Sunteți pe pagina 1din 13

AIAA 2011-3846

20th AIAA Computational Fluid Dynamics Conference


27 - 30 June 2011, Honolulu, Hawaii

Computational Model for the Dynamic Analysis of a TwoDimensional Airfoil with a Control Surface
Rodrigo A. Jimnez1 and Omar D. Lpez2
Universidad de los Andes, Bogot, Cundinamarca, Colombia

This paper introduces a computational model to simulate the aerodynamics, solid body
dynamics and control of a two-dimensional airfoil with a control surface. Inserting the
geometry as a series of points, the model uses the Hess and Smith Panel Method to solve the
associated potential flow problem and calculate the lift and moment coefficients of the
airfoil. These results are then used to solve the solid dynamic model of an airfoil. Feeding the
previous data to a controller system, an aileron deflection angle is produced, which changes
the airfoils geometry. The new geometry is inserted back to the aerodynamic model,
repeating the cycle. This technique gives fast and adequate performance information to
engineers that wish to simulate new airfoil designs. There is no need for linear
approximations to obtain the correct dynamic performances because aerodynamic forces
and moments are calculated separately. Another advantage is that other models can be
easily integrated with this tool in order to obtain aeroelastic, atmospheric, viscous,
compressible, and propulsive effects.

Nomenclature
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
1
2

panel method geometric matrix


angle of attack
time dependent factors that affect panel velocities
control surface deflection angle
wake factors that affect panel velocities
drag coefficient
lift coefficient
moment coefficient around the center of gravity
pressure coefficient at the ith panel
coefficient for the x-axis force component in the body coordinate system
coefficient for the z-axis force component in the body coordinate system
airfoils displacement in the earth axis system
x-axis length component of the ith panel
y-axis length component of the ith panel
velocity potential
gravitational acceleration
circulation
airfoils height in the earth axis system
airfoils moment of inertia
airfoils perimeter
airfoils mass
airfoils pitch velocity
air density at flight altitude
time in step k
vortex intensity in time step k
pitch angle

M.Sc. Student, Department of Mechanical Engineering, Member AIAA.


Assistant Professor, Department of Mechanical Engineering, Member AIAA.
1
American Institute of Aeronautics and Astronautics

Copyright 2011 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

=
=
=
=
=
=
=
=

Flow velocity
x-axis velocity component in the body coordinate system
Unsteady flow velocity seen by an observer on the airfoil at time step k
Tangential velocity component over the ith panel
z-axis velocity component in the body coordinate system
velocity component vector over each panel
x-coordinate of the ith panel midpoint
y-coordinate of the ith panel midpoint

I. Introduction

N the early XX century, most lift generating surfaces were designed through experimentation. This method gave
birth to various families of two-dimensional airfoils including the NACA 4 digit airfoils [1], NACA 5 digit
airfoils [2], and NACA 6 digit airfoils [3]. However, with the rapid evolution of computational methods, simulation
became an important method for aircraft design, reducing experimentation costs [4]. In fact, simulation became an
obligatory step for all aerodynamic implementation and design. Following this, researches concerning the simulation
of aerodynamic surfaces with a control mechanism have become an active field. For example, Rodriguez and Sonne
discuss the results of implementing software that evaluates the performance, control and navigation system of a
missile [5]. A freeware program, called Missile Guidance Simulator, has also been developed that allows the
simulation of a guided missile [6]. This software has been greatly used by Great Britains amateur rocket
community.
Given the complexity of the problem, the number of research projects involving the coupled simulation of airfoil
aerodynamics and dynamics with control surfaces is inferior. However, some studies in which synthetic jet actuators
are used as a means of control, have been made. Tchieu et. al. show a method in which a low order model is used to
simulate the aerodynamic forces and moments acting on an airfoil that uses synthetic jet control to ascend and
descend [7]. Results are validated with high order models, which give interesting predictions at low pitching and
plunging frequencies. Jee et. al. present a computational fluid dynamics (CFD) simulation of an airfoil with closed
loop control using synthetic jet actuators. The desired result is the pitch control of the airfoil. Results were verified
experimentally using wind tunnel data [8].
Another research area in which CFD analysis has gained importance is in animal flight dynamics. As nature has
made flying animals with a complex structure, knowledge of their aerodynamic performance is also studied by
scientists and engineers. Sriam et. al. made a first investigation in the flapping movement of an insects wing [9].
Using a time spectral computational method, the aerodynamic forces generated at different flapping frequencies
were found. Also, using a fifth order WENO method, an estimation of the optimal trajectories and the required
control inputs was made. Furthermore, Vanella et. al. show the development of a hybrid dynamic model for the
analysis of insect and small bird flight dynamics [10]. The animals body is modeled by a series of rigid bodies,
connected together with simple mechanical elements. The analysis gave out flight performance data that involved
aerodynamic, structural and control effects generated by the system. This development is oriented towards the
construction of micro air vehicles.
The present paper shows the results of a research work, the objective of which was to develop a computational
model that allowed the combination of the aerodynamics, solid body dynamics and control of a two-dimensional
airfoil. The desired applications for this implementation are aircraft that have fast control actions during flight. Only
subsonic flight will be considered, that is, Mach number ( ) will be under 0.3 since compressibility effects are not
included. First, the computational models used for aerodynamic prediction, geometry modification due to changes in
control surface deflection, airfoil dynamics and control model will be explained. Model verifications and numerical
results for various simulations will be presented in section III. Finally, the conclusions of the project will be given.

II. Computational Methods


A. Hess and Smith Panel Method (HSPM) [11]
The HSPM is a low order numerical method that is widely used to calculate the velocity field of an airfoil that is
under stationary and transient conditions in an incompressible and inviscid fluid. This allows the proper estimation
of lift, , and moment,
, coefficients. The method solves the Laplace equation of velocity potential for an
aerodynamic airfoil, (see equation (1)). Normal velocity components over the airfoil are null and tangential
components are the desired solution. After obtaining the tangential velocities over the airfoil, pressure coefficient
distribution around the airfoil surface is obtained, which leads to the calculation of the lift and moment coefficients.
2
American Institute of Aeronautics and Astronautics

The method consists on dividing the airfoil into line segments called panels, starting from the trailing edge and
advancing in a clockwise direction towards the leading edge, until finally reaching the starting point. Applying a
source and a vortex over the midpoint of every panel, the inviscid flow over of the airfoil can be calculated at a
specific flight condition. Sources are applied in order to meet the condition of zero normal velocity. Vortexes are
applied to satisfy the Kutta condition, which is given by (2), and is approximated using a finite difference model.

However, as the problem being considered is transient, the method must be applied every time step. This also
means that, as the total circulation must be conserved, any change over the surface of the airfoil must be balanced by
a change in the wake vortexes that are generated at the trailing edge. In order to do this, a set of panels must be
added at the trailing edge to represent the wake, since at every time step a vortex is shed from this point. This vortex
trail will modify the induced velocities over the airfoil panels. The result is a linear system of equations that can be
solved by Gaussian elimination or by an iterative method. The system of equations that has to be solved has the form
shown in equation (3).
(3)
Here is a coefficient matrix that only depends on
the geometry of the airfoil, is the vector that represents
the source magnitudes over the airfoils panels,
the
vortex strength at time step k,
and
represent the
transient effects due to the wake.
Even though HSPM does not take into account
viscosity effects, it can be integrated as a separate model.
Also, the fact that depends only on the geometry of the
airfoil and not on transient effects allows an easy
coupling of the geometric changes that will occur on the
airfoil once a control action takes place. Finally, another
advantage is that the HSPM is fully expandable to threedimensional analysis, so the methodology proposed in
Figure 1. Coordinate system used to calculate
this study can be further expanded to analyze a complete
aerodynamic forces and moments [12].
aircraft.
Before explaining the HSPM algorithm, the coordinate system for the aerodynamic analysis is specified. For this
particular case, a lift and drag oriented coordinate system is selected. Here the z-axis is aligned with the lift force;
that is, perpendicular to the air flow velocity. The x-axis is aligned with the air flow velocity, parallel to the drag
force. The angle of attack will be positive if the airfoils chord is above the x-axis and negative if it is below. This
coordinate system is shown in Figure 1.
A transient HSPM code consists of the following steps [11]:
1) A file with the coordinates of the airfoil is generated, starting from the trailing edge and going clockwise
around the airfoil until returning to the starting point. All points are joined together by line segments called
panels.
2) A source and a vortex are placed over the midpoint of each panel. All vortexes are assumed to have the
same strength, while sources have different strength magnitudes. Both, vortex and source magnitudes, are
initially unknown.
3) The Kutta condition in a transient condition implies that a vortex is generated at the trailing edge of the
airfoil to balance the total circulation. In case a wake is already formed, its vortexes must be convected
downstream at their specific velocity.
4) Influence coefficients that involve the effects of the wake and the airfoil on the aerodynamic performance
of the airfoil must be calculated.
5) An iterative process is started in which the magnitude of the shed vortex is solved for:
3
American Institute of Aeronautics and Astronautics

a.

6)
7)

8)

A prediction of the shed vortex panel length and orientation is made. The shed vortex must be located
over it.
b. The airfoils vortex magnitude is estimated using previously calculated influence coefficients.
c. Source magnitudes over the airfoil panels are calculated.
d. Speed of propagation of the wake vortexes are solved for, and with this information, the shed panels
length and orientation are updated using circulation values.
e. The iterative process is repeated until there is convergence for the shed vortex panels length and
orientation.
Calculate the tangential velocities over each panel and its potential.
To solve for the Kutta condition it is necessary to know the velocity potential over the airfoil, which will
lead to the calculation of the distribution of pressure coefficient. To calculate , the leading edge potential
must be known. Instead of solving the Laplace equation at this point, its potential is estimated with a series
of panels that are generated in front of the leading edge to represent the flow that comes from infinity. With
this value, the potential over all of the airfoil panels can be calculated.
Applying equation (4), the pressure coefficient over the midpoint of every airfoil panel is calculated, using
the information obtained from the previous step.

With this information,

9)

and

can be obtained with equations (5) and (6).

Repeat the procedure for the next time step, starting from step 2.

B. Control Surface Deflection


The desired type of control surface corresponds to an aileron, located at the trailing edge of the airfoil. By
selecting a pivot point over the camber line of the airfoil, a rotation of a given angle was made for all the coordinate
points that correspond to airfoils geometry from the pivot point to the trailing edge. Positive deflection angle
implies that the flap rotates counter clockwise while negative deflection angle means that the flap rotates clockwise.
Figure 2 shows the obtained geometries after a given positive and negative deflection. This type of deflection will
not cause any geometrical problems if is kept between
and
. Other values will cause a distortion of the
airfoils geometry which will lead to numerical instabilities.

A.
B.
Figure 2. Airfoil deflection forms. A. NACA 4415 airfoil with
.

. B. NACA 4415 airfoil with

C. Airfoil Dynamic Model


For the dynamic model of the airfoil, two different sets of axis coordinate systems were selected. The first is the
body axis system, which is the reference frame on the airfoil. Here, the x-axis is in the direction of the aerodynamic
chord, the z-axis is directed downward, and moments are considered positive if the airfoils nose is pitched up.
4
American Institute of Aeronautics and Astronautics

The second coordinate system to be used is the earth axis system. This is an inertial reference frame that will be
used to observe the height and displacement of the airfoil. The x-axis is taken in the north direction and the z-axis is
taken downwards. Again, positive moments will be those that generate a nose up pitch in the airfoil.
and
are oriented perpendicular and parallel to the air flow velocity, respectively. In order to characterize
the dynamic behavior of the airfoil, these forces must be transformed to the body axis system. To do this, the angle
that relates them must be known. As
and
are given in reference to the air flow velocity and the body axis
system is oriented with the airfoils chord, the angle that relates these two coordinate systems is . The result of the
transformation is given by equations (7) and (8).
does not need to be transformed because in all the coordinate
systems that are being considered, moments are positive in the same direction.
(7)
(8)
Even though flow is considered inviscid in this project, a force component in the x-axis direction due to lift will
exist. This means that there will be an aerodynamic acceleration in this direction that will modify the dynamic
behavior of the airfoil.
The second transformation that must be considered, corresponds to the velocity components, and . As flight
altitude and airfoil displacement are required, it is necessary to know the velocities in the earths x and z axes. The
pitch angle, , relates the body axis system to the earth axis system. Considering the time derivative of the altitude
as the earth z-axis velocity component and the displacement time derivative as the earth x-axis velocity component,
the necessary transformations are given by equations (9) and (10). It must be noted that, according to the earth axis
system, flying above ground level will produce a negative value of altitude.
(9)
(10)
The fact that this problem is two-dimensional means that only longitudinal equations of motion must be
considered. Other assumptions for the model presented here are:
1)
is constant.
2)
is constant.
3)
is constant.
4)
is constant.
5) No propulsive force is being applied.
6) Drag force is zero.
Taking these assumptions into account, force and moment equations about the center of gravity can be
determined. These can be reordered to obtain expressions for angular and linear accelerations, which will be the first
three state equations of the dynamic model (see equations (11) to (13)).

Equation 14 is needed to relate the pitch angle to the pitch angular velocity.
(14)
Equations (9) through (14) represent the complete set of state equations for the dynamic model of an airfoil,
which are a nonlinear system of differential equations. However, as the aerodynamic coefficients are known through
5
American Institute of Aeronautics and Astronautics

the HSPM code, there is no need to linearize them. A fourth order Runge Kutta method was used to solve this
system of equations. A diagram of the complete coupled aerodynamic and dynamic model is shown in Figure 3.

Figure 3. Coupled aerodynamic and dynamic system.


D. Control Model
A simple stability control model for the airfoils attitude used in the present work consists on maintaining the
pitch angle at a constant value with a given tolerance. Whenever the airfoil has a pitch angle that is smaller than the
desired value minus the tolerance value, the controller activates a single positive deflection angle for the control
surface in order to induce a positive moment. If the airfoil has a pitch angle that is greater than the desired value plus
its tolerance, the controller activates a single negative deflection to induce a negative moment. In the case in which
the pitch angle lies within the range of acceptable values, the deflection angle will be set to a value that cancels the
aerodynamic moment of the airfoil.

III. Numerical Results and Verifications


A. Steady State Verification of HSPM Code
The first verification process consisted in the
calculation of
and
for steady state conditions.
The selected airfoil for the test was the NACA 4415.
Results were compared with experimental data
available in NACA reports [1] and with simulation
results using CFD (ANSYS/FLUENT). Figure 4 shows
the results for the selected airfoil, with a 3 angle of
attack. It can be seen that both coefficients remain
constant through the entire simulation, as it is
expected.
Figure 5 shows the comparative results for using
the implemented HSPM code, a high order CFD and
Figure 4. Steady state simulation of a NACA 4415
experimental data. For the last two, Reynolds number
airfoil at
.
( ) was
. It can be seen that, between
angles of attack of -8 and 8 there is an accurate estimation of the lift coefficient. These range of will allow the
use of small angle approximation. The slope of the obtained curve is similar to the theoretical slope for thin airfoils,

Figure 5. Steady state

results for a NACA 4415 airfoil.

6
American Institute of Aeronautics and Astronautics

which is
. However, stall is not properly predicted by the HSPM code as expected since this is an inviscid
formulation.
Figure 6 shows the overall comparative results for
using the same methods used for . CFD simulations are
the ones that better predict moment coefficient results in comparison to experimental results. However this is
expected due to the fact that HSPM is a low order method. Taking into account that drag effects are not being
calculated and that the magnitudes of
are very small, the obtained results are considered acceptable.

Figure 6. Steady state

results for a NACA 4415 airfoil.

B. Transient State Verification of HSPM Code


Once the steady state behavior of the airfoil was correctly estimated, a transient state test was performed. It
consisted on observing the behavior of a NACA 4415 airfoil, flying at
and, after achieving a stable
condition, it was subjected to a sudden change of
in angle of attack. Initially this problem was solved with high
order CFD (see Figure 7). Figure 8 illustrates the results obtained using the HSPM code. It can be observed that the
airfoil initially starts in its steady state values of
and
. Then, after applying the sudden change in , the lift
coefficient starts increasing until arriving to its stable value in the new angle of attack. Something similar happens

A.

B.
C.
Figure 7. High order CFD results for performed transient test. A. Moment coefficient behavior. B:
Lift coefficient behavior before impulsively started flow. C: Lift coefficient behavior impulsively
started flow.

A.
Figure 8. Behavior for a sudden change in

B.
using the HSPM code. A.

B.

7
American Institute of Aeronautics and Astronautics

with the moment coefficient. The only difference is that


for the latter, there is a slight overshoot before reaching
the final value. CFD results show similar behavior. There
is a variation in the type of response. As the CFD
simulation is a high order model that includes viscous
effects,
behaves as a second order variable. This does
not happen in the HSPM code because viscosity is
ignored.
has an non-uniform behavior in CFD. While
it stabilizes, variable amplitude oscillation occurs around
the stability point. This is not observed on the HSPM
results. However, final steady results are quite similar.
Having verified the transient HSPM code, another
problem of interest, which consists on a NACA 4415
airfoil with a circular flapping motion at frequency of
, with a uniform and horizontal flow was tested.
Figure 9. Simulated circular motion of the airfoil
Figure 9 shows the trajectory followed by the airfoil and
while it travels horizontally with
.
Figure 10 the results for
and
. It can be observed
that the wing initially has a steady state period in which there is no oscillation. Then the airfoils and
start an
oscillatory behavior, similar to a sine function, in which the highest lift coefficient value corresponds to its steady
state value at
and its lowest value corresponds to the steady state result at
. Similarly, the moment
coefficient has its lowest value equal to the steady state result at
and the highest equal to the steady state
value at
. The observed phase difference between both coefficients is expected because positive changes in
increase
while decreasing
. The analytic solution for this problem corresponds to a system of equations with
complex numbers [16] (equations (15) and (16)). This solution has sinusoidal form, which also suggests that the
predicted behavior is correct. With this, it can be said that the HSPM code has a satisfactory performance for various
transient problems. Finally, after the oscillatory movement finishes, the lift coefficient stabilizes at the same value it
had at the beginning, corresponding to the unchanged air flow velocity and angle of attack. Moment coefficient has
a small overshoot before stabilizing in the value corresponding to the air flow. It is now clear that the
will always
have this small discontinuity due to the sudden stop.

Figure 10. Results for the simulated circular motion of the airfoil while it travels horizontally with

C. Aerodynamic Analysis
The next step consisted in observing the aerodynamic effects of the aileron on the wing. This was achieved by
simulating the aerodynamic performance of a NACA 4415 airfoil with a given value of . For those conditions a full
range of was applied, starting at
and finishing at
with increments. Figures 11 and 12 show the results
obtained for
and
respectively. It is shown that a positive deflection causes
to decrease. This result was
expected because the upper surface of the airfoil becomes larger than the case with no control surface deflection.
This causes the upper surface velocity to increase, reducing the pressure coefficient and hence the overall lift.
Changes due to flap deflection can be so large that flight conditions in which lift was positive can have negative lift
values. These specific locations will be of use when the desired flight condition consisted on a reduction in flight
8
American Institute of Aeronautics and Astronautics

altitude. Negative deflection has the opposite effect on lift performance. By looking closely at Figure 11, it can be
observed that the separation between curves at a
difference in angle of attack is the same. This means that the
effect of the control surface deflection on , with respect to , is linear. A 1 aileron deflection produces a -0.0734
change on .

Figure 11. Control surface effects on

for a NACA 4415 airfoil.

The deflection angle of the control surface also has important effects on the moment coefficient of the airfoil, as
seen in Figure 12. For an airfoil with no control surface the moment coefficient is almost constant. However, an
aileron deflection produces an important change in the airfoils geometry which considerably modifies the value of
. Positive values of will increase the value of the moment coefficient; a 1 deflection of the aileron will produce
a 0.0271 change in
. In fact,
can increase so much that it takes positive values. The contrary occurs for
negative deflections. The effects of the moment coefficient are particularly important when managing the stability of
an airfoil. Overall results show that both and are important variables for a proper airfoil performance.

Figure 12. Control surface effects on

for a NACA 4415 airfoil.

D. Dynamic Analysis
In order to test the behavior of the coupled aerodynamic and dynamic model of the airfoil, a NACA 4415 airfoil
was simulated at 0 angle of attack. Air density was constant at 1kg/m3. A mass of 84.57kg and moment of inertia of
6.38 x 104kg m2 [13] were used. The center of gravity is located at the airfoils quarter-chord.
Figures 13, 14 and 15 show the numerical results for the conditions previously mentioned. Aerodynamic and
dynamic performances are shown for a time period of three seconds. Initially the airfoil is at zero angle of attack.
This generates a value of
that produces a nose-down moment to the airfoil. Because of this, negative pitch and
pitch angular velocity start to occur. As time passes, the nose of the airfoil is pointing further down, which results in
a positive z-axis velocity component. This produces a change in the angle of attack, which at the same time changes
the airfoils moment and lift coefficients. The change in angle of attack is negative. This makes
have smaller
values than it had in previous time steps, which makes the airfoil lose altitude. At the same time
increases. These
9
American Institute of Aeronautics and Astronautics

result in a deceleration of pitch angular velocity. However, as it is still negative, pitch angle continues to increase.
Results are clearly unstable, as expected.

A.
Figure 13. Aerodynamic response. A.

B.
B.

Figure 14. Dynamic variable behavior. A. X-axis velocity component. B. Z-axis velocity component.
C. Pitch angular velocity. D. Pitch angle.
E. Coupled Simulation Results
The first flight conditions that were simulated
correspond to the airfoil being in equilibrium. The center
of gravity was located at a point such that the total
moment of the airfoil was zero. Also, the speed of the
airfoil was selected such that there was no vertical force
acting on it. The idea of this simulation was to observe
the numerical stability of the code. After a simulation of
600s, done in 60,000 time steps, the results of Figure 16
were obtained. It can be observed that the airfoil
maintains a constant value of
and during the entire
simulation.
Figure 15. Airfoils altitude and displacement.

A.
B.
Figure 16. Simulation results with trimmed flight conditions. A. B.
10
American Institute of Aeronautics and Astronautics

After this, the complete computational model was tested using the same flight conditions for the aerodynamic
and dynamic problem of section III.D. The controller was configured so that the airfoil maintained a steady flight,
that is, zero pitch angle with zero tolerance. In case there was positive pitch angle, the control surface would deflect
. This corresponds to a positive
value. If the pitch angle were to be negative, the deflection angle would
remain at zero, were the moment coefficient of the airfoil is negative. If the pitch angle was , the deflection angle
of the control surface would be set at , value at which the moment coefficient is approximately . The results are
shown in Figures 17 to 20.
In this case the airfoil maintains a near zero pitch angle as shown in Figure 18. It can be seen that the angle
varies in an oscillatory manner around zero. The same happens with the pitch angular velocity due to the control
action. Comparing the oscillatory movement of pitch with the moment coefficient behavior, it can be observed that
changes in its slope go hand to hand with changes in the sign of C m.
However when analyzing the altitude curve, it is clear that the airfoil is still plummeting down because the lift
coefficient is becoming negative as the control surface is deployed. This generates a positive z-axis velocity
component that continuously reduces the angle of attack of the airfoil. When the control action again raises the lift
coefficient, current speed and angle of attack values are not high enough to recover the airfoils altitude. It is evident
that if steady flight conditions are needed, another form of control has to be implemented.

A.
B.
Figure 17. Aerodynamic response including control action. A.

B.

Figure 18. Behavior of dynamic variables. A. X-axis velocity component. B. Z-axis velocity
component. C. Pitch angular velocity. D. Pitch angle.

Figure 19. Airfoils altitude and displacement.


Figure 20. Control action.
11
American Institute of Aeronautics and Astronautics

IV. Conclusions and Future Work


The present paper presents a computational model to estimate the aerodynamics and dynamics of a twodimensional airfoil that flies under stationary or transient behavior. This model couples the aerodynamics, rigid body
dynamics and control of an airfoil in one algorithm. This led to the development of a computational tool that
predicts the overall performance of two-dimensional airfoil with an aileron.
Additionally, the code allows the user to apply a simple control method, using an aileron, over the airfoil that
makes it maintain a zero degree pitch angle. Geometric changes in the airfoil, due to the deflection of the control
surface, are included in the code. This allows the low order HSPM model to estimate the aerodynamic performance
by means of the airfoils geometry, without the need to alter the flow conditions. The fact that there is an integrated
code that allows the simulation of these three different aspects makes this implementation an innovative solution for
conceptual aerodynamic design and an important technical advance in transient aeronautics.
The use of a low order model to estimate the aerodynamic performance of the airfoil leads to accurate results
with simulations that are computationally inexpensive. While CFD simulations take several minutes to solve
stationary and transient events, the implemented code only takes a few seconds, using a dual core processor of
2.66GHz and 4GB of RAM. This makes the implemented code cost effective. Estimated maximum error is of 10%
for prediction and 16% for
, under steady state conditions, in comparison with reported experimental results by
NACA. These values occur at high angles of attack and are present because the code does not predict flow
separation. For smaller angles of attack, the predicted error in estimated aerodynamic coefficients is as small as
0.5%.
Aerodynamic results of transient state flight conditions are predicted with the same accuracy as stated in the
previous paragraph. Under these conditions the code has been proved to be numerically stable. If a transient flight
condition needs to be simulated, the code also provides correct results. This has been verified by comparing two
different time varying problems with CFD results and with theoretical solutions to the studied problems. Specifically
for an untrimmed airfoil, the results obtained with the simulation show that the uncontrolled airfoil will fall with a
high pitch angle in a short period of time. After using the control system, the airfoil continues to fall in a small
period of time but with a controlled 0 pitch angle. It is believed that a correct flight path control can be obtained by
using a more complex control system or by using multivariable control systems.
In the future it may be considered to extend this code to a full aircraft. For this particular case, lateral dynamics
must also be included as should other types of control surfaces and directional thrust. Furthermore, functions that
calculate viscous effects, compressibility, aeroelasticity, mass reduction, and others can be included, without the
need to alter the already functional code.

Acknowledgements
R.A. Jimnez thanks Professor Nicols Ochoa from the Universidad de los Andes for his constant cooperation
during the development of the project. Without his help, it would have taken a lot more time to obtain such positive
results.

References
1

Jacobs, E.N., Ward, K.E., Pinkerton, R.M., The Characteristics of 78 Related Airfoil Sections from Tests in the VariableDensity Wind Tunnel. NACA Report No. 460, 1933.
2
Abbot, I.H., Von Doenhoff, A.E., Stivers Jr., L.S., Summary of Airfoil Data. NACA Report No. 824, 1945.
3
Loftin Jr., L.K., Theoretical and Experimental Data for a Number of NACA 6A-Series Airfoil Sections. NACA Report
No. 903, 1948.
4
Anderson Jr., J.D., Computational Fluid Dynamics: the Basics with Applications,5th ed., McGraw Hill, New York.
5
Rodriguez, A.A., Sonne, M.L., Evaluation of Missile Guidance and Control on a Personal Computer, Simulation, Vol. 68,
No. 6, 1997, pp 363 376.
6
Hilton, A., Missile Guidance Simulator, URL: http://www.suslik.org/Simulator/index.html [cited 24 November 2009]
7
Tchieu, A.A., Kutay, A.T., Muse, J.A., Leonard, A., Validation of a Low-Order Model for Closed-Loop Flow Control
enable Flight, AIAA 4th Flow Control Conference, AIAA 2008-3863, Seatle, Washington, 23-26 June 2008.
8
Jee, S.K., Lopez, O., Moser, R.D., Kutay, A.T., Muse, J.A., Calise, A.J., Flow Simulation of a Controlled Airfoil with
Synthetic Jet Actuators, AIAA 19th Computational Fluid Dynamics Conference, AIAA 2009-3673, San Antonio, Texas, 22-25
June 2009.
9
Sriram, A.G., Van Der Weide, E., Kim, S., Tomlin, C., Jameson, A., Aerodynamics and Flight Control of Flapping Wing
Fligt Vehicles: A Preliminary Computational Study, AIAA 43rd Aerospace Sciences Meeting, AIAA 2005-0841, Reno, Nevada,
10-13 January 2005.

12
American Institute of Aeronautics and Astronautics

10
Vanella, M., Preidikman, S., Massa, J., Estudio de Dinmica No-Lineal de Micro-Vehculos Areos de Alas Batientes
Mediante un Modelo Hbrido de Cuerpos Rgidos y Flexibles, Mecnica Computacional, Vol. 24, November 2005, pp. 21612179.
11
Cebeci, T., Platzer, M., Chen, H., Chang, K.C., Shao, J.P., Analysis of Low-Speed Unsteady Airfoil Flows, 1st Ed., Springer,
Long Beach California, 2005, Chap. 1-3.
12
Scott, J., Angle of Attack and Pitch Angle, URL: http://www.aerospaceweb.org/question/aerodynamics/q0165.shtml
[cited 10 November 2010]
13
McLean, D., Automatic Flight Control Systems 1st Ed., Prentice Hall, Cambridge, Great Britain, 1990. Chaps 1-10.
14
Blakelock, J.H., Automatic Control of Aircraft and Missiles, 2nd Ed., John Wiley and Sons, New York, 1991, Chaps. 1-2.
15
Katz, J., Plotkin, A., Low-Speed Aerodynamics 2nd Ed., Cambridge University Press, New Yoek City, 2006, Chaps. 1-2.
16
Fung, Y.C., An Introduction to the Theory of Aeroelasticity, 1st Ed., Dover, San Diego California, 1993, Chap. 13.

13
American Institute of Aeronautics and Astronautics

S-ar putea să vă placă și