Skip to main content
Log in

Hypersonic interference heating in the vicinity of surface protuberances

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

Abstract

The understanding of the behaviour of the flow around surface protuberances in hypersonic vehicles is developed and an engineering approach to predict the location and magnitude of the highest heat transfer rates in their vicinity is presented. To this end, an experimental investigation was performed in a hypersonic facility at freestream Mach numbers of 8.2 and 12.3 and Reynolds numbers ranging from Re /m = 3.35 × 106 to Re /m = 9.35 × 106. The effects of protuberance geometry, boundary layer state, freestream Reynolds number and freestream Mach numbers were assessed based on thin-film heat transfer measurements. Further understanding of the flowfield was obtained through oil-dot visualizations and high-speed schlieren videos. The local interference interaction was shown to be strongly 3-D and to be dominated by the incipient separation angle induced by the protuberance. In interactions in which the incoming boundary layer remains unseparated upstream of the protuberance, the highest heating occurs adjacent to the device. In interactions in which the incoming boundary layer is fully separated ahead of the protuberance, the highest heating generally occurs on the surface just upstream of it except for low-deflection protuberances under low Reynolds freestream flow conditions in which case the heat flux to the side is greater.

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
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23

Similar content being viewed by others

Abbreviations

α:

Protuberance deflection angle, degrees

α R :

Coefficient of resistivity, K−1

δ:

Boundary layer thickness with edge at U = 0.99U , m

μ:

Dynamic viscosity, kg m−1 s−1

θ:

Temperature relative to wall, = T aw  − T w

Ø:

Diameter, m

ρ:

Density, kg m−3

\( \left( {\sqrt {\rho c_{p} k} } \right)_{g} \) :

Thermal property of gauges, J K−1 m−2 s−0.5

c p :

Specific heat capacity at constant pressure, J kg−1 K−1

G :

System gain, = 2.06 at 0.1–5 kHz signal frequency

h :

Protuberance height, m

k :

Thermal conductivity, W m−1 K−1

l :

Characteristic linear dimension, m

L :

Separation length ahead of protuberance-plate junction, m

M :

Mach number

Nu:

Nusselt number, = StRePr

p :

Static pressure, Pa

P D :

Drive pressure, Pa

Pr :

Prandtl number, assumed = 1

q :

Heat flux, \( = {{\left( {\sqrt {\rho c_{p} k} } \right)_{g} \overline{{V_{2} }} } \mathord{\left/ {\vphantom {{\left( {\sqrt {\rho c_{p} k} } \right)_{g} \overline{{V_{2} }} } {\left( {\alpha_{R} V_{1} G} \right)}}} \right. \kern-\nulldelimiterspace} {\left( {\alpha_{R} V_{1} G} \right)}} \), W m−2

r:

Recovery factor, assumed = 1

Re :

Reynolds number, \( = {{\rho Ul} \mathord{\left/ {\vphantom {{\rho Ul} \mu }} \right. \kern-\nulldelimiterspace} \mu } \)

Re/m:

Reynolds number per unit length, \( = {{\rho U} \mathord{\left/ {\vphantom {{\rho U} \mu }} \right. \kern-\nulldelimiterspace} \mu } \)

Re L :

Reynolds number based on L, \( = {{\rho_{\infty } U_{\infty } L} \mathord{\left/ {\vphantom {{\rho_{\infty } U_{\infty } L} {\mu_{\infty } }}} \right. \kern-\nulldelimiterspace} {\mu_{\infty } }} \)

Re x,k :

Reynolds number based on x k , \( = {{\rho_{\infty } U_{\infty } x_{k} } \mathord{\left/ {\vphantom {{\rho_{\infty } U_{\infty } x_{k} } {\mu_{\infty } }}} \right. \kern-\nulldelimiterspace} {\mu_{\infty } }} \)

Re y,cl :

Reynolds number based on y cl , \( = {{\rho_{\infty } U_{\infty } y_{cl} } \mathord{\left/ {\vphantom {{\rho_{\infty } U_{\infty } y_{cl} } {\mu_{\infty } }}} \right. \kern-\nulldelimiterspace} {\mu_{\infty } }} \)

St :

Stanton number, \( = {q \mathord{\left/ {\vphantom {q {\left[ {\rho_{\infty } U_{\infty } c_{p} \theta } \right]}}} \right. \kern-\nulldelimiterspace} {\left[ {\rho_{\infty } U_{\infty } c_{p} \theta } \right]}} \)

t :

Time, s

T :

Static temperature, K

U :

Axial velocity, m s−1

V 1 :

Initial voltage across gauge, V

\( \overline{{V_{2} }} \) :

Average output voltage of integrated signal across effective run duration, V s−0.5

W :

Protuberance width, m

x :

Longitudinal distance, m

y :

Lateral distance from centreline, m

z :

Normal distance from flat plate, m

∞:

Freestream conditions

*:

Reference value

aw :

Adiabatic wall

cl :

Relative to centreline

e :

Conditions at boundary layer edge

h :

Based on protuberance height

i :

Incipient conditions

k :

Relative to protuberance leading edge

le :

Relative to flat plate leading edge

o :

Total or stagnation conditions

s :

Shock wave

u :

Undisturbed conditions at protuberance location

w :

Conditions on the wall

x :

Based on local values

References

  • Arrington JP (1968) Heat transfer and pressure distributions due to sinusoidal distortions on a flat plate at Mach 20 in helium. NASA TN D-4907

  • Bertram MH, Wiggs MM (1963) Effect of surface distortions on the heat transfer on a wing at hypersonic speeds. AIAA J 1(6):1313–1319

    Article  Google Scholar 

  • Chapman DR, Kuehn DM, Larson KH (1958) Investigation of separated flows in supersonic and subsonic streams with emphasis on the effects of transition. NACA 1356

  • Coleman HW, Lemmon EC (1973) The prediction of turbulent heat transfer and pressure on a swept leading edge near its intersection with a vehicle. AIAA 73-677

  • Davis SR (2008) Ares I-X flight test—the future begins here. AIAA paper 2008-7806

  • Eckert ERG (1956) Engineering relations for heat transfer and friction in high-velocity laminar and turbulent boundary layer flows over surfaces with constant pressure and temperature. Trans ASME 78(6):1273–1283

    Google Scholar 

  • Elfstrom GM (1971) Turbulent separation in hypersonic flow. I.C. Aero Report 71-16

  • Estruch D (2009) Hypersonic interference aerothermodynamics. PhD thesis, Department of Aerospace Sciences, Cranfield University, Bedford, UK

  • Estruch D, Lawson NJ, MacManus DG, Garry KP, Stollery JL (2008) Measurement of shock wave unsteadiness using a high-speed schlieren system and digital image processing. Rev Sci Instrum 79(12):126108–126108-3

    Google Scholar 

  • Estruch D, Lawson NJ, Garry KP (2009a) Application of optical measurement techniques to supersonic and hypersonic aerospace flows. ASCE J Aero Eng 22(4):383–395

    Article  Google Scholar 

  • Estruch D, Lawson NJ, MacManus DG, Garry KP, Stollery JL (2009b) Schlieren visualization of high-speed flows using a continuous LED light source. J Vis 12(4):289–290

    Article  Google Scholar 

  • Estruch D, MacManus DG, Richardson DP, Lawson NJ, Garry KP, Stollery JL (2009c) Experimental study of unsteadiness in supersonic shock-wave/turbulent boundary-layer interactions with separation. Aero J (in press)

  • Fay JA, Riddell FR (1958) Theory of stagnation point heat transfer in dissociated air. J Aero Sci 25(2):73–85

    MathSciNet  Google Scholar 

  • Guoliang M, Guiqing J (2004) Comprehensive analysis and estimation system on thermal environment, heat protection and thermal structure of spacecraft. Acta Astron 54(5):347–356

    Article  Google Scholar 

  • Hakkinen RJ, Greber I, Trilling L (1959) The interaction on an oblique shock wave with a laminar boundary layer. NASA Mem 2-18-59W

  • Holden MS (1964) Heat transfer in separated flow. PhD thesis, Imperial College, UK

  • Hung FT, Clauss JM (1980) Three-dimensional protuberance interference heating in high-speed flow. AIAA-80-0289

  • Hung F, Patel D (1984) Protuberance interference heating in high-speed flow. In: Proceedings of the 19th thermophysics conference. AIAA-84-1724

  • Jones RA (1964) Heat-transfer and pressure investigation of a fin-plate interference model at a Mach number of 6. NASA TN D-2028

  • Kuehn DM (1959) Experimental investigation of the pressure rise required for the incipient separation of turbulent boundary layers in two-dimensional supersonic flow. NASA Memo 1-21-59W

  • Lakshmanan B, Tiwari SN, Hussaini MY (1988) Control of supersonic intersection flowfields through filleting and sweep. In: Proceedings of the 1st national fluid dynamics congress, Cincinanati, Ohio, Part 2, pp 746–759

  • Meador WE, Smart MK (2005) Reference enthalpy method developed from solutions of the boundary-layer equations. AIAA J 43(1):135–139

    Article  Google Scholar 

  • Needham DA (1963) Progress report on the Imperial College hypersonic gun tunnel. Report 118

  • Needham DA, Stollery JL (1966a) Hypersonic studies of incipient separation and separated flows. Aeronautical Res Council paper ARC 27752 January 1966

  • Needham DA, Stollery JL (1966b) Boundary layer separation in hypersonic flow. AIAA 66-455

  • Nestler DE (1985) The effects of surface discontinuities on convective heat transfer in hypersonic flow. AIAA Paper 85-0971

  • Neumann RD, Hayes JR (1981) Protuberance heating at high Mach numbers, a critical review and extension of the database. AIAA 81-0420

  • Olivier H (2009) Thin film gauges and coaxial thermocouples for measuring transient temperatures. SWL, RWTH

  • Price EA, Stallings RL (1967) Investigation of turbulent separated flows in the vicinity of fin-type protuberances at supersonic Mach numbers. NASA TN D-3804

  • Prince SA (1995) Hypersonic turbulent interaction phenomena and control flap effectiveness. MSc thesis, Cranfield University, Bedford, UK

  • Rogers GFC, Mayhew YR (1980) Engineering thermodynamics work and heat transfer. Longman Scientific & Technical, John Wiley, New York

    Google Scholar 

  • Schultz DL, Jones TV (1973) Heat-transfer measurements in short-duration hypersonic facilities. AGARD-AG-165

  • Simmons JM (1995) Measurement techniques in high-enthalpy hypersonic facilities. Exp Therm Fluid Sci 10(4):454–469

    Article  Google Scholar 

  • Stainback PC (1969) Effect of unit Reynolds number, nose bluntness, angle of attack, and roughness on transition on a 5° half-angle cone at Mach 8. NASA TN D-4961

  • Sterret JR, Emery JC (1960) Extension of boundary layer separation criteria to a Mach number of 6.5 by utilizing flat plates with forward facing steps. NASA TN D-618

  • Sterret JR, Morrissete EL, Whitehead AH Jr, Hicks RM (1967) Transition fixing for hypersonic flow. NASA TN D-4129

  • Stollery JL, MacManus DG, Estruch D (2008) Hypersonic research and its application to a real vehicle. Seminar presentation

  • Vannahme M (1994) Roughness effects on flap control effectiveness at hypersonic speeds. MSc thesis, Cranfield University, Bedford, UK

  • Wang SF, Ren ZY, Wang Y (1998) Effects of Mach number on turbulent separation behaviours induced by blunt fin. Exp Fluids 25:347–351

    Article  Google Scholar 

  • Wang XY, Yuko J, Motil B (2009) Ascent heating thermal analysis on the spacecraft adaptor (SA) fairings and the interface with the crew launch vehicle (CLV). NASA TM-2009-215474

  • Weinstein LM (1970) Effects of two-dimensional sinusoidal waves on heat transfer and pressure over a flat plate at Mach 8. NASA TN D-5937

  • White FM (2006) Viscous fluid flow, 3rd edn. McGraw Hill, New York

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to D. Estruch.

Appendix 1: Correlation of peak heat flux in fully separated interactions

Appendix 1: Correlation of peak heat flux in fully separated interactions

Buckingham-Pi analysis is used to derive a correlation of the hot spot magnitude in fully separated interactions. This approach is commonly used in the interpretation of experimental data and has previously led to the development of predictive methods for heat transfer in attached flows (Rogers and Mayhew 1980). The main variables responsible for the heat flux need to be determined. It is well established that the main parameters to be considered are viscosity μ, density ρ, thermal conductivity k, specific heat c p , temperature relative to wall θ, fluid velocity U and a characteristic linear dimension l. As expected from the previous literature (Sect. 1), the characteristic linear dimension l which has a dominant effect in fully separated interactions is the height of the protuberance (h) whereas the effect of width is known to be negligible. The effect of boundary layer thickness is also considered negligible throughout the present experimental study as shown by the similitude between the peak heating at laminar and turbulent conditions (Sect. 3.5). Based on a dimensional analysis considering mass, length, time, thermal energy and temperature as the five fundamental units and grouping them in terms of two main non-dimensional groups (Re and Pr), the relation in Eq. 9 is derived, where a and b are exponents, C is a constant and Re h is Reynolds number based on protuberance height. Since in experimental measurements of this type accurate knowledge of the Prandtl number Pr is generally not feasible, common predictive approaches developed to date consider a constant Prandtl number. For this reason, a common simplification is to assume Pr = 1 and Eq. 9 thus reduces to Eq. 10. For more details on the derivation of this relation, refer to Rogers and Mayhew (1980). The experimental results obtained at a freestream Mach number of M  = 8.2 and at Reynolds numbers of Re /m = 6.57 × 106, 8.06 × 106 and 9.35 × 106 are used to obtain a correlation of Nu h with Re h . The highest heat flux in terms of Nu h are plotted against Re h for the α = 45°, 60° and 90° cases showing a correlation of the measurements with a slope of a = 1.6. Since Pr is considered equal to 1, Nu h  = St max Re h and a = 1.6 are introduced in Eq. 10 to obtain the relation in Eq. 11 (Fig. 24). The recovery factor is also considered to be r = 1.

$$ Nu_{h} = C\text{Re}_{h}^{a} \Pr{^b} $$
(9)
$$ Nu_{h} = C\text{Re}_{h}^{a} $$
(10)
$$ St_{\max } \text{Re}_{h}^{ - 0.6} = C $$
(11)
Fig. 24
figure 24

Correlation of St max Re −0.6 h with protuberance deflection angle α at different freestream Reynolds numbers in logarithmic scale

The term St max Re −0.6 h is constant for different Reynolds numbers but it increases with higher deflection angles (Fig. 24). The effect of α is introduced by considering the deflection experienced by the incoming flow before it reattaches to the surface ahead of the protuberance. This is expressed as shown in Eq. 12. Figure 25 shows the constant trend obtained for the M  = 8.2 tests at all the different conditions while the same does not apply in the cases where the hot spot takes place to the side of the protuberance—i.e. in unseparated or weak separated interactions. This is because the side heat flux is caused by corner effects (Sect. 4.5). The effect of Mach number is subsequently investigated through a similar correlation between the M  = 8.2 and the M  = 12.3 measurements. Its effect is found to be proportional to the square root of Mach number as shown in Eq. 13. A suitable constant C for the final relation in Eq. 13 is found to be C = 5.2 × 10−5 as shown in Fig. 19.

$$ {{St_{\max } Re_{h}^{ - 0.6} } \mathord{\left/ {\vphantom {{St_{\max } Re_{h}^{ - 0.6} } {\left( {1 - \cos \alpha } \right) = C}}} \right. \kern-\nulldelimiterspace} {\left( {1 - \cos \alpha } \right) = C}} $$
(12)
$$ {{St_{\max } Re_{h}^{ - 0.6} M^{0.5} } \mathord{\left/ {\vphantom {{St_{\max } Re_{h}^{ - 0.6} M^{0.5} } {\left( {1 - \cos \alpha } \right) = C}}} \right. \kern-\nulldelimiterspace} {\left( {1 - \cos \alpha } \right) = C}} $$
(13)
Fig. 25
figure 25

Correlation of St max Re −0.6 h /(1 − cosα) with protuberance deflection angle α at different freestream Reynolds numbers

Rights and permissions

Reprints and permissions

About this article

Cite this article

Estruch, D., MacManus, D.G., Stollery, J.L. et al. Hypersonic interference heating in the vicinity of surface protuberances. Exp Fluids 49, 683–699 (2010). https://doi.org/10.1007/s00348-010-0844-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00348-010-0844-x

Keywords

Navigation