Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-05-22T11:18:25.069Z Has data issue: false hasContentIssue false

Inward turning base-bleed technique for base drag reduction

Published online by Cambridge University Press:  12 August 2022

S. Paul*
Affiliation:
MIT Campus, Anna University, Department of Aerospace Engineering, Chennai-600044, TN, India
A. Vinoth Raj
Affiliation:
MIT Campus, Anna University, Department of Aerospace Engineering, Chennai-600044, TN, India
C. Senthil Kumar
Affiliation:
MIT Campus, Anna University, Department of Aerospace Engineering, Chennai-600044, TN, India
*
*Corresponding author. Email: pswagata82@yahoo.com

Abstract

This paper consists of CFD and experimental study for shell projectile at angle of attack 00 and at various Mach numbers (0.7, 0.9, 1.0, 1.5 and 2.0), without spin effect. Passive method of base modification is used to reduce base drag. The goal of this study is to reduce base drag by utilising a base bleed approach called the Inward Turning Base Bleed Method (IWTB). The concept of IWTB is to draw relatively high-pressure air behind the driving band and allow it to pass through the bleed holes and direct it into the low-pressure base. This will in turn raise the base pressure and lower the base drag. The study comprises of (i) basic model with a boattail angle of $8^{\circ}$ , (ii) eight different cavity models having different lip thickness and depth thickness ratios and (iii) nine different IWTB configurations for the optimised cavity model. At first, cavity model is optimised by varying the lip thickness and depth thickness ratio. Later IWTB parameters such as the bleed hole entry angles $10^{\circ}$ , $12^{\circ}$ & $15^{\circ}$ , exit bleed hole angles $30^{\circ}$ , $45^{\circ}$ & $60^{\circ}$ , bleed hole diameters 3mm, 4mm, 5mm & 6mm and number of bleed holes i.e., 4 & 8 were studied and optimised. Based on this study, the model with 8 holes, 3mm base bleed hole diameter, $15^{\circ}$ entry and $60^{\circ}$ exit angle gives comparatively lower base drag. CFD result shows, in supersonic region the base drag reduction for optimised shell projectile is $3.08\%$ ; at Mach number 0.9 the base reduction is $75.63\%$ ; and in subsonic region it reduces to $23.53\%$ . CFD results were compared with experimental result, is found to be good and the differences lies within $3.85\%$ .

Type
Research Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Royal Aeronautical Society

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

Gujarathi, S.Y. and Joshi, S.N., A review on active and passive flow control techniques, IJRMEE, 2016, 3, (4), pp 01–06.Google Scholar
Tripathi, A., Manisankar, C. and Verma, S.B., Control of base pressure for a boat-tailed axisymmetric afterbody via base geometry modifications, Aerosp. Sci. Technol. 2015, 45, pp 284–293, doi: 10.1016/j.ast.2015.05.021.CrossRefGoogle Scholar
Damljanović, D. and Isaković, J., T-38 wind-tunnel data quality assurance based on testing of a standard model, J. Aircraft, July–August 2013, 50, (4).CrossRefGoogle Scholar
Liu, H., Yan, C., Zhao, Y. and Qin, Y., Uncertainty and sensitivity analysis of flow parameters on aerodynamics of a hypersonic inlet, Acta Astronaut., 2018, 151, doi: 10.1016/j.actaastro.2018.07.011.CrossRefGoogle Scholar
Jiajan, W., Chue, R.S.M., Nguyen, T. and Yu, S., Optimisation of round bodies for aerodynamics performance and stability at supersonic speeds, Aeronaut. J., 1193, 177, pp 661–685, doi: 10.1017/S0001924000008368 CrossRefGoogle Scholar
Mirzaei, M., Karimi, M.H. and Vaziri, M.A., An investigation of a tactical cargo aircraft aft body drag reduction based on CFD analysis and wind tunnel tests, Aerosp. Sci. Technol., 2012, 23, (1), pp 263–269. doi: 10.1016/j.ast.2011.08.001 CrossRefGoogle Scholar
Morel, T., Effect of base cavities on the aerodynamic drag of an axisymmetric cylinder, Aeronaut. Q., 1979, pp 400–412, doi: 10.1017/S0001925900008611 CrossRefGoogle Scholar
Nicolás-Pérez, F., Velasco, S.F.J., García-Cascales, R.J., Otón-Martínez, A.R., López-Belchí, A., Moratilla, D., Ray, F. and Laso, A., On the accuracy of RANS, DES and LES turbulence models for predicting drag reduction with Base Bleed technology, Aerosp. Sci. Technol., 2017, 67, pp 126–140, doi: 10.1016/j.ast.2017.03.031 CrossRefGoogle Scholar
Platou, A.S., J. Impr. Proj. Boattail Spacecr. Rockets, 1975, 12(12), pp 727732, doi: 10.2514/3.57040 CrossRefGoogle Scholar
Regodić, D., Jevremović, A. and Jerković, D., The prediction of axial aerodynamic coefficient reduction using base bleed, Aerosp. Sci. Technol., 2013, 31, (1), pp 24–29, doi: 10.1016/j.ast.2013.09.001 CrossRefGoogle Scholar
Jamaz, S. and Amin Dali, M., Optimization of artillery projectiles base drag reduction using hot base flow, Therm. Sci., 2019, 23, (1), pp 353–364, doi: 10.2298/TSCI180413210D CrossRefGoogle Scholar
Tanner, M., Reduction of base drag, Prog. Aerosp. Sci., 1975, 16, (4), pp 369384, doi: 10.1016/0376-0421(75)90003-2 CrossRefGoogle Scholar
Mathur, T. and Craig Button, J., Base-bleed experiments with a cylindrical afterbody in supersonic flow, J. Spacecr. Rockets, 1996, 33, (1), pp 30–37, https://doi.org/10.2514/3.55703 CrossRefGoogle Scholar
Van Dam, P.C., Nikfetrat, K., Wong, K. and Vijgen, W.H.P.M., Drag prediction at subsonic and transonic speeds using Euler methods, J. Aircr., 1995, 32, (4), pp 839–845, doi: 10.2514/3.46799 CrossRefGoogle Scholar
Viswanath, R.P., Flow management techniques for base and afterbody drag reduction, Prog. Aerospace Sci., 1996, 32, pp 79129, doi: 10.1016/0376-0421(95)00003-8 CrossRefGoogle Scholar
Viswanath, P.R., and Patil, S.R., Effectiveness of passive devices for axisymmetric base drag reduction at Mach 2, J. Spacecr. Rockets, 27, (3) (1990), pp 234–237, doi: 10.2514/3.26130 CrossRefGoogle Scholar
Viswanath, R.P. and Patil, R.S., Zero-lift drag characteristics of afterbodies with a square base, J. Spacecr. Rockets, 1997, 34, (3), pp 290–293, doi: 10.2514/2.3231 CrossRefGoogle Scholar
Viswanath, R.P., Passive devices for axisymmetric base drag reduction at transonic speeds, J. Aircr., 1988, 25 (3), pp 258–262, doi: 10.2514/3.45586 CrossRefGoogle Scholar
McCoy, L.R., “MC Drag – A Computer Program for Estimating the Drag Coefficient of Projectiles,” US Army Ballistics Research Laboratory, Aberdeen Proving Ground, MD, ARBRL-PR-02293, February 1981.Google Scholar
ESDU 77020, ESDU 78041, ESDU 79022.Google Scholar
Lobb, K.R., Experimental measurement of shock detachment distance on spheres fired in air at hyper velocities, Defense Technical Information Center, 1962.Google Scholar
Shankara, T.S. and Sreekanth, A.K., Shock Standoff distance for a sphere, J. Appl. Phys. 1977, doi: 10.1063/1.323828 CrossRefGoogle Scholar
Saiprakash, M., Senthil Kumar, C., Kadam Sunil, G., Singh Prakash, Rampratap., Shanmugam, V., and Balu, G., Effects of angle of attack and bluntness on heating rate distribution of blunt models at hypersonic speeds, Fluid dynamics, published by Pleiades Publishing, Ltd. 2019, 54 (6), pp 850–862.Google Scholar
CFL3D User’s Manual.Google Scholar
Roache, J.P., Conservatism of the grid convergence index in finite volume computations on steady-state fluid flow and heat transfer, August 27, 2003.CrossRefGoogle Scholar
Batill, S.M., Experimental Uncertainty and Drag Measurements in the National Transonic Facility, NASA Contractor Report 4600 June 1994.Google Scholar