ArticlePDF Available

A new strategy for solving store separation problems using OpenFOAM

Authors:

Abstract and Figures

The ability of OpenFOAM to solve the problem of a store separating from an air vehicle (store separation problem) has been evaluated using a dynamic mesh (Overset/Chimera) technique for an industry-class (transonic and generic) benchmark test case. The major limitations of the standard libraries have been determined. To tackle these challenges, a new strategy has been proposed and implemented using only open-source libraries and tools. The strategy combines porting, modifying, and adapting an overset library from the OpenFOAM fork platform (foam-extend) to the standard OpenFOAM platform (ESI). Furthermore, in order to overcome the well-known weakness of the standard OpenFOAM compressible solvers, the newly adapted overset library was integrated with an open-source, density-based, and coupled solver ( HiSA), which uses the OpenFOAM technology. Additionally, a force restrained model was developed to consider the externally applied forces on the store by the store ejectors. The accuracy of the developed strategy has been compared with wind tunnel tests and the solutions of two well-known commercial codes, showing good agreements with them. While the study has focused on simulations with inviscid Euler equations (typical of the test case considered here), the viscosity effect on the solution has also been studied with Navier–Stokes equations and compared with other results in the literature, showing minor differences. To the best of the authors’ knowledge, this is the first work which studies and validates the store separation problem in transonic regime with OpenFOAM.
Content may be subject to copyright.
Original Article
Proc IMechE Part G:
J Aerospace Engineering
2022, Vol. 0(0) 115
© IMechE 2022
Article reuse guidelines:
sagepub.com/journals-permissions
DOI: 10.1177/09544100221080771
journals.sagepub.com/home/pig
A new strategy for solving store separation
problems using OpenFOAM
Saleh Abuhanieh
1,2
, Hasan U. Akay
1
and Barıs¸ Bicer
2
Abstract
The ability of OpenFOAM to solve the problem of a store separating from an air vehicle (store separation problem) has
been evaluated using a dynamic mesh (Overset/Chimera) technique for an industry-class (transonic and generic)
benchmark test case. The major limitations of the standard libraries have been determined. To tackle these challenges,
a new strategy has been proposed and implemented using only open-source libraries and tools. The strategy combines
porting, modifying, and adapting an overset library from the OpenFOAM fork platform (foam-extend) to the standard
OpenFOAM platform (ESI). Furthermore, in order to overcome the well-known weakness of the standard OpenFOAM
compressible solvers, the newly adapted overset library was integrated with an open-source, density-based, and coupled
solver (HiSA), which uses the OpenFOAM technology. Additionally, a force restrained model was developed to consider
the externally applied forces on the store by the store ejectors. The accuracy of the developed strategy has been compared
with wind tunnel tests and the solutions of two well-known commercial codes, showing good agreements with them. While
the study has focused on simulations with inviscid Euler equations (typical of the test case considered here), the viscosity
effect on the solution has also been studied with NavierStokes equations and compared with other results in the literature,
showing minor differences. To the best of the authorsknowledge, this is the rst work which studies and validates the
store separation problem in transonic regime with OpenFOAM.
Keywords
trajectory prediction, large mesh movement, overset method, compressible ow, parallel computing, open-source tools
Date received: 12 July 2021; revised: 11 December 2021; accepted: 25 January 2022
Introduction
Safe separation of stores is a crucial mission for many air
vehicles. Predicting the trajectory of the store (to decide
the safe separation envelop) can be done using wind tunnel
testings
1
and ight tests.
2
However, both methods are
costly and raise some safety issues. The recent develop-
ments in computer hardware as well as parallel Compu-
tational Fluid Dynamics (CFD) algorithms have made
computer simulations possible and more feasible. This
reduced the need for these methods.
There are two CFD approaches to solve this problem,
which are inherited from the wind tunnel testing.
3
The rst
approach is the off-line (or grid survey)
4
and the second
approach is the on-line (or captive trajectory simulation).
5
In the off-line approach, an aerodynamic database of
forces and moments are obtained by solving the ow elds
(as a steady-state case) over a static mesh for different
scenarios. The scenarios (or grid test matrix) are mixture
of different Mach numbers, altitudes, angle of attacks, side
angles, store positions, and orientations. Finding an op-
timum set of scenarios is normally obtained using the
Modern Design of Experiments (MDOE) method.
3
After
creating the aerodynamic database, a six degrees of
freedom (6-DOF) solver is used to obtain the linear dis-
placements, linear velocities, angular displacements, and
angular velocities. For the on-line approach, a transient
simulation which utilizes a dynamic mesh technique is re-
quired. Since the resulting deformations are normally large,
the suitable dynamic mesh techniques are either based on
a deformable mesh (with re-meshing capabilities)
6,7
or an
overset/chimera
811
technique. At each time step, the ow
elds are computed, the forces and moments are calculated,
and using a 6-DOF solver, the displacements and velocities
are calculated. Finally, the mesh is moved (or deformed/
re-meshed) according to these displacements.
In general, if the required number of cases to decide the
safe separation envelop is enormous for the same ge-
ometry, using the rst approach by generating large da-
tabase will be efcient. However, if the number of cases
1
Department of Mechanical Engineering, Atilim University, Turkey
2
Turkish Aerospace, OpenSource CFD Group, Turkey
Corresponding author:
Saleh Abuhanieh, Turkish Aerospace, OpenSource CFD Group, Ankara,
Turkey.
Email: salehkhairisaleh.abuhanieh@tai.com.tr
are not so many, and for different geometries, the second
approach can be more efcient. Moreover, the second
approach is time-accurate. In this study, the second ap-
proach is used.
Although OpenFOAM
1
is a very popular and suc-
cessful open-source platform for CFD, it has been rarely
used for solving the store separation problem. The work of
Wadibhasme
12
can be mentioned, where the Mesquite
dynamic mesh library of Menon
13
was used, which was
available under OpenFOAM in earlier releases. However,
Wadibhasme solved only a sample projectile case using an
incompressible solver (pimpleDyMFoam) for demon-
stration. The reasons that OpenFOAM is not used widely
in store separation problems are twofold. Firstly, store
separation analysis is normally required in transonic and
supersonic regimes, whereas in OpenFOAM, there is
a well-known limitation in its standard compressible
solvers, such that there are no density-based coupled
compressible solvers, which are normally used in these
regimes. Secondly, the OpenFOAM dynamic mesh li-
braries have limited capabilities in this area.
Because of these limitations, OpenFOAM was not used
in the past for store separation simulations, neither was
compared before to commercial codes which are used
extensively for this kind of problems. Therefore, this work
carried out to address this gap in the existing literature.
Accordingly, the main objects of this study are as follows:
(a) To develop effective and accurate strategy in
OpenFOAM for solving the store separation problem.
(b) To compare the developed strategy with the standard
OpenFOAM.
(c) To compare the developed strategy with commercial
codes.
(d) To analyze the viscosity effect on store separation at
the transonic regime.
The Eglin
1
test case has been used for code validation
and comparisons due to the availability of the wind tunnel
results.
Problem formulation
Governing equations. The problem can be described as
solving a transient, compressible, and viscous ow over
a moving body. The governing equations can be written as
ρ½
tþ=ρvðÞ¼0 (1)
ðρvÞ
tþ=fρvvg¼ð=pÞþ=fτg(2)
½ρe
tþ=ðρveÞ¼=qþfσg:f=vg(3)
where ρis the density, tis the time, vis the velocity vector,
pis the pressure, τis the viscous stress tensor, eis the
specic internal energy, qis the heat ux vector, and σis
the mechanical stress tensor.
Additionally, a dynamic mesh model driven by the
6-DOF equations of motion, described below, is to be
integrated with the CFD model
XF¼d
dt mVðÞ (4)
XM¼d
dt HðÞ (5)
where Fand Mare the applied force and moment vectors
computed at the center of gravity of the store from the
CFD analysis, mVis the linear momentum vector, and His
the angular momentum vector. The linear and angular
displacement vectors of the store are then computed by
integrating equations (4) and (5), respectively, at each time
step.
The overset/chimera method
Since the rst time it was proposed to the aerospace
community four decades ago,
14
the overset method proved
to be a very useful technique in CFD. Its basic idea is to
assemble the computational domain using separate
meshes (sub-domains). It has been mainly used to solve
two problems. Firstly, to simplify the meshing of complex
geometries. In this case, each mesh can be prepared alone,
and that allows for generation of high-quality meshes
much easier including the block-structured meshes. Sec-
ondly, to simulate the cases where a solid body is moving
inside a computational domain. The latter capability has
been used widely to solve the store separation problems.
An example is shown in Figure 1, where a pitching airfoil
case is presented. The airfoil mesh (the overset mesh is in
red) is prepared separately and merged/overlapped with
a background mesh (in blue). This conguration allows
the airfoil to pitch or to move signicantly during the
simulation without any need for re-meshing.
The main task in the overset method is to connect the
multiple meshes in a single computational domain where
the discretized governing equations are solved. This
process has been commonly termed in the literature as the
Overset Grid Assembly (OGA).
1518
In OpenFOAM, the
same has been called as the cell-to-cell mapping.
2
The OGA process can be divided into the following steps:
Figure 1. The pitching airfoil case. The background mesh is
in blue and the overset mesh is in red.
2Proc IMechE Part G: J Aerospace Engineering 0(0)
(i) Hole identication: Finding the cells which rep-
resent the solid bodies or the cells which are located
outside the computational domain. These cells are
excluded from the calculation (hole cells
2
in
OpenFOAM).
(ii) Fringe construction: Deciding the cells which shall
receive the information from the other mesh(s).
These cells are called the fringe, receptors, acceptors,
or in OpenFOAM, they are called the interpolated
cells
2
.
(iii) Donors search: Identifying the cells which deliver
the information to the interpolated cells (the donor
cells
2
in OpenFOAM). Normally, those are the cells
which their centers are the closest to the interpolated
cells centers; however, other criteria can be used too.
The cells which are not hole cells and not interpolated
are considered as calculated cells.
2
For these cells, the
governing equations is solved. Thus, the donor cells are
under the calculated cells category.
The output information from this process is usually
called the Domain Connectivity Information (DCI).
15,18
In
the OpenFOAM code and literature, the names cellTypes
and cell classications
19
have been used as well.
The DCI for the pitching airfoil case is presented in
Figure 2. For the overset mesh (Figure 2(a)), all the cells
are calculated (in blue) except the cells at the outer
boundary (overset patch). These cells are interpolated (in
yellow) from the background mesh. For this case, there are
no hole cells in the overset mesh.
For the background mesh (Figure 2(b)), the hole cells
(in red), which are the projection of the airfoil, are ex-
cluded from the computation. The neighbor cells of the
hole cells are considered as interpolated cells. The re-
maining cells are the calculated cells.
Once the DCI are computed, the donor cells (e.g.,
D
0
D
4
in Figure 3) for each interpolated cell (e.g., I
0
in
Figure 3) are known. This is often called the interpolation
stencil.
The next step, which is normally at the linear system
solver level, is to use the interpolation stencil to evaluate
the required elds values at each interpolated cell. Dif-
ferent interpolation schemes can be used, for instance, the
inverse distance is expressed as:
ξI0¼X
ND
i¼1
ðωDiÞξDi(6)
where ξI0is the value of the eld variable ξat the in-
terpolated cell I
0
,N
D
is the total number of donor cells for
cell I
0
,ωDiis weight contribution by the donor cell D
i
, and
nally, ξDiis the value of ξat the donor cell D
i
. The weight
is dened as:
ωDi¼
1
dDj

P
ND
j¼1
1
dDj

(7)
where d
Di
is the distance between the donor D
i
cell center
to the interpolated cell center, the denominator is the sum
of all the donorsinverse distances.
The test case
Eglin
1
is by far the most dominating test case for code
validations for the store separation problem. That can be
attributed to the availability of the geometry, the experi-
mental test reports, and many published numerical results.
Basically, it is a wind tunnel test which was conducted in
Arnold Engineering Development Center (AEDC) in 1990.
The test object includes wing, pylon, and nned store
(Figure 4). The test results are available at two free-stream
Mach number (Ma): 0.95 and 1.2, while the angle of attack
was xed to 0
o
for the two tests. In this work, the transonic
(Ma = 0.95) test case has been used where the store tra-
jectory data is available from time = 0.0 s to time = 0.33 s.
Geometry
The full-scale geometry as per the test report
1
has been
created using OpenVSP
3
(for the airfoils) and Salome.
4
Both codes are free and open-source. The combined wing,
pylon, and store geometry are shown in Figure 4.
Figure 2. The overset DCI for the pitching airfoil case. The
overset/store mesh (a) and the background mesh (b). The
calculated cells are in blue, the interpolated cells are in yellow,
and the hole cells are in red.
Figure 3. The interpolation stencil of cell I
0
consists of the
donors D
0
D
4
.
Abuhanieh et al. 3
The proposed strategy
The preliminary results of this work, which have been
presented in Ref. 20, indicate clearly that the standard
OpenFOAM implementation has two main limitations:
(i) The ow solver: More accurate and stable com-
pressible solver is required.
(ii) The OGA algorithm: A more efcient, faster and
robust algorithm is required that can classify the cells
accurately for complex cases. Having an accurate
classication/DCI would improve the accuracy of the
results as well.
More details on the observed limitations may be seen in
Appendix II.
The ow solver. HiSA
5
is an external (non-standard
OpenFOAM) open-source and free solver which utilizes
the OpenFOAM libraries. It is a density-based and cou-
pled solver which can solve both transient and steady-state
cases. It has been developed at the Council for Scientic
and Industrial Research, South Africa (CSIR).
6
HiSA
solver has been veried and used here together with the
overset method. The code is also parallelizable through
OpenFOAM parallel library.
Similar to most density-based coupled solvers, it solves
the following coupled vector form of conservation of
mass, momentum, and energy equations:
W
tþ=FWðÞ¼QWðÞ (8)
where Wis the conservative variables vector, F(W) is the
ux vector, and Q(W) is the source terms vector. More
information about the theory, code, and many validation
cases can be found in Refs. 21 and 22.
The adapted OGA algorithm. Under the OpenFOAM
platform, there is an alternative open-source overset li-
brary in the foam-extend fork.
7
Its algorithm is more
robust and faster than the standard OpenFOAM overset
library. The rst step was to test the capability of this
library. After that, the library was ported to OpenFOAM,
since many data structures were not compatible between
the two forks of OpenFOAM. The third step was to im-
plement this library as a new cellCellStencilinside the
openFOAM overset library as shown in the block diagram
in Figure 5. For this purpose, the interface class foa-
mExtendStencilhas been developed.
The implementation in this way allows the selection of
the adapted OGA at run time, without changing any line of
code in the standard OpenFOAM. The developed interface
class contains oversetMeshobject from the ported
overset library as shown in Listing 1. After reading all
necessary data from the case folder (e.g., fringes selecting
settings, hole patches, and the cells in each overlapped
mesh), the object is initialized (Listing 2). Any Open-
FOAM solver, which supports the dynamic meshes, calls
the update()function at each time step; thus, the OGA-
related part is written inside the inherited update()
function. The oversetMeshobject provides the required
DCI information for each mesh (Listing 3). Finally, the
interpolation stencil depicted in Figure 3 and the inter-
polation weights from each donor (according to the se-
lected overset interpolation scheme) are returned to the
ow solver (Listing 4).
Listing 1: Declaration of the adapted overset object
(foamExtendStencil.H).
Listing 2: Initialization of the adapted overset object
(foamExtendStencil.C).
Listing 3: Sample code for obtaining the DCI infor-
mation from the adapted overset object
(foamExtendStencil.C).
Figure 4. Eglin test case full-scale geometry.
Figure 5. Block diagram showing the adapted foam-extend
overset library as a new cellCellStencilinside OpenFOAM in
blue color.
4Proc IMechE Part G: J Aerospace Engineering 0(0)
Listing 4: Sample code for transferring the interpolation
weights to the ow solver (foamExtendStencil.C).
Listing 5: Sample code for using the globalCellCells
function (foamExtendStencil.C).
A shortcoming has been discovered during the testing
is that foam-extend library was not able to collect all the
donors from all the processors in case of parallel run (step
(iii) in The overset/chimera method section). As a conse-
quence, the results were changing according to the number
of the used cores.
This problem has been resolved by getting only the
main donors (D
0
in Figure 3) from the oversetMesh
object. Then, for collecting the remaining/extended do-
nors (D
1
D
4
in Figure 3) for each main donor, the
globalCellCellsfunction from the main class cell-
CellStencilhave been utilized properly (Listing 5). This
function uses the faces of the cell to nd its neighbors
(face-walk). Furthermore, it implements all the necessary
synchronization tasks between the processors in case of
parallel executions.
The nal step was to integrate the new library with the
HiSA solver. During the early stages of the HiSA code
investigation, it was observed by the authors that the solver
in principle shall be able to run any dynamic mesh library
from the standard OpenFOAM platform. Since the standard
OpenFOAM overset class (dynamicOversetFvMesh)is
derived from the dynamic mesh class, and it uses the
cellCellStencilobjects for performing the OGA op-
erations only, a proper implementation for the new OGA
was sufcient. Thus, the HiSA solver works with new
library without any modication on its source code. That
was an advantage of using the standard OpenFOAM as
the common development platform.
The differences between the adapted OGA library
which was developed in this work and the original foam-
extend library can be summarized as follows:
(i) The functionality for nding the extended donors
cells is different as explained previously.
(ii) In order to maintain the compatibility with the
standard OpenFOAM, the overset interpolation
functionality was implemented in a way that the
function is called for each acceptor. Thus, the foam-
extend interpolation methodology, where one func-
tion call calculates the weights for all acceptors, was
not used.
(iii) In some cases, the overset is used only to assemble
the computational domain from different meshes,
andnomotionisinvolved.Inthiscase,runningthe
OGA algorithm every time step is unnecessary.
Thereisanoptioninthefoam-extendlibraryto
prevent the assembly again after the rst time
(cacheFringe); however, it works only with
a specic fringe type (overlap). This option was
not used in this work; instead, this functionality was
implemented at a higher level inside the foa-
mExtendStencilclass. In this way, the overall
OGA algorithm works only at the rst time step,
while the interpolation (computationally much
cheaper than the OGA) is done at each time step as
usual. Furthermore, this functionality has been
extended by introducing a user-dened variable
(oversetFrequency). This extension can be used
to reduce the overhead for the transient cases too
by executing the OGA every N time steps instead
of each time step. Lijewski
23
compared the
overhead and the accuracy of assembling the OGA
every 2 and 10 time steps. According to his
conclusion, the reduction in accuracy was small.
Comparison with the standard OpenFOAM
After checking all the standard solvers and the dynamic
mesh libraries in OpenFOAM, the only available option
to solve the store separation problem has been by using
the overRhoPimpleDyMFoam solver. This solver is
a segregated, pressure-based compressible solver which
supports the overset/chimera dynamic mesh library.
Mesh. To create the mesh, different open-source tools
have been used, including Tetgen,
24
SUMO,
25
cfMesh,
8
and snappyHexMesh.
9
After several trials, snappyHexMesh
(unstructured, octree-based, and hexa-dominant) has
been selected for mesh generation. Compared to the other
meshing tools mentioned above, it retrieves the surface
mesh with good quality while keeping the minimum cell
volume relatively high and the number of cells relatively
low. The minimum cell size has a direct effect on the
computation time of the overRhoPimpleDyMFoam solver
since it is a segregated solver. Although it is an implicit
solver, the maximum Courant number, which can guarantee
a stable run is still limited because of the complexity of the
overset scheme. Increasing the minimum cell size reduces
the maximum Courant number, which guarantees a more
stable run. The Courant number dened in OpenFOAM is
expressed as:
Co ¼Δtλ;λ¼1
2VX
faces
χfacei
(9)
Abuhanieh et al. 5
where Co is the Courant number, Δtis the time step size, λ
is the characteristic time scale, Vis the cell volume, and χ
is the volumetric ux at each face. The maximum Courant
number is the highest value of Co among all cells.
For comparison purposes with the developed strategy,
and considering the high computational time of the
overRhoPimpleDyMFoam solver, including the overset
overheads (see Appendix II), a very coarse mesh with
0.58 Mcells has been generated and tested. A slice from
this mesh is shown in Figure 6.
Case setup and the standard OGA algorithm. For the sake of
conciseness, the case setup and the details of the used
OGA algorithm are included in Appendix II. The case
parameters such as mass, center of mass, and moment of
inertia have been taken from the test report.
1
Results
Steady-state studies
First, the steady-state Euler solution has been obtained.
Since the overset in OpenFOAM can be used only with
transient cases, a transient simulation (without motion) is
driven to steady-state by running the case until the re-
siduals reach convergence. The residuals are plotted in
Figure 7, where the density-based HiSA solver used with
the developed strategy yields the density residuals, while
the pressure-based solver overRhoPimpleDyMFoam used
with the standard OpenFOAM yields the pressure
residuals.
The pressure coefcient (C
p
) at roll angle is plotted
in Figure 8 for the standard OpenFOAM (over-
RhoPimpleDyMFoam plus the cellVolumeWeight OGA)
and for the developed strategy (HiSA plus the adapted
OGA) compared with the experiment. Although the mesh
is coarse, the developed strategy solution is able to match
the C
p
trend of the experiment. However, for the standard
OpenFOAM solution, there is a signicant deviation
specially in the middle section. That can be attributed to
the inaccuracy of the OGA algorithm, since the roll
angle resides in the small region between the store and the
pylon. Examining Figure 33(a) shows that this region was
not solved by the ow solver; the cells there are either
interpolated or holes. Thus, the value of the elds there
take the free-stream values. That explains the value of zero
for the (C
p
) at this region in Figure 8. This example il-
lustrates the direct effect of the OGA accuracy on the
resultsaccuracy.
Transient simulations
After initializing the eld variables by the obtained steady-
state solution, the actual transient run started by applying
6-DOF motion solver. The Mach number at time = 0.17 s
over a plane is plotted in Figure 9 for the standard
OpenFOAM. A diffused solution can be observed in the
store mesh near the outer patch (the outer boundaries of
the red box in Figure 6). This can be attributed to the rst-
order overset interpolation scheme which has been used in
this case.
Unlike the standard OpenFOAM results, with the
developed strategy shown in Figure 10, the outer patch of
the store cannot be distinguished, primarily due to the
Figure 6. A slice from the 0.58 Mcells mesh. The background
mesh is in blue and the overset mesh is in red.
Figure 7. The residuals convergence for the developed
strategy and the standard OpenFOAM.
Figure 8. Pressure coefcient versus X/L for store at time =
0.0 s and roll angle = 5
o
for the standard OpenFOAM and the
developed strategy using the 0.58 Mcells mesh.
Figure 9. Mach number plot at time = 0.17 s (plane: y normal
at the initial center of mass) for the standard OpenFOAM.
6Proc IMechE Part G: J Aerospace Engineering 0(0)
higher order interpolation scheme used (inverse distance).
Furthermore, the shocks are more observable, thanks to
the used second-order upwind scheme (AUSM
+
-up).
26
A
brief description of this scheme is presented in Appendix III.
The linear displacements of the store versus time are
plotted in Figures 1113. Compared with the experiment,
both results are in good agreement. However, the results of
the developed version of the code are better in all three
directions.
The angular displacements of the store versus time are
plotted in Figures 1416, where a general agreement with
the experiment can be observed for the two results. The
developed strategy results are better for pitch and yaw
angles. However, for roll angle, the standard OpenFOAM
shows a better agreement with the experiment. That may
be attributed to the smaller value of the moment of inertia
for rolling, compared to the pitching and yawing (27 kg.m
2
versus 488 kg.m
2
for the Eglin case store at full-scale
1
).
That makes the roll angle too sensitive to the errors in
aerodynamic force calculations. Furthermore, with the
absence of results with ner meshes for the standard
OpenFOAM, it is hard to evaluate the reliability of this
Figure 10. Mach number plot at time = 0.17 s (plane: y
normal at the initial center of mass) for the developed strategy.
Figure 11. Linear displacement in the x-direction for the
store with comparison to the experiment for the standard
OpenFOAM and the developed strategy.
Figure 12. Linear displacement in the y-direction for the
store with comparison to the experiment for the standard
OpenFOAM and the developed strategy.
Figure 13. Linear displacement in the z-direction for the
store with comparison to the experiment for the standard
OpenFOAM and the developed strategy.
Figure 14. Roll angle of the store with comparison to the
experiment for the standard OpenFOAM and the developed
strategy.
Figure 15. Pitch angle of the store with comparison to the
experiment for the standard OpenFOAM and the developed
strategy.
Figure 16. Yaw angle of the store with comparison to the
experiment for the standard OpenFOAM and the developed
strategy.
Abuhanieh et al. 7
particular behavior. Unlike the linear displacements, the
accuracy of the angular displacements depends on the
accuracy of the calculated moments from CFD, which in
turn depends signicantly on nding the accurate shock
location. That is why obtaining accurate results for angular
displacements is more challenging.
Code performance
The developed strategy improves the speed of solving the
store separation problems in two ways:
(i) The adapted OGA algorithm is faster. Table 1 shows
a comparison between the two algorithms (the
standard OpenFOAM cellVolumeWeight and the
adapted one) using the same node on TRUBA HPC
10
(mesh size 0.58 M) for one time step. It is clear that
the improved algorithm is much faster for this mesh
conguration (at least 7X). However, the effective
speedup with larger size meshes may need further
investigation.
(ii) The used ow solver is coupled and is able to solve the
problem with a higher Courant numbers, which allows
obtaining solutions with larger time steps.
Accuracy studies for the developed strategy. In this section,
the accuracy of the developed version of OpenFOAM and
its mesh independency have been studied by considering
different levels of mesh renements and comparing the
results with the Eglin test case experiments. The results are
also compared with the solutions of two popular com-
mercial CFD codes, starCCM,
10
and Fluent,
11
that have
also used an overset algorithm for this test case.
Four different meshes (coarse with 1.04 Mcells, me-
dium with 1.62 Mcells, ne with 4.48 Mcells, and very-
ne with 6.20 Mcells) have been generated for this study
using snappyHexMesh. A slice from the ne mesh (at
center of mass and y normal) is shown in Figure 17.
Case setup and the OGA algorithm results
For the sake of conciseness, the case setup details and the
used OGA algorithm results are included in Appendix II.
Results
Steady-state simulations
First, the steady-state (Euler) solution has been obtained.
The pressure coefcient (C
p
) at roll angle is plotted in
Figure 18, and compared with the experiment, the
starCCM code results in Ref. 10, and the Fluent code
results in Ref. 11.
It can be seen that the developed strategy results for C
p
agree quite well with the experimental results and with the
two referenced numerical solutions. Apparently, the two
shocks at X/L; 0.27 and at X/L; 0.67 have been captured
well. The three Euler numerical results (and many others,
for example, Refs. 6,8, and 9) show deviation in the small
region (X/L; 0.87 X/L; 0.94) or the last two measurement
points in the experiment report. This deviation can be
attributed to the viscosity effect (see The viscosity effect),
since in literature, for example, Refs. 6and 27, the Navier
Stokes equations have been solved and there was an
improvement in that particular region for the same Mach
number and roll angle. This roll angle (f=5
o
) is of special
interest, since it is located in the tiny gap between the store
and pylon.
28
Unlike the viscosity effect, adding the sting
(shown in Figure 19), which was gaged to the store to
measure the pitching and yawing moments
1
during the
testing, apparently did not improve the results
appreciably.
8,9
There is a third shock which appears at the tail of the
store (at X/L; 0.98), which can be observed clearly in
Figure 20; however, there is no measurement point at this
location in the experimental data.
Table 1. Timing comparison between the standard
OpenFOAM cellVolumeWeight OGA algorithm and the adapted
foam-extend one.
The standard
OpenFOAM
The adapted
foam-extend Standard/adapted
time ratio
# Of cores Time [s) # Of cores Time [s) [-]
1 48.0 1 5.75 8.35
4 33.0 4 4.54 7.27
8 21.0 8 3.02 6.95
16 19.0 16 2.19 8.68
Figure 17. Slice from the ne mesh (4.48 Mcells), the
background mesh in blue and the overset mesh in red.
Figure 18. Pressure coefcient versus X/L for store at time =
0.0 s and roll angle = 5
o
for the developed strategy (ne mesh),
starCCM, and Fluent.
8Proc IMechE Part G: J Aerospace Engineering 0(0)
For the small oscillations observed in the present re-
sults, the authors have noticed similar oscillations in other
cases too while using snappyHexMesh for mesh gener-
ation. The octree concept used by snappyHexMesh does
not preserve the input surface mesh; instead, it snaps the
surface of the cells after cutting them. That can be the
reason for such oscillations which may be considered as
a minor post-processing issue.
In Figure 20, the Mach number contour for steady case
is plotted at the lower side of the wing showing the store as
well. It may be observed that multiple shocks are ap-
pearing there and they are not symmetrical around the
store. From this view (normal to z-direction), it can be
observed that the un-symmetry of the shocks around the
store is the main reason of yawing the store during
separation.
Transient simulations
The next step, is to start the transient simulation by
applying the 6-DOF motion solver. The effect of the
mesh renement level on the solution has been studied
here. The angular displacements obtained by solving the
Euler equations for the four meshes (coarse, medium,
ne, and very-ne) are compared in Figures 2123,
respectively. As may be observed, in general, the solution
is converging monotonically toward the experimental
resultsasthemeshisrened. That indicates that the
discretization error is decreasing. Additionally, minor
differences between the solutions of the ne and very-
ne meshes indicate that the mesh convergence of the
solution is also reached.
The change of store location with time is shown in
Figure 24. The blue color represents the experimental
results and the Mach number contours represent the ob-
tained CFD results for the ne mesh. Minor deviations
Figure 20. Mach number contour at the lower side of the
wing and store at time = 0.0 s for the ne mesh.
Figure 21. Roll angle of the store. Comparison between the
Euler solutions for the four different meshes.
Figure 22. Pitch angle of the store. Comparison between the
Euler solutions for the four different meshes.
Figure 23. Yaw angle of the store. Comparison between the
Euler solutions for the four different meshes.
Figure 19. The used sting to measure the pitching and
yawing moments attached to the tail of the tested/metric store.
1
Dimensions are in inches (1/20 scale).
Figure 24. Store location with time. The Mach number
contours represent the ne mesh results and the experimental
results are in blue.
Abuhanieh et al. 9
from the experiment can be noticed, which reects that
accurate displacements are obtained.
To highlight the importance of the applied external
forces to the store by the ejectors during separation, the
Center of Pressure (CoP) has been calculated at time =
0.05 s using the following relation:
CoP ¼Zxp
Zp
(10)
where CoP is the center of pressure coordinate vector, xis
the point coordinate vector, and pis the static pressure.
Figure 25 shows the location of the calculated CoP (the
red sphere). Since the center of mass for the store (the
black sphere) is located after the CoP and the lift force
direction is upward at this time step, the store is expected
to pitch nose-down. However, as per the results (e.g.,
Figure 22), the store is pitching nose-up. That is simply
due to the external forces, and the same can be veried
easily by calculating the total moments on the store with
and without the ejector forces. To incorporate the external
forces by the ejectors, a new force restrained model
(ejectorExternalForce) has been developed in this work.
This model applies a xed force on a specic point which
moves with the body up to a certain length (the stroke
length specied in the experiment).
Since the angular displacements are normally con-
sidered the critical results for this kind of analysis, the
developed strategy results for the ne mesh are compared
to the experiment, the starCCM code results in Ref. 10 and
the Fluent code results in Ref. 11 as shown in Figures 2628,
for roll, pitch and yaw angles, respectively.
It may be noticed that for the three angles, the de-
veloped strategy results are comparable to the results of
starCCM and Fluent. Both codes used the overset as
a dynamic mesh technique to obtain the results presented
here.
The viscosity effect. To evaluate the viscosity effect on this
particular test case, the Euler solution of the ne mesh
(4.48 M) has been compared to the NavierStokes solution.
The average y
+
(the non-dimensional wall distance) for
this mesh is around 200; thus, wall functions were used.
Furthermore, the kωSST (shear stress transport) tur-
bulence model
29
has been used.
Figure 25. The calculated center of pressure (the red sphere)
and the center of mass (the black sphere) at time = 0.05 s. The
contour plot shows the static pressure on the store.
Figure 26. Roll angle of the store with comparison to the
experiment and two commercial codes for the ne mesh.
Figure 27. Pitch angle of the store with comparison to the
experiment and two commercial codes for the ne mesh.
Figure 28. Yaw angle of the store with comparison to the
experiment and two commercial codes for the ne mesh.
Figure 29. Pressure coefcient versus X/L for store at time
= 0.0 s and roll angle = 5
o
for the ne mesh. Comparison
between the Euler and the NavierStokes solutions.
10 Proc IMechE Part G: J Aerospace Engineering 0(0)
For the steady-state solution, Figure 29 shows the
pressure coefcient (C
p
) comparison at roll angle 5°.
Compared to the experiment, for the NavierStokes so-
lution, the second shock at X/L; 0.67 is slightly shifted to
the right. That can be attributed to the high y
+
mesh used.
Aner mesh in the viscous region can provide a better
result. With comparison to the Euler solution, the shock
near the tail of the store at X/L; 0.98 is detected more
sharply. However, since no measurement points are avail-
able at the test report in this region, improved compar-
isons of the NavierStokes solution with experiments
cannot be conrmed at this time.
Unlike the developed strategy results for both Euler and
NavierStokes solutions, the results obtained by Sunay
et al.
6
and Pandya et al.
27
matched better with the ex-
periment, for the last two measurement points. Further-
more, in their solutions, the second shock at X/Lx0.67
and the third shock at X/Lx0.98 are weaker, but still
respect the experimental results. Consequently, it is
expected that solving the NavierStokes equations with an
appropriate viscous mesh can provide better results for the
steady-state solution.
For the transient run, the angular displacements results
are shown in Figures 3032 for roll, pitch, and yaw angles,
respectively. The work of Huang et al.
30
can be considered
as a comprehensive study which analyzed the effect of the
mesh (size and topology), the model scale, the wing
symmetry, and the turbulence models. Their best Navier
Stokes result for angular displacements (using Spalart
Allmaras turbulence model
31
) has been included along
the work of Sunay et al.
6
in these gures and compared
with the present results for both Euler and NavierStokes
simulations. Unlike the steady-state simulation, for the
trajectory prediction, it can be observed that solving the
viscous ow even with very well prepared meshes may not
necessarily provide better results than the Euler solution.
A similar observation can be made by reviewing the re-
sults of Figure 19 in Huang et al.
30
Conclusions
A new strategy which involves modifying and adapting
the foam-extend overset library, and integrating it with the
density-based coupled solver HiSA for solving the store
separation problem within the OpenFOAM platform has
been proposed and implemented. The new strategy is more
accurate and efcient than the standard OpenFOAM so-
lution for this problem; additionally, it is capable of
solving industry-scale cases/geometries. Furthermore, it is
faster at least by 7 times. A mesh renement study has
been conducted, and an improvement in the results have
been observed with increasing the number of cells, which
is a good sign about the accuracy of the developed
strategy. The obtained results agree well with the exper-
imental results and are comparable with the results of the
two well-known commercial codes. A NavierStokes
solution has been compared to the Euler solution. It is
shown that for the transonic regime, at least for this case,
the Euler solution for predicting the trajectory can provide
satisfactory results, which in practice can save time and
effort. Similar observations have been noticed in the
literature.
Acknowledgments
The numerical calculations reported in this paper were partially
performed on the resources of the ULAKBIM High Performance
and Grid Computing Center of The Scientic and Technological
Research Council of Turkey. Wethank the Turkish Aerospace for
the support provided for this study.
Declaration of Conicting Interests
The author(s) declared no potential conicts of interest with
respect to the research, authorship, and/or publication of this
article.
Funding
The author(s) received no nancial support for the research,
authorship, and/or publication of this article.
Figure 30. Roll angle of the store. Comparison between the
Euler and NavierStokes solutions for the ne mesh.
Figure 31. Pitch angle of the store. Comparison between the
Euler and NavierStokes solutions for the ne mesh.
Figure 32. Yaw angle of the store. Comparison between the
Euler and NavierStokes solutions for the ne mesh.
Abuhanieh et al. 11
ORCID iD
Saleh Abuhanieh https://orcid.org/0000-0002-3620-8546
Notes
1. www.openfoam.com
2. www.openfoam.com/documentation/guides/latest/doc/
guide-overset.html
3. www.openvsp.org
4. www.salome-platform.org
5. hisa.gitlab.io
6. www.csir.co.za
7. www.sourceforge.net/p/foam-extend
8. www.cfmesh.com
9. www.openfoamwiki.net/index.php/SnappyHexMesh
10. https://www.truba.gov.tr/index.php/en/main-page/
References
1. Heim ER. CFD Wing/pylon/nned Store Mutual Interference
Wind Tunnel Experiment. ADB152669 Technical Report 1991.
2. Arnold RJ and Epstein CS. AGARD ight test techniques
series on store separation ight testing. AGAR Dograph
1986; 5(300).
3. Jamison KA. Grid-mode transonic store separation analysis
using modern design of experiments. In: 31st Congress of
the International Council of the Aeronautical Sciences.
Brazil: Belo Horizonte, 2018.
4. Davids S and Cenko A. Grid based approach to store
separation. In: 19th AIAA Applied Aerodynamics Confer-
ence. Anaheim, CA, USA, 2001.
5. Panagiotopoulos EE and Spyridon DK. CFD transonic store
separation trajectory predictions with comparison to wind
tunnel investigations. Int J Eng 2010; 3(6): 538553.
6. Sunay YE, Gulay E and Akgul A. Numerical simulations of
store separation trajectories using the eglin test. Scientic
Tech Rev 2013; 63(1): 1016.
7. Dehghan M, Davari AR and Manshadi MD. Numerical
investigation on the weight, speed, and installation location
effects on fuel tank separation trajectory. Proceedings of the
Institution of Mechanical Engineers, J Aerospace Eng 2017;
231(13): 23312344.
8. Lijewski L and Suhs N. Chimera-eagle store separation. In:
AIAA Atmospheric Flight Mechanics Conference. SC,
USA: Hilton Head Island, 1992.
9. Demir H, Alemdaroglu N and Ozveren V. External store
separation from ghter aircraft. In: RTO AVT Symposium on
Functional and Mechanical Integration of Weapons and
Land and Air Vehicles. Williamsburg, VA, USA: RTO-MP-
AVT-108, 2004.
10. MacLucasa D and Gledhillb I. Time-accurate transonic CFD
simulation of a generic store release case. R D J South Afr
Inst Mech Eng 2018; 34: 916.
11. Khaware A, Sivanandham A and Gupta VK. Numerical
simulation of store separation trajectory for Eglin test case
using overset mesh. In: AIAA SciTech Forum. FL, USA:
AIAA Aerospace Sciences MeetingKissimmee; 2018.
12. Wadibhasme R. Exploration and Implementation of Various
Dynamicmesh in OpenFOAM (Master thesis). Centre for
Modeling and Simulation, Savitribai Phule Pune University;
2016.
13. Menon S. A numerical study of droplet formation and be-
havior using interface tracking methods interface tracking
method (PhD thesis). Mechanical and Industrial Engi-
neering. University of Massachusetts Amhers; 2011.
14. Benek JA, Steger JL and Dougherty FC. A exible grid
embedding technique with applications to the Euler equa-
tions. In: 6th Computational Fluid Dynamics Conference.
MA, USA: Danvers; 1983.
15. Martin JE, Noack RW and Carrica PM. Overset grid as-
sembly approach for scalable computational uid dynamics
with body motions. J Comput Phys 2019; 390: 297305.
16. Roget B and Sitaraman J. Robust and efcient overset grid
assembly for partitioned unstructured meshes. J Comput
Phys 2014; 260: 124.
17. Jingjing F and Chao Y. Enhancement and application of
overset grid assembly. Chin J Aeronautics 2010; 23(6):
631638.
18. Noack RW, Boger DA, Kunz RF, etal.Suggar++:Animproved
general overset grid assembly capability. In: AIAA Computa-
tional Fluid Dynamics. San Antonio, TX, USA; 2009.
19. Chandar D. Development of a parallel overset grid frame-
work for moving body simulations in OpenFOAM. J Appl
Computer Sci Mathematics 2015; 9(2): 2230.
20. Abuhanieh S, Akay HU and Bicer B. A new strategy for
solving store separation problems using OpenFOAM
(abstract), In: 32nd International Conference on Parallel
Computational Fluid Dynamics. Nice: France, 2021.
(virtual).
21. Heyns JA, Oxtoby OF and Steenkamp A. Modelling high-
speed ow using a matrix-free coupled solver. In: 9th
OpenFOAM Workshop. Zagreb, Croatia, 2014.
22. Abuhanieh S, Bicer B and Sahin M. A validation for the
OpenFOAM-Hisa solver for drag prediction (abstract), In:
32nd International Conference on Parallel Computational
Fluid Dynamics. Nice: France, 2021. (virtual).
23. Lijewski L. Comparison of transonic store separation tra-
jectory predictions using the Pegasus/DXEAGLE and
Beggar codes. In: 15th Applied Aerodynamics Conference.
Atlanta, GA, USA, 1997.
24. Si H. Tetgen, a delaunay-based quality tetrahedral mesh
generator. ACM Trans Math Softw 2015; 41(2): 136.
25. Tomac M and Eller D. Towards automated hybrid-prismatic
mesh generation. Proced Eng 2014; 82: 377389.
26. Liou M. A sequel to AUSM, part ii: AUSM
+
-up for all
speeds. J Comput Phys 2006; 214(1): 137+170.
27. Pandya MJ, Frink NT and Noack RW. Progress toward
overset-grid moving body capability for USM3D un-
structured ow solver. In: 17th AIAA Computational Fluid
Dynamics Conference. Toronto, ON: Canada, 2005.
28. Snyder DO, Koutsavdis EK and Anttonen JSR. Transonic
store separation using unstructured CFD with dynamic
meshing. In: 33rd AIAA Fluid Dynamics Conference and
Exhibit. Orlando, FL, USA, 2003.
29. Menter F, Kuntz M and Langtry R. Ten years of industrial
experience with the sst turbulence model. In: Proceedings of
the fourth international symposium on turbulence, heat and
mass transfer. Turkey: Antalya, 2003.
30. Huang H, Blyth RH, Prior MA, et al. A comparison of
approaches to multi-body relative motion using the Kestrel
solver. In: AIAA Scitech 2019 Forum. San Diego, CA, USA,
2019.
31. Spalart P and Allmaras S. 30th Aerospace Sciences Meeting
and Exhibit. RenoUSA: NV, 1992.A one-equation turbu-
lence model for aerodynamic ows
32. Liou M and Steffen C. A new ux splitting scheme.
J Comput Phys 1993; 107: 2339.
33. Liou M. A sequel to AUSM: AUSM
+
.J Comput Phys 1996;
129(2): 364382.
12 Proc IMechE Part G: J Aerospace Engineering 0(0)
Appendix I
Notations
Co. Courant number [-]
CoP center of pressure coordinate [m]
ddistance vector magnitude [m]
D
i
donor cell i [-]
especic internal energy [J/kg]
Fapplied forces vector [N]
F(W)ux vector (convective ux and viscous ux)
Hangular momentum vector [kg m
2
/s]
I
i
interpolated cell i [-]
Mapplied moments vector [Nm]
mVlinear momentum vector [kg m/s]
N
D
number of donors [-]
pstatic pressure [Pa]
qheat ux vector [W/m
2
]
Q(W) source terms vector
ttime [s]
vuid velocity vector [m/s]
Vcell volume [m
3
]
Wconservative uid ow variables vector
xcoordinate vector [m]
y
+
non-dimensional wall distance [-]
θpitch angle (around y) [
o
]
λthe characteristic time scale [1/s]
ρdensity [kg/m
3
]
σmechanical stress tensor [N/m
2
]
τviscous stress tensor [N/m
2
]
froll angle (around x) [
o
]
χface volumetric ux [m
3
/s]
ψyaw angle (around z) [
o
]
ξ
i
eld value at cell i
ω
i
interpolation weight of cell i [-]
Δttime step size [s].
Appendix II
overRhoPimpleDyMFoam case setup
The details of the setup for the case solved with over-
RhoPimpleDyMFoam are summarized in Tables 24.
The standard OGA
For this test case (Eglin 0.58 Mcells) only the rst-order
cellVolumeWeight scheme worked with the over-
RhoPimpleDyMFoam solver. In the OpenFOAM im-
plementation, the cellVolumeWeight is not only an
interpolation scheme, to be used by the interpolated cell to
get the value from the donors cells, but it includes the
OGA process, which has been described in The overset/
chimera method section. A slice showing the overset DCI
for the overset/store mesh (Figure 33(a)) and the back-
ground mesh (Figure 33(b)) is shown in Figure 33.
In this work, studying the current/available capabilities
of the standard OpenFOAM to solve the store separation
problem reveals the following limitations:
(A) The solution obtained by the overRhoPimpleDyMFoam
is reasonable. However, the accuracy is not good
enough.
(B) A converged solution with a second-order upwind-
ing scheme was not obtained. This can be attributed
to the limited numerical stability of the used seg-
regated ow solver.
(C) The computational time is high (842 core-hours for
the Eglin case with 0.58Mcells), and more than the
half of each time step execution time is consumed
by the overset processes. Thus, running a ne mesh
may easily turn the case to be non-practical to be
solved.
(D) Only the rst-order overset interpolation scheme was
used (cellVolumeWeight), since the other schemes
(inverseDistance and leastSquares) OGA algorithms
failed to classify the cells (either calculated, hole or
interpolated) correctly. A rst-order overset in-
terpolation scheme is not enough to obtain accurate
results.
Table 2. Solver settings.
Solver overRhoPimpleDyMFoam
OpenFOAM version v2006
Flow Transient inviscid
Discretization schemes First-order upwinding in space
and rst-order implicit in time
(backward euler)
Time step 1 × 10
4
s (maximum)
Maximum courant number 5.0
Pseudo time (inner loop)
settings
Maximum iterations = 30,
tolerance = 1 × 10
4
, there
is no dual time scheme option
Free-stream Mach number 0.95
Overset interpolation
scheme
cellVolumeWeight (rst-order)
Table 3. Boundary conditions.
Patch/eld Pressure Velocity Temperature
Fareld Free-stream
pressure
Free-stream
velocity
inletOutlet
wing_and_pylon zeroGradient Slip zeroGradient
Store zeroGradient movingWall
velocity
zeroGradient
oversetPatch Overset Overset Overset
Symm Symmetry Symmetry Symmetry
Table 4. dynamicMeshDict settings.
dynamicFvMesh dynamicOversetFvMesh
Solver sixDoFRigidBodyMotion
Abuhanieh et al. 13
The developed strategy case setup.
The details of the setup for the case solved with the
HiSA solver and the developed strategy are summarized in
Tables 5 and 6.
The adapted OGA results
For the adapted overset library, only one OGA algorithm is
available. However, for selecting the fringe (interpolated
cells) from each mesh, different methods can be selected
(manual, overlap, faceCells, etc.).
Unlike the standard OpenFOAM overset library, in this
implementation, the OGA algorithm is independent from the
interpolation scheme. In the present study, the interpolation
schemes which have been implemented in the overset library
are: inverseDistance and the leastSquares as shown in Figure 5
in blue under the interface class. The prexfoamExtendhas
been added to distinguish them from the ones used by the
standard library to be able to select them during run time. A
slice showing the DCI for the overset/store mesh (Figure
34(a)) and the background mesh (Figure 34(b))isshownin
Figure 34. The overlap fringe has been used here.
Appendix III
AUSM
+
-up Scheme
AUSM
+
-up is a second-order upwinding scheme for shock
capturing, which is an extension to the AUSM (Advection
Upstream Splitting Method) family of schemes.
32,33
It has
been developed to improve the prediction of ows at low
Mach speeds, making the original AUSM scheme work ef-
ciently at all speeds. The algorithm can be described as
follows
26
:
Table 5. Solver settings.
Solver HiSA (version 1.4.2)
OpenFOAM version v2006
Flow Transient inviscid
Discretization schemes Second-order upwinding in
space (AUSM
+
-up) and
rst-order implicit in time
(backward euler)
Time step 1 × 10
4
s (maximum)
Maximum courant number 60.0
Pseudo time (inner loop)
settings
Maximum iterations = 100,
tolerance = 5 × 10
3
,
using dual time scheme
Free-stream Mach number 0.95
Overset interpolation scheme foamExtend_inverseDistance
(rst-second-order)
dynamicMeshDict settings Same as the
overRhoPimpleDyMFoam
settings
Table 6. Boundary conditions.
Patch/eld Pressure Velocity Temperature
Fareld Characteristic fareld pressure Characteristic fareld velocity Characteristic fareld temperature
wing_and_pylon Characteristic WallPressure Slip Characteristic WallTemperature
Store Characteristic WallPressure movingWall velocity Characteristic WallTemperature
oversetPatch Overset Overset Overset
Symm Symmetry Symmetry Symmetry
Figure 33. The overset DCI for the overset/store mesh (a) and the background mesh (b). The calculated cells are in blue, the
interpolated cells are in yellow and the hole cells are in red (plane: y normal at the initial center of mass). The observed limitations for the
standard OpenFOAM.
14 Proc IMechE Part G: J Aerospace Engineering 0(0)
(i) For each interface, calculate the Mach numbers left
and right states
ML=R¼uL=R
a1=2
(11)
where M
L
/R are the Mach number left and right states, u
L
/
R are the convective velocities (vn), and a
1/2
is the speed
of sound at the interface (can be obtained by averaging the
a
L
and a
R
).
(ii) Calculate the mean Mach number
M2¼u2
Lþu2
R
2a2
1=2
(12)
(iii) Calculate the reference Mach number
M2
o¼min1,maxM2,M2
0;1(13)
where M
is the free-stream Mach number.
(iv) Find the scaling function value
fo¼Moð2MoÞ2½0;1(14)
(v) Evaluate the split Mach numbers (M
±
(4)
) for the left
Mþ
ð4ÞðMLÞand right M
ð4ÞðMRÞMach number states.
The used split Mach number is a fourth-order
polynomial function.
(vi) Calculate the interface Mach number
M1=2¼M
þ
ð4ÞðMLÞþM
ð4ÞðMRÞ
Kp
fa
max1σM2,0pRpL
ρ1=2a2
1=2
(15)
where K
p
and σare constants and p
R
and p
L
are the
pressure right and left states, respectively. ρ
1/2
is the av-
erage of the density left and right states.
(vii) Calculate the mass ux at the interface
_
m1=2¼a1=2M1=2
ρLif M1=2>0
ρRotherwise
(16)
(viii) Calculate the pressure ux
p1=2¼P
þ
ð5ÞðMLÞpLþP
ð5ÞðMRÞpR
kuPþ
ð5ÞðMLÞP
ð5ÞðMRÞðρLþρRÞfaa1=2ðuLuRÞ
(17)
where K
u
is constant. P±
ð5Þis a fth-order polynomial.
(ix) Finally, evaluate the total ux at the interface
f1=2¼_
m1=2
ψLif _
m1=2>0,
ψRotherwise þp1=2
(18)
where ψis a vector quantity convected by _
m. In 1-D, ψ=
(1,u,H)
T
, where His the total enthalpy. pis the pressure
ux vector, which contains only one pressure term, p= (0,
p,0)
T
in 1-D.
Figure 34. The overset DCI for the overset/store mesh (a) and the background mesh (b) for the ne mesh. The calculated cells are in
blue, the interpolated cells are in yellow, and the hole cells are in red (plane: y normal at the initial center of mass).
Abuhanieh et al. 15
... The details of the numerical implementation for all the computational tools used herein can be found in the first author's thesis. 40 For solving the cavity flow without the store, the "empty cavity" case, the HiSA CFD solver is used to solve the URANS system of equations (Equation (1)) with the IDDES / k-ω in Equations (8) and (9) as a closure turbulence model over a single domain/mesh. The used discretization schemes and boundary conditions are mentioned in the Results section. ...
Article
Full-text available
In this work, the ability of open-source CFD tools to conduct store separation simulations from cavities is evaluated and validated using a generic test case from the literature. Firstly, the ability and accuracy of these tools for solving cavity flows at high speeds are evaluated. Secondly, their competence in predicting the trajectory of a generic store from a generic deep cavity is checked. Finally, and in order to reduce the associated computational costs, a release-time dependability factor from the recent literature is studied and evaluated. The obtained results using the selected open-source CFD tools matched quite well with the wind tunnel results. Furthermore, the results show that predicting the release-time dependability using a quantified index/factor can be a potential remedy for reducing the computational cost for this type of CFD simulations.
Conference Paper
Full-text available
The analysis of the separation of a Precision Guided Munition (PGM) from many configurations of an advanced jet trainer was performed using aerodynamic data from wind-tunnel tests characterised using the grid method. As strong aerodynamic mutual interference is present due to transonic shockwaves between the wing of the aircraft and the tail of PGM the loads on the store changes significantly for different combinations of PGM position and orientation relative to the aircraft. This means that the grid method must sample a wide range of positions and orientations. If this is done in usual manner, the grid test matrix is large and costly. There is another method for efficiently characterising phenomena with a number of mutually interacting variables known as the Modern Design of Experiments (MDOE) which can significantly reduce the number of grid samples required. The possibility of developing the grid test matrix using the MDOE method is investigated using a simple panel code model. The correct approach to implement the MDOE grid method is identified and the relative interpolation errors are characterised. The application of the MDOE method to the trainer jet/PGM separation wind-tunnel test is described.
Article
Full-text available
Store release from a parent aircraft in the transonic regime is a complex transient interaction to simulate, due to both compressibility effects and the strong interference flow field generated between the parent and store bodies. This work presents the results from a time-accurate transonic numerical simulation of a generic store release case, with detailed attention to ejector force profiles. Confidence limits for time-accurate trajectory calculations with STAR-CCM+® have been improvedfor this standard benchmark case.
Thesis
Full-text available
OpenFOAM is an opensource solver written in C++ for solving fluid dynamics problems.In this thesis, the framework of OpenFOAM is explored through directory tree, solvers, utilities and simple case setup. After general introduction, concept of dynamicMesh is introduced and its importance in modeling the CFD problems has been discussed.Discussion on dyanamicMesh includes various types of dynamicMesh capabilities available in OpenFOAM like Mesh Motion, GGI and dynamicTopoFVMesh with their working principles.The dynamicMeshDict file provided with each type explains the various algorithms implemented to ensure the mesh motion during runtime eg. AdaptiveMeshReconnection,Edge swiping, solidBodyMotion.The case setup provided with dyanmicMesh explains the importance of various parameters of dynamicMesh and constrains of dynamicMeshDict file. Further explanation on setup of dynamicMesh explains the available predefined solid body motions like linearMotion, Oscillating Rotating Motion MultiMotion etc. and explains the setup of subsequent motion coefficients associated with each motion.Further insight on dynamicTopoFvMesh,which has capability of mesh motion with topology change is given and general understanding about the setup of dynamicMeshdict file is built. With the lesson learned from algorithms and file setup, Series of problems setup like ballTranslation and projectile are run. The results of ballTranslation case shows mesh topology changes due to motion of solid ball are handled with great accuracy.The projectile case implements pimpleDymFoam solver along with dynamicTopoFvMesh and also uses the 6-DOF motion solver to compute projectile motion at runtime. Projectile case also paves the way for sabot separation problem which need the dynamicMesh with topology change and 6-DOF motion solver to do the CFD analysis.
Conference Paper
Full-text available
The overset, or chimera, grid methodology utilizes a set of overlapping grids to discretize the solution domain. Parallel computation of the overset domain connectivity information (DCI) pose unique problems relative to the typical ow solver. The present paper investigates the parallel computation of the DCI within the context of Suggar++, a general capability for obtaining the overset domain connectivity information. Significant improvements to the donor search process are also presented that reduce the time required during the donor search phase. The effect of grid partitioning on the overset work is investigated and demonstrated to cause the work to increase. A new approach to partitioning for the overset domain connectivity assembly process is proposed and demonstrated to avoid the increase in work of the conventional partitioning. The parallel execution of Suggar++ is examined in detail and is found to provide a significantly improved parallel performance relative to the current capability in SUGGAR. Finally, a new approach to eliminate orphans in tight gap regions by having the overset grid assembly process identify solver locations to be treated as immersed boundary points is also presented.
Article
A methodology for the decomposition of the overset grid assembly problem is presented and evaluated. The method is based on identifying grid subsets of a computational domain for which each domain connectivity is calculated individually, and combined at run time to obtain the final complete solution. The proposed method is a framework for the partition of the problem based on the characteristics of the simulation and relies on the user's knowledge of the case, avoiding some of the pitfalls of automatic methods. The methodology was implemented in a computational fluid dynamics solver, allowing the scalable computation of very large simulations of moving bodies, with efficient use of the computational resources. Several examples illustrating the method and the performance improvements with respect to the standard assembly approach are presented.
Article
A numerical survey coupled with six degree-of-freedom flight simulation have been undertaken to study the fuel tank separation trajectory, released from a trainer airplane. Two different spanwise release points for the tank, near and farther from the fuselage under the starboard wing with full and empty fuel were considered. The studies were performed at two free stream Mach numbers of 0.23 and 0.42 at zero angle of attack. Dynamic unstructured tetrahedral mesh approach combined with spring-based smoothing and local remeshing was applied with an implicit, second-order upwind accurate Euler solver. A six degree-of-freedom routine using a fourth-order multi-point time integration scheme was coupled with the flow solver to update the payload trajectory information at each time step. According to the results, the payload installed farther from the fuselage falls down with a higher forward velocity than that located closer, once released from the wing. The spanwise installation point was also found to have a strong impact on the pitch attitude of the released payload. The payload weight has been shown to play a vital role in longitudinal-lateral coupling behavior and the associated moments on the released payload.
Article
TetGen is a C++ program for generating good quality tetrahedral meshes aimed to support numerical methods and scientific computing. The problem of quality tetrahedral mesh generation is challenged by many theoretical and practical issues. TetGen uses Delaunay-based algorithms which have theoretical guarantee of correctness. It can robustly handle arbitrary complex 3D geometries and is fast in practice. The source code of TetGen is freely available. This article presents the essential algorithms and techniques used to develop TetGen. The intended audience are researchers or developers in mesh generation or other related areas. It describes the key software components of TetGen, including an efficient tetrahedral mesh data structure, a set of enhanced local mesh operations (combination of flips and edge removal), and filtered exact geometric predicates. The essential algorithms include incremental Delaunay algorithms for inserting vertices, constrained Delaunay algorithms for inserting constraints (edges and triangles), a new edge recovery algorithm for recovering constraints, and a new constrained Delaunay refinement algorithm for adaptive quality tetrahedral mesh generation. Experimental examples as well as comparisons with other softwares are presented.
Conference Paper
Two time-accurate computational fluid dynamic prediction techniques are demonstrated for a store separation event at transonic speeds. Both overlapping and blocked grid approaches are used, coupled with an implicit Euler flow solver with a flux-differenc e split scheme based on Roe's approximate Riemann solver and a six-degree-of- freedom trajectory component. All major trends of the trajectory are captured and surface pressure distributions are predicted. Computational execution times for both approaches are discussed and code improvements are proposed. NOMENCLATURE