See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/257378615 Numerical heat transfer study around a spiked blunt-nose body at Mach 6 Article in Heat and Mass Transfer · April 2013 DOI: 10.1007/s00231-012-1095-6 CITATIONS READS 9 226 1 author: R. C. Mehta Noorul Islam University 102 PUBLICATIONS 569 CITATIONS SEE PROFILE Some of the authors of this publication are also working on these related projects: Reentry aerodynamics, Venting, Inverse heat conduction problem View project Drag Reduction for Spike Attached to Blunt-Nosed Body at Mach 6, Experimental investigation on spiked body in hypersonic flow View project All content following this page was uploaded by R. C. Mehta on 23 September 2015. The user has requested enhancement of the downloaded file. Numerical heat transfer study around a spiked blunt-nose body at Mach 6 R. C. Mehta Heat and Mass Transfer Wärme- und Stoffübertragung ISSN 0947-7411 Heat Mass Transfer DOI 10.1007/s00231-012-1095-6 1 23 Your article is protected by copyright and all rights are held exclusively by SpringerVerlag Berlin Heidelberg. This e-offprint is for personal use only and shall not be selfarchived in electronic repositories. If you wish to self-archive your work, please use the accepted author’s version for posting to your own website or your institution’s repository. You may further deposit the accepted author’s version on a funder’s repository at a funder’s request, provided it is not made publicly available until 12 months after publication. 1 23 Author's personal copy Heat Mass Transfer DOI 10.1007/s00231-012-1095-6 ORIGINAL Numerical heat transfer study around a spiked blunt-nose body at Mach 6 R. C. Mehta Received: 3 September 2011 / Accepted: 27 November 2012 Ó Springer-Verlag Berlin Heidelberg 2012 Abstract An aerospike attached to a blunt body significantly alters its flowfield and influences aerodynamic drag at high speeds. The dynamic pressure in the recirculation area is highly reduced and this leads to the decrease in the aerodynamic drag. Consequently, the geometry of the aerospike has to be simulated in order to obtain a large conical recirculation region in front of the blunt body to get beneficial drag reduction. Axisymmetric compressible Navier–Stokes equations are solved using a finite volume discretization in conjunction with a multistage Runge– Kutta time stepping scheme. The effect of the various types of aerospike configurations on the reduction of aerodynamic drag is evaluated numerically at a length to diameter ratio of 0.5, at Mach 6 and at a zero angle of incidence. The computed density contours are showing satisfactory agreement with the schlieren pictures. The calculated pressure distribution on the blunt body compares well with the measured pressure data on the blunt body. Flowfield features such as formation of shock waves, separation region and reattachment point are examined for the flatdisc spike and on the hemispherical disc spike attached to the blunt body. One of the critical heating areas is at the stagnation point of a blunt body, where the incoming hypersonic flow is brought to rest by a normal shock and R. C. Mehta (&) School of Mechanical and Aerospace Engineering, Nanyang Technological University, Singapore 639798, Singapore e-mail: atulm@md4.vsnl.net.in Present Address: R. C. Mehta Department of Aeronautical Engineering, Noorul Islam University, Kumaracoil 629180, India adiabatic compression. Therefore, the problem of computing the heat transfer rate near the stagnation point needs a solution of the entire flowfield from the shock to the spike body. The shock distance ahead of the hemisphere and the flat-disc is compared with the analytical solution and a good agreement is found between them. The influence of the shock wave generated from the spike is used to analyze the pressure distribution, the coefficient of skin friction and the wall heat flux facing the spike surface to the flow direction. List of symbols Cf Skin friction coefficient CD Aerodynamic drag Cp Pressure coefficient D Cylinder diameter e Specific energy h Enthalpy H Source vector F, G Flux vector k Thermal conductivity L Length of the spike M Mach number q Wall heat flux Pr Prandtl number p Pressure pa Ambient pressure s Distance along the surface of the spike T Temperature t Time u, v Velocity components W Conservative vector x, r Coordinate direction c Ratio of specific heats l Viscosity q Density 123 Author's personal copy Heat Mass Transfer r D Stress vector Shock stand-off distance Subscripts e Boundary layer edge F Flat-disc o Stagnation S Hemispherical disc w Wall ? Freestream condition Fig. 1 Schematic sketch for flowfield over spiked blunt body 1 Introduction A high-speed flow over a blunt body generates a bow shock wave in front of it, which causes a rather high surface pressure and, as a result, high aerodynamic drag. The surface pressure on the blunt body can be substantially reduced if a conical shock wave is generated by attaching a forward-facing spike. Thus, the introduction of the spike decreases the aerodynamic drag. The spike produces a region of recirculating separated flow that shields the bluntnosed body from the incoming flow. It is advantageous to have a vehicle with a low drag coefficient in order to minimize the thrust required from the propulsive system during the supersonic and hypersonic regime. Many researchers have investigated flowfield simulations over the blunt body attached with forward facing aero-spike. A typical flow over a spike attached to a blunt body is based on experimental investigations. A schematic of the flowfield over the aerodisc spiked blunt body at zero angle of attack is shown in Fig. 1. The flow field around a spiked blunt body appears to be very complex and contains a number of interesting flow phenomena and characteristics, which are yet to be investigated. The recirculating region is formed around the root of the spike up to the reattachment point of the flow at the shoulder of the hemispherical body. Due to the recirculating region, the pressure at the stagnation region of the blunt body will decrease. Flow behind the conical shock wave separates on the spike and creates a conical shaped recirculation zone appears in the vicinity of the stagnation region. Due to the formation of the flow recirculation, the pressure and wall heat flux are reduced in the forward facing region of the blunt body. However, the reattachment of the shear layer on the shoulder of the hemispherical body increases the local heat flux and pressure. The reattachment shock is moved downstream as depicted in the diagram, which is a function of the geometrical parameter of the spike and the shape of the spiked nose body. The prime focus has been with regard to drag characteristics where effects due to spike length, spike head geometry, forward body geometry and relative spike 123 diameter have been explored. Most of the experimental studies on spiked bodies are carried out to get effects on aerodynamic heating [1–3] and the surface pressure distributions [4–6] at supersonic and hypersonic Mach numbers [7–10]. Kubota [11] has experimentally investigated the overall characteristics of the spiked blunt body configuration at hypersonic Mach numbers. Crawford [12] experimentally investigated the effects of the spike length on the nature of the flow field for a freestream Mach number 6.8 and Reynolds number 0.12 9 106 - 1.5 9 106 based on the cylinder diameter. The spike geometry drastically influences the aerodynamic drag of the blunted body at high-speeds. Motoyama et al. [13] have experimentally investigated the aerodynamic and heat transfer characteristics of conical, hemispherical, flat-faced aerospikes, and hemispherical and flat-faced discs attached to the aerospike for a freestream Mach number 7, freestream Reynolds number 4 9 105/m, for L/D = 0.5 and 1.0, and angle-of-attack 0 to 8°. They found that the aerodisc spike (L/D = 1.0 and aerodisc diameter of 10 mm) has a superior drag reduction capability as compared to the other aerospikes. Yamauchi et al. [14] have numerically investigated the flowfield around a spiked blunt body at freestream Mach numbers of 2.01, 4.14 and 6.80 for different ratio of L/D. Shoemaker [15], Fujita and Kubota [16], Asif et al. [17] used a numerical approach to solve the compressible Navier–Stokes equations. Their investigation shows that the reattachment point can be moved backward or removed, which depends on the spike length or the nose configuration. However, because of the reattachment of the shear layer on the shoulder of the hemispherical body, the pressure near that point becomes large. Milicev et al. [18] have experimentally investigated the influence of four different types of spikes attached to a hemisphere-cylinder body at Mach number 1.89, Reynolds number 0.38 9 106 based on the cylinder diameter, and at an angle-of-attack of 2°. They observed in their experimental studies that a reliable estimation of the aerodynamic effects of the spike can be made in conjunction with flow visualization techniques. Numerical simulations [19–21] have been carried out to obtain the comparative studies of Author's personal copy Heat Mass Transfer the flowfield over the spike. Peak heating for the reattachment of the separated flow on a spiked blunt body is numerically studied for freestream Mach numbers 2.01, 4.15 and 6.80. Recently, axisymmetric numerical simulations [22] have been performed for different types of spikes attached to a blunt nose cone at Mach numbers of 5.0, 7.0 and 10.0 employing CFD-FASTRAN flow solver. It is found in the numerical simulations that a spike with an aerodisc outperformed the plain pointed aerospike in terms of drag and aerodynamic heating reduction [23, 24]. Heat transfer rate on spike surface depends strongly upon the stagnation point velocity gradient which is difficult to estimate for the flat-nose spike. A flat nose spike reduces stagnation point heat transfer rate [25] by up to 50 % as compared to the hemispherical nose. The present paper presents a numerical simulation of the flowfield over a hemispherical and a flat-disc aerospike attached to a blunt body. The focus of the present numerical analysis is to investigate the mechanism of drag reduction and flowfield pattern at a freestream Mach number 6. However, the aerodisc and aeroflat spiked blunt body include phenomena of flowfield facing the flow direction that have not been described previously. The flowfield features captured by the density, Mach and pressure contours are used to understand the mechanism of the drag reduction. The influence of the shock wave generated from the spike is also studied to understand the pressure distribution, the coefficient of skin friction and wall heat transfer over the spike surface. The pressure ratio and the heat transfer rate at the stagnation point are computed and compared with the analytical solution. The numerical results of the research on a hemispherical and a flat-disc spike at L/D ratio of 0.5 compared with the available experimental data are found to be in good agreement. H ¼ ½0; 0; rþ ; 0T where the superscript T represents transpose matrix. The viscous and heat flux terms in the above equations are 2 ou rxx ¼ lr U þ 2l 3 ox 2 ov rrr ¼ lr U þ 2l 3 or ou ov þ rxr ¼ l or ox 2 v rþ ¼ p lr U þ 2l 3 r ou ov v þ þ rU ¼ ox or r oT qx ¼ k ox oT qr ¼ k or where CP is specific heat at constant pressure, U is the mean stream velocity. l is calculated according to Sutherland’s law. The flow is assumed to be laminar, which is also consistent with Bogdonoff and Vas [1], Yamauchi et al. [14], Fujita and Kubota [16] and Ahmed and Qin [23]. T is related to p and q by the perfect gas equation of state as 1 p ¼ ðc 1Þq e u2 þ v2 ð3Þ 2 The ratio of the specific heats is assumed constant and is equal to 1.4. 2 Governing fluid dynamics equations 3 Numerical scheme A numerical simulation is carried out to solve the unsteady, compressible, axisymmetric Navier–Stokes equations over a forward facing spike attached to a blunt body. The governing equations can be written in the following strong conservation form as 3.1 Spatial discretization oW oF 1 oðrGÞ H þ þ ¼ ot ox r or r where ð1Þ To facilitate the spatial discretization in the numerical scheme, the time dependent axisymmetric compressible Navier–Stokes equation (1) can be written in the integral form over a finite volume as Z Z Z o WdX þ ðFdr GdxÞ ¼ HdX ð4Þ ot X T W ¼ ½q; qu; qv; qe F ¼ qu; qu2 þ p rxx ; quv rxr ; ðqe þ pÞu urxx vrxr þ qx T G ¼ qv; quv rxr ; qv2 þ p rrr ; ðqe þ pÞv T vrrr urxr þ qy ð2Þ C X where X is the computational domain, C is the boundary domain. The contour integration around the boundary of the cell is taken in the anticlockwise sense. The numerical technique uses a finite volume method on a structured nonoverlapping quadrilateral mesh. The spatial and temporal terms are decoupled using the method of lines. The flux vector is divided into the inviscid and viscous components. 123 Author's personal copy Heat Mass Transfer A cell-centered scheme is used to store the flow variables. The discretization of inviscid fluxes is performed using the cell average scheme. The discretization of viscous fluxes involves the centre difference scheme of the second order. Thus, the discretized solution to the governing equations results in a set of volume-averaged state variables of mass, momentum, and energy which are in balance with their area-averaged fluxes (inviscid and viscous) across the cell faces [26]. The finite volume code constructed in this manner reduces to a central difference scheme and is second-order accurate in space provided that the mesh is smooth enough [27]. The cell-centered spatial discretization scheme is non-dissipative; therefore, artificial dissipation terms are included as a blend of a Laplacian and biharmonic operator in a manner analogous to the second and fourth difference. The artificial dissipation term [28] was added explicitly to prevent numerical oscillations near the shock waves to damp high-frequency modes. 3.2 Time marching scheme Temporal integration was performed using the three-stage time-stepping scheme of Jameson et al. [28] based on the Runge–Kutta method. The artificial dissipation is evaluated only in the first stage. The details of this flowfield technique are further described in Ref. [29]. The solver uses a time-marching procedure to compute the flow. The flow is defined to be steady because the flowfield is converging to a steady state. The steady options use local time stepping, which leads to a faster convergence to the steady-state flow field. 3.3 Initial and boundary conditions Conditions corresponding to a freestream Mach number 6.0 were given as initial conditions. All the variables were extrapolated at the outer boundary, and the no-slip wall condition was used to implement the boundary conditions. An isothermal wall condition was prescribed for the surface of the model, that is, a wall temperature of 300 K. The symmetric condition was applied on the centerline. 4 Model and grid generation 4.1 Spike geometry The dimensions of the spiked blunt body considered in the present analysis are depicted in Fig. 2. The model is axisymmetric, the main body has a hemispherical-cylinder 123 Fig. 2 Dimensions of the spiked blunt body. a Flat-disc aerospike, b hemispherical disc aerospike nose, and diameter D is 4.0 9 10-2 m. The spike consists of an aerodisc part and a cylindrical part. The hemispherical cap of the spike has diameter 0.2 D attached to a stem of diameter of 0.1 D and L = 1.5 D as shown in Fig. 2a. The flat-disc has an identical dimension of the hemispherical spike as shown in Fig. 2b. 4.2 Computational grid One of the controlling factors for the numerical simulation is the proper grid arrangement. The following procedure is used to generate grid in the computational region of the spiked blunt body. The computational domain is divided into a number of non-overlapping zones. The mesh points are generated in each zone using finite element methods [30] in conjunction with the homotopy scheme [31]. The spiked blunt nosed body is defined by a number of grid points in the cylindrical coordinate system. Using these surface points as the reference nodes, the normal coordinate is then described by the exponentially stretched grid points extending outwards up to an outer computational boundary. Grid independence tests [32] were carried out, taking into consideration the effect of the computational domain, the stretching factor to control the grid intensity near the wall, and the number of grid points in the axial and normal directions. The outer boundary of the computational domain was varied from 2.5 to 3.0 times the cylinder diameter D and the grid-stretching factor in the radial direction is varied from 1.5 to 5. These stretched grids were generated in an orderly manner. To verify the chosen grid delivers an accurate solution, the number of grid cells was increased until a steady state solution Author's personal copy Heat Mass Transfer Fig. 3 Enlarged view of computational grid. a Hemispherical disc aerospike, b flat-disc aerospike occurred, that is, the resulting axial force on the investigated shape did not change anymore. Several test runs were made with a total doubled grid cell number. Therefore, the grid was highly refined in both directions. Grids were chosen with the number of grid points in the axial direction ranging from 187 for the shortest blunt spike to up to 220 for the longest spike configuration, and the number in the radial direction ranging from 52 to 82. The present numerical analysis was performed on 187 9 62 grid points. The downstream boundary of the computational domain is maintained at 4–6 times the cylinder diameter. This grid arrangement was found to give a relative difference of about ±1.5 % for the drag coefficient. A convergence criterion of 10-5 is used, based on the difference in the density values at any grid point between two successive iterations. The minimum spacing for the fine mesh is dependent upon the Reynolds number. The finer mesh near the wall helps to resolve the viscous effects. The coarse-mesh helps reducing the computer time. A close-up view of the computational grid over the hemispherical and the flat-faced aerospike is shown in Fig. 3. The structured grid generation and the mono-block are suitable to accommodate the spike shape. As seen in the figures, these types of grid use quadrilateral cells in 2-D in the computational array. The quadrilateral cells, which are very efficient at filling space, support a high amount of skew and stretching before the solution will be significantly affected. Additionally, the grid can be aligned with the flow, thereby yielding greater accuracy within the solver. Several grid arrangements are considered to verify the grid independency. The numerical results are validated with the available experimental data in the next section. 5 Results and discussion 5.1 Flowfield visualization and characteristics Characteristic features of the flowfield around the hemispherical and the flat-disc aerospike attached to blunt body at high speeds were investigated with the help of velocity vector, density, pressure and Mach contours plots. Figure 4 depicts the velocity vector plots over the hemispherical aerospike and the flat-disc aerospike for L/D = 0.5 at M? = 6 and Reynolds number 0.979 9 107 based on the cylinder diameter. The spiked body is completely enveloped within the recirculation region. The bow shock interacts with the reattachment shock generated by the blunt body. The interaction of the shock wave produced by the hemispherical aerospike differs significantly with the flat disc spike. The flow separation on the spike and recirculation zone formed on the blunt body cap depends on the shape of the spike. The contour plots explain the cause of the drag reduction due to increase of the separation region over the spike. Figure 5 shows the enlarged view of the nondimensional pressure contour plots over the hemispherical and the flat disc spike attached to the blunt body. It can be observed from the pressure contour plots low and high pressure region over the spike. Figure 6 depicts the close-up view of the Mach contour plots over the spiked-blunt body. The contour plots will help to locate the high Mach zones. Figure 7 displays the zoomed region of the vector plot on the hemispherical and flat-disc spike configurations. The bow shock wave follows the aerospike contour and the fore body is entirely subsonic up to the corner tangency point of the flat-faced and the hemispherical aerospike where the sonic line is located. The 123 Author's personal copy Heat Mass Transfer Fig. 4 Velocity vector plot over spiked-blunt body. a Hemispherical disc aerospike, b flat-disc aerospike effects of the subsonic flow on the hemispherical and the flat disc bodies have been investigated by Truitt [33]. 5.2 Shock stand-off distance The computed density contour plots with schlieren pictures [34] are shown in Fig. 8. The separated shear layer and the recompression shock from the reattachment point on the shoulder of the hemispherical body are visible in the contour plot. The shock wave in front of the spike cap will reduce the drag as compared to the case without the spike. In the fore region of the aerodisc, the fluid decelerates through the bow shock wave. At the shoulder of the aerodisc or hemispherical cap, the flow turns and expands rapidly, the boundary layer detaches, forming a free shear layer that separates the inner recirculating flow region behind the base from the outer flowfield. The corner expansion over aerodisc process is a Prandtl– Meyer distorted by the presence of the approaching boundary layer. The computed flowfields show good agreement with the schlieren photographs. The flowfields are very different between the hemispherical and the flat- 123 faced spike as seen in the contour plots and are presented using schematic sketch in Fig. 8. For the case of a flat-nosed spike flying at hypersonic speeds, a detached bow wave is formed in front of the nose which is practically normal to the body axis [33]. Since the flow behind the normal shock is always subsonic, simple continuity considerations show that the shock-detachment distance and stagnation-velocity gradient are essentially a function of the density ratio across the shock. The flow behind the shock wave is subsonic, the shock is no longer independent of the far-downstream conditions. A change of the spike shape (geometry) in the subsonic region affects the complete flow field up to the shock. Figure 8 also represent schematic shock stand-off distance and the location of the sonic line. The shockdetachment distance becomes smaller with increasing density ratio. Probstein [35] gives an expression for the shock detachment distance DF (Fig. 8a) with diameter of the flat-disc DS ratio as rffiffiffiffiffiffiffi DF q1 ¼ 2:8 ð5Þ DS q0 Author's personal copy Heat Mass Transfer Fig. 5 Enlarged view of pressure contours. a Hemispherical disc aerospike, b flat-disc aerospike Fig. 6 Enlarged view of Mach contours. a Hemispherical disc aerospike, b flat-disc aerospike where the density ratio across the normal shock [36] is e¼ 2 1ÞM1 q1 ðc þ2 ¼ 2 ðc þ 1ÞM1 q0 ð6Þ The ratio of shock stand-off distance DS with hemispherical spike of diameter, DS (Fig. 8b) is DS 2e qffiffiffi ¼ DS 1 þ 8e 3 ð7Þ The values of DF/DS and DS/DS are found to be 0.1898 and 0.1109, respectively. The numerical values of the ratio of shock stand-off to spike cap diameter are calculated from the velocity vector and pressure contour plot and they are 0.19 and 0.11 which show good agreement with the analytical values. The spherical spike shows the greatest change in velocity gradient as compared to the flat disc. The shock wave stands in front of the blunt body and forms a region of subsonic flow around the stagnation region. 123 Author's personal copy Heat Mass Transfer Fig. 7 Close-up view of velocity vector plot over spikedblunt body. a Hemispherical disc aerospike, b flat-disc aerospike 5.3 Surface pressure distribution The pressure coefficient [Cp = 2{(p/p?) - 1}/cM2?] variation on the blunt-nosed body with different spike configurations are shown in Fig. 9. The computed pressure 123 distributions compare well with the experimental pressure measurement data [34]. The x/R = 0 location is the spike/ nose tip junction, where R is radius of the cylinder. The location of the maximum pressure on the surface of the spiked blunt body is at a body angle of about 40°. Author's personal copy Heat Mass Transfer Fig. 8 Flowfield for spiked blunt body and schematic shock location. a Density contour plot and schlieren picture for hemispherical disc aerospike, b density contour plot and schlieren picture for flat-disc aerospike The location corresponds to the reattachment point. It is interesting to note that the maximum pressure is found on the same location on the blunt body. The low pressure ahead of the blunt body shows the cause of the reduction in the aerodynamic drag. Figures 10, 11, 12 depict the variation of non-dimensional pressure p/pa, skin friction coefficient and heat flux over the spike surface facing the flow direction along the spike. The s/DS = 0 is the location of the stagnation point. DS is diameter of the spike as shown in Fig. 8. The sonic line position on the flat and the hemispherical spike disc are calculated using the computed pressure, isentropic and normal shock relations [33] and are found to be 0.095 and 0.1, respectively. It shows that the sonic line appears on the shoulder of the hemispherical spike. The pressure ratio p/pa on the stagnation point is 48.84 and 38.23 for the flat-disc and the hemispherical disc spike, respectively. The pressure ratio across the normal shock [37] is 41.83. It shows 123 Author's personal copy Heat Mass Transfer 2.00 1.5x10 7 1.0x10 7 0.5x10 7 Experimental flat disk spike Experimental hemispherical spike Numerical hemispherical spike Numerical flat-disk spike 1.75 2 qw (W/m ) 1.50 Cp 1.25 1.00 0.75 Flat-disc spike Hemispherical spike 0.50 0.25 0 0 0 0.5 1.0 1.5 2.0 2.5 3.0 0 0.02 0.04 0.06 0.08 0.10 s/DS 3.5 x/R Fig. 12 Heat flux variation on the spike facing the flow direction Fig. 9 Pressure distribution along the spike attached to blunt nosed body Table 1 Calculated aerodynamic drag for L/D = 0.5 50 Spike geometry CD Hemispherical spike 0.576 Flat face spike 0.458 40 Flat-disc spike Hemispherical spike p/pa 30 20 10 0 0 0.02 0.04 0.06 0.08 0.10 s/DS Fig. 10 Pressure distribution on the spike facing the flow direction -4 1.25x10 -4 Flat-disc spike Hemispherical spike 1.00x10 -4 Cf 0.75x10 -4 0.50x10 -4 0.25x10 0 0 0.02 0.04 0.06 0.08 0.10 s/DS Fig. 11 Skin friction variation on the spike facing the flow direction that the percentage pressure ratio difference of the order of 16.77 and -8.59 % for the flat-disc and the hemispherical disc spike, respectively. The difference is attributed to the 123 finite compressibility in the shock and the spike surface. The aerodynamic drag is given in Table 1. Newtonian flow [33] and similarity method [38] use the given shock radius, freestream velocity, and density ratio across the shock to determine the body surface where the normal component of the velocity vanishes. These methods do not take into consideration the finite compressibility that exists between the shock wave and the spike surface. The numerical analysis is able to take into consideration the compressibility effects in the subsonic region. 5.4 Wall heat flux The inviscid flow field in the vicinity of the stagnation point is described in a fluid dynamics sense as the conversion of a unidirectional high velocity stream by a normal shock wave into a high temperature subsonic layer, which is taken to be inviscid and incompressible [35]. At the stagnation point of a blunt body, the incoming hypersonic flow brought to rest by a normal shock and adiabatic compression. The heat transfer rate is directly proportional to the enthalpy gradient at the wall and square root of the tangential velocity gradient at the edge of the boundary layer. The inviscid flow field in the vicinity of the stagnation point is described as the conversion of the unidirectional, high velocity stream by a normal shock wave into a high temperature subsonic layer. The enthalpy gradient is determined by the shape of the velocity profile in the boundary layer and by the variation of the air properties with temperature. Heat flux at the stagnation point can be calculated using the following expression of Fay and Riddell [39] Author's personal copy Heat Mass Transfer s ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffi qw lw ðqe le K Þ ð he h w Þ qe le 0:6 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qw ¼ 0:763 Pr ð8Þ The value of stagnation point velocity gradient K is taken as 0.3 [25]. The hemispherical spike shows the greatest changes in the velocity gradient as compared to the flat disc spike. The magnitude of the stagnation-velocity gradient indicates the maximum heat transfer over the hemispherical spike. The computed values of the stagnation point heat flux are 0. 556 9 107 and 1.45 9 107 W/m2 for the flatdisc and the hemispherical disc spike, respectively. The stagnation point calculated using Eq. (8) is 0. 831 9 107 and 1.55 9 107 W/m2 for the flat-disc and the hemispherical disc spike, respectively. The discrepancy is due to the value of K. The value of K is difficult to calculate for the flat disc spike [25]. The results show reasonably good agreement between them. The close-up view of velocity vector plot over the spiked-blunt body gives a comparative velocity gradient in the s direction of the spike. The velocity vector is turning in the stream line fashion on the hemispherical spike whereas on the flat-disk spike it appears the flow is impinging. 6 Conclusion The flowfield around a forward facing hemispherical and flat disc spike attached to blunt body has been numerically simulated at a freestream Mach number 6, at length to diameter ratio of 0.5 and at zero angle of attack. The flow visualizations were done using the velocity vector and contour plots in order to analyze the influence of the shape of the spike on the drag reduction. The computed contour plots agree well with the schlieren pictures of the experiments and the computed pressure distribution compares well with measured pressure on the blunt body. The bow shock wave formation is found over the spherical and the flat spike which generate different separation zones over the blunt body. The shock wave stand-off distances for the hemispherical and the flat spiked are compared with the analytical results and found in good agreement. The pressure distribution, the coefficient of skin friction and the wall heat flux variation along the surface of the spike facing the flow direction are influenced by the spike shape. The pressure and the heat transfer rate at the stagnation point are computed and compared with the analytical solution. The numerical analysis gives complete flowfield information over the spike surface including the shock stand-off distance, sonic line, and velocity gradient along the surface of the spike. The hemispherical disc spike gives high aerodynamic drag and heat flux as compared to the flat-faced disc spike. Acknowledgments The author expresses his sincere gratitude to the Editor and Referees for giving their valuable comments, suggestions, and encouragement towards the improvement of the present work. References 1. Bogdonoff SM, Vas IE (1959) Preliminary investigations of spike bodies at hypersonic speeds. J Aerosp Sci 26(2):65–74 2. Maull DJ (1960) Hypersonic flow over symmetric spiked bodies. J Fluid Mech 8:584 3. Wood CJ (1961) Hypersonic flow over spiked cones. J Fluid Mech 12:614 4. Menezes V, Saravanan S, Jagdeesh G, Reddy KP (2003) Experimental investigation of hypersonic flow over highly blunted cones with aerospikes. AIAA J 41(10):1955–1966 5. Mehta RC (2000) Pressure oscillations over a spiked blunt body at hypersonic Mach number. Comput Fluid Dyn J 9(2):88–95 6. Mehta RC (2002) Numerical analysis of pressure oscillations over axisymmetric spiked blunt bodies at Mach 6.8. Shock Waves 11:43–440 7. Milicv SS, Pavlovic MD, Ristic S, Vitic A (2001) On the influence of spike shape at supersonic flow past blunt bodies. Fac Univ Ser Mech Autom Control Robot 3(12):371–382 8. Calarese W, Hankey WL (1985) Modes of shock wave oscillations on spike tipped bodies. AIAA J 23(2):185–192 9. Hahn M (1966) Pressure distribution and mass injection effects in the transitional separated flow over a spiked body at supersonic speed. J Fluid Mech 24:209–223 10. Milicev SS, Pavlovic MD (2002) Influence of spike shape at supersonic flow past blunt nosed bodies experimental study. AIAA J 40(5):1018–1020 11. Kubota H (2004) Some aerodynamic and aerothermodynamic considerations for reusable launch vehicles. AIAA 2428 12. Crawford DH (1959) Investigation of the flow over a spiked-nose hemisphere at a Mach number of 6.8. NASA TN-D 118 13. Motoyama N, Mihara K, Miyajima R, Watanuki W, Kubota H (2001) Thermal protection and drag reduction with use of spike in hypersonic flow. AIAA paper 2001-1828 14. Yamauchi M, Fujjii K, Tamura Y, Higashino F (1993) Numerical investigation of hypersonic flow around a spiked blunt body. AIAA paper 93-0887 15. Shoemaker JM (1990) Aerodynamic spike flowfields computed to select optimum configuration at Mach 2.5 with experimental validation. AIAA paper 90-0414 16. Fujita M, Kubota H (1992) Numerical simulation of flowfield over a spiked blunt nose. Comput Fluid Dyn J 1(2):187–195 17. Asif M, Zahir S, Kamran N, Kan MA (2004) Computational investigations aerodynamic forces at supersonic/hypersonic flow past a blunt body with various forward facing spikes. AIAA paper 2004-5189 18. Milicev SS, Pavlovic MD, Ristic S, Vitic A (2002) On the influence of spike shape at supersonic flow past blunt bodies. Mech Autom Control Roboti 3(12):371–382 19. Mehta RC (2000) Peak heating for reattachment of separated flow on a spiked blunt body. Heat Mass Transf 36:277–283 20. Mehta RC, Jayachandran T (1997) Navier-stokes solution for a heat shield with and without a forward facing spike. Comput Fluids 26(7):741–754 21. Mehta RC (2000) Heat transfer study of high speed over a spiked blunt body. Int J Numer Meth Heat Fluid Flow 10(7):750–769 22. Gauer M, Paull A (2008) Numerical investigation of a spiked nose cone at hypersonic speeds. J Spacecr Rocket 45(3):459–471 23. Ahmed MYM, Qin A (2010) Drag reduction using aerodisks for hypersonic hemispherical bodies. J Spacecr Rocket 47(1):62–80 123 Author's personal copy Heat Mass Transfer 24. Mehta RC (2010) Numerical simulation of the flow field over conical, disc and flat spiked body at Mach 6. Aeronaut J 114(1154):225–236 25. White FM (1991) Viscous fluid flow, 2nd edn. McGraw Hill International Edition, Singapore 26. Peyret R, Vivind H (1993) Computational methods for fluid flows. Springer, Berlin, pp 109–111 27. Blazek J (2001) Computational fluid dynamics: principles and applications, 1st edn. Elsevier Science Ltd, Oxford 28. Jameson A, Schmidt W, Turkel E (1981) Numerical simulation of euler equations by finite volume methods using Runge–Kutta time stepping schemes. AIAA paper 81-1259 29. Mehta RC (1998) Numerical investigation of viscous flow over a hemisphere-cylinder. Acta Mech 128(1–2):48–58 30. Mehta RC (2011) Block structured finite element grid generation method. Comput Fluid Dyn J 18(2):140–144 31. Shang JS (1984) Numerical simulation of wing-fuselage aerodynamic interference. AIAA J 22(10):1345–1353 32. Mehta RC (2000) Numerical heat transfer study over spiked-blunt body at Mach 6.8. AIAA paper 2000-0344 123 View publication stats 33. Truitt RW (1959) Hypersonic aerodynamic. Ronald Press Co., New York 34. Kalimuthu R (2009) Experimental investigation of hemispherical nosed cylinder with and without spike in a hypersonic flow. Ph. D. thesis, Department of Aerospace Engineering, Indian Institute of Technology, Kanpur, India, April 2009 35. Probstein RF (1956) Inviscid flow in the stagnation region of very blunt-nosed bodies at hypersonic flight speeds. WADC–TN 56-395, Sept 1956 36. Ames Research Staff (1953) Equations, tables and charts for compressible flow, NACA report 1135 37. Liepmann HW, Roshko A (2007) Elements of gas dynamics, 1st South Asian Edition. Dover Publications Inc, New Delhi 38. Hayer WD, Probstein RF (1959) Hypersonic flow theory. Academic Press, New York 39. Fay JA, Riddell FR (1958) Theory of stagnation point heat transfer in dissociated air. J Aeronaut Sci 25:73–85