You are on page 1of 18

Evaluation of Simplified Hypersonic Flight-Vehicle Geometries

This paper includes computational simulation of simplified flight-vehicle geometry in


supersonic and hypersonic airflow. Wedge, cone, and cone-flare geometry is examined as test
cases to demonstrate the use of the commercially available FLUENT computational fluid
dynamics software. Flow velocities in the range of Mach 2.35-20 are examined and verified by
comparison to widely available experimental, numeric, and computational results. The test
cases show good agreement with the available data for instances where the ideal gas
assumption is valid. Included are results which account for the effect of grid spacing and
initial condition assumptions. Some effort was made to expand upon the available results for
the cone-flare geometry to test the limits of the numeric solution.

Nomenclature hypersonic wind tunnels typically run in short bursts.


Also, the air must be preheated to prevent change in state
from gas to liquid as the temperature drops across the
nozzle. Hypersonic flight data is also extremely rare.
The only real sustained hypersonic flight currently
performed by a vehicle is the Space Shuttle during re-
= coefficient of pressure entry, which re-enters the atmosphere at velocities around
K = hypersonic similarity parameter Mach 25. The longest hypersonic, powered flight
L = characteristic length recorded is a flight of approximately 200 seconds
M = Mach number performed by the X-51 test vehicle in early 2010. As
= free-stream Mach number experimental hypersonic data can only be collected at
p = pressure extreme cost, computational methods are extremely
T = temperature valuable to the design of hypersonic flight vehicles. With
U = velocity computational fluid dynamics the effects of varying the
geometry of a vehicle can be examined at a relatively
= cone half-angle
small cost.
τ = slenderness ratio, cone diameter / length
Δx = distance between nodes in flow axial direction This project examined a small set of geometries
Δy = distance between nodes in flow radial / commonly found on vehicles designed for hypersonic
transverse direction flight in both supersonic and hypersonic flow conditions.
ρ = density Of special interest in this project is the pressure at the
= kinematic viscosity surface as hypersonic vehicles commonly use a lifting-
body approach where the lift is created by the geometry of
the body as opposed to having wings to create lift. There
Introduction was also some examination of the shape of the shock on
the geometry. As the Mach number increases the shock is
The hostile conditions that accompany hypersonic
formed closer to the body so that at high Mach the shock
flight make it an ideal application of computational fluid
can actually impinge on the body if the body widens past
dynamics. Steady, hypersonic flow is nearly impossible
the shock angle, as with a cone-flare geometry. High
to achieve in a wind-tunnel environment. The pressure
temperature effects are also of concern in hypersonic
differential required to accelerate air to speeds in excess
flow, but the scope of this project includes only a cursory
of Mach 5 requires such a large amount of energy that
look at temperature effects.
For the final test case, a cone-flare geometry with a
blunted tip was examined. This geometry is of interest in
the design of hypersonic flight vehicles as it is likely that
there are locations on a vehicle that could use a ramp-like
surface. This project compares against a specific case
Case Selection which contains numerical results from the FLUENT code
used for this project as well as experimental results from a
There are four separate cases examined for this Mach 6 test [5]. Similar geometry has been examined in
project, with the intent of increasing complexity as the other hypersonic studies ([6] [7] [8]), which leads one to
project progressed. Each case selected has either a well- conclude that this is a useful geometry for use in
established analytical solution or experimental data that validation of computational solutions.
may be used for validation. Special care was taken to find
sources widely considered reliable.

The first test case is that of an oblique shock on a 15 Constant Input Parameters
degree wedge at Mach 2.5 [1]. In this case the author
There are some solution parameters that remained
compares the results of a simulation using the NASA
constant throughout the project. This is a listing of all
fluid dynamics simulation code WIND with the exact
parameters set by the user in FLUENT that purposely
analytical solution in Anderson (Compressible) [2]. The
remained the same for all data sets. Gravity was
exact solution for supersonic flow over an infinite wedge
neglected. All solutions are for steady state; there were
is known for certain parameters including pressure,
no transient effects included. The density-based solver
density, and temperature ratios across the oblique shock.
was used, as was absolute velocity formulation. The
The author examines several different two-dimensional,
FLUENT material ‘air’ was changed from constant
structured grids varying the spacing.
density to ideal gas to allow for compressible flow
The second test case is a shock on a 10 degree cone at solutions as all cases included in the project were above
Mach 2.35 [3]. While the differing Mach numbers and Mach 0.3. All solutions used the implicit, Roe-FDS,
half angles make a direct comparison difficult, it can be least-squares cell based solver.
seen in the exact results that the shock on a wedge is
Also constant throughout the project was the inlet and
stronger than the shock on a cone at the same Mach and
outlet boundary condition specification. While the
half angle as the ratio of properties across the shock is
operating condition changed, the inlet and outlet of the
notably lower for the cone. This is due to the relief of
mesh were set to the FLUENT ‘Pressure Far-Field’
energy around the cone that is not present in the wedge
boundary type. This boundary is intended for use when
case. The change in geometry is reflected with a change
the boundary is assumed to be at infinity. Using the
in the boundary conditions. Where the wedge is modeled
Pressure Far-Field allows for the direct input of Mach
with a 2-D geometry with an assumed infinite depth, the
number as a solution variable, as opposed to needing to
cone is modeled in 2-D but with an axi-symmetric
input the velocity as calculated based on the desired Mach
boundary at the cone axis of symmetry.
number and free-stream atmospheric conditions. Also,
In the third test case, the 10 degree cone case is the intent of the project is to examine the flow
expanded to a range of cones with a half-angle of 5, 10, characteristics of a vehicle traveling through a quasi-
15, 20, 25, and 30 degrees run at Mach 10. The purpose infinite atmosphere, which is what the Pressure Far-Field
of this case is to compare the pressure coefficient for a is intended to simulate. Throughout this project there are
computational data set versus a slender body references to FLUENT results. Unless otherwise stated,
approximation from Anderson (Hypersonic) [4]. Also FLUENT data is the data produced by the user for this
included in [4] is an experimental data set which matches project.
well with the slender-body approximation up to a cone
half-angle of around 30 degrees.
Supersonic 15 degree Wedge
Literature Review = Δy). Table 1 gives the Δx and Δy values for each grid,
along with the shock wave width (linear measurements in
The literature for this case is a National Aeronautics inches). The maximum Δy is the Δy at the inlet and the
and Space Administration (NASA) National Program for minimum at the inlet and the minimum Δy is the Δy at the
Applications-Oriented Research in CFD (NPARC) outlet (the same number of nodes are along both the inlet
website designed to demonstrate the validation and and outlet). Each grid was run for both a 1st order and 2nd
verification of the WIND CFD code for a 15 degree half- order upwind scheme. The pressure ratio across the shock
angle wedge at Mach 2.5 [1]. As this website is intended was the initial parameter for comparison, but there was no
to provide an example of CFD code verification, it is a appreciable difference in the pressure ratio for any of the
good example of a thorough examination of a CFD code. grids used. Each grid was able to predict the pressure
There are no experimental results referenced at the ratio well, but the size of the shock wave varied greatly;
website, but the exact analytic solution for the case is as the shock width should be zero (for practical
simple enough to be used directly. The only feature in the measurement purposes), this can be used as a direct
flow is the presence of a shock wave; the remainder of the measurement of the error in the solution.
flow is steady and undisturbed. The flow parameters are
constant on either side of the shock and the relationship Table 1: Grid Size and Shock Width for Grid Convergence Study
between the flow characteristics on the two sides of the
max min shock Time
shock is given by a collection of ratios, where the grid size Δx
Δy Δy width req’d
subscript 1 indicates free-stream conditions ahead of the to
inches solve
shock and the subscript 2 indicates conditions behind the
1st <5
shock. order 36x24 0.500 0.500 0.3660 3.100 sec
360x240 0.050 0.050 0.0366 0.310 8 min
Mesh Characteristics and Solution Methods 400x200 0.045 0.060 0.0439 0.189 4 min
2nd <5
This problem used a 400x200 mesh (200 from the order 36x24 0.500 0.500 0.3660 1.240 sec
ramp to the far wall, 400 from the inlet to the outlet), 12
360x240 0.050 0.050 0.0366 0.124 min
which is in the range between the NASA author’s Grid C 8 min
400x200 0.045 0.060 0.0439 0.126
and Grid D (from Wedge Study #2). A structured 2-D
mesh was created by meshing each edge of the domain
with the given number of nodes and allowing Gambit to Figure 1 is a plot of the shock width versus the
mesh the face using the ‘map’ method. The 15 degree longitudinal grid spacing, Δx. Figure 2 is a plot of the
wedge allowed for a mesh with limited skew in the shock width versus the transverse grid spacing Δy. In
elements. The mesh was then brought into FLUENT both cases the error seems to be decreasing at nearly the
where the inviscid assumption was applied. Using a same rate for both the 1st order and 2nd order solutions.
Courant number of 1.0, the solution converged at 1190 This is because the 1st or 2nd order solver was 1st or 2nd
iterations for Mach 2.5 (default FLUENT convergence order in time, both being the same order in space. Using
criteria were used). the 2nd order in time solver does, however, give more
accurate results. In order to estimate the order of
Using almost all the same solver parameters, an error accuracy in time the Courant number should have been
study on the chosen mesh was performed to ensure that varied. This case is steady in time, but changing the
the mesh was created in such a manner as to provide Courant number while holding constant Δx and flow
accurate results. Two grids, aside from the one used for velocity would have the effect of changing the Δt. For the
this case, were created to illustrate error convergence with steady case the Δt is the time between iterations while
decreasing Δx and Δy (x being the axial or longitudinal solving.
coordinate and y being the transverse coordinate). The
first grid was a 36x24 grid, which is approximately an The effect of cell skew can be seen when comparing
order of magnitude coarser than the grid used for the data Figure 1 and 2; the longitudinal grid spacing decreases
collection (400x200), and the second grid was a 360x240 from the 360x240 to the 400x200 grid, but the transverse
grid, which is the same order of magnitude as the grid spacing increases, so the overall accuracy is less
400x200 grid but allows for a square grid in the entry (Δx when using the 2nd order solution. The grid should be
refined in both directions (in 2-D) to gain real benefit analytic solution is labeled as ‘% error’ whereas the
from increasing the number of nodes. This study gives difference between the FLUENT and the WIND code is
some credibility to the use of the 400x200 grid with the labeled as ‘% difference’ as it is a comparison between
2nd order upwind solver for this case. The ~2% gain in two computational solutions that may or may not be
accuracy is not significant for the increase in CPU time accurate. The results show good agreement with both
required when compared to the 360x240 grid. data sets (<5% difference). However, the comparison
between WIND and the analytic solution shows that the
WIND results are <1% different from the analytic
solution. This demonstrates that the FLUENT solution,
while good, could be done better. Given the behavior of
the residuals at the time of convergence, it is quite likely
that a further reduction of the convergence threshold
would yield a more accurate solution. This would lead to
a higher computational cost that may or may not be
worthwhile depending on the accuracy needed. For this
current activity, 5% error is considered sufficiently
accurate.

Table 2: Comparison of FLUENT Results to Analytic and NASA


Solution

Mach
Figure 1: Shock Width vs Longitudinal Grid Spacing for Mach 2.5 number
15 deg Wedge Pressure Density Temperature
behind
ratio ratio ratio
the
(p2/p1) (ρ2/ ρ1) (T2/T1)
oblique
shock

Anderson [2] 1.8735 2.4675 1.8665 1.3220

FLUENT data 1.8370 2.5776 1.9214 1.3417

% error 1.95% 4.46% 2.94% 1.49%

NASA WIND
1.8684 2.4696 1.8561 1.3305
data [1]

FLUENT data 1.8370 2.5776 1.9214 1.3417

% difference 1.68% 4.37% 3.52% 0.84%

Further Results
Figure 2: Shock Width vs Transverse Grid Spacing for Mach 2.5 15
deg Wedge As the intent of this project is to examine hypersonic
flow, it was decided to re-run this case at higher Mach
Comparison to Analytic or Experimental Results numbers to see the results. This specific case had low
computational cost, even for higher Mach numbers. The
At convergence, the values for Mach behind the shock
same mesh and initial conditions (with the exception of
as well as the ratio of pressure, density, and temperature
Mach number) were solved for Mach numbers of 3, 4, 5,
across the shock were calculated from the FLUENT
6, and 7. The solution at Mach 3 took 1300 iterations to
results. Table 2 summarizes the results from the solution
solve, which is expected as this is a higher velocity case
at Mach 2.5 with a comparison to both the analytic
and therefore higher property gradients across the shock.
solution from Anderson [2] and the NASA WIND code
It is interesting to note that the solutions at Mach 4, 5, and
results [1]. The difference between the FLUENT and the
6 required fewer iterations to converge to the same
accuracy. Only at Mach 7 were more iterations required, The literature for this case is a National Aeronautics
and at Mach 7 the solution would not converge to the and Space Administration (NASA) National Program for
same value of residuals as the other cases. Table 3 shows Applications-Oriented Research in CFD (NPARC)
the Mach behind the shock and ratio of temperatures website designed to demonstrate the validation and
across the shock for the Mach 2.5 as well as Mach 3-7 verification of the WIND and NPARC CFD codes for a
solutions with the percent error as compared to the 10 degree half-angle cone at Mach 2.35 [3]. This website
Anderson analytic solution [2]. is not nearly as thorough as the 15 degree wedge website
[1], so for a code validation study reference [1] should be
Table 3: Comparison of FLUENT Results to Analytic Solution at
Higher Mach Numbers
utilized. There are no experimental results referenced at
the website, but the exact analytic solution for the case is
simple enough to be used directly. As compared to the
M2 M2 % error
M1 wedge, there is more variability in the flow behind the
(Anderson) (FLUENT) M2
shock. For this reason the flow parameters at the cone
2.5 1.874 1.837 1.98%
surface differ from those just across the shock. For this
3 2.255 2.222 1.48%
case the subscript 1 indicates free-stream conditions,
4 2.930 2.899 1.05% subscript 2 indicates conditions behind the shock, and
5 3.505 3.468 1.05% subscript 3 indicates conditions at the cone surface.
6 3.993 3.953 1.01%
7 4.402 4.348 1.21%
Mesh Characteristics and Solution Methods

T2/T1 T2/T1 This problem uses a 200x200 node mesh. In order to


(Anderson) (FLUENT)
allow for more accurate calculation in the region near the
2.5 1.322 1.340 1.35% cone, a node bias of 1% toward the cone axis of symmetry
3 1.388 1.409 1.47% was used. The GAMBIT quad map feature was then used
4 1.546 1.564 1.16% to mesh the analysis domain with a structured grid.
5 1.736 1.756 1.17%
The key difference between a wedge and a cone
6 1.958 1.987 1.49%
simulation, when using a 2-D representation of the
7 2.215 2.258 1.95% geometry, is that a wedge uses a 2-D planar geometry
whereas a cone requires the use of an axi-symmetric
boundary. One of the errors that led to incorrect results
Figure 3 shows the converged residuals for the Mach
early in the project was the improper use of the ‘Axis’
7 case. The solution is considered converged for this
boundary condition. The ‘Axis’ boundary condition in
analysis as the error is not increasing with time.
FLUENT must be applied to an axis of rotation. It is
possible in GAMBIT to give the Axis boundary condition
to a line that is not an axis of rotation. For the initial
solution in the project the Axis boundary condition was
applied to both the leading flat wall (labeled A in Figure
4) and to the cone surface (labeled B in Figure 4). This
led to the incorrect results in Table 4.

Figure 3: Residuals for Mach 7 FLUENT Solution

Supersonic 10 degree Cone B


A
Literature Review
increased accuracy gained from using a 3-D grid is not
Figure 4: Representative 10 deg Cone Mesh worth the computational cost. The Table 6 shows a
comparison of the FLUENT results compared to the
NASA WIND and NPARC code results [3] for the results
given on the source website. For the selected properties,
the computational solutions compare very well to each
other.
Table 4: Results for the Incorrect 10 deg Cone Boundary Conditions
Table 5: Comparison of Results for the 10 deg Cone to Analytic
FLUENT Analytic [2] % error Solution

behind the shock


FLUENT Analytic [2] % error
M 2.2677 2.2677 0.00%
behind the shock
p / p1 1.2257 1.1781 4.04%
M 2.2471 2.2677 0.91%
T / T1 1.0407 1.0481 0.71%
p / p1 1.1680 1.1781 0.86%
ρ / ρ1 1.0000 1.1240 11.03%
T / T1 1.0481 1.0481 0.00%
at the cone
ρ / ρ1 1.1273 1.1240 0.29%
M 2.0590 2.1469 4.09%
at the cone
p / p1 2.0119 1.4234 41.34%
M 2.1442 2.1469 0.13%
T / T1 1.1047 1.1063 0.14%
p / p1 1.3734 1.4234 3.52%
ρ / ρ1 0.5381 1.2867 58.18%
T / T1 1.0963 1.1063 0.91%
ρ / ρ1 1.2545 1.2867 2.50%

Further examination of the FLUENT User’s Guide led


to the realization that the ‘Axis’ boundary condition Table 6: Comparison of FLUENT Results to the NASA WIND and
should be applied to the axis about which the axi- NPARC Code Results
symmetric solution should be calculated, which in this
case is the line ‘A’ in Figure 4. The correct boundary % difference

condition to use for line ‘B’ in Figure 4 is ‘Wall’, which FLUENT FLUENT
is the same boundary condition used on the ramp for the FLUENT WIND NPARC vs vs
15 degree wedge. Using a Courant number of 1.0, the WIND NPARC

solution converged after approximately ten minutes for


Mach 2.35 (default FLUENT convergence criteria were Mcone 2.1442 2.1468 2.1467 0.122% 0.120%
used).
pcone
1.3734 1.3741 1.3741 0.051% 0.051%
Comparison to Analytic or Experimental Results /pfree-stream

At convergence, the values for Mach behind the shock Tcone


1.0963 1.0951 1.0951 0.103% 0.102%
/Tfree-stream
and at the cone as well as the ratio of pressure, density
(rho), and temperature across the shock and at the cone
were calculated from the FLUENT results. Table 5
summarizes the results from the solution at Mach 2.35
with a comparison to the analytic solution from Anderson Hypersonic Cone Pressure Coefficient
[2]. Most of the error is <1%, which is a better result than
Literature Review
the 15 degree wedge case. The two properties showing
the greatest error are the pressure and density at the cone The literature for this case is the text Hypersonic and
surface. For this case it is likely that the error comes from High Temperature Gas Dynamics [4], specifically Section
the coarseness of the grid near the cone surface and the 4.6 on small-disturbance theory. In this part of the text,
use of an axi-symmetric 2-D grid. Alternately to the 2-D approximations for properties of engineering interest (ie
axi-symmetric solution, a 3-D cone geometry could be pressure coefficient) are derived from the “hypersonic
created. Given the low amount of error in this case, the small disturbance equations” [4]. These equations assume
high Mach number and neglect the slenderness ratio two supersonic test cases allowed for the intelligent
squared (slenderness ratio τ = diameter/length). Given tailoring of the mesh as the general properties of the flow,
this approximation, an expression for the pressure ie the location of the shock, are known. Figure 5
coefficient on a cone is presented in equation 4.126, illustrates the boundaries of the analysis region with each
shown below. boundary given a label that will be referenced throughout
this section.

where far wall

The literature states that this approximation is valid


“even for a reasonably large cone semiangle of 30 quad A
quad B outlet
degrees” based on a comparison to experimental results inlet
[4], which would correspond to τ2 = 0.33. This is the type quad D
quad C
of approximation that was used in the early days of
hypersonic analysis in the 1950’s and 60’s before cone
axis
computational fluid dynamic simulations were available.
Figure 5: Hypersonic Cone Mesh Delineation
This approximation is somewhat limited in its usefulness
to us today as it only applies to a cone, but given that a
majority of hypersonic bodies at the time were missiles
with a conical nosecone, it was extremely useful at the This case includes results for 5, 10, 15, 20, 25, and 30
time. degree cones. The 5-25 degree cones all used a similar
mesh along the edges of the domain, but the 30 degree
Mesh Characteristics and Solution Methods
cone needed a separate set of nodes along the edges as
This is the first case in the project that analyzes flow GAMBIT was not able to mesh quad B (see Figure 5) due
at hypersonic velocity. ‘Hypersonic’ is to be flow with a to the high level of skew in the elements. For the 5-25
velocity greater than Mach 5 for the sake of the cases degree cone mesh, the number of nodes along each edge
included in this project, so a flow velocity of Mach 10 was as follows: In quad A, the quadrant containing only
was selected for this case to ensure the flow is soundly free-stream flow, there were 20 nodes along each edge.
within the realm of hypersonic flow. With the higher In quad B there were 20 nodes in the boundary shared
velocity comes a stronger, thinner shock and greater with quad A, 100 nodes along the far wall, 20 nodes along
possibility of instability in the numeric solution due to the the outlet, and 200 nodes along the boundary with quad
rapid change in flow characteristics. Because of this, the D. As the flow velocity is perpendicular to the inlet, the
mesh generated for the hypersonic cone analysis is more number of nodes along the inlet and outlet is identical to
complex than the mesh generated for the supersonic 10 allow the velocity vectors to flow generally along a node
degree cone, even though the physical geometry modeled line. In quad C there were 20 nodes along both the
is the same or similar. A handful of different meshes boundary with quad A and the axis, and 400 nodes along
were created in an attempt to find a good balance between the inlet and the boundary with quad D. In quad D, the
the accuracy of the solution and the CPU time required area of greatest interest (where the shock will form), there
for the solution. were 400 nodes along the boundary with quad C and on
the outlet and 200 nodes along the boundary with quad B
After review of the various meshes, it was decided to and the cone. The 30 degree mesh required 40 nodes
break the mesh down into quadrants in an attempt to along the boundaries which contained 20 nodes for the
accurately capture the flow characteristics in the region of other meshes.
interest and to avoid an unnecessarily tight mesh in
regions where the flow is at free-stream conditions. Also, With the intent of concentrating the elements in the
it was desired to maintain a structured mesh for the sake region of the domain of greatest interest, for the 5-25
of accuracy and computational cost. The results from the degree cases there were 80,000 elements in quad D, 8,000
in quad C, 400 in quad A, and somewhere between 2,000
and 4,000 elements in quad B depending on the solution
found by the meshing algorithm in GAMBIT. Figures 6-8
show the mesh used for the 10 degree cone and the 30
degree cone. Between Figures 6, 7, and 8 the zones of
structured and unstructured mesh can clearly be seen.
The unstructured mesh in quad B was created using the
GAMBIT ‘pave’ meshing algorithm. Quad elements
were used when possible, but for some of the meshes it
was necessary to use a mixture of quad and tri elements.
Also seen in Figures 6-8, and especially in Figure 8, is
that quad D was created such that it would be a rectangle Figure 8: 30 degree Hypersonic Cone Mesh, Full View
to minimize the amount of skew in the quad D elements.
An attempt was made to use the grid refinement
feature in FLUENT for the 5 degree cone grid. The
solution was run to convergence, after which a refinement
based on pressure gradient was performed. The shock
size after the first run of the solution was approximately
the width of one row of cells, and since the cell rows were
at a different angle from the shock angle, there was a
stair-step (as opposed to linear) shape to the shock. The
shock wave would follow one row of cells for a length
and then ‘jump’ to the next row of cells. Since the
gradients followed the stair-step shape, the refinement
added grids in the stair-step shape. The solution was run
again after the grid refinement, and there was little to no
Figure 6: 10 degree Hypersonic Cone Mesh, Full View
change in the flow properties behind the shock. Based on
this, for this analysis case the most effective grid
refinement would be to add a region of structured grid
such that the cell rows are parallel to the shock. The
built-in FLUENT grid refinement seems best suited to
unstructured grids in flow regimes where the flow
properties are known with less certainty.

For each of the hypersonic cone solutions inviscid


flow was assumed. Also, a Courant number of 0.5 was
used for the final solutions. Early runs were attempted
with a Courant number of 1.0, but the convergence of
Figure 7: 10 degree Hypersonic Cone Mesh, Close-up at Quadrant residuals was found to be steadier with the Courant
Intersection number at 0.5 such that the penalty of increased CPU time
per iteration was worthwhile. A second-order solver was
used for each solution as it produced accurate results for
the supersonic wedge and cone cases. The third-order
MUSCL solver was used for one test case, but the
solution did not converge within a reasonable time as
compared to the second-order solution (any solution
requiring greater than 20-25 minutes to solve was
considered unreasonable for this project).
Comparison to Analytic or Experimental Results computational solution for the 30 degree laminar case,
resulting in even more inaccurate results.
The results for the hypersonic cone cases covered only
a small portion of the curve shown as Figure 4.8 in Table 7: Comparison of Hypersonic Cone Results to Analytic
Solution
Anderson (Hypersonic) [4], but the results show good
agreement with the analytic solution for cones with a half- Cone Half- Analytic
FLUENT Solution
angle of 5-25 degrees. Figure 9 is a co-plot of the Angle θ Cp / sin2θ % error
Cp / sin2θ Method
(deg) [4]
pressure coefficient divided by the sine squared of the
5 2.531 2.446 3.36%
cone half-angle for the analytic solution and the FLUENT
10 2.228 2.206 1.00%
data for the 5-25 degree cones. The 30 degree FLUENT
15 2.157 2.217 2.75%
solution does not match well at all with the analytic
20 2.131 2.106 1.15%
solution, which could be due to a number of factors.
25 2.118 2.091 1.28%
Based on the relations given in [9], the expected shock
30 2.111 27.92 1222.11% inviscid
angle for a 30 degree cone at Mach 10 is 33.7 degrees,
30 2.111 92.01 4257.65% laminar
meaning the shock is a mere 3.7 degrees off of the cone.
At that angle the shock and the boundary layer at the body
are likely interacting, which dramatically increases the
Hypersonic Cone-Flare
complexity of the flow. Given that the computational
solution was for the inviscid case, though, there is no Literature Review
boundary layer in the computational solution.
The main paper examined for the hypersonic cone-
flare geometry was Blunted Cone-Flare in Hypersonic
Flow [5]. The specific geometry with dimensions is
shown in Figure 10; linear dimensions are given in
millimeters. With this geometry the flow is considerably
more complex than with the wedge or cone geometry as
there are now two shock waves forming on the body, one
a bow shock at the leading edge and an oblique shock at
the vertex of the flare where the flow is directed into
itself. The added complexity means that the analytic
solution exists only for specific positions in the flow, not
for the flow as a whole. Analytic solutions exist only for
locations such as the stagnation point at the leading edge
of the cone-flare.
Figure 9: Replication of Figure 4.8 from Anderson [4] with
FLUENT Data

Table 7 shows the Cp / sin2θ values for the analytic


solution [4] and the FLUENT solution along with the
percent error. As the percent error is <5% for the 5-25
degree cones, this is considered a good solution. The
solution method is given for the two 30 degree cone cases
analyzed; all other results were for inviscid flow. It was
noted that the boundary layer and shock interaction that
could occur for the 30 degree case was not relevant to the
inviscid solution. A second attempt was made at solving
the 30 degree cone assuming laminar flow, and this
Figure 10: Hypersonic Cone-Flare Geometry Analyzed
yielded more inaccurate results. It is assumed that the
boundary layer and shock interaction does affect the This specific paper was chosen because of the large
number of qualitative plots for comparison and the fact
that the author used FLUENT as the computational In [7], the authors explore some possible sources of
solver. The thoroughness with which the author of the error in numeric solutions where the agreement with
paper presents the results lends credibility to the results. experimental data is off by a greater amount than can be
Examining the effect of the temperature of the test article explained by the known experimental uncertainty.
is a good step in improving the quality of the analysis. According to their results, taking into account the effects
Also, with more complex geometry it becomes desirable of vibration non-equilibrium in the test setup can account
to have experimental results for verification, and the for some of the loss of heat transfer. Also, vibration in
selected paper contains referenced experimental data. As the test article can lead to a small amount of slip at the
the temperature of the moving fluid at the test article test article wall, which further reduces the heat transfer
surface is greater than 300 K (the initial assumed test rate [7]. While these error factors are certainly worth
article temperature), it stands to reason that setting up the examining, the scope of this project does not allow for
computational simulation with a test article temperature their inclusion here.
greater than 300 K could yield more accurate results.
Reference [8], “Validation of Fully Implicit, Parallel
One interesting aspect of the free-stream conditions in Finite Element Simulations of Laminar Hypersonic
the experimental and computational setups is the free- Flows,” is exactly that. The authors developed a fully
stream temperature of 67.07 K and pressure of 673.67 Pa implicit numerical solution usable for a certain hypersonic
[5]. It was noted in the Introduction of the project that flow regime, and the paper gives a validation by
wind tunnels with hypersonic capability preheat the comparing numeric results to experimental data. This is
inflowing air so that the air does not reach liquid state another example of laminar flow over hypersonic
while expanding to hypersonic velocity. The low geometry which is similar to other literature. The hollow
temperature and pressure (relative to standard sea-level flare used is more similar to the cone-flare geometry in
atmosphere) are indicators of the unique challenges in this project. The cone-flare in [8] has a more severe flare
simulating hypersonic flight conditions. At standard angle, similar to that in [6], which yields much different
atmospheric pressure, nitrogen condenses at around 77 K, flow characteristics in the so-called ‘separation region’
and oxygen around 90 K. Only the low pressure prevents where the flare turns the flow back onto itself.
the air in the experiment from liquefying.
The similarities between the results in references [5],
Other literature for similar geometry was included in [6], [7], and [8], along with the number and variety of
the background research for this project. In the literature sources between them, would lead one to trust them as
found on this or similar geometry, all of the experimental reliable. In addition, great effort was made to ensure that
setups were chosen so that there would be laminar flow in the sources used for this project fit well with the whole of
the test section. It is assumed that this is due to the the existing data for hypersonic flow over simplified
complex nature of hypersonic flow by itself. Additional geometries. While each source has a unique strength,
complexity added by turbulent flow would further they fit together as a unified set as opposed to being
increase the numeric accuracy needed to produce a discordant pieces each demonstrating a completely
worthwhile computational solution and is itself an area for separate solution.
study.
The only real issue with the literature surveyed is that
The study performed in [6] found that using a second- all the cases are for laminar flow. This is good for
order scheme yielded reasonably accurate results academic research and validation of computational
compared to higher order schemes, and actually methods, which is the stated goal of the majority of the
performed nearly identically to the higher order schemes reviewed literature. However, real flight conditions
at velocities above Mach 10. The course text explicitly would almost certainly involve turbulent flow. For
states that “second-order approximations usually offer a example, the Reynolds number for the hollow flare
good combination of ease of use, accuracy, and cost experiment in [8] is 2.5E4, but the same test article at the
effectiveness in engineering applications,” and the results same Mach number in standard atmosphere at 60,000 feet
in [6] bear this out for the case of hypersonic flow [11] would be at a Reynolds number of 1.4E7 (assuming
specifically.
), which only increases as the length is expanded to
more realistic flight vehicle dimensions (the test article is
8.7 inches in length). At this high of a Reynolds number,
it is expected that the flow over the vehicle will be fully
turbulent. Therefore the results of this project are not
directly applicable to hypersonic flight vehicles, but are
useful for understanding the general characteristics of
hypersonic flow.

Mesh Characteristics and Solution Methods

There were a handful of separate meshes used for this Figure 12: Cone-Flare Final Mesh, Close-up
case. The initial mesh included a region behind the cone-
flare, extending approximately 50 mm aft of the body.
This resulted in a large number of iterations (~3000) A structured mesh was created close to the body,
where the low-end absolute pressure limit set in FLUENT while an unstructured mesh was used to fill the remainder
was reached. Even using the FLUENT solution steering of the solution space. The second mesh created (first one
set for hypersonic flow (Courant number variable from ending at the aft end of the cone-flare) used only one
0.5-4.0) the solution took a considerable time to converge. layer of structured mesh near the body. After running the
It was decided that accurately computing the wake in this solution, it was decided that the transition between the
case would require a much longer mesh. The drag finer, structured mesh at the body and the sparser,
increase due to the extremely low pressure region at the unstructured mesh further away from the body was too
aft of the hypersonic geometry is of great interest in the abrupt. A second layer of structured mesh near the body
overall design of hypersonic vehicles, but this specific was added to create a more smooth transition between the
cone-flare geometry is intended to model the forward end structured and unstructured mesh sections. This third
of a vehicle. It was mentioned in the literature review for mesh was the mesh used for the final results for this case.
this case that actual hypersonic vehicle flight conditions Using this mesh, the shock at the leading edge was
would almost certainly be in the turbulent regime. The completely within the structured mesh near the body.
wake properties would depend greatly on a turbulent flow
solution, and the cases chosen for comparison in this Initially the assumed boundary condition at the far-
project are for laminar flow. Taking into account all these wall was set as ‘Wall’ in this analysis. The far-wall was
factors, it was decided to place the mesh outlet at the aft created at a distance away from the cone-flare assumed
end of the cone-flare. Figures 11 and 12 show the mesh sufficient to prevent interaction between the flow at the
used for this case, which used an axi-symmetric 2-D body and any wall effects at the far-wall. After viewing
solver. The minimum node spacing in the mesh at the the contours from the initial solution, however, interaction
body was approximately 0.015-0.02 inches. between the flow around the body and the flow at the far-
wall could be seen. It was decided that a boundary
condition of ‘Symmetry’ at the far-wall would be more
appropriate. A Symmetry boundary condition allows the
flow properties to remain at free-stream at the far-wall as
the properties are assumed symmetric about the far-wall
boundary. Alternatively a slip condition at the far-wall
could have been used to reduce or eliminate the boundary
layer at the far-wall. The desire for this project is to more
directly simulate flight conditions as opposed to exactly
replicating experimental data, so the symmetry condition
is more appropriate for this case.
Figure 11: Cone-Flare Final Mesh, Full View
Three separate simulations were run for the cone-flare
case – inviscid, laminar (to match the experiment and
literature), and k-epsilon (all using the same grid). The
inviscid condition was used for the supersonic wedge,
supersonic cone, and hypersonic cone cases so it was
desired to examine the use of the inviscid case for the
more complex flow. Up to this point there has been no
mention of turbulence models, so a k-epsilon case was run
to see how the results would compare to the laminar case.
Figure 13 shows the residual value versus iteration for the
inviscid solution.

Figure 15: Residuals for the Turbulent (k-epsilon) Case (Mach 6


Cone-Flare)

Upon examination of the residual plots in Figures 13-


15, it can be seen that the inviscid and turbulent solutions
took longer to solve (7000-8000 iterations) than the
laminar solution (~3500 iterations) and that the inviscid
and turbulent simulations seem to converge more
smoothly. The instability in the laminar solution residuals
is likely due to the flow in the separation zone where the
Figure 13: Residuals for the Inviscid Case (Mach 6 Cone-Flare) flow recirculates. This recirculation is not present in
either the inviscid or turbulent solutions, hence the
Figure 14 shows the residual value versus iteration for the velocity values are more stable. The laminar case never
laminar case. The experimental data and main literature reached the convergence criteria set in FLUENT, but the
used for reference [5] analyzed laminar flow. residuals did reach a state of iterative equilibrium. It is
worth noting that the time for solution of the laminar case
is nearly half of that for the inviscid or turbulent case.
The Reynolds number remained constant across the three
solutions, which should lead to errors with the turbulent
case. The Reynolds number for this solution setup is well
within the laminar flow regime, meaning the use of a
turbulence model is not physically reasonable for this
setup.

The recirculation in the separation zone for the


laminar case can be seen in Figure 16. This is similar to
the result predicted in [5], which is shown in the
Comparison to Analytic or Experimental Results section.
Figure 14: Residuals for the Laminar Case (Mach 6 Cone-Flare) Figures 17 and 18 show the same region for the laminar
(15) and turbulent (16) cases. Figures 16-18 each use a
Figure 15 shows the residual values versus iteration for
separate scale for the velocity contour arrow size as was
the turbulent (k-epsilon model) case.
deemed appropriate for the specific solution used. It is
assumed the turbulence model case shows no
recirculation because the numerical factor allowing the
fluid to travel transverse to the velocity streamlines is
allowing for more of a transverse component in the
velocity, resulting in undisturbed flow at the corner.
Because the fluid is not energetic enough to be turbulent, solution space for this case, it is possible to directly
no eddies form. calculate the stagnation temperature and pressure. These
values are of great concern as these are the peak values of
temperature and pressure in the flow. Table 8 shows the
stagnation temperature and pressure results for this
project (the FLUENT values) compared to the analytic
and numeric results presented in [5]. As the numeric
results in [5] are from FLUENT, it is not surprising that
the results from this project match more closely to the
numeric results than the analytic results from [5]. The
error for this case is on the same order as the error
calculated for the preceding cases in this project.

Table 8: FLUENT Results Compared to Analytic and Numeric


Results from Ref [5]

Stagnation Stagnation
Figure 16: Velocity Vectors Showing Recirculation for Laminar Temperature Pressure
Case (K) (kPa)
FLUENT 547 30.90
Analytic [2] 550 31.54
% error 0.55% 2.03%
FLUENT 547 30.90
Ref [5] 549 31.48
% difference 0.36% 1.84%

Qualitative comparison also shows good agreement


between the results of this project and those in [5].
Figures 19a and 19b show the absolute pressure contours
Figure 17: Velocity Vectors Showing No Recirculation for the (in Pascals) near the leading edge for the literature [5]
Inviscid Case (19a) and the FLUENT results (19b). Figures 20a and
20b show the static temperature contours (in Kelvin) near
the leading edge for the literature [5] (20a) and the
FLUENT results (20b). Both sets of data show good
agreement in values and contour shape.

Figure 18: Velocity Vectors Showing No Recirculation for the


Turbulent Case

Comparison to Analytic or Experimental Results


Figure 19a: Pressure Contours from [5] (Pa)
While the flow is complex enough that an exact
analytic solution does not exist for the majority of the
FLUENT solution for this project in 21b. While the
existence of the separation zone is apparent in the
FLUENT solution, and the FLUENT solution fairly
accurately predicts the start of the separation zone at
around 73 mm (lateral position along body, seen in Figure
22b, laminar case), the separation length is under-
predicted (the onset of the separation zone in [5] is at
72.73 mm). It is assumed that this is due to the differing
grid characteristics at the cone-flare junction. The
referenced grid in [5] is scaled such that the nodes are
more concentrated at the junction and are sparser away
from the junction, where the grid used for this project
uses uniformly spaced nodes along the length of the cone-
Figure 19b: Pressure Contours from FLUENT Solution (Pa) flare body. This same characteristic is present when
looking at the nodes near the body. The nodes near the
body are denser and become sparser as the distance from
the body increases. Again, the grid used for this project
uses uniformly spaced nodes in the region near the body.
The under-prediction of pressure at the surface illustrated
by comparing Figures 22a and 22b is likely due to the
lack of grid fidelity near the body. It seems the more
complex flow around the body for the cone-flare requires
a more dense mesh near the body to truly capture the flow
characteristics than is necessary for the simpler wedge
and cone geometries examined in the earlier portion of
this project.

Figure 20a: Temperature Contours from [5] (K)

Figure 21a: Velocity Vectors in Separation (or Recirculation) Zone


from [5]

Figure 20b: Temperature Contours from FLUENT Solution (K)

The referenced literature predicts a region where the


flow reverses direction at the junction between the cone
and the flare portions of the body. Figures 21a and 21b Figure 21b: Velocity Vectors in Separation (or Recirculation) Zone
from FLUENT Solution
show the existence of a recirculation zone where the flow
reverses, with the results from [5] in 21a and the
nearly the velocity of the Space Shuttle during re-entry
into the Earth’s atmosphere. Figures 23, 24, 25, and 26
show the Mach contours for each solution. The stronger
shock along with the lower shock angle with increasing
velocity can clearly be seen.

Figure 22a: Pressure at the Body versus X Position from [5]

Figure 23: Mach Contours for Mach 6 Case

Figure 22b: Pressure at the Body versus X Position from FLUENT


Solution

Much like in the pressure and temperature contour


plots (Figures 19, 20) the shape of the pressure graph is
similar between [5] and this project. The greatest
difference is in the separation zone where the FLUENT
results do not show the pressure plateau mentioned in [5].
Figure 22b also shows the pressure values for the inviscid
and turbulent (k-epsilon) solutions run for this case. The Figure 24: Mach Contours for Mach 10 Case
absence of a separation zone, noted with Figures 16-18, is
the only significant area of difference between the three
solution methods.

Further Results

The results in [5] are at Mach 6, which is on the low


end of the hypersonic regime definition. For this project
it was desired to see the effects at higher Mach numbers
for the same geometry. The following results used the
same initial conditions as the Mach 6 case detailed above
with the only exception being the Mach number. Results
were calculated at Mach 10, 15, and 20. Mach 20 is
Figure 27: Temperature Contours for Mach 10 Case

Figure 25: Mach Contours for Mach 15 Case Table 9: Stagnation Temperature Comparison for Mach 6-20

Stagnation
Temperature
(K)
ideal gas
Mach FLUENT % difference
[2]
6 550 547 0.55%

10 1408 1400 0.57%

15 3085 3070 0.49%

20 5433 5410 0.42%

These test cases illustrate one of the potential issues


when using a computational simulation for engineering
analysis. The computational solution seems good, it
Figure 26: Mach Contours for Mach 20 Case shows good agreement with analytic results, the residuals
converge, and the results seem to make physical sense.
The temperature contours in the stagnation region for
However, some of the assumptions made in deriving the
Mach 10, 15, and 20 all had the same basic shape (which
equations used to calculate the stagnation temperature
was slightly different from the contour shape for the
break down at temperatures around 4000 K [4], which is
Mach 6 case shown in Figure 20b), with the temperature
somewhere between Mach 15 and 20 in this solution set.
values increasing with increasing velocity. Figure 27
The assumption of constant ratio of specific heats no
shows the temperature contours for the Mach 10 case.
longer holds as the specific heat is now a function of
Table 9 shows the stagnation temperature for each
temperature. Also, the assumption of a continuous fluid
solution versus the analytically expected stagnation
is no longer valid as there are chemical reactions
temperature based on ideal gas assumptions [2].
occurring in the fluid (air) at this temperature. While the
results here are good from a numerical stability standpoint
in that they demonstrate the ability of the simulation to
cover the breadth of the flow regime, they are of
questionable value from an engineering standpoint at
values above Mach 15.

Figure 28 is a plot of the normalized pressure at the


body versus the position along the body, similar to Figure
22b, for Mach 6-10 (all cases were run with the laminar
solver). The pressure is normalized by the minimum the body as opposed to having wings to create lift. It was
pressure on the body. It can be seen that not only does the found that the use of the FLUENT computational fluid
pressure increase with increased velocity, the peak dynamic software could accurately predict the flow
pressure value increases dramatically at higher velocities. around simple geometry for supersonic and hypersonic
Looking at Figures 23-26, the shock layer at the location airflow for Mach 2-15 within about 5% for the parameters
where the leading-edge shock impinges on the flare examined. All results in this project assumed airflow
portion of the body is increasingly thinner as the velocity under ideal gas conditions, which is a strong limiter in the
increases. At Mach 15 and 20, there is a very thin shock applicability of this study. Use of the pressure far-field
later at this peak pressure location, resulting in very high boundary condition and the inlet and outlet of the solution
gradients. space allowed for the direct specification of flow Mach
number and yielded accurate results.

It was found that the 2nd order upwind solver is


necessary for a reasonably accurate solution, and that this
solver is sufficiently accurate for understanding the
general behavior of the flow in the regime studied. For a
simple geometry (ie wedge or cone) a grid with minimum
node spacing on the order of 0.04-0.06 inches was found
to be acceptable, but for a more complex geometry where
the shock impinges on the body (ie cone-flare), a grid
with node spacing on the order of 0.015-0.02 inches was
necessary. This grid spacing can be expanded further
away from the body, especially once the locations of the
flow disturbances within the solution space are known.
Use of the inviscid flow model was found to be
Figure 28: Pressure at the Body, Normalized by the Minimum appropriate for cases where the shock was not near the
Pressure, for Mach 6-20
surface of the geometry analyzed; for conditions where
the shock impinges on or is near the surface of the body, a
laminar flow model is necessary to accurately predict the
Conclusions flow as the conditions at the body effect other regions of
the flow. Overall, it was found that a simulation using the
This project examined a small set of geometries FLUENT software using a mesh within the parameters
commonly found on vehicles designed for hypersonic used in this project can yield an accurate solution for
flight in both supersonic and hypersonic flow conditions. simple a supersonic or hypersonic flight vehicle
Of special interest in this project is the pressure at the geometric feature within approximately 20 minutes,
surface as hypersonic vehicles commonly use a lifting- which allows for the rapid evaluation of vehicle geometry
body approach where the lift is created by the geometry of from an engineering standpoint.
References

[1] Slater, J. W., “Oblique Shock on a 15 Degree Wedge at Mach 2.5,”


http://www.grc.nasa.gov/WWW/wind/valid/wedge/wedge.html [retrieved 4 December 2010].
[2] Anderson, J. D., Modern Compressible Flow with Historical Perspective, 2nd ed., McGraw-Hill, Inc, New York,
1990.
[3] Slater, J.W., “10 Degree Cone at Mach 2.35,” http://www.grc.nasa.gov/WWW/wind/valid/cone10/cone10.html
[retrieved 4 December 2010].
[4] Anderson, J. D., “Hypersonic Inviscid Flowfields: Approximate Methods,” Hypersonic and High Temperature Gas
Dynamics, 1st ed., McGraw-Hill, Inc., New York, 1989, pp. 76-147.
[5] Savino, R., and Paterna, D., “Blunted Cone-Flare in Hypersonic Flow,” Elsevier Computers & Fluids, published 23
November 2004; Vol. 34, pp. 859-875.
DOI:10.1016/j.compfluid.2004.05.012
[6] Tissera, S., Drikakis, D., and Birch, T., “Computational Fluid Dynamics Methods for Hypersonic Flow Around
Blunted-Cone-Cylinder-Flare,” Journal of Spacecraft and Rockets, published July 2010; Vol. 47, No. 4, 2010, pp.
563-570.
DOI: 10.2514/1.46722
[7] Nompelis, I., Candler, G. V., and Holden, M. S., “Effect of Vibrational Nonequilibrium on Hypersonic Double-
Cone Experiments,” AIAA Journal, published November 2003; Vol. 41, No. 11, pp. 2162-2169.
[8] Kirk, B., and Carey, G. F, “Validation of Fully Implicit, Parallel Finite Element Simulations of Laminar Hypersonic
Flows,” AIAA Journal, published June 2010; Vol. 48, No. 6, pp. 1025-1036.
DOI: 10.2514/1.40860
[9] Bertram, M. G., “Correlation Graphs for Supersonic Flow Around Right Circular Cones at Zero Yaw in Air as a
Perfect Gas,” NASA TN D-2339, June 1964.

You might also like