistic aspects. Crit. Rev. Ther. Drug Carrier Syst. 14 (4), 395453.CrossRefGoogle ScholarPubMed
Zamankhan, P., Helenbrook, B. T., Takayama, S. & Grotberg, J. B. 2012 Steady motion of Bingham liquid plugs in two-dimensional channels. J. Fluid Mech. 705, 258279.CrossRefGoogle Scholar
Zhang, Y. L., Matar, O. K. & Craster, R. V. 2003 Atheoretical study of chemical delivery within the lung using exogenous surfactant. Med. Engng Phys. 25 (2), 115132.Google Scholar
Zheng, Y., Anderson, J. C., Suresh, V. & Grotberg, J. B. 2005 Effect of gravity on liquid plug transport through an airway bifurcation model. J. Biomed. Engng 127 (5), 798806.CrossRefGoogle ScholarPubMed
Zheng, Y., Fujioka, H. & Grotberg, J. B. 2007 Effects of gravity, inertia, and surfactant on steady plug propagation in a two-dimensional channel. Phys. Fluids 19 (8), 082107.Google Scholar
Zheng, Y., Fujioka, H., Grotberg, J. C. & Grotberg, J. B. 2006 Effects of inertia and gravity on liquid plug splitting at a bifurcation. J. Biomed. Engng 128 (5), 707716.Google Scholar