Skip to main content
Log in

Effect of Reynolds number on flow behavior and pressure drag of axisymmetric conical boattails at low speeds

  • Research Article
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

The effect of Reynolds number on flow behaviors and pressure drag around axisymmetric conical boattails was experimentally investigated at low-speed conditions. Four conical boattails with slant angles of 12°, 16°, 20°, and 22° were studied. The Reynolds number ranged from 4.34 × 104 to 8.89 × 104 based on the model diameter. The global-luminescent-oil-film skin-friction measurement was employed to analyze the surface skin-friction topology. Quantitative skin-friction values at the centerline were obtained in this study. The results show that a separation bubble can be formed on boattail surfaces at angles from 12° to 20°. However, at a boattail angle of 22°, flow is fully separated near the boattail shoulder. The integrated afterbody pressure drag indicated that, at angles of 12°, 16°, and 22°, the Reynolds number has very small effect on the afterbody drag, while, at 20° the drag coefficient decrease was relatively large with increasing Reynolds number. We believe that this study provided the first results for a boattail angle of 20° and we observed that the size of the separation bubble decreased as the Reynolds number increased. The effect of the separation bubble on the pressure distribution was also examined in detail.

Graphical abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Bader P, Pscherning M, Sanz W, Woisetschlager J, Heitmeir F, Meile W, Brenn G (2016) Flat-plate boundary layers in accelerated flow. In: Proceedings of ASME Turbo Expo 2016: turbine technical conference and exposition, Seoul, South Korea, GT2016-56044

  • Boutilier MSH, Yarusevych S (2012) Parametric study of separation and transition characteristics over an airfoil at low Reynolds numbers. Exp Fluids 52:1491–1506

    Article  Google Scholar 

  • Braslow AL, Knox EC (1958) Simplified methods for determination of critical height of distributed roughness particles for boundary-layer transition at Mach numbers from 0 to 5. National Advisory Committee for Aeronautics NASA technical Note no 4363

  • Britcher CP, Alcorn CW (1990) Interference-free measurements of the subsonic aerodynamics of slanted-base ogive cylinders. AIAA J 29(4):520–525

    Article  Google Scholar 

  • Brunn A, Nitsche W (2006) Active control of turbulent separated flows over slanted surfaces. Inter J Heat Fluid flow 27:748–755

    Article  Google Scholar 

  • Choi H, Lee J, Park H (2014) Aerodynamics of heavy vehicles. Ann Rev Fluid Mech 46:441–468

    Article  MathSciNet  Google Scholar 

  • Compton WB (1974) Effect on base drag of recessing the bases of conical afterbodies at subsonic and transonic speeds. NASA Technical note D-4821

  • Compton WB, Runckel JF (1970) Jet effects on the boattail axisal force of conical afterbodies at subsonic and transonic speeds

  • Dresz DA (1989) Drag measurements on a laminar-flow body of revolution in the 13-inch magnetic suspension and balance system. NASA Technical paper 2895

  • Driver DM, Drake A (2008) Skin-friction measurements using oil-film interferometry in NASA’s 11-Foot transonic wind tunnel. AIAA J 46(10):2401–2407

    Article  Google Scholar 

  • Genc MS, Karasu I, Acikel HH (2012) An experimental study on aerodynamics of NACA2415 aerofoil at low Re numbers. Exp Thermal Fluids Scien (39):252–264

  • Gillieron P, Chometon F (1999) Modelling of stationary 3D separated air flows around an Ahmed reference model. ESAIM Proc 7:173–182

    Article  Google Scholar 

  • Howard FG, Goodman WL (1984) Axisymmetric bluff-body drag reduction through geometrical modification. J Aircraft 20(6):516–522

    Article  Google Scholar 

  • Ilday O, Acar H, Elbay MK, Atli V (1992) Wakes of three axisymmetric bodies at zero angle of attack. AIAA J 31(6):1152–1154

    Article  Google Scholar 

  • Joseph P, Amandolese X, Aider JL (2012) Drag reduction on the 25 slant angle Ahmed reference body using pulsed jets. Exp Fluids 52:1169–1185

    Article  Google Scholar 

  • Karman Th V (1930) Mechanical similitude and turbulence. NACA Technical Memorandums No 611

  • Kendall A, Koochesfahani M (2008) A method for estimating wall friction in turbulent wall-bounded flows. Exp Fluids 44:773–780

    Article  Google Scholar 

  • Kentfield JAC (1984) Short multi-step afterbody fairings. J Aircraft 21(5):351–352

    Article  Google Scholar 

  • Kim D, Lee H, Yi W, Choi H (2016) A bio-inspired device for drag reduction on a three-dimensional model vehicle. Bioinspir Biormim 11:026004

    Article  Google Scholar 

  • Kremheller A, Fasel H (2010) Water tunnel experiments of three-dimensional separation bubbles on a flat plate. AIAA Paper 2010–4738

  • Kumar R, Viswanath PR, Prabhu A (2002) Mean and fluctuating pressure in boat-tail separated flows at transonic speeds. J Spacecraft 39(3):430–438

    Article  Google Scholar 

  • Lavrukhin GN, Popovich KF (2009) Aerogasdynamics of jet nozzles vol 2 Flow around the base. Fizmatlit, Moscow (in Russian)

    Google Scholar 

  • Lee D, Nonomura T, Anyoji M, Aono H, Oyama A, Asai K, Fujii K (2015) Mechanisms of surface pressure distribution within a laminar separation bubble at different Reynolds numbers. Phys Fluids 27:023602

    Article  Google Scholar 

  • Lee T, Nonomura T, Asai K, Liu T (2018) Linear least-squares method for global luminescent oil film skin-friction field analysis. Rev Sci Instrum 89:065106

    Article  Google Scholar 

  • Liu T, Sullivan JP (1998) Luminescent oil-film skin-friction meter. AIAA J 36(8):1460–1465

    Article  Google Scholar 

  • Liu T, Montefort J, Woodiga S, Merati P, Shen L (2008) Global luminescent oil-film skin-friction meter. AIAA J 46(2):476–485

    Article  Google Scholar 

  • Liu T, Woodiga S, Ma T (2011) Skin-friction topology in a region enclosed by penetrable boundary. Exp Fluids 51:1549–1562

    Article  Google Scholar 

  • Mair WA (1969) Reduction of base drag by boat-tailed afterbodies in low speed flow. Aeronaut Q 20:307–320

    Article  Google Scholar 

  • Mariotti A, Buresti G, Gaggini G, Salvetti MV (2017) Separation control and drag reduction for boat-tailed axisymmetric bodies through contoured transverse grooves. J Fluid Mech 832:514–549

    Article  Google Scholar 

  • McQuilling M, Wolff M, Fonov SD, Crafton J, Sondergaard R (2008) An experimental investigation of a low-pressure turbine blade suction surface using a shear and stress sensitive film. Exp Fluids 4(1):73–88

    Article  Google Scholar 

  • Naughton JW, Sheplak M (1996) Modern developments in shear-stress measurement. Prog Aerosp Scie 38:515–570

    Article  Google Scholar 

  • Olson DA, Katz AW, Naguib AM, Koochesfahani MM, Rizzetta DP, Visbal MR (2013) On the challenges in experimental characterization of flow separation over airfoils at low Reynolds number. Exp Fluids 54:1470–1481

    Article  Google Scholar 

  • Presz WM, Pitkin ET (1974) Flow separation over axisymmetric afterbody models. J Aircraft 11(11):677–682

    Article  Google Scholar 

  • Presz WM, Pitkin ET (1976) Analytical model of axisymmetric afterbody flow separation. J Aircraft 13(7):500–505

    Article  Google Scholar 

  • Reda DC, Wilder MC (2001) Shear-sensitive liquid crystal coating method applied through transparent test surfaces. AIAA J 39(1):195–197

    Article  Google Scholar 

  • Roberts WB (1980) Calculation of laminar separation bubbles and their effect on airfoil performance. AIAA J 18(1):25–31

    Article  Google Scholar 

  • Schloemer HH (1967) Effect of pressure gradient turbulent boundary-layer wall-pressure fluctuations. J Acoust Soc Am 42(1):93–113

    Article  Google Scholar 

  • Selig MS, Deters RW, Villiamson GA (2011) Wind tunnel testing airfoils at low Reynolds numbers. AIAA Paper 2011 – 875

  • Silhan FV, Cubbage JM (1957) Drag of conical and circular-arc boattail afterbodies at Mach numbers from 0.6 to 1.3. NASA Technical Memorandum RM L56K22

  • Silton SI, Dinavahi SPG (2008) Base drag considerations on a 050-caliber Spinning Projectile. AIAA Paper 2008–6739

  • Strachan RK, Knowles K, Lawson NJ (2007) The vortex structure behind an Ahmed reference model in the presence of a moving ground plane. Exp Fluids 42:659–669

    Article  Google Scholar 

  • Suliman MA, Mahmoud OK, Al-Sanabawy MA, Abdel-Hamid OE (2009) Computational investigation of base drag reduction for a projectile at different flight regimes. In 13th international conference on aerospace sciences and aviation technology Cairo Egypt ASAT-13-FM-05 May 24–26

  • Tani I (1964) Low speed flows involving bubble separations. Progr Aerosp Sci 5:70–103

    Article  Google Scholar 

  • Tran TH, Ambo T, Lee T, Chen L, Nonomura T, Asai K (2018) Effect Boattail angles on the flow pattern on an axisymmetric afterbody at low speed. Exp Thermal Fluid Sci 99:324–335

    Article  Google Scholar 

  • Tripathi A, Manisankar C, Verma SB (2015) Control of base pressure for a boat-tailed axisymmetric afterbody via base geometry modifications. Aerosp Sci Technol 45:284–293

    Article  Google Scholar 

  • Venning J, Jacono DL, Burton D, Thompson M, Sheridan J (2015) The effect of aspect ratio on the wake of the Ahmed body. Exp Fluids 56(6):126

    Article  Google Scholar 

  • Vino G, Watkins S, Mousley P, Watmuff J, Prasad S (2005) Flow structures in the near wake of Ahmed model. J Fluids Struct 20:673–695

    Article  Google Scholar 

  • Viswanath PR, Ramesh G, Madhavan KT (2000) Separation control by tangential blowing inside the bubble. Exp Fluids 29:96–102

    Article  Google Scholar 

  • Westerweel J, Geelhoed PF, Lindken R (2004) Single-pixel resolution ensemble correlation for micro-PIV applications. Exp Fluids 37:375–384

    Article  Google Scholar 

  • Wolfe WP, Oberkampf WL (1985) Drag predictions for projectiles and finned bodies in incompressible flow. J Spacecraft 24(1):14–22

    Article  Google Scholar 

  • Woodiga SA (2013) Global skin-friction diagnostics: the Glof technique and measurements of complex separated flows. PhD Thesis Department of Mechanical and Aeronautical Engineering Western Michigan University USA

  • Woodiga SA, Liu T (2009) Skin-friction fields on delta wings. Exp Fluids 47:897–911

    Article  Google Scholar 

  • Zhong H, Woodiga S, Wang P, Shang J, Cui X, Wang J, Liu T (2015) Skin-friction topology of wing-body junction flows. Euro J Mech B/Fluids 53:55–67

    Article  Google Scholar 

  • Zilliac GG (1996) Further development of the fringe-imaging skin-friction technique. NASA TM-110425

Download references

Acknowledgements

This work was supported by Kakenhi Grant 16H04582 from the Japan Society for the Promotion of Science and by Presto Grant JPMJPR1678 from the Japan Science and Technology Agency.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to The Hung Tran.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1. Velocity fields in vertical symmetrical plane and boundary-layer measurement

To examine the characteristics of velocity fields in a vertical symmetrical plane and boundary layer before the boattail shoulder, particle image velocimetry (PIV) tests were performed. The experimental setup for PIV measurement is shown in Fig. 13.

Fig. 13
figure 13

Experimental setup for PIV measurement. a Side view and b back view

In this experiment, a smoke generator (LSG-500S) was placed before the inlet of the wind tunnel to realize a flow-containing particles with mean diameter of 1 µm. The flow-field illumination was provided by a double-pulsed Nd-YLF Laser (LDY-303, 527 nm, Litron Lasers). The thickness of the laser sheet was approximately 1 mm. The time interval between pulses was changed between 4 µs and 8 µs depending on the free-stream velocity. The movement of particles in wind-tunnel tests was captured by a high-speed camera with maximum resolution of 1,280 × 800 pixels (Phantom V611, Vision Research); the camera was equipped with a Nikon objective lens with 100 mm focal length and an extension tube of 36 mm. The sampling rate of the camera was 600 fps and the time length of measurement was approximately 9 s. The camera angle was 40 × 25 mm2 to provide an image resolution of approximately 32 pixels/mm.

Experimental conditions were the same as skin-friction measurement. Velocity profiles of the boundary layer were measured for all boattail angles. At each boattail model, the experiment was performed for four Reynolds numbers from Re = 4.34 × 104 to Re = 8.89 × 104.

To obtain the average velocity field with high spatial resolution and detailed boundary layer profile, a single-pixel resolution ensemble correlation algorithm was applied. This algorithm was developed by Westerweel et al. (2004), and it provides a powerful tool for measuring the velocity profile with high spatial resolution. In the algorithm, the spatial correlation between a single pixel in the first frame and an interrogation window in the second frame is calculated for a set of double frames. This calculation of cross-correlation uses the time evolution of a single-pixel intensity instead of the spatial interrogation window. Therefore, only the averaged velocity fields can be available. To reach a satisfactory convergence of the calculation, a large number of images are required. In this study, 5,400 double-frame images which correspond to the 9 s were processed.

The streamwise velocity and flow field on the symmetric vertical plan at boattail angle of 20° and Reynolds numbers of Re = 4.34 × 104 and Re = 8.89 × 104 are shown in Fig. 14. The black lines show the position of the zero stream velocity. Clearly, a reversed flow region is observed near the shoulder for both Reynolds numbers. However, the separation bubble is much smaller and the streamlines above the bubble are more severely bent with the larger Reynolds number. The length of separation bubble at Re = 4.34 × 104 determined by PIV measurement was longer than the result of GLOF measurement. It is probably because the amount of particles was not enough near the wall surface. In addition, the turbulent intensities are very high near the reattachment position, which could lead to some error in the measurement.

The boundary layer was examined at x/Lb = − 0.12 (2.52 mm before the shoulder) and was averaged by the surrounding 10 pixels (0.31 mm) in the streamwise direction. The mean boundary-layer profile is shown in Fig. 15. The thickness of the boundary layer was determined by the distance from the wall to a point where the flow velocity reached 99% of the free-stream velocity. The boundary-layer thickness δ, displacement thickness δ*, momentum thickness θ, and shape factor H are listed in Table 3. The boundary-layer thickness was slightly changed at different boattail configurations. It should be noted that the shape of boundary layer does not monotonously change with boattail angle. In fact, the velocity profile is less steep at β = 22° due to the fully separated flow with the low-pressure gradient magnitude near the shoulder (Fig. 11d).

Fig. 14
figure 14

Streamwise velocity fields on symmetrical vertical plane at boattail angle of 20°. a Re = 4.34 × 104 and b Re = 8.89 × 104

Fig. 15
figure 15

Boundary-layer profile at x/Lb = − 0.12. a Re = 4.34 × 104 and b β = 20°

Table 3 Characteristics of boundary layer at x/Lb = − 0.12 and Re = 4.34 × 104
Table 4 Boundary-layer thickness at different Reynolds numbers (x/Lb = − 0.12, β = 20°)

Obviously, boattail angle has a very small effect on boundary-layer profile before the shoulder. The shape factor H was around 1.3, which indicates that the boundary layer before the shoulder was fully turbulent.

The boundary-layer thickness at a boattail angle of 20° changed slightly (δ/D = 0.095–0.109) with the Reynolds numbers tested in this study (Table 4). The change of boundary-layer thickness in this study was much smaller than it was in a previous study (δ/D = 0.132–0.395; Mair 1969). As reported by Mair (1969), the boundary-layer thickness has a small effect on the boattail pressure drag. Consequently, the effect of boundary-layer thickness on pressure distribution and pressure drag should be limited in this study.

Appendix 2. Wall shear stress estimation

In the turbulent boundary layer of a smooth wall, there is a region where the mean stream velocity is a logarithmic function of distance from that point to the wall. It is referred as log-law region and was first reported by Karman (1930). The relation between the mean velocity and the distance to the wall in the log-law region can be expressed as follows:

$${u^+}=\frac{1}{\kappa }\ln {z^+}+{C^+},$$
(6)

where \({u^+}=\frac{u}{{{u_\tau }}}\) is dimensionless velocity and \({z^+}=\frac{{z{u_\tau }}}{\nu }\)is the wall coordinate. u is mean velocity, z is distance to the wall, ν is kinematic viscosity, and uτ is shear velocity. The shear velocity is defined as\({u_\tau }=\sqrt {\frac{{{\tau _{\text{w}}}}}{\rho }}\), where τw is wall shear stress and ρ is density of air. The empirical constants in Eq. 6 are called vor Karman constant κ and additive constant C+. In this study, they are assumed as κ = 0.41 and C+ = 5.0 (Kendall and Koochesfahani 2008).

When the boundary layer profile is given, the wall shear stress τw can be calculated to match with the Eq. 6. Consequently, the local skin coefficient Cfx can be calculated as \({C_{{\text{fx}}}}=2{\tau _{\text{w}}}/(\rho {u^2})\). To obtain the quantitative mean skin-friction values at the centerline of boattails, skin-friction coefficient Cfx at x/Lb = − 0.12 was calculated for all the testing cases. An example of logarithmic fitted line at boattail angle of 20° is shown in Fig. 16. At z+ < 20 and Re ≥ 5.92 × 104, slight divergences of data occur. In addition, at z+ > 100, the experimental data begins to deviate from the log-law region. The reason for these trends is from favorable pressure gradient near the shoulder (Schloemer 1967; Bader et al. 2016). However, the experimental data were fitted well in the log-law region and the effect of pressure gradient on the wall flow can be neglected. Experimental results also showed that the skin-friction coefficient slightly decreases with increasing Reynolds number from Re = 4.34 × 104 to Re = 8.89 × 104 (Table 5).

Fig. 16
figure 16

Velocity profiles and fitted line at β = 20°

Table 5 Skin-friction coefficient (x/Lb = − 0.12, β = 20°)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tran, T.H., Ambo, T., Lee, T. et al. Effect of Reynolds number on flow behavior and pressure drag of axisymmetric conical boattails at low speeds. Exp Fluids 60, 36 (2019). https://doi.org/10.1007/s00348-019-2680-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-019-2680-y

Navigation