Hostname: page-component-5c6d5d7d68-txr5j Total loading time: 0 Render date: 2024-08-16T13:41:07.671Z Has data issue: false hasContentIssue false

Dynamics and mixing of vortex rings in crossflow

Published online by Cambridge University Press:  14 May 2008

RAJES SAU
Affiliation:
Aerospace Engineering & Mechanics, University of Minnesota, Minneapolis, MN 55455, USA
KRISHNAN MAHESH
Affiliation:
Aerospace Engineering & Mechanics, University of Minnesota, Minneapolis, MN 55455, USA

Abstract

Direct numerical simulation is used to study the effect of crossflow on the dynamics, entrainment and mixing characteristics of vortex rings issuing from a circular nozzle. Three distinct regimes exist, depending on the velocity ratio (ratio of the average nozzle exit velocity to free-stream crossflow velocity) and stroke ratio (ratio of stroke length to nozzle exit diameter). Coherent vortex rings are not obtained at velocity ratios below approximately 2. At these low velocity ratios, the vorticity in the crossflow boundary layer inhibits roll-up of the nozzle boundary layer at the leading edge. As a result, a hairpin vortex forms instead of a vortex ring. For large stroke ratios and velocity ratio below 2, a series of hairpin vortices is shed downstream. The shedding is quite periodic for very low Reynolds numbers. For velocity ratios above 2, two regimes are obtained depending upon the stroke ratio. Lower stroke ratios yield a coherent asymmetric vortex ring, while higher stroke ratios yield an asymmetric vortex ring accompanied by a trailing column of vorticity. These two regimes are separated by a transition stroke ratio whose value decreases with decreasing velocity ratio. For very high values of the velocity ratio, the transition stroke ratio approaches the ‘formation number’. In the absence of trailing vorticity, the vortex ring tilts towards the upstream direction, while the presence of a trailing column causes it to tilt downstream. This behaviour is explained. In the absence of crossflow, the trailing column is not very effective at entrainment, and is best avoided for optimal mixing and entrainment. However, in the presence of crossflow, the trailing column is found to contribute significantly to the overall mixing and entrainment. The trailing column interacts with the crossflow to generate a region of high pressure downstream of the nozzle that drives crossflow fluid towards the vortex ring. There is an optimal length of the trailing column for maximum downstream entrainment. A classification map which categorizes the different regimes is developed.

Type
Papers
Copyright
Copyright © Cambridge University Press 2008

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

REFERENCES

Acarlar, M. S. & Smith, C. R. 1987 A study of hairpin vortices in a laminar boundary layer. Part 2. Hairpin vortices generated by fluid injection. J. Fluid Mech. 175, 4383.CrossRefGoogle Scholar
Blossey, P., Narayanan, S. & Bewley, T. R. 2001 Dynamics and control of a jet in crossflow: direct numerical simulation and experiments. Proc. IUTAM Symp. Turbulent Mixing Combustion (ed. Pollard, A. & Candel, S.), Kluwer., pp. 45–56.Google Scholar
Chang, Y. K. & Vakili, A. D. 1995 Dynamics of vortex rings in crossflow. Phys. Fluids 7, 15831597.CrossRefGoogle Scholar
Eroglu, A. & Briedenthal, R. E. 2001 Structure, penetration and mixing of pulsed jets in crossflow. AIAA J. 39, 417423.CrossRefGoogle Scholar
Fric, T. F. & Roshko, A. 1994 Vortical structure in the wake of a transverse jet. J. Fluid Mech. 279, 147.CrossRefGoogle Scholar
Gharib, M., Rambod, E. & Shariff, K. 1998 A universal time scale for vortex ring formation. J. Fluid Mech. 360, 121140.CrossRefGoogle Scholar
Gopalan, S., Abraham, B. M. & Katz, J. 2004 The structure of a jet in cross flow at low velocity ratios. Phys. Fluids 16, 20672087.CrossRefGoogle Scholar
Johari, H. 2006 Scaling of fully pulsed jets in crossflow. AIAA J. 44, 27192725.CrossRefGoogle Scholar
Karagozian, A. R., Cortelezi, L. & Soldati, A. 2003 Manipulation and control of jets in crossflow. CISM Courses and Lectures No. 439, International Center for Mechanical Sciences, Springer Wien, New York.CrossRefGoogle Scholar
Mahesh, K., Constantinescu, G. & Moin, P. 2004 A numerical method for large–eddy simulation in complex geometries. J. Comput. Phys. 197, 215240.CrossRefGoogle Scholar
M'Closkey, R. T., King, J. M., Cortelezzi, L. & Karagozian, A. R. 2002 The actively controlled jet in crossflow. J. Fluid Mech. 452 325335.CrossRefGoogle Scholar
Muppidi, S. 2006 Direct numerical simulations and modeling of jets in crossflow. PhD Thesis, University of Minnesota.CrossRefGoogle Scholar
Muppidi, S. & Mahesh, K. 2008 Direct numerical simulation of passive scalar transport in transverse jets. J. Fluid Mech. 598, 335360.CrossRefGoogle Scholar
Sau, R. & Mahesh, K. 2007 Passive scalar mixing in vortex rings. J. Fluid Mech. 582, 449491.CrossRefGoogle Scholar
Shapiro, S. R., King, J. M., Karagozian, A. R. & M'Closkey, R. T. 2006 Optimization of controlled jets in crossflow. AIAA J. 44, 12921298.CrossRefGoogle Scholar
Ting, L. & Tung, C. 1965 Motion and decay of a vortex in a nonuniform stream. Phys. Fluids 8, 10391051.CrossRefGoogle Scholar
Wu, J. M, Vakili, A. D. & Yu, F. M. 1988 Investigation of interacting flow of nonsymmetric jets in crossflow. AIAA J. 26, 940947.CrossRefGoogle Scholar