Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-19T21:23:58.979Z Has data issue: false hasContentIssue false

Self-similar decay and mixing of a high-Schmidt-number passive scalar in an oscillating boundary layer in the intermittently turbulent regime

Published online by Cambridge University Press:  05 June 2013

Carlo Scalo
Affiliation:
Department of Mechanical and Materials Engineering, Queen’s University, 130 Stuart Street, Kingston, Ontario K7L 3N6, Canada
Ugo Piomelli
Affiliation:
Department of Mechanical and Materials Engineering, Queen’s University, 130 Stuart Street, Kingston, Ontario K7L 3N6, Canada
Leon Boegman
Affiliation:
Department of Civil Engineering, Queen’s University, 58 University Avenue, Kingston, Ontario K7L 3N6, Canada

Abstract

We performed numerical simulations of dissolved oxygen (DO) transfer from a turbulent flow, driven by periodic boundary-layer turbulence in the intermittent regime, to underlying DO-absorbing organic sediment layers. A uniform initial distribution of oxygen is left to decay (with no re-aeration) as the turbulent transport supplies the sediment with oxygen from the outer layers to be absorbed. A very thin diffusive sublayer at the sediment–water interface (SWI), caused by the high Schmidt number of DO in water, limits the overall decay rate. A decomposition of the instantaneous decaying turbulent scalar field is proposed, which results in the development of similarity solutions that collapse the data in time. The decomposition is then tested against the governing equations, leading to a rigorous procedure for the extraction of an ergodic turbulent scalar field. The latter is composed of a statistically periodic and a steady non-decaying field. Temporal averaging is used in lieu of ensemble averaging to evaluate flow statistics, allowing the investigation of turbulent mixing dynamics from a single flow realization. In spite of the highly unsteady state of turbulence, the monotonically decaying component is surprisingly consistent with experimental and numerical correlations valid for steady high-Schmidt-number turbulent mass transfer. Linearly superimposed onto it is the statistically periodic component, which incorporates all the features of the non-equilibrium state of turbulence. It is modulated by the evolution of the turbulent coherent structures driven by the oscillating boundary layer in the intermittent regime, which are responsible for the violent turbulent production mechanisms. These cause, in turn, a rapid increase of the turbulent mass flux at the edge of the diffusive sublayer. This outer-layer forcing mechanism drives a periodic accumulation of high scalar concentration levels in the near-wall region. The resulting modulated scalar flux across the SWI is delayed by a quarter of a cycle with respect to the wall-shear stress, consistently with the non-equilibrium state of the turbulent mixing.

Type
Papers
Copyright
©2013 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

Akhavan, R., Kamm, R. D. & Shapiro, A. H. 1991a An investigation of transition to turbulence in bounded oscillatory stokes flows. Part 1. Experiments. J. Fluid Mech. 225, 395422.Google Scholar
Akhavan, R., Kamm, R. D. & Shapiro, A. H. 1991b An investigation of transition to turbulence in bounded oscillatory Stokes flows. Part 2. Numerical simulations. J. Fluid Mech. 225, 423444.Google Scholar
Antonia, RA & Orlandi, P. 2003 Effect of Schmidt number on small-scale passive scalar turbulence. Appl. Mech. Rev. 56 (6), 615632.Google Scholar
Bergant, R. & Tiselj, I. 2007 Near-wall passive scalar transport at high Prandtl numbers. Phys. Fluids 19, 065105.Google Scholar
Boudreau, B. P. & Jørgensen, B. B. 2001 The Benthic Boundary Layer: Transport Processes and Biogeochemistry. Oxford University Press.CrossRefGoogle Scholar
Bouldin, D. R. 1968 Models for describing the diffusion of oxygen and other mobile constituents across the mud-water interface. J. Ecol. 56 (1), 7787.Google Scholar
Bryant, L. D., Lorrai, C., McGinnis, D. F., Brand, A., Wüest, A. & Little, J. C. 2010 Variable sediment oxygen uptake in response to dynamic forcing. Limnol. Oceanogr. 55 (2), 950964.Google Scholar
Burns, N. M. & Ross, C. 1972 Oxygen-nutrient relationships within the Central Basin of Lake Erie. Paper no. 6. USEPA  Tech. Rep. TS-05-71-208-24. Canada Centre for Inland Waters, Burlington, Ontario.Google Scholar
Calmet, I. & Magnaudet, J. 1997 Large-eddy simulation of high-Schmidt-number mass transfer in a turbulent channel flow. Phys. Fluids 9, 438455.CrossRefGoogle Scholar
Campbell, J. A. & Hanratty, T. J. 1983 Turbulent velocity fluctuations that control mass transfer to a solid boundary. AIChE J. 29 (2), 215221.Google Scholar
Committee on Environment and Natural Resources 2010 Scientific assessment of hypoxia in US coastal waters. Tech. Rep. Interagency Working Group on Harmful Algal Blooms, Hypoxia, and Human Health of the Joint Subcommittee on Ocean Science and Technology, Washington, DC.Google Scholar
Costamagna, P., Vittori, G. & Blondeaux, P. 2003 Coherent structures in oscillatory boundary layers. J. Fluid Mech. 474, 133.CrossRefGoogle Scholar
Diaz, R. J. & Rosenberg, R. 2008 Spreading dead zones and consequences for marine ecosystems. Science 321 (5891), 926929.Google Scholar
Gayen, B., Sarkar, S. & Taylor, J. R. 2010 Large eddy simulation of a stratified boundary layer under an oscillatory current. J. Fluid Mech. 643, 233266.Google Scholar
Germano, M., Piomelli, U., Moin, P. & Cabot, W. H. 1991 A dynamic subgrid-scale eddy viscosity model. Phys. Fluids A 3, 17601765.Google Scholar
Hanes, N. Bruce & Irvine, Robert L. 1968 New technique for measuring oxygen uptake rates of benthal systems. J. (Water Pollut. Control Fed.) 40 (2), 223232.Google Scholar
Hanratty, T. J. 1956 Turbulent exchange of mass and momentum with a boundary. AIChE J. 2 (3), 359362.Google Scholar
Higashino, M., Clark, J. J. & Stefan, H. G. 2009 Pore water flow due to near-bed turbulence and associated solute transfer in a stream or lake sediment bed. Water Resour. Res. 45 (12), W12414.Google Scholar
Higashino, M., O’Connor, B. L., Hondzo, M. & Stefan, H. G. 2008 Oxygen transfer from flowing water to microbes in an organic sediment bed. Hydrobiologia 614, 219231.CrossRefGoogle Scholar
Higashino, M. & Stefan, H. G. 2011 Dissolved oxygen demand at the sediment–water interface of a stream: near-bed turbulence and pore water flow effects. J. Environ. Engng 137 (7), 531.Google Scholar
Hino, M., Kashiwayanagi, M., Nakayama, A. & Hara, T. 1983 Experiments on the turbulence statitistics and the structure of a reciprocating oscillatory flow. J. Fluid Mech. 131, 363400.Google Scholar
Jensen, B. L., Sumer, B. M. & Fredsøe, J. 1989 Turbulent oscillatory boundary layers at high Reynolds numbers. J. Fluid Mech. 206, 265297.Google Scholar
Jiménez, J. & Moin, P. 1991 The minimal flow unit in near-wall turbulence. J. Fluid Mech. 225, 213240.CrossRefGoogle Scholar
Jimenez, J. & Pinelli, A. 1999 The autonomous cycle of near-wall turbulence. J. Fluid Mech. 389, 335359.Google Scholar
Keating, A., Piomelli, U., Bremhorst, K. & Nešić, S. 2004 Large-eddy simulation of heat transfer downstream of a backward-facing step. J. Turbul. 5 (20), 127.Google Scholar
Lilly, D. K. 1992 A proposed modification of the Germano subgrid-scale closure method. Phys. Fluids A 4, 633635.Google Scholar
Lorke, A., Müller, B., Maerki, M. & Wüest, A. 2003 Breathing sediments: the control of diffusive transport across the sediment–water interface by periodic boundary-layer turbulence. Limnol. Oceanogr. 48 (6), 20772085.Google Scholar
Moser, R. D., Kim, J. & Mansour, N. N. 1999 Direct numerical simulation of turbulent channel flow up to $Re= 590$ . Phys. Fluids 11, 943945.CrossRefGoogle Scholar
Na, Y. 2004 On the large-eddy simulation of high Prandtl number scalar transport using dynamic subgrid-scale model. J. Mech. Sci. Technol. 18 (1), 173182.Google Scholar
O’Connor, B. L. & Hondzo, M. 2007 Enhancement and inhibition of denitrification by fluid-flow and dissolved oxygen flux to stream sediments. Environ. Sci. Technol. 42 (1), 119125.CrossRefGoogle Scholar
O’Connor, B. L. & Hondzo, M. 2008 Dissolved oxygen transfer to sediments by sweep and eject motions in aquatic environments. Limnol. Oceanogr. 53 (2), 566578.Google Scholar
Patterson, J. C., Allanson, B. R. & Ivey, G. N. 1985 A dissolved oxygen budget model for Lake Erie in summer. Freshwat. Biol. 15, 683694.Google Scholar
Pinczewski, W. V. & Sideman, S. 1974 A model for mass (heat) transfer in turbulent tube flow. Moderate and high Schmidt (Prandtl) numbers. Chem. Engng Sci. 29 (9), 19691976.Google Scholar
Pope, S. B. 2000 Turbulent Flows. Cambridge University Press.Google Scholar
Rabalais, N. N., Diaz, R. J., Levin, L. A., Turner, R. E., Gilbert, D. & Zhang, J. 2010 Dynamics and distribution of natural and human-caused hypoxia. Biogeosciences 7 (2), 585619.Google Scholar
Rao, Y. R., Hawley, N., Charlton, M. N. & Schertzer, W. M. 2008 Physical processes and hypoxia in the central basin of Lake Erie. Limnol. Oceanogr. 53 (5), 20072020.CrossRefGoogle Scholar
Røy, H., Huettel, M. & Joergensen, B. B. 2004 Transmission of oxygen concentration fluctuations through the diffusive boundary layer overlying aquatic sediments. Limnol. Oceanogr. 49 (3), 686692.Google Scholar
Salon, S., Armenio, V. & Crise, A. 2007 A numerical investigation of the stokes boundary layer in the turbulent regime. J. Fluid Mech. 570, 253296.Google Scholar
Scalo, C., Boegman, L. & Piomelli, U. 2013 Large eddy simulation and low-order modelling of sediment oxygen uptake in a transitional oscillatory flow. J. Geophys. Res. 118 (1–14).Google Scholar
Scalo, C., Piomelli, U. & Boegman, L. 2012a High-Schmidt-number mass transport mechanisms from a turbulent flow to absorbing sediments. Phys. Fluids 24 (8), 085103.Google Scholar
Scalo, C., Piomelli, U. & Boegman, L. 2012b Large-eddy simulation of oxygen transfer to organic sediment beds. J. Geophys. Res. 117 (C6), C06005.Google Scholar
Shaw, D. A. & Hanratty, T. J. 1977 Turbulent mass transfer to a wall for large Schmidt numbers. AIChE J. 23 (1), 2837.Google Scholar
Sleath, J. F. A. 1987 Turbulent oscillatory flow over rough beds. J. Fluid Mech. 182, 369409.Google Scholar
Spalart, P. R. & Baldwin, B. S. 1989 Direct simulation of a turbulent oscillating boundary layer. In Turbulent Shear Flows (ed. André, J. C. et al. ). vol. 6. pp. 417440. Springer.Google Scholar
Trowbridge, J. H. & Agrawal, Y. C. 1995 Glimpses of a wave boundary layer. J. Geophys. Res. 100 (C10), 2072920743.Google Scholar
Veenstra, J. N. & Nolen, S. L. 1991 In-Situ sediment oxygen demand in five Southwestern US lakes. Water Res. 25 (3), 351354.Google Scholar
Verzicco, R. & Vittori, G. 1996 Direct simulation of transition in Stokes boundary layers. Phys. Fluids 8, 1341.Google Scholar
Vittori, G. & Verzicco, R. 1998 Direct simulation of transition in an oscillatory boundary layer. J. Fluid Mech. 371, 207232.Google Scholar
Wang, L. & Lu, X. 2004 An investigation of turbulent oscillatory heat transfer in channel flows by large eddy simulation. Intl J. Heat Mass Transfer 47 (10–11), 21612172.Google Scholar
Warhaft, Z. 2000 Passive scalars in turbulent flows. Annu. Rev. Fluid Mech. 32, 203240.CrossRefGoogle Scholar
Zang, Y., Street, R. L. & Koseff, J. R. 1993 A dynamic mixed subgrid-scale model and its application to turbulent recirculating flows. Phys. Fluids A 5, 3186.Google Scholar