You are on page 1of 22

Solution to industry benchmark problems with the

lattice-Boltzmann code XFlow


David M. Holman1 , Ruddy M. Brionnaud1 , and Zaki Abiza1
1

Next Limit Technologies, Angel Cavero 2, 28043, Madrid (Spain)


david.holman@nextlimit.com

Abstract
This contribution presents some of the capabilities of the Computational
Fluid Dynamics (CFD) code XFlow, which uses a proprietary particle-based
kinetic solver based on the Lattice-Boltzmann Method. Using traditional
CFD software, industrial problems require time consuming meshing process
which often leads to errors or even divergence of the simulation. Due to its
particle-based and fully Lagrangian approach, the complexity of the geometry
surfaces is not a limiting factor in XFlow even in the presence of moving
parts, allowing to solve real industrial problems. The performance of XFlow
will be demonstrated for different industry benchmarks. The first example is
the Ahmed body which is a classical benchmark in the automotive industry.
The second benchmark presented will be the NASA trapezoidal wing. XFlow
results will be described and show good agreement with experimental data.
Keywords: Lattice-Boltzmann, Lagrangian, particle-based, Ahmed body,
NASA trapezoidal wing
1. Introduction
For the past 20 years, the field of Computational Fluid Dynamics (CFD)
has reached a high level of maturity, but it has only been recently that
CFD has been broadly applied to the improvement of several processes at
different stages: research, design, manufacturing, optimization, etc. The
need for robust and reliable analysis tools is therefore growing rapidly, in
proportion to the increasing complexity of simulations. To provide quick,
accurate feedback to realistic engineering problems is consequently essential
for companies to be competitive.
Preprint submitted to ECCOMAS 2012

The traditional numerical methodologies employed so far are based on


methods involving finite volumes and finite elements, applied to NavierStokes equations. However, even though such methods have been widely
investigated, they still hold major drawbacks, limiting their capacity to
solve real industrial problems: uncertainties induced by the meshing process; highly empirical approaches to the turbulence modeling (RANS); the
treatment of the nonlinear convective term; artificial stabilization parameters; and so on. Because of this, in most cases engineers are not able to model
real systems; they are forced to fall back on simplified models and approximations. These methods require a time-consuming meshing process, are not
tolerant to moving parts, and are usually limited to steady-state analysis,
ignoring transient dynamics.
Particle-based methods have been in development for several decades, and
are now starting to come to the fore. Among them, the promising LatticeBoltzmann Method (LBM) surmounts many of the drawbacks of traditional
CFD methods. XFlow CFD uses a particle-based and fully Lagrangian approach based on LBM. With this method, classic fluid-domain meshing is
not required and surface complexity is not a limiting factor.
XFlow has been validated in several benchmarks, demonstrating the validity of the method to solve industrial problems. The first example presented
in this paper is the Ahmed body, a classic benchmark for the automotive industry. The cars geometry has a variable slant angle and is a challenging
test case in terms of turbulence modelling and drag estimation. The NASA
trapezoidal wing is the second benchmark presented in this paper, a three
element airfoil composed of a slat, a main blade and a flap. The goal is
to assess the aerodynamic coefficients on a large range of incidence angles,
including the post-stall region.
2. Numerical methodology
Over the last few years, schemes based on minimal kinetic models for the
Boltzmann equation are becoming increasingly popular as a reliable alternative to conventional CFD approaches.
The Lattice Boltzmann method (LBM) was originally developed as an improved modification of the Lattice Gas Automata to remove statistical noise
and achieve better Galilean invariance [1, 2]. Due to the flexibility afforded
by its close connection to kinetic theory, the LBM can be adapted to model
several physical phenomena. Recent research has led to major improvements,
2

including physically consistent models for multiphase and multicomponent


flow and fully compressible flow [3, 4, 5].
2.1. Lattice Gas Automata
The Lattice Gas Automata (LGA) is a simple scheme for modeling the
behavior of gases. The basic idea behind the LGA is that particles with
specific velocities (ei , i = 1, ..., b) propagate through a d-dimensional lattice,
at discrete times t = 0, 1, 2, ... and collide according to specific rules designed
to preserve the mass and the linear momentum when different particles reach
the same lattice position.
The simplest LGA model is the HPP approach, introduced by Hardy,
Pomeau and de Pazzis, in which particles move in a two-dimensional square
lattice and in four directions (d = 2, b = 4). The state of an element of the
lattice at instant t is given by the occupation number ni (r, t), with ni = 1
being presence and ni = 0 absence of particles with velocity ei .
The stream-and-collide equation that governs the evolution of the system
is
ni (r + ei , t + dt) = ni (r, t) + i (n1 , ..., nb ), i = 1, ..., b,
(1)
where i is the collision operator that computes a post-collision state conserving mass and linear momentum. If one were to assume i = 0, only an
streaming operation would be performed.
From a statistical point of view, the system is made up of a large number
of elements which are macroscopically equivalent to the problem investigated.
The macroscopic density and linear momentum can be computed as:
=
v =

b
1X
ni
b i=1

(2)

b
1X
ni ei
b i=1

(3)

2.2. Lattice Boltzmann method


While the LGA schemes use boolean logic to represent the occupation
stage, the LBM method makes use of statistical distribution functions fi
with real variables, preserving by construction the conservation of mass and
linear momentum.
The Boltzmann transport equation is defined as follows:
fi
+ ei fi = i , i = 1, ..., b,
t
3

(4)

where fi is the particle distribution function in the direction i, ei the corresponding discrete velocity and i the collision operator.
The stream-and-collide scheme of the LBM can be interpreted as a discrete approximation of the continuous Boltzmann equation. The streaming
or propagation step models the advection of the particle distribution functions along discrete directions, while most of the physical phenomena are
modeled by the collision operator which also has a strong impact on the
numerical stability of the scheme.
In the most common approach, a single-relaxation time (SRT) based on
the Bhatnagar-Gross-Krook (BGK) approximation is used
BGK
=
i

1 eq
(f fi ),
i

(5)

where is the relaxation time parameter, related to the macroscopic viscosity


as follows
1
= c2s ( ).
(6)
2
fieq is the local equilibrium function usually defined as
fieq

= wi

ei u u u
1+ 2 +
cs
2c2s

ei ei

c2s

!!

(7)

Here cs is the speed of sound, u the macroscopic velocity, the Kronecker


delta and the wi are weighting constants built to preserve the isotropy. The
and subindexes denote the different spatial components of the vectors appearing in the equation and Einsteins summation convention over repeated
indices has been used.
By means of the Chapman-Enskog expansion the resulting scheme can be
shown to reproduce the hydrodynamic regime for low Mach numbers [5, 6, 7].
The single-relaxation time approach is commonly used because of its simplicity. However it is not well-posed for high Mach number applications and
it is prone to numerical instabilities. Some of the limitations of the BGK are
addressed with multiple-relaxation-time (MRT) collision operators where the
collision process is carried out in moment space instead of the usual velocity
space
(8)
MRT
= Mij1 Sij (meq
i mi ),
i
where the collision matrix Sij is diagonal, meq
i is the equilibrium value of the
moment mi and Mij is the transformation matrix [8, 9].
4

An alternative method that aims to overcome the limitations of the BGK


approach is the entropic lattice Boltzmann (ELBM) scheme, which may rely
on a single-relaxation-time where the attractors of the particle distribution
functions are based on the minimization of a Lyapunov-type functional enforcing the H-theorem locally in the collision step. However, this method is
expensive from the computational point of view [10] and thus not used in
practical engineering applications.
The collision operator in XFlow is based on a multiple relaxation time
scheme. However, as opposed to standard MRT, the scattering operator is
implemented in central moment space. The relaxation process is performed
in a moving reference frame by shifting the discrete particle velocities with
the local macroscopic velocity, naturally improving the Galilean invariance
and the numerical stability for a given velocity set [11].
Raw moments can be defined as
xk y l z m =

N
X

fi ekix eliy em
iz

(9)

fi (eix ux )k (eiy uy )l (eiz uz )m

(10)

and the central moments as

xk y l z m =

N
X
i

2.3. Turbulence modeling


The approach used for turbulence modeling is the Large Eddy Simulation
(LES). This scheme introduces an additional viscosity, called turbulent eddy
viscosity t , in order to model the subgrid turbulence. The LES scheme we
have used is the Wall-Adapting Local Eddy viscosity model, that provides a
consistent local eddy-viscosity and near wall behavior [12].
The actual implementation is formulated as follows:
(Gd Gd )3/2
(S S )5/2 + (Gd Gd )5/4
g + g
=
2
1 2
1
2
2
=
(g + g
) g
2
3
u
=
x

t = 2f
S
Gd
g

(11)
(12)
(13)
(14)

where f = Cw x is the filter scale, S is the strain rate tensor of the resolved
scales and the constant Cw is typically 0.325.
A generalized law of the wall that takes into account for the effect of
adverse and favorable pressure gradients is used to model the boundary layer
[13]:
U1 + U2
u U1 up U2
U
=
=
+
uc
uc
uc u
uc up




w u
dpw /dx up
+ u
+ up
=
f1 y
f2 y
+
u2 uc
uc
|dpw /dx| uc
uc
uc y
+
y =

uc = u + up
u =

up =

(15)
(16)
(17)
(18)

|w | /

(19)


!
dp 1/3
w


.
dx

(20)

Here, y is the normal distance from the wall, u is the skin friction velocity,
w is the turbulent wall shear stress, dpw /dx is the wall pressure gradient, up
is a characteristic velocity of the adverse wall pressure gradient and U is the
mean velocity at a given distance from the wall. The interpolating functions
f1 and f2 given by Shih et al. [13] are depicted in figure 1.
25

45

f1 (y + u/uc )

40

20

f2 (y + up /uc )

35
30

15
f2

f1

25

10

20
15

10

5
100

101

y + u/uc

102

100

Figure 1: Unified laws of the wall

101

y + up /uc

102

3. Ahmed body benchmark


The Ahmed Body is a classic benchmark for the automotive industry. It
was first defined and its characteristics described in the experimental work
of Ahmed [14]. The car geometry was studied at various slant angles from
0 to 40 degrees. The experimental measurements were conducted by Ahmed
in the DFVLR subsonic wind tunnels at Braunschweig and Gottingen which
have a square nozzle of (3 x 3) m and a length of 5.8 m.
The first goal of this study is to validate the curve of the drag coefficient
against the slant angle obtained by Ahmed in [14], and the second one is to
analyze the mean recirculation structures on the slant surface of the Ahmed
body and in the downstream region.
3.1. Simulation setup
A strictly identical geometry to the one used by Ahmed was imported
into the virtual wind tunnel featured in XFlow. This virtual wind tunnel
consists of a rectangular domain and was set to dimensions of (8 x 2 x 2)
m. A far-field velocity boundary condition was used at the inlet and the
top boundaries, and zero gauge pressure was imposed at the outlet. Periodic
boundary conditions were set on the side walls, and a free-slip wall with no
velocity was imposed at the bottom boundary.
The geometry of the Ahmed body was separated into two parts in order
to simplify the setup modification for variable slant angles. The first part is
the fore body that has an invariable geometry. The second part is the rear
body which is replaced when the slant angle changes. These two parts are
shown on figure 2.

Figure 2: Fore body geometry and rear geometry

The simulation settings are gathered in table 1, and correspond to a


Reynolds number based on the car length equal to 4.29 million. The sim7

Table 1: Simulation specifications of the Ahmed body benchmark

Inlet velocity
Density
Dynamic viscosity
Car length
Reynolds number
Slant angles
Turbulence intensity

60 m/s
1 kg/m3
1.46014 105 Pa.s
1044 mm
4.29 106
0 ; 5 ; 10 ; 12.5 ; 15 ; 20 ; 25 ; 30 ; 40 degrees
0.5%

ulation time was two seconds and the time step t = 7.69231 105 s is
automatically estimated by XFlow to ensure the numerical stability.
3.2. Spatial discretization
Since XFlow is a particle based technology it does not require a timeconsuming meshing process. The preprocessor generates the initial octree
lattice structure based on the input geometries and the user-specified resolution for each geometry. The lattice may have several levels of detail which
are hierarchically arranged. Each level solves spatial and temporal scales
two times smaller than the previous level, thus forming the aforementioned
octree structure.
The lattice structure may be modified later by the solver if the computational domain changes (due to the presence of moving parts) or if the
resolution changes dynamically in order to adapt to the flow patterns (adaptive wake refinement). The adaptive wake refinement feature in XFlow is
based on the module of the vorticity field: in the lattice elements where
the vorticity reaches a threshold value the lattice is automatically refined.
Similarly, when the vorticity is lower than another threshold, eight adjacent
lattice elements are merged to form a coarser lattice element. This saves
computational resources and removes the need to refine your solution in advance. Consequently, as in illustrated figure 3, three resolutions are required
by the user: the far field, the wake and the near wall resolutions.
In order to select the best resolution near the walls and within the wake
that allows us to get good results in an acceptable time, a resolution dependency study is conducted before starting the validation of the Ahmed body.
This preliminary study consists in refining the resolutions and seeing how this
affects the accuracy of the results, but also checking if the code is converging
8

Figure 3: Example of lattice structure using the near wall and adaptive wake refinement

Table 2: Near walls and wake resolutions used in the resolution dependency study

Resolution (m)
# of Elements at t = 0.3 s

h
0.04
88,316

h/2
0.02
222,337

h/22
0.01
1,132,292

h/23
0.005
8,316,626

to the right solution. It is done by measuring the drag coefficient predicted


by XFlow for a slant angle of 35 degrees which is a reference angle for this
benchmark. The far field is taken constant as 0.08 m, and four resolutions
are considered for the walls and the wake as described table 2.
The drag coefficient is computed for the four cases and compared with
the experimental value measured by Ahmed [14]. The drag points from the
simulations are plotted in figure 4 in function of the number of elements at
t = 0.3 s. The point corresponding to the resolution h/22 = 0.01 m gives
good results and in an acceptable time for a slant angle = 35 , and will
therefore become the reference near wall resolution for the rest of the study.
The figure 4 also confirms the convergence of the code to the correct solution.
A second question arises regarding the value of the wake resolution. As
the wake refinement algorithm creates a significant number of elements as it
develops, its importance in the drag contribution must be assessed accurately
to get a good compromise between solution quality and computational time.
Hence, a second study is conducted on the wake resolution starting from the
elected near wall resolution (0.01 m) and then increasing by multiples of two,
due to the lattice structure. The figure 5 demonstrates the importance of
solving the wake accurately: using the same resolution near the walls and
within the wake the drag coefficient history shows a nice prediction, but
9

0.65 h
0.60
0.55

Drag Coefficient, Cx

0.50
0.45
0.40

h/2

0.35
0.30
0.25
0.200

h/22

1000

h/23

2000

3000

4000
5000
N (103 nodes)

6000

7000

8000

9000

Figure 4: Drag coefficient against the number of lattice nodes for different resolutions at
= 35

as soon as the wake resolution is the double or quadruple of the near wall
resolution affects the results quite dramatically. Hence, for all our runs, the
spatial discretization chosen for all the different slant angles is done with an
automatic wake refinement with a resolution of 0.08 m for the far field, and
0.01 m around the Ahmed body and within the wake.
3.3. Numerical results
The time required in XFlow to set up the case is about 10 minutes and
mainly consists in geometry importation, the flow and boundary specifications, and the resolution setup. The calculation time is almost the same for
all the slant angles and varies between 6 and 8 hours with the previously
selected resolutions on two Intel Xeon E5620 (2.4GHz).
The first result given by Ahmed is the curve representing the drag coefficient against the slant angle , and gives the drag contributions of every
part of the Ahmed body: the front Ck , the rear vertical surface Cb , the rear
slant surface Cs and the friction drag Cr . The total drag Ahmed found was
Cw and was the sum of the different contributions. Hence, the total drag
obtained from XFlow for the different slant angles is superimposed with the
Cw from Ahmed, as shown in figure 6.
From the figure 6 we observe a good overall drag prediction by the code:
the drag breakdown occurs right after 30 degrees and the minimum drag point
10

0.50

Wake 0.01m
Wake 0.02m
Wake 0.04m
Experimental

0.45

Drag Coefficient, Cx

0.40

0.35

0.30

0.25

0.200.0

0.1

0.2

Time (s)

0.3

0.4

0.5

Figure 5: Drag coefficient history for different wake refinement resolution at = 35

is the critical angle 12.5 degrees, as measured by Ahmed. The absolute drag
values predicted by XFlow are accurate and the relative error varies from
only 0.4% to 3.2% for most of the angles, except around the drag breakdown
and at 0 degree angle where it reaches a maximum of only 7.1%. These small
discrepancies can be explained, on the one hand, by the complexity around
the flow around 30 degrees of slant angle which is switching from a massive
3D separation in the near-wake region to an almost 2D attached structure at
higher angles [15], and, on the other hand, by stronger gradients produced
by the rear of the car at 0 degree angle.
3.4. Flow field results
The second part of the results analysis is done by analyzing the main
recirculation structures resulting from the flow around the Ahmed body. For
this study, the averaging of the flow fields is required in order to filter the
temporal fluctuations and to identify the main structures of the turbulent
wake. The averaging of the fields started from t = 0.3 s when the flow was
established, as indicated for example by figure 5, to cut off the transient
period.
Ahmed provides pictures of the oil flow on the slanted surface for =
12.5, 25 and 30 degrees. It can be compared with XFlow which features
Line Integral Convolution (LIC) that approximates the surface streamlines
on a body. The figure 7 shows similar structure for the three angles: a quite
11

smooth and attached flow at 12.5 degrees, smooth flow patterns with two
small and symmetric fringes on the sides at 25 degrees, and two large and
symmetric separation bubbles at 30 degrees.
Ahmed also provides different velocity vectors plots in the symmetry
plane of the car, showing the near-wake region. This allows the study of
the separation bubble on the rear slant and within the wake for different
slant angles.
Figure 8 compares the near-wake region for a slant angle of 5 degrees
between the experimental results measured by Ahmed and results obtained
by XFlow at the same scale. This allows us to check the length of the bubble
separation located around the non-dimensional coordinate x/Lref = 0.375,
predicted in an extremely similar way in the two pictures. Two main eddy
structures are detected - highlighted in red boxes on figure 8 - which are
symmetrical from the top and bottom of the separation bubble. The code
tends to locate them slightly further downstream, though with reasonable
overall flow patterns.
The near-wake structure for a slant angle of 25 degrees also show good
similarities. This figure 8 shows an equivalent triangular separation bubble,
ending around the non-dimensional coordinate x/Lref = 0.2 for both cases.
4. NASA trapezoidal wing benchmark
The NASA trapezoidal wing benchmark comes from the 1st AIAA CFD
High Lift Prediction Workshop (HiLiftPW-1), sponsored by the Applied
Aerodynamics Technical Committee, which took place in June 2010 in Chicago,
IL. The challenge was to simulate a half aircraft configuration composed of
a body and a 3-element airfoil with a plane of symmetry as shown in figure
9 for a wide range of angles of attack. The trapezoidal wing is composed
of slat, main element and flap. The latter can be in two different configurations: Configuration 1 at 25 degrees and Configuration 8 at 20 degrees of
angle-of-attack.
The objectives of the benchmark are multiple [16]:
Assess the prediction capability of CFD codes in landing/taking-off
configuration,
Develop practical modeling guidelines for the analysis of high-lift configurations,

12

Table 3: Resolutions used for the resolution-dependency at 13 degrees incidence

Near wall (m)


Wake (m)
# of Elements at t = 0.3 s

h
0.04
0.08
201,513

h/2
0.02
0.04
653,211

h/22
h/23
0.01
0.005
0.02
0.01
2,893,687 21,880,186

Provide an impartial forum for evaluating the effectiveness of existing


CFD codes and modeling techniques,
Identify areas that require additional research and development.
4.1. Simulation setup
XFlow simulations were run for the Configuration 1 with no brackets. The
Mach number was 0.2, the Reynolds number based on the mean aerodynamic
chord (MAC) was 4.3 million. The angles of attack run for this benchmark
were: -4, 1, 6, 13, 21, 25, 28, 32, 34 and 37 degrees. The hardware used in
all the computations was a single workstation with two Intel Xeon E5620 @
2.4 GHz processors (8 cores) and 12GB of RAM.
A resolution dependency study has also been performed for this benchmark using the four resolutions described in Table 3 and a constant far field
resolution of 1.28 m. An incidence angle of 13 degrees which is one of the
reference angles of the first workshop was employed.
The drag coefficient obtained with each of the four simulations is plotted
in figure 10 as a function of the number of elements at t = 0.3 s. The
point corresponding to resolution h/23 gives the best estimation of the drag
compared to the experimental data, with only 1% of relative error. This
value will therefore be used as the reference near wall resolution for the rest
of the study.
However, two different wake resolutions have been used depending on the
incidence of the NASA trapezoidal wing. Indeed, for large angles of attack,
a significant wake develops and the number of lattice elements introduced by
the adaptive wake refinement increases. At 32 degrees, the simulation reaches
25 million lattice elements, which is the maximum number of elements that
can fit in the 12 GB of RAM available on the workstation. Special care is
thus required in order to keep this number within the memory constraints
for higher angles. The wake resolution has been limited to double the normal
value for those cases (Resolution 2 in table 4).
13

Table 4: Resolutions used for the 1st High Lift Prediction Workshop

Resolution 1
Resolution 2

Walls (m)
0.005
0.005

Wake (m)
0.01
0.02

Far Field (m)


1.28
1.28

Max. # of Particles
Angles
6
25 10
[-4 ; 32 ]
10 106
[34 ; 37 ]

4.2. Numerical results


The experimental data were produced at the 14x22 wind-tunnel at the
well-known NASA Langley. Forces, moments, and Cp distribution were provided with free transition [17]. Data were provided as lower and upper values
which are assumed to be the range of uncertainty in the wind tunnel measurements.
On figure 11, the drag coefficient against the angle of attack is shown.
XFlow results show very good agreement with the experimental data along
the whole range of angles. The drag slope is accurate and still behaves
correctly at both low and high incidences, with a slight slope decrease.
The lift coefficient is also very well predicted for the whole range of angles.
Within the range [1, 28] degrees, XFlow predicts accurately both slope and
absolute lift coefficient values. Starting from 32 degrees, the critical angle
is reached and the code also succeeds in predicting this: the wind tunnel
data indicates the maximum lift point at around 33 degrees, and it happens
between the point of 32 degrees and 34 degrees. Starting from that point,
the lift drops, due to a large bubble of separation on the wing. The bubble
of separation grows on the tip of the wing, as shown in the Figure 12.
Since both drag and lift coefficients are quite well predicted, the polar
curve on Figure 11 is hence matching the experimental results, especially in
the pre-stall region.
The pitching moment coefficients also lie between the upper and lower
limits of the experimental results within almost the whole range.
5. Conclusions
The CFD code XFlow features a kinetic particle-based solver that differs
from the traditional approaches, which are usually mesh-based. The latticeBoltzmann method employed is able to solve advanced industrial problems
even in the presence of complex geometries or moving parts.
14

The methodology has demonstrated it can solve industrial benchmarks


efficiently. For instance the Ahmed body is a classic benchmark for the automotive industry that XFlow solved with a high degree of accuracy. XFlow
did not face convergence issues even for extreme slant angles, and changing
the rear of the car did not add additional workload. The code has been
demonstrated to be robust and accurate in terms of drag and flow pattern
prediction, and closely matches the data measured by Ahmed in the DFVLR
subsonic wind tunnel of Braunschweig including the drag breakdown around
30 degrees and the low slant angles where gradients are stronger.
The High Lift Prediction Workshop benchmark has also been successfully
validated by XFlow. The NASA trap wing geometry was tested within a
range of incidence between -4 and 37 degrees, which includes the post-stall
region. The drag, lift and pitching moment coefficients predicted by the code
are in good agreement with the experimental tests conducted in the NASA
Langley 14x22 wind tunnel. The stall angle is also accurately predicted
around 33 degrees.
XFlow has therefore demonstrated its robustness and accuracy in different
benchmarks. The method is well-suited for external aerodynamics and shows
strong potential for more advanced topics, such as analysis involving complex
geometries, the presence of moving parts and fluid-structure interaction.
References
[1] U. Frisch, B. Hasslacher, Y. Pomeau, Lattice-gas automata for the
navier-stokes equation, Physical review letters 56 (14) (1986) 15051508.
[2] G. R. McNamara, G. Zanetti, Use of the Boltzmann equation to simulate lattice-gas automata, Physical Review Letters 61 (1988) 23322335.
doi:10.1103/PhysRevLett.61.2332.
[3] S. Chen, G. Doolen, Lattice boltzmann method for fluid flows, Annual
review of fluid mechanics 30 (1) (1998) 329364.
[4] S. Succi, The lattice boltzmann equation, For Fluid Dynamics and Beyond.
[5] Z. Ran, Y. Xu, Entropy and weak solutions in the thermal model for
the compressible euler equations, Arxiv preprint arXiv:0810.3477.

15

[6] Y. H. Qian, D. DHumi`eres, P. Lallemand, Lattice BGK models for


Navier-Stokes equation, EPL (Europhysics Letters) 17 (1992) 479.
doi:10.1209/0295-5075/17/6/001.
[7] F. J. Higuera, J. Jimenez, Boltzmann approach to lattice gas simulations, EPL (Europhysics Letters) 9 (1989) 663. doi:10.1209/02955075/9/7/009.
[8] X. Shan, H. Chen, A general multiple-relaxation-time boltzmann collision model, International Journal of Modern Physics C 18 (4) (2007)
635643.
[9] D. dHumi`eres, Multiplerelaxationtime lattice boltzmann models in
three dimensions, Philosophical Transactions of the Royal Society of
London. Series A: Mathematical, Physical and Engineering Sciences
360 (1792) (2002) 437451.
[10] P. Asinari, Entropic multiple-relaxation-time lattice boltzmann models,
Tech. rep., Politecnico di Torino, Torino, Italy (2008).
[11] K. Premnath, S. Banerjee, On the three-dimensional central moment
lattice boltzmann method, Journal of Statistical Physics (2011) 148.
[12] F. Ducros, F. Nicoud, T. Poinsot, Wall-adapting local eddy-viscosity
models for simulations in complex geometries, in: Proceedings of 6th
ICFD Conference on Numerical Methods for Fluid Dynamics, 1998, pp.
293299.
[13] T. Shih, L. Povinelli, N. Liu, M. Potapczuk, J. Lumley, A generalized
wall function, NASA Technical Report.
[14] S. Ahmed, G. Ramm, G. Faitin, Some salient features of the timeaveraged ground vehicle wake, Tech. rep., Society of Automotive Engineers, Inc., Warrendale, PA (1984).
[15] G. Franck, N. Nigro, M. Storti, J. DEla, Numerical simulation of the
flow around the ahmed vehicle model, Latin American applied research
39 (4) (2009) 295306.
[16] C. Rumsey, The 1st aiaa cfd high lift prediction workshop (Jun. 2010).
URL http://hiliftpw.larc.nasa.gov/index-workshop1.html
16

[17] C. McGinley, L. Jenkins, R. Watson, A. Bertelrud, 3-d high-lift flowphysics experimenttransition measurements, AIAA Paper 5148 (2005)
2005.

17

Figure 6: Drag coefficient against the slant angle

18

Figure 7: Averaged Line Integral Convolution (LIC) on the slanted surface from Ahmed
(left) and XFlow (right)

19

Figure 8: Near-wake structure at scale for: a) = 5 , b) = 25

Figure 9: NASA trapezoidal wing geometry

20

0.40

XFlow
Experiment

Drag Coefficient, CD

0.38
0.36

h/2

0.34
h/23

0.32

h/22

0.300

15
10
Number of Lattice Nodes (106 )

20

25

Figure 10: Drag coefficient against the number of lattice nodes for different resolutions at
= 13

0.8

3.5

Experimental
Experimental Lower
Experimental Upper
XFlow

3.0
Lift Coefficient, CL

Drag Coefficient, CD

1.0

0.6
0.4
0.2
0.0 10
3.5

10

(deg)

20

30

0.0 10

40

10

10

(deg)

20

30

20

30

40

0.1

2.5

Pitching Moment, Cm

Lift Coefficient, CL

(b)
0

0.0

3.0
2.0
1.5
1.0
0.5
0.00.0

1.5
1.0
0.5

(a)
0

2.5
2.0

(c)
0.2

0.4
0.6
Drag Coefficient, CD

0.8

1.0

0.2
0.3
0.4
0.5

(d)
10

(deg)

40

Figure 11: Drag (a) and lift (b) coefficients against the angle of attack, the polar curve
(c), and the pitching moment coefficient (d)

21

Figure 12: Averaged Line Integral Convolution (LIC) at 37 degrees incidence

22

You might also like