PHD 2011 Aswift
PHD 2011 Aswift
Declaration
I confirm that the thesis is my own work, that I have not presented anyone elses
work as my own and that full and appropriate acknowledgement has been given
where reference has been made to the work of others.
...............................................................
Adam Swift
April 2011
ii
Abstract
Aeroelastic phenomena such as flutter can have a detrimental effect on aircraft
performance and can lead to severe damage or destruction. Buffet leads to a reduced fatigue life and therefore higher operating costs and a limited performance
envelope. As such the simulation of these aeroelastic phenomena is of utmost
importance. Computational aeroelasticity couples computational fluid dynamics
and computational structural dynamics solvers through the use of a transformation method. There have been interesting developments over the years towards
more efficient methods for predicting the flutter boundaries based upon the stability of the system of equations.
This thesis investigates the influence of transformation methods on the flutter
boundary predition and considers the simulation of shock-induced buffet of a
transport wing. This involves testing a number of transformation methods for
their effect on flutter boundaries for two test cases and verifying the flow solver for
shock-induced buffet over an aerofoil. This will be followed by static aeroelastic
calculations of an aeroelastic wing.
It is shown that the transformation methods have a significant effect on the
predicted flutter boundary. Multiple transformation methods should be used to
build confidence in the results obtained, and extrapolation should be avoided.
CFD predictions are verified for buffet calculations and the mechanism behind
shock-oscillation of the BGK No. 1 aerofoil is investigated. The use of steady
calculations to assess if a case may be unsteady is considered. Finally the static
aeroelastic response of the ARW-2 wing is calculated and compared against experimental results.
iii
iv
Acknowledgement
I would like to thank my supervisors Professor Ken Badcock and Professor George
Barakos and all the members of the CFD lab, both past and present, at the
University of Liverpool. It has been a privilege to work with you all.
I would like to thank the members of the R&T group at BAE Systems Warton
for their support and encouragement that helped make my annual placements so
productive. I am especially grateful to Mark Lucking for his guidance and support
during these placements.
I would like to thank my family and friends for their their patience and support. I am especially grateful to Maryam for her unwavering love and patience.
I am grateful for the financial support of the Engineering and Physical Sciences Research Council (EPSRC) and BAE SYSTEMS through a case award. I
would also like to extend my gratitude to the 2nd UK Applied Aerodynamics
Consortium for providing computing time on HPCx and HECTOR.
vi
Publications
Computational Study of the BGK No. 1 Aerofoil Buffet, A. Swift and K.J.
Badcock, submitted to the Aeronautical Journal.
Investigation into the Effect of Transformation Methods on Flutter Boundary
Predictions, A. Swift and K.J. Badcock, submitted to the Aeronautical
Journal.
Transonic Aeroelastic Simulation for Envelope Searches and Uncertainty Analysis, K.J. Badcock, S. Timme, S. Marques, H. Khodaparast, M. Pradina,
J.E. Mottershead, A. Swift and A. Da Ronch, Progress in Aerospace Sciences (online), June 2011.
Framework for Establishing the Limits of Tabular Aerodynamic Models for
Flight Dynamics Analysis, M. Ghoreyshi, K. Badcock, A. Da Ronch, S.
Marques, A. Swift and N. Ames, Journal of Aircraft, 48(1), 2011.
Inter-grid Transfer Influence on Transonic Flutter Predictions, A. Swift and
K. Badcock, AIAA-2010-3049, April 2010.
Framework for Establishing the Limits of Tabular Aerodynamic Models for
Flight Dynamics Analysis, M. Ghoreyshi, K. Badcock, A. Da Ronch, S.
Marques, A. Swift and N. Ames, AIAA-2009-6273, August 2009.
vii
viii
Contents
Declaration
Abstract
iii
Acknowledgement
Publications
vii
Table of Contents
xii
List of Tables
xiii
List of Figures
xvii
Nomenclature
xxiv
1 Introduction
1.1 Aeroelasticity . . . . . . . . .
1.2 Prediction of Flutter . . . . .
1.3 Computational Aeroelasticity
1.4 Thesis Organisation . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
2 Formulation
2.1 Navier-Stokes Equations . . . . . . . . . . . .
2.2 Turbulence . . . . . . . . . . . . . . . . . . . .
2.3 The Closure Problem . . . . . . . . . . . . . .
2.3.1 Reynolds Averaging . . . . . . . . . . .
2.3.2 Favre Mass Averaging for Compressible
2.3.3 Boussinesqs Approximation . . . . . .
2.3.4 Vector Form . . . . . . . . . . . . . . .
2.3.5 Non-dimensional Form . . . . . . . . .
2.3.6 Curvilinear Form . . . . . . . . . . . .
2.4 Turbulence Modelling . . . . . . . . . . . . . .
ix
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
. . . .
. . . .
. . . .
. . . .
Flow .
. . . .
. . . .
. . . .
. . . .
. . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
1
7
8
10
.
.
.
.
.
.
.
.
.
.
13
13
15
18
19
19
21
23
23
24
24
2.5
2.6
2.7
2.8
3 Transformation Methods
3.1 Introduction . . . . . . . . . . . . . . . .
3.1.1 The Transformation Problem . .
3.1.2 Notation . . . . . . . . . . . . . .
3.1.3 Requirements . . . . . . . . . . .
3.1.4 Review . . . . . . . . . . . . . . .
3.2 Types of Structural Models . . . . . . .
3.3 Existing Transformation Methods . . . .
3.3.1 Multiquadric-Biharmonic . . . . .
3.3.2 Infinite Plate Spline . . . . . . . .
3.3.3 Thin-Plate Splines . . . . . . . .
3.3.4 Inverse Isoparametric Mapping .
3.3.5 Finite Plate Spline . . . . . . . .
3.3.6 Non-Uniform Rational B-Splines .
3.3.7 Boundary Element Method . . .
3.3.8 Radial Basis Functions . . . . . .
3.3.9 Constant Volume Tetrahedron . .
3.4 Rigid Section Method . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
25
27
27
27
28
29
29
30
31
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
35
35
35
36
37
40
41
42
42
44
46
47
48
48
49
50
52
55
.
.
.
.
.
.
.
.
.
.
57
57
57
57
58
61
62
64
65
67
67
4.3.6 Summary . .
4.4 Flutter Evaluation .
4.4.1 Goland Wing
4.4.2 MDO Wing .
4.5 Summary . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
5 Aerofoil Buffet
5.1 Introduction . . . . . . . . . . . . . . . .
5.2 Test Case . . . . . . . . . . . . . . . . .
5.2.1 Other Aerofoil Buffet Cases . . .
5.3 BGK No. 1 Aerofoil Results . . . . . . .
5.3.1 Grids . . . . . . . . . . . . . . . .
5.3.2 Probe Locations . . . . . . . . . .
5.3.3 Reduced Frequency Calculation .
5.3.4 Convergence Studies . . . . . . .
5.3.5 Pseudo Steps Study . . . . . . . .
5.3.6 Time Step Study . . . . . . . . .
5.3.7 Grid Refinement Study . . . . . .
5.3.8 Buffet Boundary Estimate . . . .
5.3.9 Mach 0.71 Angle of Attack Sweep
5.4 Summary . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
. . .
. . .
. . .
. . .
. . .
. . .
0.0
1.0
0.0
2.0
. . .
. . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
69
69
69
73
75
.
.
.
.
.
.
.
.
.
.
.
.
.
.
77
77
77
82
86
86
88
88
90
90
92
94
94
97
110
.
.
.
.
.
.
.
.
.
.
.
.
117
117
119
121
121
121
124
125
126
135
135
135
137
7 Conclusions
139
7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
xi
153
155
157
xii
List of Tables
2.1 Summary of Solution Method Costs (from Ref. [1]) . . . . . . . .
19
51
71
74
5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
1.396
. . . .
. . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
. 80
. 86
. 89
. 90
. 92
. 94
. 99
. 107
. 110
xiii
xiv
List of Figures
1.1
1.2
1.3
1.4
4
5
6
4.1
4.2
4.3
4.4
4.5
. . . . . . . . . . . . . .
. . . . . . . . . . . . . .
using IPS (from ref. [10])
. . . . . . . . . . . . . .
. . . . . . . . . . . . . .
. . . . . . . . . . . . . .
36
43
46
49
49
53
59
60
61
61
63
64
65
66
4.9
4.10
4.11
4.12
4.13
4.14
4.15
4.16
4.17
5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10
5.11
5.12
5.13
5.14
5.15
5.16
5.17
Case 3: Sections at 98% of the span for the refined extended plate
model, rigidly translated and rotated back to the original orientation.
Case 4: Sections at 98% span through the mode 3 mode shape
rotated and translated back to the original orientation. . . . . . .
Case 5: Vertical displacement of sections at 98% span for the mode
5 mode shape. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Goland wing mode tracking for all modes using CVT . . . . . . .
Goland wing mode tracking for mode 1 for the different structural
models (real part) . . . . . . . . . . . . . . . . . . . . . . . . . . .
Goland wing mode tracking for mode 1 for the different structural
models (real part) . . . . . . . . . . . . . . . . . . . . . . . . . . .
MDO wing mode tracking for all modes using CVT . . . . . . . .
MDO wing mode tracking for mode 1 for the different structural
models (real part) . . . . . . . . . . . . . . . . . . . . . . . . . . .
MDO wing mode tracking for mode 1 for the different structural
models (real part) . . . . . . . . . . . . . . . . . . . . . . . . . . .
66
67
68
70
71
72
73
74
74
5.18 Mach number contours at different instants in time for one period
of oscillation for angle of attack 4.905 , M = 0.095 . . . . . . .
5.19 Streamlines at different instants in time for one period of oscillation
for angle of attack 4.905 . . . . . . . . . . . . . . . . . . . . . . .
5.20 Cross-covariance and wave speed for angle of attack 4.905 . . . .
5.21 BGK No. 1 angle of attack 6.97 results . . . . . . . . . . . . . .
5.22 Pressure wave propagation . . . . . . . . . . . . . . . . . . . . . .
5.23 Unsteady shock movement during one period for angle of attack
6.97 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.24 Mach number contours at different instants in time for one period
of oscillation for angle of attack 6.97 , M = 0.095 . . . . . . . .
5.25 Streamlines at different instants in time for one period of oscillation
for angle of attack 6.97 . . . . . . . . . . . . . . . . . . . . . . .
5.26 Cross-covariance and wave speed for angle of attack 6.97 . . . . .
5.27 BGK No. 1 angle of attack 9.0 results . . . . . . . . . . . . . . .
105
106
108
109
110
111
112
113
114
115
xvii
xviii
Nomenclature
Acronyms
Definition
AGARD
AMB
ARW-2
BEM
BGK
Bauer-Garabedian-Korn aerofoil.
BILU
BSM
CAE
Computational Aeroelasticity.
CFD
CFL
Courant-Friedrichs-Lewy number.
CPU
CSD
CVT
DAST
DES
DNS
DOF
Degrees of Freedom.
FFT
FPS
GVT
HIRENASD
IIM
IMQ
Inverse Multiquadrics.
IPS
LANTRIN
LCO
LES
LIF
MDO
Multi-Disciplinary Optimisation.
MQ
Multiquadrics.
MUSCL
NAE
NASA
NUBS
Non-Uniform B-Spline.
NURBS
ONERA
RANS
RBF
SDM
TDT
TFI
Transfinite Interpolation.
TPS
URANS
USA
Symbols
Definition
qx , qy , qz
Heat fluxes.
, u ,
dx
Displacements.
fi
Blending functions.
Forces.
F, G, H
Complex eigenvector.
Vector of residuals.
Vector of unknowns.
Eigenvectors.
Coordinates.
x, x,
x
Chord.
CN
Ck
Kolmogorov constant.
CL
Lift coefficient.
cp
cv
Total energy.
E()
Frequency.
xxi
fi
Body forces.
Total enthalpy.
Identity matrix.
Jacobian.
Stiffness matrix.
Reduced frequency.
kij
kx , ky , kz
Length.
Mass matrix.
M , C , K
Ni (, )
Shape functions.
Pressure.
p(x)
Linear polynomial.
Pr
Prandtl number.
P rT
Rotation matrix.
r()
Blending function.
r2
Re
Reynolds number.
ReT
Temperature.
xxii
Transformation matrix.
Time.
T0
Tp
Velocity.
u, v, w
ud
uu
Virtual work.
x, y, z
Cartesian coordinates.
xs
s(x)
Interpolant.
Greek Symbols
Definition
, ,
CVT constants.
Coefficients.
, ,
Modal coordinate.
, u ,
Thermal conductivity.
Wavenumber.
Eigenvalues.
Dynamic viscosity.
Kinematic viscosity.
Modal matrix.
, ,
Blending functions.
xxiii
Density.
Support radius.
ij
Shear Stresses.
ijR
Reynolds stresses.
Rate of dissipation.
, ,
Subscripts
Definition
Freestream value.
Fluid.
ff
Fluid-fluid.
fs
Fluid-structure.
Inviscid.
i, j, k
Indices=1,2,3,. . . .
Laminar.
Structure.
sf
Structure-fluid.
ss
Structure-structure.
Turbulent.
Viscous.
Superscripts
Definition
Mapped quantity.
Current time-step.
n+1
Next time-step.
Reynolds stress.
Dimensional variable.
xxiv
Chapter 1
Introduction
1.1
Aeroelasticity
out of five fasteners missing in the left wing, this reduced the torsional stiffness
of the elevator. The reduced torsional stiffness caused a limit cycle oscillation of
the elevator, which caused the wing to flutter [19]. A history of flutter can be
found in ref. [16].
Limit cycle oscillation is a type of flutter, which is a self sustaining limited
amplitude oscillation with the amplitude limited due to non-linearities. Both the
F-16 and F/A-18 have encountered limit cycle oscillations at high subsonic and
transonic speeds for certain store configurations with AIM-9 missiles at the wing
tips and heavy stores on the outboard pylons [20]. Limit cycle oscillations can
damage stores, reduce their fatigue life and targeting effectiveness [20].
Control surface buzz is similar to limit cycle oscillation in that the amplitude
of the oscillations are limited, however the mechanism behind it is different. It is
caused by the interaction of a shock wave, boundary layer and control surface rotation mode. There are two main types of control surface buzz, the first is caused
by buffet of the control surface by the separated flow due to shock-boundary layer
interaction upstream and the second type is due to shock wave oscillation on the
control surface itself with no separation involved.
Buffeting is defined as the structural response to the aerodynamic excitation
produced by separated flows. Buffet can be caused by massive separation over a
lifting surface (stall buffet), separation due to shock/boundary-layer interaction
(Mach number buffet) or vortex breakdown. Related to buffet is transonic buffet
(also refered to as shock buffet), this is an aerodynamic phenomenon in which a
shockwave is generated over a lifting surface that interacts with the boundarylayer causing partially or fully separated flow. This results in a self-sustained
oscillating flow field and occurs without structural motion, however the frequency
of shock oscillation is usually in the range of the low-frequency modes of an elastic
structure which would lead to buffeting.
One of the first records of buffet dates back to 21st July 1930 where a commercial aeroplane crashed at Meopham, Kent. In this case the aircraft flew into
a cloud and shortly afterwards came down in pieces, the subsequent investigation found that there was a strong updraft in the area. This caused the angle
of attack to increase sharply leading to massive separation over the wing. The
tail was situated in this region of highly unsteady separated flow and therefore
subjected to intense forced vibrations that ultimately caused the tail to fail [21].
This accident was an example of stall buffet. A more modern example was on the
American F-15 fighter jet caused by leading edge separation at angles of attack
of over 22 degrees, which excited the first torsion mode [22]. In 1998 the cost of
repairing and replacing F-15 vertical tails due to fatigue was estimated to be $5-6
million per year [4], in addition to the loss of flight readiness. This type of buffet
3
is caused by the separation over wings, the F-16 had severe damage to ventral
fins positioned at the underside of the rear fuselage caused by the upstream separation of the airflow over the LANTRIN targeting pods [4]. This is shown in the
Figure 1.1 and the extensive damage to the ventral can be seen.
This type of buffet is very much an unsolved problem for modern fighter jets with
the EF-2000 type fighter [26, 27], F-22 [28, 29] and the F-35 [30] suffering from
different levels of buffet.
Figure 1.2: Vortex breakdown over the F/A-18 (from ref. [5])
Buffet effects all aircraft, both civil and military, however the causes and flight
regimes are different. Figure 1.3 gives a comparison of the flight envelope for the
buffet criteria for a civil transport aircraft and a military fighter jet. Buffet is
split into different levels of severity denoted as light, moderate and heavy. Civil
aircraft are designed to cruise below buffet onset and maneuver to at least 1.3g
without any buffeting [6]. It is allowed to enter the buffet region for short periods
during rare severe gusts, but it should not create any structural or control issues.
Military fighters fly much further below the buffet boundary, but for performance
reasons will frequently maneuver deep inside the buffet region. This can lead to
a significant reduction in the fatigue life of the aircraft [6].
Over the years many potential solutions to buffet have been proposed, the
simplest being to limit the flight envelope. On the F/A-18, streamline fences
were introduced on the upper surface of the leading edge extension to reduce the
buffet loads, which although significantly reducing the buffet loads also caused a
slight loss in lift. This solution was less effective at higher angles of attack [24].
Piezoelectric actuators have been fitted to the tail of aircraft to actively control
the tail. There are other solutions such as active control surfaces, sucking and
blowing to control vortices and finally stiffening the structure. There have been
hybrid approaches such as combining active control surfaces and piezoelectric
5
Figure 1.4: Comparison between the flutter boundaries predicted using the linear
method and the wing-only configuration for the weakened structural model. (from
ref. [7])
higher operating costs. This all adds to the complexity of clearing an aircrafts
flight envelope. However in more recent years aeroelasticity has been used in
multidisciplinary optimisation routines to attempt to take advantage of the reduced structural weight possible when aeroelasticity is taken into account [36].
As mentioned previously, the X-53 Aeroelastic Research Wing has attempted to
make use of aeroelasticity to increase the control authority of the aircraft. Also
by taking aeroelasticity into account the drag produced can be reduced through
the definition of a jig shape.
1.2
Prediction of Flutter
The standard methods for calculating flutter have been the k and p-k methods.
Both of these methods work by coupling linear aerodynamics and linear structural
mechanics and then solving an eigenvalue problem. In the k method the aerodynamics are introduced into a vibration analysis using complex inertial terms,
the p-k method introduces the aerodynamics as frequency dependent stiffness
and damping terms. These methods are explained in reference [37]. The linear
aerodynamics usually come from the doublet lattice method for subsonic flow.
The doublet lattice method is based upon potential flow, which assumes the flow
is incompressible, inviscid, irrotational and that the angle of attack is small with
7
thin lifting surfaces. Therefore this method does not capture non-linearities in
the flow due to shockwaves, boundary-layers and flow separation. This methodology is implemented in commercial packages such as NASTRAN and performs
well inside its region of applicability and as such is the industry standard that all
other methods are compared against.
However in the transonic regime as previously mentioned there is the transonic
dip in the flutter boundary. This cannot be captured using linear aerodynamics
due to the nonlinearities caused by the formation of shockwaves. To overcome
this problem there has been a move towards unsteady nonlinear aerodynamics,
which can be predicted by, for example, the Euler equations. This has focused
on time-marching calculations and the technology which has been developed to
allow these calculations will be covered in more detail in the next section. In
these time-marching calculations the system is given an initial perturbation and
the temporal response is calculated to determine its growth or decay to infer the
system stability. This has been shown to be effective at calculating the flutter
boundary, for the example of a wing [7] and for a full aircraft [38]. However
this approach is unfeasible for the use of clearing a flight envelope due to the
computational expense.
In recent years there has been an effort to reduce the cost of flutter calculations
at the University of Liverpool and this has lead to the development of the Schur
code [39, 40, 41]. This solver computes the onset of flutter as a stability problem
for the steady state of the coupled system. It assumes Hopf Bifurcation, which
is when stability is lost as the coupled systems Jacobian matrix has a pair of
eigenvalues that cross the imaginary axis. This system is solved using the Schur
complement and has been parallelised to allow realistically large problems to
be computed. The method has been found to be about two orders of magnitude
faster than the time domain approaches. For the flutter calculations in this thesis
the Schur solver will be used and the solver is presented in more detail in Section
2.8.
1.3
Computational Aeroelasticity
[42].
There are two methodologies for computational aeroelasticity, the monolithic
and the partitioned approaches. In the monolithic approach the governing equations are reformulated to combine the fluid and structural equations, which are
then solved and integrated in time simultaneously. The partitioned approach is
more commonly used and uses separate aerodynamic and structural solvers. This
necessitates a method of transferring the aerodynamic loads to the structure and
the displacements to the aerodynamic solver. Since there are issues with solving
the system of equations in the monolithic approach, the partitioned approach has
seen considerable effort in solving the various technological challenges required
to make this approach work. The partitioned approach can be extended to include additional disciplines in a straight-forward manner, whereas the monolithic
approach requires the governing equations to be reformulated to include the additional discipline. The main technological challenges for the partitioned approach
are,
how to deform the CFD mesh
how to sequence the aerodynamic and structural solvers
how to transfer the loads and displacements between solvers
1.4
Thesis Organisation
The objective of this work is to investigate the influence of transformation methods on the flutter boundary and to consider the simulation of shock-induced buffet
of a transport wing. This involves testing a number of transformation methods
for their effect on flutter boundaries, verifying the flow solver for shock-induced
10
buffet over an aerofoil and then for static aeroelastic calculations of an aeroelastic
wing. Reviews of these topics can be found in their respective chapters (Chapter
3 for transformation methods, Chapter 5 for shock-induced buffet of an aerofoil
and Chapter 6 for shock-induced buffet of a transport wing).
Chapter 2 introduces the formulation of the solvers to be used in this work.
This includes the Aeroelastic Multiblock (AMB) flow solver and the Schur solver
for fast flutter boundary calculations.
Chapter 3 introduces transformation methods and a description of their formulation. A different approach to transformation is proposed called the rigid ribs
approach that is used for beam stick models.
Chapter 4 presents the evaluation of the transformation methods, both for
mode shape reconstruction and their effect on flutter boundary predictions. This
is shown using two test case, the Goland and the multi-disciplinary optimisation
(MDO) wings. The effect of changing the direction of extrapolation for a beam
stick model is also presented.
Chapter 5 is devoted to the calculation of buffet over the BGK No. 1 aerofoil
to verify the AMB solver for shock-induced buffet. In this chapter a number of
convergence studies are presented to find the effect of various parameters on the
solver and solution. This is followed by an estimate of the buffet boundary using
steady state calculations. The estimated buffet boundary is then verified using
unsteady calculations and the mechanism for the shock oscillation is investigated.
Chapter 6 presents the static aeroelasticity of the Aeroelastic Research Wing
(ARW-2) to verify that the aeroelastic deformation and aerodynamics can be
captured by the AMB solver. This is followed by a buffet search using steady
state calculations to find possible buffet cases.
Chapter 7 concludes the thesis and offers suggestions for future work. The appendices offer more detailed formulations of some of the transformation methods
presented in Chapter 3.
11
12
Chapter 2
Formulation
2.1
Navier-Stokes Equations
(2.1)
Energy Equation
This equation is derived from the principle of conservation of energy.
qx qy qz
(E) (uH) (vH) (wH)
+
+
+
= q
+
+
t
x
y
z
x
y
z
(uxx ) (uyx ) (uzx ) (vxy ) (vyy ) (vzy )
+
+
+
+
+
+
x
y
z
x
y
z
(wxz ) (wyz ) (wzz )
+
+
+
+ (fx u + fy v + fz w)
(2.5)
x
y
z
where q is the rate of volumetric heat addition per unit mass, qi are the heat
2
2
2
fluxes, the total energy is given by E = e + u +v2 +w and the total enthalpy is
given by H = E + p .
Additional Equations
Since air is a Newtonian fluid the shear stresses are given by
xx = 2 u
+ ( u
+
x
x
+ ( u
+
zz = 2 w
z
x
xz = zx =
u
z
v
y
v
y
+
+
w
)
z
w
)
z
w
x
yy = 2 v
+ ( u
+ v +
y
x y
v
xy = yx = x
+ u
y
w
yz = zy = y + v
z
w
)
z
(2.6)
T
T0
23
T0 + 110
T + 110
(2.7)
T
x
qy =
T
y
qz =
T
z
(2.8)
where is the thermal conductivity. So for a given time and location, the NavierStokes equation have the variables , u, v, w, p, e and T to solve for and only
five equations. By assuming that air is a perfect gas an additional equation can
be specified
p = RT
(2.9)
h = cp T
14
(2.10)
where cv is the specific heat at constant volume and cp is the specific heat at
constant pressure. Then the heat fluxes can be rewritten as
qx =
h
P r x
qy =
h
P r y
qz =
h
P r z
(2.11)
2.2
Turbulence
When the Reynolds number is sufficiently low the fluid flows with no disruption
between layers. This type of flow is called laminar. As the Reynolds number
is increased the flow becomes unstable and transitions to a turbulent flow. A
turbulent flow has rapid fluctuations in velocity and pressure and is inherently
three-dimensional and unsteady. Turbulence has a range of scales that are linked
to structures in the flow called turbulent eddies. This can be seen in Figure 2.1
of a laser induced florescence image of a submerged turbulent jet. There are
large scale eddies of the order of the jet width as well as much smaller eddies.
Turbulent eddies are local swirling motions where the vorticity is large. The
largest eddies interact with and extract energy from the mean flow by a process
called vortex stretching. The mean velocity gradients distort the eddies if they
are aligned in a direction in which the mean velocity gradients can stretch them.
Each turbulent eddy has a characteristic length, , velocity, u and time, scale
associated to it. For the larger eddies their Reynolds number tends to be large
(on the order of the mean flow) suggesting that they are dominated by inertial
forces. The smaller eddies are also stretched, but mainly by slightly larger eddies
as opposed to the mean flow. This leads to a cascade of energy from the larger
eddies to the smaller eddies. An energy spectrum of turbulence is shown in
Figure 2.2. It can be seen that most of the energy is contained in the larger
eddies and as such are referred to as the energy containing eddies. The largest
eddies are unaffected by viscosity and depend on their velocity and length scales.
Dimensional analysis gives the spectral energy content as
E() u2
(2.12)
Figure 2.1: False-colour image of the far-field of a submerged turbulent jet using
laser induced fluorescence (LIF). (from ref. [8])
Figure 2.2: Energy spectrum for a turbulent flow, log-log scales. (from ref. [9])
16
smallest eddies contain the least energy and are dominated by viscous forces.
The smallest of these scales of turbulence are called the Kolmogorov scales of
length, , velocity, u and time, . The Kolmogorov scales can be derived using
dimensional analysis and are only dependent on the rate of dissipation, and the
kinematic viscosity, and are given by
=
1/4
u = ()
1/4
1/2
(2.13)
The smallest of the turbulent scales is fixed by viscous dissipation of energy, but is
still many times larger than any molecular scale, as such turbulence is a continuum
phenomenon. However the rate of dissipation is controlled by the rate at which
these smallest eddies receive energy from the larger eddies. Through dimensional
analysis the ratio of the scales can be obtained based upon the turbulent Reynolds
1/2
number, ReT = k l ,
3/4
ReT ,
u
1/4
ReT ,
u
1/2
ReT
(2.14)
This shows that as the Reynolds number increases so does the range of scales in
the flow. Kolmogorov argued that the spectral energy of the smallest scales should
be a function of the rate of dissipation and the kinematic viscosity. Through
dimensional analysis
E( = 1/) 5/4 1/4
(2.15)
The diffusive action of the viscous stress tends to smear out any anisotropic
behaviour in the smallest eddies leading to the smallest eddies being isotropic.
In Figure 2.2 the smallest eddies are referred to as the viscous range. In Figure
2.2 there is an intermediate range of eddies referred to as the inertial subrange.
These eddies were assumed by Kolmogorov to be large enough to unaffected by
viscosity, but small enough to be expressed as a function of energy dissipation,
so through dimensional analysis
E() = Ck 2/3 5/3
for
1
1
(2.16)
is dominated by the large eddies. These large eddies cause increased mixing of
mass, momentum and energy and is called turbulent mixing.
The rate of dissipation, has been used throughout this section and its definition is
=
dk
dt
(2.17)
2.3
Table
Method
URANS
DES
LES
DNS
2.3.1
2.1: Summary
Empiricism
Strong
Strong
Weak
None
of Solution
Grid Size
107
108
1011.5
1016
Reynolds Averaging
(2.18)
1
=
t
t+t
f dt
(2.19)
f dt 0
(2.20)
t+t
The time average of the product of two flow variables f g is given by f g + f g , because the average of the product of two fluctuating components is not necessarily
zero.
2.3.2
The above Reynolds averaging approach can be used for compressible flow, however it leads to a very complex system of equations. To avoid this Favre mass
averaging is used. Favre mass averaging defines mass averaged variables, so for a
generic flow variable f the mass averaged variable, f is
f
f =
(2.21)
where is the time averaged density. In a similar way to Reynolds averaging the
flow variables can be decomposed into mass averaged f and fluctuating f parts.
f = f + f
19
(2.22)
It is important to note that the time averages of the double primed fluctuating
quantities are not equal to zero, however the time average of f is equal to
zero. This procedure is a mathematical simplification that removes the density
fluctuations from the averaged equations, but it does not remove the effect of the
density fluctuations from the turbulence. To get to the Favre averaged NavierStokes equations the following flow variables are substituted into the NavierStokes equations given by Equations 2.1 to 2.5 and then time averaged.
u = u + u
e = e + e
qx = qLx
+ qx
= +
v = v + v
w = w + w
+h
h=h
T = T + T
qy = qLy
+ qy qz = qLz
+ qz
p = p + p
Continuity Equation
(
u) (
v ) (w)
+
+
+
=0
t
x
y
z
(2.23)
Momentum Equations
In the x direction
R
R
)
u) (
uu) (
uv) (
uw)
p (xx + xx
) (yx + yx
(
+
+
+
=
+
+
t
x
y
z
x
x
y
R
(zx + zx )
+
(2.24)
z
In the y direction
R
R
) (yy + yy
)
p (xy + xy
(
v ) (
v u) (
v v) (
v w)
+
+
+
=
+
+
t
x
y
z
y
x
y
R
(zy + zy )
+
(2.25)
z
In the z direction
R
R
)
(w
p (xz + xz
u) (w
v ) (w w)
) (yz + yz
(w)
+
+
+
=
+
+
t
x
y
z
z
x
y
R
(zz + zz )
+
(2.26)
z
20
Energy Equation
uH) (
v H) (wH)
(E) (
+
+
+
=
t
x
y
z
[qLx
+ qT x ] +
[qLy
+ qT y ] +
[qLz
+ qT z ]
x
y
z
R
R
R
u(xx + xx
+
) + v(yx + yx
) + w(
zx + zx
)
x
R
R
R
) + v(yy + yy
) + w(
zy + zy
)
+
u(xy + xy
y
R
R
R
+
u(xz + xz
) + v(yz + yz
) + w(
zz + zz
)
z
zz = 2 zw 32 ( xu +
xz = zx =
v
y
v
y
w
+
+
)
z
w
)
z
2 +
v 2 +w
2
+ u
+ k and H = h
v
v
yy = 2 y
32 ( xu + y
+
v
+ yu
xy = yx = x
yz = zy = yw + zv
R
R
xx
= u u
yy
= v v
R
R
R
zz = w w
xy = yx
= u v
R
R
R
R
xz
= zx
= u w yz
= zy
= v w
(2.27)
(2.28)
2 +
v 2 +w
2
+ k.
)
z
(2.29)
(2.30)
and the heat flux components (subscripts L and T are for laminar and turbulence
respectively) are
h
qLx = Pr x
qT x = PrTT xh
(2.31)
qLy = Pr yh qT y = PrTT yh
T h
h
qLz = P r z qT z = P rT z
where T is the turbulent eddy viscosity and P rT is the turbulent Prandtl number.
This introduces six more unknowns that need to be modeled to close the problem.
For general three-dimensional flows there are 6 unknown mean-flow properties,
density, pressure, enthalpy and the three velocities, as well as six Reynolds-stress
unknowns, which brings the total number of unknowns to twelve. There are only
five equations to solve and so more equations are required to close the problem.
This is referred to as the closure problem.
2.3.3
Boussinesqs Approximation
One common approach to solve the closure problem is using the Boussinesq approximation. This is based on an analogy between the Reynolds stresses and
21
the viscous stresses. This leads to the equations for the Reynolds stresses being
rewritten in the form
R
xx
= u u = 2T
v
y
1
3
u
2
x
u
2
x
v2
x
v2
x
w
2
x
2
23 k
23 k
2
u
v2
w
2
R
zz
= w w = 2T zw 31 x
+ x
+ x
2 k
2 3 2
v
u
v
w
2
R
R
13 x
32 k
+ x
+ x
xy
= yx
= u v = 2T 12 yu + x
u
2
v2
w
2
R
R
23 k
+ x
+ x
xz
= zx
= u w = 2T 21 zu + xw 31 x
2
u
v2
w
2
R
R
yz
= zy
= v w = 2T 21 zv + yw 31 x
+ x
+ x
32 k
(2.32)
where k is the specific turbulent kinetic energy given by
R
yy
= v v = 2T
1
3
1
k = u u
2
(2.33)
The two main assumptions used in the Boussinesq approximation are that the
Reynolds stresses can be defined at each point in space and time by mean velocity
gradients and that the turbulent eddy viscosity is a scalar property of the flow
with the relationship between the Reynolds stresses and the mean velocity gradients being linear. This approximation reduces the number of additional unknowns
from six to one. The majority of turbulence models calculate this turbulent eddy
viscosity as a function of velocity and length scales. There are a number of limitations of using this approximation. When a turbulent boundary-layer is perturbed
from its equilibrium state a new equilibrium state is not reached for at least ten
boundary-layer thicknesses downstream [9]. This observation means that the
Boussinesq approximation along with the equilibrium approximations implicit in
algebraic models will not provide an accurate description of separated flow. The
Boussinesq approximation is not valid for flows over curved surfaces due to significant streamline curvature, which gives rise to uneven normal Reynolds stresses.
Flows that are three-dimensional can cause the Boussinesq approximation not
to be valid, this can be demonstrated in straight non-circular ducts where the
turbulence model using this approximation fails to predict secondary motions.
Flows which have large changes in mean strain rate such as separation cannot be
described by the Boussinesq approximation. Although the Boussinesq approximation may not be valid in these flow conditions many turbulence models using
this approximation have been successfully applied to these flow conditions.
22
2.3.4
Vector Form
The Navier-Stokes equations (in this case the Favre mass averaged form) can be
written in vector form,
W (Fi Fv ) (Gi Gv ) (Hi Hv )
+
+
+
=0
t
x
y
z
where W is the solution vector of conserved variables given by
u
v
W=
w
E
The inviscid vectors are given by
uv
uu + p
v u Gi =
v v + p
Fi =
w
w
u
v
uH
vH
The viscous vector are given by
Fv =
Re
Gv =
Re
Hv =
Re
Hi =
uw
v w
w w + p
w H
(2.34)
(2.35)
0
R
xx + xx
R
xy + xy
R
xz + xz
R
R
R
u(xx + xx ) + v(xy + xy ) + w(
xz + xz
) (qLx
+ qT x )
0
R
yx + yx
R
yy + yy
R
yz + yz
R
R
R
u(yx + yx
) + v(yy + yy
) + w(
yz + yz
) (qLy
+ qT y )
0
R
zx + zx
R
zy + zy
R
zz + zz
R
R
R
u(zx + zx
) + v(zy + zy
) + w(
zz + zz
) (qLz
+ qT z )
2.3.5
(2.36)
(2.37)
Non-dimensional Form
x
,
L
y=
p=
y
,
L
p
2 ,
U
z = Lz ,
u = Uu , v =
t
t = L /U
T = TT , e =
,
23
v
,
U
e
2 ,
U
w = Uw ,
(2.38)
2.3.6
Curvilinear Form
The Navier-Stokes equations can be rewritten in curvilinear form [10], this makes
solving the equations on a body-fitted grid easier. If the physical region of interest
is described by a Cartesian co-ordinate system x, y, z then a mapping is defined
onto a computational domain , , (a unit cube), the transformation can be
written as
= (x, y, z),
= (x, y, z),
= (x, y, z),
t=t
(2.39)
(, , )
(x, y, z)
(2.40)
i F
v ) (G
i G
v ) (H
i H
v)
W
(F
+
+
+
=0
t
(2.41)
= 1W
W
J
i = 1 (x Fi + y Gi + z Hi )
F
J
i = 1 (x Fi + y Gi + z Hi )
G
J
i = 1 (x Fi + y Gi + z Hi )
H
J
v = 1 (x Fv + y Gv + z Hv )
F
J
v = 1 (x Fv + y Gv + z Hv )
G
J
v = 1 (x Fv + y Gv + z Hv )
H
J
(2.42)
where
2.4
Turbulence Modelling
If the Boussinesq approximation has been used then the eddy viscosity needs to
be calculated using a turbulence model. Wilcox defines an ideal turbulence model
24
as a model that introduces the minimum complexity whilst capturing the essence
of the relevant physics. The models for the Favre mass averaged Navier-Stokes
equations can be split into two categories, those that use the Boussinesq approximation and those that do not. This section will concentrate on a Boussinesq
approximation model. The turbulence model that will be presented here is SST
two equation model.
2.4.1
Eddy Viscosity
T =
a1 k
,
= max {a1 ,
|
{z
Stress Limiter
25
F2 }
}
(2.43)
(k)
|t{z }
Rate of change of k
ui
+
uj k) = ij
(
k +
| {z }
xj
xj
xj
| {z } | {z } Dissipation |
Convection
Production
k k
+ k
xj
{z
}
Diffusion
(2.44)
()
|t {z }
Rate of change of
ui
2 +
(
+
uj ) = ij
| {z }
xj
k
xj
xj
| {z } | {z } Dissipation |
Convection
Production
k
+ 2(1 F1 )w2
xj xj
|
{z
}
k
+ w
xj
{z
}
Diffusion
(2.45)
Cross-diffusion
Closure Coefficients
9
(2.46)
100
the remaining closure coefficients are a blend of the k- and k- coefficients using
the equation
a1 = 0.31,
= 0.41,
= F1 1 + (1 F1 )2
(2.47)
where is a quantity to be blended and the subscripts 1 and 2 stand for k- and
k- coefficients respectively.
!
!
k
4
k
500
w2
F1 = tanh(arg14 ), arg1 = min max
,
,
y y 2
CDkw y 2
k
20
CDkw = max 2w2
(2.48)
, 10
xj xj
The coefficients for the k- part are
k1 = 0.85,
w1 = 0.5,
1 = 0.075,
1 =
1 w1 2
(2.49)
w2 = 0.856,
2 = 0.0828,
and finally
F2 = tanh(arg22 ),
arg2 = max
26
2 =
2 w2 2
2 k 500
,
y y 2
(2.50)
(2.51)
Additional improvements have been suggested by Menter et al. [54], they are a
production limiter
Pk = ij
ui
xj
Pk = min (Pk , 10 k)
(2.52)
and vorticity is replaced with strain rate in Equation 2.43. However these improvements are not implemented in the current solver.
2.5
Flow Solver
All the computations to be presented were performed using the Aeroelastic MultiBlock (AMB) in-house flow solver [55]. A wide variety of problems have been
studied using this code including cavity flows, delta-wing aerodynamics, rotorcraft problems, flutter[7] and control surface buzz[10]. The governing equations
are discretised using a cell-centred finite volume solver combined with an implicit
dual-time stepping method. In this manner, the solution marches in pseudotime for each real time-step to achieve fast convergence. The discretisation of
the convective terms uses Oshers upwind scheme. Monotone upstream-centred
schemes for conservation laws (MUSCL) interpolation is used to provide nominally second-order accuracy and the van Albada limiter is also applied to remove
any spurious oscillations across shock waves. Central differencing is used to discretise the viscous terms, with the resulting non-linear system of equations generated being solved by integration in pseudo-time using a second-order backward
difference. A Generalised Conjugate Gradient method is then used in conjunction
with a Block Incomplete Lower-Upper (BILU) factorisation as a preconditioner
to solve the linear system of equations, which is obtained from a linearisation in
pseudo-time. A number of turbulence models are available in the solver as well as
large-eddy simulation (LES) and detached eddy simulation (DES), however for
the calculations presented throughout the RANS equations were solved using a
turbulence model.
2.6
Mesh Movement
The mesh is deformed using a hybrid approach where the block vertices are
deformed using the spring analogy and the mesh is deformed inside each block
using transfinite interpolation (TFI).
2.6.1
The spring analogy has been used by many researchers over the years to deform
unstructured [43, 45] and structured meshes. However for structured grids there
27
are more efficient methods of mesh deformation, such as TFI. With this in mind
the spring analogy was used to deform the block vertices and TFI was used for
the edges, faces and volumes. The spring analogy works by assuming each block
edge is a spring with a stiffness that is the inverse of its length. To ensure that
the blocks do not skew or invert additional springs were added diagonally inside
the blocks. For the ith vertex connected to the jth vertex by the edge ij, the
stiffness is given by
kij = p
(xi xj
)2
1
+ (yi yj )2 + (zi zj )2
(2.53)
(2.54)
where the ith vertex is connected to m other vertices (denoted by j) using the
current displacements (superscript n) for the update the displacement (superscript n + 1). The Dirichlet boundary condition is used (fixed displacements at
the boundaries). Once the displacements have been converged adequately then
the updated vertex coordinates are given by
converged
xnew
= xold
i
i + dxi
(2.55)
Once the block vertices have been updated the block edges can be updated
using TFI with the block vertices as boundary conditions. Then the block faces
are interpolated using TFI with the updated block edges as boundary conditions.
Finally the block volumes are updated via TFI from the new block faces.
2.6.2
(2.56)
and therefore the displacements along the edge can be calculated from the edge
vertices,
dx() = dx(0)[1 r()] + dx(1)r()
(2.57)
28
2.6.3
(2.58)
Now the updated positions of the block edges are known, they can now be used
as the boundary conditions for updating the block faces [10]. The faces can be
transformed to a unit square with local coordinates (,) to allow the displacement
to be interpolated using linear blending functions.
dx(, ) = f1 (, ) + f2 (, )
(2.59)
f1 (, ) = [1 r1 ()]dx(0, ) + r3 ()dx(1, )
(2.60)
where
and
f2 (, ) = [1 r4 ()] [dx(, 0) f1 (, 0)] + r2 () [dx(, 1) f1 (, 1)]
(2.61)
The blending functions ri along each edge are calculated in the same manner as
before, using Equation 2.56. Finally the updated coordinates of a given node are
obtained by
x(, ) = xinitial (, ) + dx(, )
(2.62)
2.6.4
Each block is bounded by six updated faces with the displacements known at all
their nodes [10]. The face numbering are given in Figure 2.3. Linear blending is
used between pairs of faces.
dx(, , ) = f1 (, , ) + f2 (, , ) + f3 (, , )
(2.63)
f1 (, , ) = (1 )dx(0, , ) + dx(1, , )
(2.64)
where
f2 (, , ) = (1 ) [dx(, 0, ) f1 (, 0, )] + [dx(, 1, ) f1 (, 1, )] (2.65)
f3 (, , ) = (1 ) [dx(, , 0) f2 (, , 0)] + [dx(, , 1) f2 (, , 1)] (2.66)
The blending functions, , and are given by
= (1 r )(1 r )r1 + (1 r )r r2 + r (1 r )r3 + r r r4
(2.67)
(2.69)
29
(2.68)
(2.70)
(2.71)
(2.72)
(2.73)
(2.74)
and the blending functions along each edge ri are calculated as before using
Equation 2.56. Finally the update coordinates of a given node is calculated using
x(, , ) = xinitial (, , ) + dx(, , )
2.7
(2.75)
Structural Solver
(2.76)
where M is the mass matrix, C is the viscous damping matrix and K is the
stiffness matrix, all of size n n. Also x is the time dependent displacements
and f is the time dependent external force vector both of size n. To calculate the
undamped free vibration characteristics, Equation 2.76 is rewritten as
Mx
+ Kx = 0
30
(2.77)
This equation can be solved subsituting x = Xeit into Equation 2.77. The
solution is
[K 2 M]X = 0
(2.78)
This is the premultiplied by M 1 to get
[M 1 K 2 M 1 M]X = 0
(2.79)
(2.80)
This can be solved to give the eigenvalues and the eigenvectors X, which
are the mode shapes. The mode shapes are usually put in to a modal matrix
= [X1 , X2 . . . , XN ] and are mass normalised to give T M = I. Because the
system is assumed to be linear, its characteristics can be determined once prior
to any aeroelastic calculations.
Equation 2.76 can be transformed into modal space by using x = where
is the modal coordinate. First by premultiplying by T and then substituting
x = into Equation 2.76 to get
M + C + K = T f
(2.81)
p
X
i i
(2.83)
i=1
2.8
Schur Code
Aeroelastic calculations can be simulated by coupling the fluid and the structural
solvers together and marching through time. This method is computationally
expensive. The Schur solver views the problem of computing flutter onset as a
31
stability problem for a steady state of the coupled problem [41]. The semi-discrete
form of the coupled CFD-CSD system is written as
dw
= R(w, )
dt
(2.84)
w = [wf , ws ]T
(2.85)
where
is a vector containing the fluid unknowns (wf ) and the structural unknowns (ws ),
and
R = [Rf , Rs ]T
(2.86)
is a vector containing the fluid residual (Rf ) and the structural residual (Rs ).
The residual also depends on a parameter (in this case is altitude) which is
independent of w. An equilibrium w0 of this system satisfies R(w0 , ) = 0.
The linear stability of equilibria of Equation 2.84 is determined by eigenvalues
of the Jacobian matrix J = R/w. The Schur solver does a stability analysis
based on the coupled system Jacobian which includes the Jacobian of the CFD
residual with respect to the CFD and structural unknowns. The calculation of
the Jacobian J is most conveniently done by partitioning the matrix as
" R R #
f
f
Jf f Jf s
wf
ws
J = Rs Rs =
(2.87)
Jsf Jss
w
w
f
The details of the Jacobian calculation are given in references [39] and [40].
It is conventional in aircraft aeroelasticity for the structure to be modelled
by a small number of modes, which leads to the number of the fluid unknowns
being far greater than the structural unknowns. This means that the Jacobian
matrix has a large sparse block Jf f surrounded by thin strips for Jf s and Jsf . As
described in reference [41] the stability calculation is formulated as an eigenvalue
problem, focussing on eigenvalues of the coupled system that originate from the
uncoupled block Jss .
The coupled system eigenvalue problem is written as
Jf f Jf s
p = p
(2.88)
Jsf Jss
where p = [pf , ps ]T and are the complex eigenvector and eigenvalue respectively. The eigenvalue (assuming it is not an eigenvalue of Jf f ) satisfies [56] the
nonlinear eigenvalue problem
S()ps = ps
where S() = Jss Jsf (Jf f I)1 Jf s .
32
(2.89)
The nonlinear Equation 2.89 is solved using Newtons method. Each iteration
requires the formation of the residual, S()ps ps and its Jacobian matrix.
The calculation of the correction matrix, Jsf (Jf f I)1 Jf s , is required to form
the Jacobian matrix with respect to ps and . This can be achieved through
2n solutions of a linear system against Jf f I, one for each column of Jsf
with n being the number modes retained. These solutions are then multipled
against Jsf . Now, for each value of the bifurcation parameter, there are multiple
solutions of the nonlinear system in Equation 2.89, and so the cost of forming
the correction matrix at each Newton step, for each solution and for a range of
structural parameters becomes too high. To overcome this the expansion
2
2 3
(Jf f I) = Jf1
f + Jf f + Jf f + . . .
(2.90)
is used where must be small for the series to converge. Note that this assumption
is not restrictive since we assume that the eigenvalue we are calculating is a
small change from the eigenvalue 0 of Jss . Then 0 can be used as a shift to
the full system eigenvalue by replacing Jf f by Jf f 0 I and Jss by Jss 0 I.
This modifies the nonlinear eigenvalue problem in Equation 2.89 by redefining
S() = (Jss 0 I) Jsf (Jf f 0 I I)1 Jf s . The series approximation then
becomes
(Jf f 0 I I)1 = (Jf f 0 I)1 +(Jf f 0 I)2 +2 (Jf f 0 I)3 +. . . (2.91)
When the shifted problem is solved for , the eigenvalue of the original system is
then 0 +. The terms (Jf f 0 I)1 Jf s , (Jf f 0 I)2 Jf s can be pre-computed to
yield the series approximation which can then be evaluated for any at virtually
no computational cost.
This method is referred to as the Schur method. The series approximation
is used for approximating the Jacobian matrix of the residual from Equation
2.89. For the residual the evaluation of S()ps p s can be made based on
a series approximation at virtually no additional cost after the series matrices
are formed. This formulation leads to a very efficient method of tracing the
aeroelastic eigenvalues as functions of altitude, which in turn provides stability
boundaries.
33
34
Chapter 3
Transformation Methods
3.1
3.1.1
Introduction
The Transformation Problem
3.1.2
Notation
- xs = (xs1 xsns )T is the set of structural grid points (used for either
x, y or z)
- dxs = (dxs1 dxsns )T are the displacements calculated at the grid
points by the structural solver (used for either dx, dy or dz)
- Fs = (Fs1 Fsns )T are the forces calculated during the transformation
process (used for either Fx , Fy or Fz )
CFD model
3.1.3
Requirements
Conservation of energy
Conservation of energy can be achieved through the use of the principle of virtual
work. The virtual work W is defined by
W = dxTs Fs = dxTa Fa
(3.1)
(3.2)
(3.3)
Using this method the global conservation of energy is satisfied regardless of the
method that is used to fill the transformation matrix [62].
Conservation of total force and moment
Conservation of the global forces can be written as
ns
X
Fsi =
na
X
Fai
(3.4)
i=1
i=1
Smoothness
If the deformed structural grid is smooth then the aerodynamic surface grid once
deformed should also be smooth. If the transformation method artificially generates surface distortions (i.e. ripples, ridges, spikes etc.) it could lead to unexpected results like premature separation or additional shockwaves. There is also
the possibility of the CFD solver failing, especially if the grid folds inside the
aircraft for example at the wing-fuselage junction. If the load distribution on the
aerodynamic surface grid is smooth then the transformed load distribution on the
structural grid should also be smooth. However if there are discontinuities in the
flow solution (i.e. shock-waves) then these should be accurately transferred. Here
smooth is defined in a general sense, but for the CFD solution the surfaces need
to be C 2 continuous to avoid pressure blips. However the local transformation
methods in general are only C 0 continuous, the effect of this lack of sufficient
continuity will be explored in a later chapter.
Complex geometries
When applied to a complex geometry the scheme must not introduce any holes
at component junctions. The scheme should also be able to cope with control
surfaces and stores.
Memory requirements
The methods should use a minimal amount of memory, since high fidelity simulations can involve large CFD and structural grids. For example a full aircraft CFD
surface grid can have in the region of na = 2 105 and a high fidelity structural
model can have in the region of ns = 1.8 104 . Although the use of beam stick
models for aeroelastics is common, which reduces the number of structural nodes
to the region of ns = 200. Using a beam stick model and a global transformation
method can lead to a matrix of around 6 ns na = 2.4 108 non-zero values.
The larger the memory requirements for the transformation method the smaller
the simulation size is that can be run.
CPU time
The CPU time that the transformation method requires should be small in comparison to the CFD calculation, to make the scheme feasible for use in a closely
coupled system.
39
3.1.4
Review
Transformation started to gain interest in the late 1960s as panel methods and
finite element structural models usage became widespread. The initial methods
used were beam and surface splines such as those published by Done [63].
In 1972 infinite plate splines (IPS) were proposed by Harder and Desmarais
[52] and use the solution of a multiply-supported infinite plate for the data transfer. This is still a popular method and is used in commercial finite element
packages such as NASTRAN [64] and ASTROS [65]. This method was an attempt to step away from previous methods, which required a large amount of
user input to choose the best structural and aerodynamic points to get an accurate solution. Also in the seventies a number of methods were proposed in
the field of mathematics to solve the problem of scattered data interpolation,
including Hardys multiquadrics (MQ) [66] and inverse multiquadrics (IMQ) and
Duchons thin plate splines (TPS) [67]. However these new methods were not
applied to the aeroelastic transformation problem until much later.
Franke published a paper in 1982 [68] that reviewed all the methods known
at the time for scattered data interpolation. From this paper the MQ and TPS
were noted to perform very well. Murti and Valliappan published a paper in 1986
on inverse isoparametric mapping (IIM) [69] for the application of remeshing for
crack propagation and this would later be used as a transformation method. At
the end of the eighties Appa published a paper on finite plate splines (FPS)
[70], which was aimed at improving on the infinite plate splines of Harder and
Desmarais. This paper also introduced the concept of virtual surfaces.
The nineties saw an explosion in publications on the transformation problem.
In 1990 Kansa [71] published a series of papers on MQ for use in CFD. In 1992
Pidaparti [72] published the first application of IIM to an aeroelastic problem. A
significant contribution to the field of transformation methods in the nineties was
a paper by Smith et al. [60]. This paper reviewed the transformation methods
IPS, FPS, MQ, TPS, IIM, and NUBS through the use of 260 test cases. The
main conclusions were that TPS was the most accurate, robust and cost-efficient
of all methods tested and IIM performed very well, but requires an extension
to 3D and the elimination of its dependency on structured grids. It is also recommended that multiple methods should be available in any software package.
In 1996 Samareh [73] presented a method for load transfer based upon NURBS.
Cebral and Lohner [74] published a conservative, monotonic, adaptive Gaussian
quadrature load transfer method. In 1998 Chen [47] presented a new boundary
element method (BEM).
In the next decade the new methods that were introduced were radial basis
functions (RBF) and constant volume tetrahedron (CVT). Radial basis functions
40
have been used for transfer for a long time, since IPS, TPS and MQ can be
written in this form. Beckert and Wendland [75] presented radial basis functions
to aeroelastics in 2001 using a number of different basis functions using compact
support. This methodology has been improved upon in a series of papers [2, 50,
76, 77, 78]. Constant volume tetrahedron was proposed by Goura [79] in 2001 and
has subsequently been improved upon [7, 39, 62]. This is a local method where
each aerodynamic node is attached to a structural triangle forming a tetrahedron.
As the structural nodes move the out-of plane component is scaled to keep the
volume of the tetrahedron constant.
3.2
There are a number of different structural models that can be used in aeroelastic
calculations. The simplest are the beam stick models (BSM), these models use
a number of beam elements with lumped masses connected together using stiff
springs. These models are assumed to be chordwise rigid. An example BSM is
shown in Figure 3.2 (a) for the XML12 aircraft test case. The beam elements
are shown in black, the rigid bars are shown in magenta and the lumped masses
are shown as blue triangles. The rigid bars are used for visualisation and to
define additional points to aid the transformation methods. These BSM are still
widely used in industry due to the small number of degrees of freedom (DOF) and
subsequently the ease of matching the model to ground vibration tests (GVT).
These models are also used in the conceptual phase of design because the more
complex models are not available until further along in the design process. These
models are usually used for transport-type configurations with high aspect ratio
wings. Plate models model the lifting surfaces as a plates. An example is shown
in Figure 3.2 (b) for the Standard Dynamics Model (SDM) fighter test case.
These models are usually used for fighter aircraft with low aspect ratio wings.
Wing-box models represent a wing as a wing-box, which only consists of the
spars, ribs and skins above and below the ribs and spars, the rest of the structure
is treated as dead weight and modeled as lumped masses. These models are
also assumed to be chordwise rigid. An example is shown in Figure 3.2 (c) for
the Goland wing test case. These are typically used for transport-type aircraft
whose wings are constructed in this way. Shell models represent the aircraft using
shell elements and usually model the spars, ribs and the full skin. An example
is shown in Figure 3.2 (d) for the ARW-2 wing test case. Full finite element
models using bricks can be used, but are quite rare due to the large number
of degrees of freedom that they contain. These models accurately capture the
structure, but at a high computational cost and for aeroelastic calculations the
41
other types of structural models can provide the required accuracy at a more
reasonable cost. An example is shown in Figure 3.2 (e) for the High Reynolds
Number Aero-Structural Dynamics (HIRENASD) wing test case.
This thesis concentrates on the beam stick and wing-box models. Most transfer methods cannot handle beam stick models directly, since beam elements provide displacements and rotations at the beam nodes. It is common to use rigid
bars attached to the beam elements to provide additional points with displacements. Depending on the transfer method to be used these additional points can
be triangulated or used directly.
3.3
3.3.1
Multiquadric-Biharmonic
The multiquadric (MQ) was first derived by Hardy [66] for the approximation of
geographical surfaces and was applied to CFD by Kansa [71]. Hardy assumed that
any function s(x) could be written as an expansion of n continuously differentiable
basis functions ,
n
X
s(x) =
i (kx xi k)
(3.5)
i=1
1/2
42
43
2
rmin
2
rmax
2
rmin
j1
n1
j = 1, 2, ..., n
(3.6)
2
2
where rmin
and rmax
are input parameters. By carefully choosing the values
2
2
of rmin and rmax the transformation can be made more accurate by changing
the condition number of the Css matrix.
Unit Sub-domain
Kansa [71] also recommended that the data should be mapped onto a unit
sub-domain to reduce the errors. One-dimensional problems are mapped
on to a unit line, two-dimensional problems are mapped on to a unit square
and three-dimensional problems are mapped on to a unit cube. Then if
necessary additional rotation and shear transformations were introduced to
make the distances more distinct.
Domain Decomposition
Kansa [71] also recommended using domain decomposition to split the large
coefficient matrix into smaller quasi-local problems. This increases the accuracy and computational efficiency.
Observations
MQ performs better than inverse multiquadric (IMQ) where the basis function
1/2
is (kxk) = [kxk2 + r 2 ]
[68]. The main advantages of MQ are that it is
infinitely differentiable function and it is accurate in regions where the surfaces
are steep. However in regions where surfaces are relatively flat MQ produces a
surfaces that is not as smooth as the original surface, this can be improved using
a hybrid scheme. Smith et al. [60] found that MQ accuracy is case dependent
and performs poorly interpolating highly oscillatory functions. It was also found
that the parameter r should be kept within certain limits to insure a stable linear
system of equations. Finally it was found that the need for scaling was case
dependent and that MQ was sensitive to grid resolution.
3.3.2
Infinite plate spline (IPS) was proposed by Harder and Desmarais [52] and is used
by programs such as NASTRAN [64] and ASTROS [65]. IPS is a special 2D case
44
of TPS [60]. It is based upon the superposition of the solutions for the partial
differential equation of equilibrium of an infinite plate. The deflections normal to
the plate surface due to n point forces Fi at given locations (xi , yi ) on the plate
can be written as
dz(x, y) = a0 + a1 x + a2 y +
n
X
Fi ri2 ln ri2
(3.7)
i=1
where ri2 = x2i + yi2 . The unknowns, ak and Fi are obtained from the equilibrium
conditions
n
n
n
X
X
X
yi Fi = 0
(3.8)
Fi =
xi Fi =
i=1
i=1
i=1
N
X
Fi rij2 ln rij2
(3.9)
i=1
dzj = a0 + a1 xj + a2 yj +
N
X
Fi rij2 ln(rij2 + )
(3.10)
i=1
and a minimum of three nodes are required [60]. It is a global method and as
such generates a large transformation matrix. Extrapolations to the edges of
the planform from the interior structural grid points do not always appear to be
reliable. Figure 3.3 shows that IPS is unable to recover rigid rotations exactly. A
circle is driven by a rigid bar and as the bar is rotated the circle is skewed. Smith
et al. [60] found IPS to be sensitive to the grid resolution and that the accuracy
was only adequate.
Figure 3.3: A circle of points rigidly rotated by a bar using IPS (from ref. [10])
3.3.3
Thin-Plate Splines
Thin-plate splines (TPS) of Duchon [67] provide a means to characterise an irregular surface by using functions that minimise an energy functional. This method
is very similar to RBF and MQ, however uses a different basis function. IPS is a
2D special case of TPS. The basis function for TPS is
(kxk) = kxk2 ln kxk
(3.11)
where x = (x, y, z)T in this case. This is solved in the same way as RBF with the
linear polynomial (see Section 3.3.8).
Improvements
Sub-domaining
Smith et al. [60] stated that TPS outperformed IPS when the data be
mapped onto a number of sub-domains. Then if necessary additional rotation and shear transformations were introduced to make the distances
more distinct. This mapping on to a number of sub-domains also causes
the transformation to be more localised.
46
Observations
Kamakoti et al [61] observed that since splines are invariant to translation and
rotation, it is a useful tool for moving and flexible surfaces. Smith et al [60]
commented that if TPS is used globally in 2D it has all the same limitations as
IPS, but when used in 3D outperforms IPS.
3.3.4
Inverse Isoparametric Mapping (IIM) was proposed by Murti and Valliappan [69]
and was applied by Pitaparti [72]. The same shape functions Ni are used for both
the aerodynamic grid points and the structural deformation. Each aerodynamic
node is projected to a quadrilateral element. A pair of generalised coordinates
(, ) are defined on the quadrilateral using shape functions so that
xa =
n
X
Ni (, )xesi
(3.12)
Ni (, )ysei
(3.13)
i=1
ya =
n
X
i=1
where n is the number of nodes for the element. The element is assumed to lie in
the x y plane and the superscript e indicates the structural node is a member
of that element. The shape function for a quadrilateral element [80] is given by
N1 (, ) = (1 )(1 )/4
N2 (, ) = (1 + )(1 )/4
N3 (, ) = (1 + )(1 + )/4
N4 (, ) = (1 )(1 + )/4
, [1, 1]
(3.14)
Observations
IIM is the most accurate among all methods for interpolation [61] and can be
derived for different types of elements such as 8-noded quadrilaterals. IIM is only
valid for 2D interpolations and it is commented that no extrapolation is possible
purely using IIM [81]. However Pidaparti [72] showed that it was possible to use
this transformation method with common extrapolation techniques to extrapolate
data to control surfaces. This method is only C 0 continuous in displacement
between quadrilateral elements, but since the interpolation is bilinear it is an
improvement over weighted triangles.
47
3.3.5
The Finite Plate Spline (FPS) proposed by Appa [70] uses a finite plate instead
of the infinite plate of Harder and Desmarais [52]. This was applied to a fighter
aircraft wing by Guruswamy and Byun [57]. This method introduced the concept
of a virtual surface, which is a surface that passes through the aerodynamic
and structural points with the same planform as the aerodynamic surface. The
virtual surface is discretised into finite bending elements. A set of constraints
are established such that when deformed, the virtual surface passes through the
deformed structural points. Appa [70] suggested that C 1 shape functions should
be used, since the C 0 did not give satisfactory results. As such the following
transformation matrix can be derived [60, 70]
1 T
T = a 1 K + Ts s
s
(3.15)
3.3.6
The method proposed by Samareh [82] uses NURBS to transfer the forces and
displacements between the CFD and structural models. The concept behind this
method is to have one main representation of the aircraft in the form of NURBS
surfaces. This single representation can be discretised into a number of other
domains (CFD,CSD, etc.). Each domain is solved and transfers data to and from
the NURBS representation, shown in Figure 3.4. More details can be found in
the papers by Samareh [11, 82].
Observations
This requires the structural model to be a full finite element model that shares
the same surface as the CFD, which will add to the cost. This method requires
access to a CAD package.
48
3.3.7
The Boundary element method (BEM) was proposed by Chen and Jadic [47].
BEM is based upon the elastostatic boundary element method with a BEM solver
being devised to generate a transformation matrix. In this approach the fluid
surface mesh forms the external boundary of an elastic homogeneous body with
the structural nodes as internal points as shown in Figure 3.5.
49
Observations
Sadeghi et al [62] observed that BEM recovers the exact transformation of rigid
body motion and produces a smooth transformation, because all grid points are
connected within continuous elastic bodies, and because of the conditions of minimum strain applied for the inverse BEM. The linear approach allows for the
formulation of a global transformation matrix, which is used for conservative displacement and force transformations. BEM can also be used for CFD volume
grid deformation. Rampurawala [10] states that BEM requires more memory
than IPS and as such is memory intensive.
3.3.8
N
X
i=1
i (kx xi k) + p(x)
(3.16)
where s(x) is the function to be evaluated at x, (kxk) is the basis function, the
index i identifies the centres for the RBF and xi is the location of that centre
(centres are usually the structural points). The linear polynomial p(x) is used to
ensure that translations and rotations are recovered. The coefficients i are found
by requiring the exact recovery of the structural node positions. The polynomial
in the x-direction is given by
px (x) = x0 + xx x + xy y + xz z
(3.17)
When the polynomials are included in the system, the additional requirement is
N
X
i q(x) = 0
(3.18)
i=1
for all polynomials q(x) with degree less than or equal to p(x). Equation 3.16 can
be recast into a transformation matrix, as shown in Appendix C. Then equations
3.2 and 3.3 can be used to calculate the displacements and forces respectively.
Basis Functions
There is an infinite choice of basis functions that can be used. A selection of basis
functions from the literature are given in Table 3.1, where (.)+ = max(0, .) and
in the Euclids Hat basis function r = 2 [76].
50
expkxk
kxk2 ln kxk
1/2
(c2 + kxk2 )
1/2
(c2 + kxk2 )
(1 kxk)2+
(1 kxk)4+ (4kxk + 1)
(1 kxk)6+ (35kxk2 + 18kxk + 3)
2
+ 8kxk + 1)
(1 kxk)8+(32kxk3 + 25kxk
1
4 3
3
2
12 kxk r kxk + 3 r
Improvements
Norm Biasing
The definition of the norm has a significant impact upon the interpolation,
which is the input for the basis function. Typically, the Euclidean norm is
used, shown below
p
(3.19)
kx xi k = (x xi )2 + (y yi )2 + (z zi )2
Support Radius
The support radius is defined as (kxk/) where is the support radius
and is any basis function. This allows the control of the area of influence
of each centre. A very large support radius smooths the deformation over
a large area, whereas a small support radius makes it much more localised.
The larger the support radius the higher the condition number of the Css
matrix, which leads to higher accuracy when there the polynominals are not
used. However when the polynomials are used the matrix sparsity pattern
is more complex and the accuracy is increased as the support radius is
decreased. From numerical tests the wing used effects the whether the
condition number increases or decreases as the support radius decreases,
even though the error decreases for both wings tested. The reason behind
this need further work.
51
Data Storage
Rendall and Allen [2] suggest using a threshold value above which an element of the matrix is stored. This reduces the storage cost associated with
this method.
Pointwise Partition of Unity
Rendall and Allen [78] proposed the pointwise partition of unity, which is
an improvement of the partition of unity of Wendland [83] and Ahrem et al
[76]. The global solution is generated from a weighted combination of local
solutions. The local solutions are for each aerodynamic point deformed by
a small number of the closest structural points. This results in a much
faster solution with lower storage requirements and a more physical force
distribution.
Data Reduction
Rendall and Allen [50] developed a method that uses a greedy algorithm to
minimise the number of points used as centres whilst minimising the error.
This works by starting with a small number of centres and then applying
a unit translation in the x, y and z directions. This gives a measure of the
error introduced by using a subset of the centres. If the error is above a
threashold amount then an additional centre is added at the loction of the
maximum error, this is repeated until the error is below the target error.
Observations
The great strength of RBF is the fact that the same code can be used to perform
the coupling and/or the mesh movement on structured and unstructured meshes.
RBF is an interpolation scheme for use on scattered data sets and as such is not
based upon the structural dynamics or connectivity. Unfortunately this method
is computationally and memory intensive, since the matrix Css is n2s in size and if
mesh motion is also required the memory required is ns nv . Another disadvantage is the large number of options that can be changed, such as basis functions,
norm biasing and support radius. There is a problem reported by Rendall and
Allen [2] and Ahrem et al [76] where camber is induced at the trailing edge of
a wing especially near the tip, this is probably due to extrapolation. The improvements given above address most of the issues at the cost of a more complex
implementation.
3.3.9
three structural grid points, shown in Figure 3.6. The position of the aerody-
(3.21)
where a = xs,j xs,i , b = xs,k xs,i and d = a b. The constants are calculated
using
(b.b)(a.c) (a.b)(b.c)
=
(3.22)
(a.a)(b.b) (a.b)(a.b)
(a.a)(b.c) (a.b)(a.c)
(3.23)
=
(a.a)(b.b) (a.b)(a.b)
(c.d)
=
(3.24)
(d.d)
The volume of the tetrahedron is held constant during the calculation by recalculating . The method can be linearised in the structural displacements [79],
giving
dxal = Adxs,i + Bdxs,j + Cdxs,k
(3.25)
where
A= I BC
B = I UV (b)
C = I UV (a)
U =I
2
D(d)S(d)
d2
53
(3.26)
d is the magnitude of d.
0 z3 z2
0 z1
V (z) = z3
z2 z1
0
z1 0
D(z) =
0 z2
0 0
z1 z2
S(z) = z1 z2
z1 z2
0
0
z3
z3
z3
z3
(3.27)
(3.28)
(3.29)
Equation 3.25 can be written as a transformation matrix, which can be recalculated each time step to reduce error. Then equations 3.2 and 3.3 can be used to
calculate the displacements and forces respectively.
Improvements
Selection of the Structural Elements
For this method the selection of the correct structural element to map the
aerodynamic node to is critical. This mapping is done as a preprocessing
step and is provided as an input to the solver. The original method of
selecting the structural element was to minimise the distance between the
projected fluid point and the structural element centroid. This could result
in the fluid point being above a structural triangle that it was not mapped
to, since that neighbouring centroid was closer [7].
To minimise the amount of extrapolation required the following modification was made. Each structural element was split into three subtriangles.
The area of the jth structural element ABC (ABC ) along with the subtriangles APB (AP B ), BPC (BP C ), APC (AP C ) are calculated where P
is the projected aerodynamic node and the following function is minimised
Sj = |ABC AP B BP C AP C |
(3.30)
54
Observations
CVT is fully 3D and accounts for the out-of-plane component. It is easy to
implement and has low memory and CPU requirements. However smoothness
is not guaranteed because the structural triangles used are only C 0 continuous.
CVT is sensitive to the resolution of the structural grid (the structural grid is
required to have a similar resolution to the aerodynamic grid [62]). It is also
highly dependent on the quality of the mapping routine used.
3.4
A different approach is taken for beam models. Since one of the assumptions is
that the chordwise section is rigid perpendicular to the beam, a direction is defined
along which the wing section is assumed to be rigid. These directions would be
defined by the ribs in the wing structure. The beam is defined by the points xi
and along with the rigid direction, the corresponding leading xLEi and trailing
xT Ei edge points can be calculated. Then, the motion of this section is defined
by the translation dxi and rotation di of the beam point. The displacement of
any point y on the section can then be derived as
dy = dxi + R (y xi )
where the rotation matrix R is given by
2
(3.31)
55
56
Chapter 4
Evaluation of Transformation
Methods
4.1
Introduction
There are some questions of interest that this chapter aims to answer. First,
what influence does C0 continuity (CVT and IIM) have on the section shapes?
Secondly what is the contribution of the out-of-plane component on the section
shapes? Finally, what is the influence of any shape distortion on the flutter
predictions? Two test cases are used to shed light on these questions and are
described below.
4.2
4.2.1
Test Cases
Goland Wing
The Goland wing, shown Figure 4.1, is a rectangular wing that has a chord of 6
feet and a span of 20 feet. The aerofoil section is a 4% thick parabolic section.
The CFD grid used was a coarse version with 35 thousand points and is blockstructured using an O-O topology. This allows points to be focused in the tip
region, which is the most critical region for the aerodynamic contribution to
the aeroelastic response. This was solved using the Euler equations. The Euler
equations can be used because the Goland wing is a slender wing and therefore has
a thin boundary layer and along with a high Reynolds number the viscous effects
can be safely neglected. Throughout the flow regime being investigated the flow is
expected to be attached also allowing the use of the Euler equations. At the time
of the calculations the Schur solver was only capable of using the Euler equations
to calculate the flutter boundaries, which was still a significant improvement over
linear methods due to the ability to capture shockwaves. For LCO the paper by
Beran et al. [84] found that the effect of viscosity has a minimal effect on the onset
57
4.2.2
MDO Wing
The second test case is the multi-disciplinary optimisation (MDO) wing [85],
shown in Figure 4.3, this is a commercial transport wing with a span of 36m.
The wing was optimised to fly at a certain altitude and Mach number and has
a thick supercritical section. The CFD grid used was a coarse grid with 81
thousand points and was solved using the Euler equations. The use of the Euler
58
(a) Geometry
(a) Original
model
NASTRAN
60
pendicular to the beam and the position as well as the original section is shown
in Figure 4.3.
(a) Geometry
4.3
The transfer methods were implemented and are tested in this section in terms of
the mapped mode shapes. To test the transformation methods shape comparisons
are shown for several cases in increasing complexity. The test section in each case
is translated and rotated back to the original orientation to allow evaluation of
the shape distortion introduced by the transformation.
61
4.3.1
Case 1 - In-plane
For this case the fluid grid is defined on a plane which is on the same surface
as the structural grid and with the same planform as the Goland wing. The
fluid grid is chosen to be finer than the structural grid and this case tests the
in-plane treatment of the transformation methods only. Both the original plate
and extended plate are compared in this case. Figure 4.5 shows the slice at 98%
of the span for the third mode shape, which is the second torsion mode.
The CVT section exhibits a saw-tooth shape for both structural models, but
passes through all of the NASTRAN points. This can be attributed to the lack
of derivative continuity between structural elements. For the original plate, once
the trailing edge of the structure has been reached, CVT extrapolates linearly.
The IIM section passes through all of the NASTRAN points and provides linear extrapolation parallel to the element edges. The fact that IIM is only C0
continuous does not show in the sections. IPS has trouble in the region of extrapolation for the original plate model where it flicks up to an unrealistically large
deflection. Also throughout both structural models IPS oscillates between the
structural model points. There is a well known problem with IPS[10, 60, 62] that
it cannot recreate rigid rotations. IPS performs much better when there is no extrapolation, but still has minor oscillations. The RBF support radius was chosen
to be 1.0 in order for the first mode shape to pass through the NASTRAN points,
and no norm-biasing was applied. The RBF result shows the same problem with
extrapolation as IPS, however it is much more severe. The observed additional
camber has been seen in other papers [76] (incorrectly attributed to fuselage
interference) and Rendall et al. [2] (where it was reduced by using norm-biasing).
This case highlights the differences between the local (CVT and IIM) and
global (IPS and RBF) methods. The local methods exhibit a problem with slope
discontinuity that the global methods do not suffer from, although IIM did not
suffer from this visibly. IIM uses bilinear elements, so although they are only
C0 continuous this higher order interpolation appears to avoid the slope discontinuities experienced by CVT. The global methods however fail to extrapolate
realistically and have oscillations that are reduced when there are no extrapolation regions. All the transformation methods performed better on the extended
plate than the original plate due to the lack of extrapolation.
62
Figure 4.5: Case 1: Section at 98% span for the third mode shape, rigidly translated and rotated back to the original orientation.
63
4.3.2
For this case the fluid grid is now defined on the correct wing profile, but the
fluid points are projected onto the structural plane. The transformation methods
are used to define the displacements at the projected points and then these displacements are applied without modification to the wing points. This tests the
discrepancy introduced by failing to calculate the out-of-plane component. Again
only the original and extended plate models are used.
Figure 4.6 shows a slice through the two plate models at 98% of the span for
the third mode shape. All of the methods show the same behaviour as seen for
case 1, but now on the wing section. CVT shows the C0 continuity effect which
leaves the aerofoil nonsmooth. Interestingly IIM shows no sign of this problem
and produces an excellent section shape. RBF and IPS both display the same
behaviour as case 1 with the trailing edge failing to be extrapolated correctly, as
well as section oscillations.
Figure 4.6: Case 2: Section at 98% span for the third mode shape, rigidly translated and rotated back to the original orientation.
64
4.3.3
Figure 4.7: Case 3: Section at 98% span for the third mode shape, rigidly translated and rotated back to the original orientation.
65
Figure 4.8: Case 3: Sections at 98% of the span through the mode 3 mode shape
for IIM comparison for cases 2 and 3, rigidly translated and rotated back to the
original orientation.
(a) CVT
(b) IIM
(c) IPS
(d) RBF
Figure 4.9: Case 3: Sections at 98% of the span for the refined extended plate
model, rigidly translated and rotated back to the original orientation.
66
4.3.4
Finally the beam stick model is used and all transfer methods have the out-ofplane contributions from the CVT method applied to them. Figure 4.10 shows
a slice at 98% for the third mode shape. All the methods are effectively using
a coarsened extended plate model, since the BSM has one less row of structural
points in the chord-wise direction that corresponds to the trailing edge of the
wing box in the other structural models. The sections are significantly worse as
the trailing edge is approached for all methods except IIM. CVT has a higher
amplitude saw-tooth in the trailing edge region, but the saw-tooth in the leadingedge region is unchanged from case 3. This underlines the local character of
this method. IIM appears to only be affected by the poor CVT out-of-plane
contributions. IPS has the largest deviation from the section shape in the coarse
region of the grid at the trailing edge, but also see an increase in the amplitude
over case 3 in the leading edge region. RBF has the same trend as IPS, but to a
worse degree.
Figure 4.10: Case 4: Sections at 98% span through the mode 3 mode shape
rotated and translated back to the original orientation.
4.3.5
Next, results are shown for the two structural models of the MDO wing. The
transformation methods are compared in terms of the vertical displacement calculated in Figure 4.11. The rigid rib results are considered exact in this comparison
since all methods only have access to the beam structural model, and the rigid rib
method exploits this information exactly by design. The displacements obtained
by the different methods for the perpendicular ribs are similar except for those
from IPS, which suffers due to the coarse structural grid used. RBF is not shown
67
because the deflections are very different compared to the other methods. There
is much more variety in the results for the parallel ribs, in this case IIM and the
rigid ribs approach match well. CVT has a slight saw-tooth pattern and IPS deviates slightly due to the coarse mesh. RBF prediction is rather poor for the MDO
wing. The reason that the discrepancies are small compared to the Goland wing
is due to the scaling. A significant difference is seen between the displacements
obtained from the different rib orientations, with larger displacements seen for
this mode from the parallel ribs shown in Figure 4.11(c). For other modes this is
the opposite way around. The reason for this difference is that the values for the
rotation of the section are obtained from different points on the structural beam.
Figure 4.11: Case 5: Vertical displacement of sections at 98% span for the mode
5 mode shape.
68
4.3.6
Summary
From these test cases it has been shown that allowing the transfer methods to
extrapolate is inadvisable. The in-plane component dominates the final displacements calculated, other factors have a larger effect on the final displacement than
accounting for the out-of-plane affects. Of the methods tested only IIM was unaffected by changing grid density. All the other methods benefited from the structural grid being refined. The local method CVT suffered from a saw-tooth pattern
of discontinuities in the slope, which could be reduced by refining the structural
grid. The other local method IIM did not suffer from this discontinuity in slope,
probably due to the higher order interpolation inside the bilinear elements used.
The RBF result could be improved using the techniques and advanced algorithms
covered in the previous chapter. Also Ahrem et al. [77] published a paper on
using RBF with BSM, however the proposed algorithm was found to be six times
more expensive than the standard algorithm. The proposed rigid ribs method
was shown to perform well producing smooth section shapes similar to IIM. As
expected the orientation of the ribs has an effect on the section shapes, which as
the orientation becomes further from the ideal perpendicular section the section
shapes get worse.
4.4
Flutter Evaluation
4.4.1
Goland Wing
The Goland wing mode tracking at a Mach number of 0.80 and zero degrees
incidence for all modes using CVT and the original plate model is shown in Figure
4.12. It can be seen that modes one and two interact, with mode one becoming
undamped at 13,216 feet as the real part of the eigenvalue becomes positive. The
evaluation of the influence of the transformation methods is presented below for
the real part of mode one.
The effect of using different structural models is shown for the Goland wing
in Figure 4.13 and Table 4.1. Consistent with the section shape results presented
above, IIM shows the least spread of results as the structural model is changed.
CVT and IPS show comparable spread in their results, particularly where the
interaction is strong as the mode becomes undamped. RBF shows a large spread
69
and has no real correlation between the different structural model results. The
original plate model did not flutter for mode one, when all the other transfer
methods and structural model combinations did. The extended and BSM models
did flutter in mode one, but their flutter points are almost 9,000 feet different.
Next the results from different methods are compared for each structural model
in turn and this is shown in Figure 4.14. The predictions from the local IIM and
CVT methods are similar for all the structural models. There is considerable
spread in the results for the global methods (IPS and RBF). For the extended
plate where the section shapes were the most reasonable, all the flutter points are
clustered including RBF. The original plate shows the CVT and IIM methods
closely clustered with the IPS flutter point close, RBF however did not flutter
for this mode. The BSM model is effectively a coarsened extended plate and
since the global methods suffered the most in the section shapes their results are
spread out compared to the local methods. If the IIM flutter results are assumed
to be correct due to the excellent performance in the section shapes, then the
other methods results can be compared with them. It would appear that the
discontinuity in slope has a smaller effect on the flutter results than induced
oscillations in the mode shapes.
Figure 4.12: Goland wing mode tracking for all modes using CVT
70
(a) CVT
(b) IIM
(c) IPS
(d) RBF
Figure 4.13: Goland wing mode tracking for mode 1 for the different structural
models (real part)
71
(c) BSM
Figure 4.14: Goland wing mode tracking for mode 1 for the different structural
models (real part)
72
4.4.2
MDO Wing
The mode tracking for the MDO wing at a Mach number of 0.85 and an incidence
of one degree based on the beam model with perpendicular ribs using CVT, is
shown in Figure 4.15. Modes one, two and four participate in the instability, with
mode one going undamped first at -2358m.
The effect of the rib orientation for the MDO wing is shown in Figure 4.16.
It can be seen that there are significant differences in damping between the two
rib orientations, even if the crossing at a similar altitude. The comparison of
the predictions from the different transformation methods is shown in Figure
4.17 and Table 4.2. The perpendicular ribs has a tight clustering of the flutter
points for all the methods except RBF which fails to flutter for this mode. This
is not a suprise since its section shape was so different to the other methods it
was omitted. The parallel ribs shows more variation between the methods with
the local methods clustered tightly. The rigid ribs predicts flutter at a slightly
higher altitude compared to the local methods, but the global methods predict
the flutter point at even higher altitudes. It is not obvious from the sections why
the rigid ribs and IIM are not more similar in this case, although it could be due
to the other modes that interacts with mode one having a different section shapes
leading to a different flutter points.
Figure 4.15: MDO wing mode tracking for all modes using CVT
73
Figure 4.16: MDO wing mode tracking for mode 1 for the different structural
models (real part)
Figure 4.17: MDO wing mode tracking for mode 1 for the different structural
models (real part)
Table 4.2: Flutter altitude for MDO wing (all values in meters)
Perpendicular Parallel
CVT
-2358
-2612
IIM
-2175
-2481
IPS
-1848
-915
RBF
N/A
4198
Rigid ribs
-2125
-1743
Spread
510
6810
74
4.5
Summary
75
76
Chapter 5
Aerofoil Buffet
5.1
Introduction
5.2
Test Case
perimental data points is shown in Figure 5.1. The shock oscillation region was
obtained by fixing the Mach number and varying the angle of attack. At each
angle a power spectra for the normal force was computed and the presence of
shock oscillation was deduced from whether a peak at 70-80Hz was present. The
buffet boundary was calculated from a plot of rms values of the normal-force
fluctuations, CN vs. CL and was determined as the point that has a slope of
dCN /dCL = 0.1 [87]. This was an arbitrarily chosen value that matched the
buffet boundary calculated by the trailing-edge pressure divergence criterion.
Figure 5.1: BGK No. 1 buffet boundary with the experiment locations (circles)
The experiments covered the Mach number range of 0.501 to 0.805, angle
of attack range of -0.36 to 11.74 and Reynolds numbers between 15 million
and 21 million. The wind tunnel had a floor and ceiling porosity of 19.3% or
20.5% depending on the reference. The aerofoil used in the experiments had a
chord of 10in. and a span of 15in., and had 50 pressure orifices on the upper
surface and 20 on the lower surface. It also had 16 miniature fast transducers
on the upper surface for unsteady measurements. The locations of these pressure
orifices and fast transducers are given in ref. [90] and are shown in Figure 5.2.
For experiments at Mach number 0.688 skin friction was also measured. It was
estimated that transition occurred at 10% of the chord. The results published
focused on three test cases; case 1: Mach number 0.688 with varying angle of
attack; case 2: Mach number 0.710 with varying angle of attack; case 3: angle of
attack 6 with varying Mach number.
78
The results were very similar to those of case 1. The flow at angle of attack
-0.316 was sub-critical with fully attached flow and at angle of attack 1.396
there was a weak shock formed with no separation. For angle of attack 3.017
there was a stronger shock with small oscillations which had decayed by the next
transducer 5% of the chord downstream. A separation bubble was formed behind
the strong shock at angle of attack 4.905 and then at angle of attack 6.970 the
flow has become fully separated.
For case 3 the experiments used an angle of attack of approximately 6 whilst
varying the Mach number from 0.597 to 0.772. It was found that at M=0.597
the flow was attached due to the absence of pressure fluctuations. As the Mach
number was increased to 0.688 the flow gained a stronger shock that leads to
a separation bubble. Above Mach 0.722 the flow is fully separated from the
shock to the trailing edge and there is no evidence of the separation bubble. The
maximum CN was found to be at Mach 0.733. The frequency of oscillation was
found to be a linear function of Mach number.
For the BGK No. 1 aerofoil experiments, boundary-layer suction was used and
the ceiling and floor had 19.3% or 20.5% porosity depending on the reference. Also
the span of the aerofoil was only 1.5 chords potentially effecting the accuracy of
the CFD predictions due to the uncertainty caused by the wind tunnel walls. It
is likely that the two-dimensional character of these experiments is not beyond
question.
There has been a CFD study using the BGK No. 1 aerofoil by Xiao et al.
[3]. This CFD study focused on case two, the flow conditions used are shown in
Table 5.1. The grid used had a C-topology and was 640 64 with 512 points on
the aerofoil and 128 in the wake. The initial wall spacing used was 1 106 of
the chord. A grid refinement study was carried out and it was found that this
grid was adequate for their study. The grid refinement study also found that the
dominant frequency was insensitive to the grid refinement.
Table 5.1: Xiao et al. [3]: Flow conditions
Reynolds Number
20 106
Mach Number
0.710
Angles of Attack
1.396 , 6.970 , 9.000
Non-dimensional Time Step
0.05
Non-dimensional Duration
150-200
The first case was at angle of attack 1.396 , where a weak shock is expected
with no separation. The CFD predicted the shock too far downstream, however
when the angle of attack was changed to account for wind tunnel corrections
the shock was predicted closer to the experiment. The second case was at an
80
angle of attack of 6.970 , the experimental result was a strong oscillating shock
causing full separation. The mean lift coefficient from the CFD was 1.030 and the
experimental mean lift coefficient was 1.016. For the dominant reduced frequency
the CFD prediction was 36% lower than the experimental result. For this case
no wind tunnel correction was used. The CFD failed to capture the magnitude
of the unsteady pressure, however there was better agreement for the phase. The
final case at angle of attack 9.000 showed a steady shock at the leading edge
that caused separation to the trailing edge.
An attempt has been made to understand the mechanism behind self-sustained
shock oscillations and shock-induced buffet [12, 92, 93]. This was also studied
computationally by Xiao et al. [3]. The mechanism proposed by Lee [12] is shown
in Figure 5.3. The mechanism is for self-sustained shock oscillations with fully
Figure 5.3: Model of self-sustained shock oscillation for the BGK No. 1 aerofoil
(from ref. [12])
separated flow after the shock wave. Due to the shock motion pressure waves
are formed that propagate downstream in the separated flow region at velocity
ud . When these waves reach the trailing edge they generate new waves either
from the wake fluctuation or from the trailing edge boundary layer, which travel
upstream outside the region of separated flow at velocity uu . These new waves
interact with the shock and impart energy to it to maintain the oscillation. This
loop then repeats, the period of the shock oscillation Tp should agree with the
time it takes for the waves to propagate to the trailing edge and the time it takes
the new wave to propagate upstream to the shock. This can be given by the
following equation
Z c
Z xs
1
1
dx
dx
(5.1)
Tp =
uu
c
xs u d
where ud is the speed of downstream pressure wave propagation, uu is the speed
81
5.2.1
Figure 5.4: 18% thick circular-arc aerofoil shock oscillation domains (from ref.
[13])
The mechanism for the flow over this aerofoil at zero degrees was described
by McDevitt et al. [13]. It was explained as the shock induced boundary-layer
separation on the upper surface causes a thickening of the boundary-layer on the
upper surface. This effectively creates a negative camber which has the effect of
slowing the flow on the upper surface. This tends to suppress the shock-induced
phenomenon, but at the same time induce higher velocities over the lower surface,
causing shock induced boundary-layer separation there and the flow field flips.
This mechanism was verified by McDevitt [95] by using a splitter plate at the
leading and trailing edges.
An alternate mechanism was proposed by Chen et al. [99] similar to that
of Lee for the BGK No. 1 aerofoil. The model proposed by Lee assumed the
shock oscillates above the aerofoil, the model proposed by Chen et al. assumes
the shock propagates upstream and leaves the aerofoil. The flow mechanism is
described as a series of compression waves develop in the region near the trailing
edge and move upstream. As they move upstream they coalesce to form a strong
shockwave, this shockwave continues to move upstream whilst strengthening and
induces boundary-layer separation. Once it has reached the mid-chord the shock
weakens to a weak shock and then compression waves that propagate upstream
leaving the aerofoil.
At ONERA in France wind tunnel tests were done on the OAT15A aerofoil
which is a supercritical profile. The experiments were done by Jacquin et al. [100]
and CFD calculations have also been performed by Deck [101].
The experiments covered the Mach number range of 0.7 to 0.75, angle of
83
Starting at the most downstream location the separation bubble pushes the shock
upstream strengthening it, with the wedge and dynamic effects dominating over
the curvature effect. When the shock reaches the upstream position the aerofoil
curvature effect now dominates over the wedge and dynamic effects. As the shock
weakens the boundary-layer reattaches followed by the rest of the flow field after
a lag. As the flow field reattaches the shock travels downstream strengthening
and when it reaches the downstream position it causes separation initiating a new
cycle.
85
5.3
5.3.1
Three grids were generated and their sizes are given in Table 5.2. The blocking is
of the C-H type of a similar setup to Xiao et al.[3] with 256 points on the aerofoil
upper and lower surfaces, 128 in the wake and 64 in the normal direction. The
medium grid is shown in Figure 5.5. The coarse and fine grids are a coarsening or
a refinement of the medium grid by a factor of two in each direction. To calculate
Table 5.2: Grid Sizes
Grid
Number of nodes
coarse
24,576
medium
98,304
fine
221,184
the initial cell spacing the following equations where used. They are based on
the turbulent flow over a flat plate to give a single point in the laminar sub-layer
of the boundary-layer. The inputs for the equations are the Reynolds number,
characteristic length and desired y + . First the temperature was calculated using
T =
followed by the pressure
p=
288.15
1 + 1
2
101325
1
1 + 1
2
(5.2)
(5.3)
(5.4)
(5.7)
2
cf U
2
86
(5.8)
(a) Blocking
87
u =
y=
(5.9)
y+
u
(5.10)
0.8
y+
0.6
0.4
0.2
0.2
0.4
0.6
0.8
x/c
Figure 5.6: y + distribution around the BGK No. 1 aerofoil at = 6.97, M=0.71
and Re=20 106
5.3.2
Probe Locations
Table 5.3 gives the location of the pressure taps and the probe locations used in
the analysis of the calculations.
5.3.3
f c
U
88
(5.11)
89
5.3.4
Convergence Studies
5.3.5
Pseudo steps are used to iterate to the real time solution. The number of these
steps can be changed to allow different levels of convergence. The unsteady
parameters used were 6,000 time steps using a time step of 0.05 non-dimensional
time with an unsteady tolerance of 0.001. The number of pseudo steps allowed for
each unsteady time step was varied in this study, the values used were 100, 500,
1,000 and 1,500. Table 5.4 gives the mean lift coefficient, the reduced frequency of
Table 5.4: Results for the pseudo steps study
Pseudo
Mean Cl Reduced
Percentage of
steps
Frequency Unconverged steps
100
0.967
0.161
100%
500
0.994
0.230
28%
1,000
0.994
0.230
12%
1,500
0.993
0.230
7%
Experiment
1.016
0.25
Xiao et al. [3]
1.03
0.16
oscillation and the percentage of time steps that failed to converge to the desired
unsteady tolerance. By increasing the number of pseudo steps the percentage
of unconverged time steps reduced and the mean lift coefficient and reduced
frequency converged, as shown in Figure 5.7. Figure 5.7(a) shows the lift history
and that the 100 pseudo steps caused the lift oscillations to decay to a steady
solution, whereas increasing the number of pseudo steps gives almost identical
unsteady oscillating lift histories. The time-averaged pressure distribution shown
in Figure 5.7(c) is a reasonable match to the experimental results for all numbers
of pseudo steps except for 100 pseudo steps. The skin-friction shown in Figure
5.7(d) for all numbers of pseudo steps the flow separates after the shock and on
average fails to reattach. Figure 5.7(e) shows the unsteady pressure distribution
90
91
at the experimental pressure taps, this again shows that 100 pseudo steps results
in a steady solution. The CFD results show a higher unsteadiness than the
experiment and point towards the shock oscillating further back on the aerofoil.
However this is in-line with the other CFD study by Xiao et al. [3]. For all the
calculations presented from now on 1,000 pseudo steps will be used, since this is a
good compromise between the potential benefit of a greater number of converged
time steps and the cost of the calculation. Figure 5.7(f) shows this compromise
graphically.
5.3.6
The non-dimensional time steps used were 0.05 (6,000 steps), 0.025 (12,000 steps),
0.0125 (24,000 steps) and 0.00625 (48,000 steps). All the calculations were ran
to 300 non-dimensional time. The unsteady parameters used were an unsteady
tolerance of 0.001 and 1000 pseudo steps.
A time step study was not performed by Xiao et al. [3], but previous studies
on the 18% thick circular-arc aerofoil by Rumsey et al. [96] showed that the
reducing the time step converges the reduced frequency. This trend can also be
seen in the results presented in Table 5.5, with both the reduced frequency and
the mean coefficient of lift converging.
The results of the present study are given Figure 5.8. The time-averaged
pressure distribution shows all the tested time-steps are in good agreement with
the experimental data. The skin-friction shows that all the time steps are on
average fully separated after the shock with no reattachment. The unsteady
pressure distribution shows that as the time step increases the unsteadiness also
increases with reducing time step. Since the cost of the calculation approximately
doubles when the time step is halved a time step of 0.0125 was chosen for the
remaining calculations as a compromise between cost and accuracy.
92
93
5.3.7
For the grid refinement study the three grids given in Table 5.2 where used. From
the previous studies the unsteady parameters used were 12,000 time steps using
a time step of 0.0125 non-dimensional time leading to a total run length of 150
non-dimensional time with an unsteady tolerance of 0.001 and 1000 pseudo steps.
Xiao et al.[3] found the dominant frequency to be independent of the grid
used. Other CFD studies on the 18% thick circular-arc aerofoil found that the
whilst the reduced frequency was relatively insensitive to grid density it still varied
by 2-5%. Changing the grid in the study by Xiao et al. changed the pressure
distribution and the lift coefficient. Table 5.6 shows the results from the grid
study. The mean lift is reasonably well converged, but the reduced frequency is
still converging towards the experimental result.
Table 5.6: Results for grid refinement study
Grid
Mean Cl Reduced frequency
Coarse
1.042
Medium
1.002
0.238
Fine
0.997
0.245
Experiment
1.016
0.250
Xiao et al. [3]
1.03
0.16
More results are given in Figure 5.9, it can be seen that the coarse grid was
unable to capture the unsteady flow. The medium and fine grids both produced
the expected unsteady flow with the amplitude of the lift oscillations being very
similar. The difference in the mean pressure and skin-friction distributions are
minimal between the medium and fine grids, whereas the coarse grid shows a
strong steady shock. The unsteady pressure distributions are similar for the
medium and fine grids and have a higher intensity than the experimental results.
The coarse grid has very little unsteadiness as expected from the lift history. All
the remaining calculations use the medium grid.
5.3.8
To calculate the buffet boundary using unsteady CFD calculations takes significant computational time with at least three calculations required for each point
along the buffet boundary. In order to increase the speed of the buffet boundary
prediction using steady computations would be advantageous. If the lift is plotted against iteration number useful patterns emerge. Even though these results
are not time accurate and represent no real time history, if the lift coefficient
stabilises after an initial oscillation the flow is steady. If these oscillations per94
95
sist then the flow is possibly unsteady. This was shown by Singh [105] using a
NACA0012 aerofoil.
A large number of steady state calculations were ran using the medium grid
to estimate the buffet boundary. The cases were ran to 40,000 implicit iterations
using a CFL of 2.0 and with the lift coefficient output at each iteration. The
lift coefficient history and the residual of the steady calculation can be used
to estimate whether the the flow is steady or unsteady and therefore estimate
the buffet boundary. The test region is shown graphically in Figure 5.10 for
both using the lift coefficient and residual as the estimator. In the figure the
regions of experimental shock oscillation and the experimental buffet boundary
are highlighted. Using the lift coefficient as the estimator the region of shock
oscillation is well captured, but the experimental buffet boundary is not. This is
not surprising as the most likely feature to cause the lift coefficient to oscillate in
this flow is an oscillating shock wave. Using the residual as the estimator results
in significantly more unsteady results. These cover the entire buffet region, but
also extend outside the experimentally estimated region.
4.905 oscillated. The remaining two angles of attack oscillated, but gradually
damped towards further convergence. The turbulence model residual however
gave a slightly different story. The lowest two angles of attack gave a steady
convergence, 9.0 oscillated before damping to a steady convergence. The angle
of attack of 4.905 failed to converge and its turbulence model residual oscillated.
At angle of attack of 6.97 the turbulence model residual oscillated as it converged,
but never settled down to a steady convergence. To give an indiction that the
reason that the calculations are do not converge is not a numerical issue, this
angle of attack sweep was calculated using CFL number of 5.0 and is also shown
in Figure 5.11. The results at the higher CFL number confirm that the angles of
attack of 1.396 and 3.017 are both steady. The results also confirm that angle
of attack of 4.905 fails to converge and therefore is likely to be unsteady. The
higher angles of attack of 6.97 and 9.0 are less clear, as both converge to the
desired level. The 6.97 angle of attack the turbulence model residual oscillated
as it converged and both angles of attack had oscillations as the Navier-Stokes
residual converged. Since their convergence was quite different from the steady
cases these would also be worth looking into. This method has some promise for
a fast initial estimate of whether a calculation is likely to be unsteady, but as the
results look different for each CFL number care must be taken to only compare
iteration histories of calculations with the same CFL number. This leaves some
calculations with convergence histories that are different from the rapid smooth
convergence of steady calculations giving an indication of which calculations are
worth running unsteady.
Figure 5.12 shows the lift curve for the steady calculations. It can be observed
that the CFD lift coefficient values are higher for all angles of attack tested below
6.97 . Stall is at the same point as the experiment at about 5 .
5.3.9
An angle of attack sweep for the Mach number 0.71 ran as unsteady calculations
will be presented here and will cover the region before shock-induced oscillation,
shock-induced oscillation and the region after the shock-induced oscillation. For
the unsteady runs only data from 50 to 150 non-dimensional time will be used
for statistical analysis to remove transient effects at the start of the calculations.
Following the convergence studies each angle of attack will be ran for 150 nondimensional time using a time step of 0.0125 and 1000 pseudo steps on the medium
grid.
97
2.5
-1
1.5
-2
-3
0.5
-4
-5
-0.5
10000
20000
30000
log(Residual)
CL
-6
40000
Desired
Convergence
10000
20000
30000
40000
Iteration Number
Iteration Number
Figure 5.11: Comparison of Mach 0.71 angle of attack sweep for steady calculations
98
Figure 5.12: Lift curve for the Mach 0.71 angle of attack sweep
1.396 Angle of Attack
This case is the lowest angle of attack ran and converged to a steady solution.
This agrees with the estimate in Figure 5.10 using both the lift coefficient and
residual, this also agrees with the experimental estimate. The CFD study by
Xiao et al. used an angle of attack of 1.3 to account for wind tunnel correction.
The comparison of the pressure distributions is shown in Figure 5.13 and the lift
coefficient is shown in Table 5.7. The pressure distributions for both the original
and the corrected angles of attack fail to match the experimental results. The
lower angle of attack does shift the shock towards the experimental location, but
does not predict its location correctly. This is in-line with the results by Xiao
et al. who used the lagged k- turbulence model and also failed to predict the
experimental shock location. There is no separation in this case from the skin
friction distribution as expected.
99
100
102
(a) Magnitude
Skin-
Figure 5.17: Unsteady shock movement during one period for angle of attack
4.905
104
(a)
(b) = 0.0
(c) = 0.1
(d) = 0.2
(e) = 0.3
(f) = 0.4
(g) = 0.5
(h) = 0.6
(i) = 0.7
(j) = 0.8
(k) = 0.9
(l) = 1.0
Figure 5.18: Mach number contours at different instants in time for one period
of oscillation for angle of attack 4.905, M = 0.095
105
(a)
(b) = 0.0
(c) = 0.1
(d) = 0.2
(e) = 0.3
(f) = 0.4
(g) = 0.5
(h) = 0.6
(i) = 0.7
(j) = 0.8
(k) = 0.9
(l) = 1.0
Figure 5.19: Streamlines at different instants in time for one period of oscillation
for angle of attack 4.905
106
magnitude, but are not as conclusive as the results presented by Xiao et al. [3].
Table 5.8: Reduced frequency results
Lift history Path 1 Path 2 Path 3
0.215
0.273
0.250
0.239
Figure 5.20: Cross-covariance and wave speed for angle of attack 4.905
108
109
(a) Magnitude
5.4
Summary
Skin-
Figure 5.23: Unsteady shock movement during one period for angle of attack
6.97
111
(a)
(b) = 0.0
(c) = 0.1
(d) = 0.2
(e) = 0.3
(f) = 0.4
(g) = 0.5
(h) = 0.6
(i) = 0.7
(j) = 0.8
(k) = 0.9
(l) = 1.0
Figure 5.24: Mach number contours at different instants in time for one period
of oscillation for angle of attack 6.97 , M = 0.095
112
(a)
(b) = 0.0
(c) = 0.1
(d) = 0.2
(e) = 0.3
(f) = 0.4
(g) = 0.5
(h) = 0.6
(i) = 0.7
(j) = 0.8
(k) = 0.9
(l) = 1.0
Figure 5.25: Streamlines at different instants in time for one period of oscillation
for angle of attack 6.97
113
Figure 5.26: Cross-covariance and wave speed for angle of attack 6.97
114
115
116
Chapter 6
Static Wing Aeroelasticity
6.1
Introduction
The aeroelastic research wing (ARW-2) was a full-scale right semispan wing,
which was from part of the NASA Drones for Aerodynamic and Structural Testing (DAST) program [106]. The goals of DAST were to validate, 1) system synthesis and analysis developments for active control of aeroelastic response and 2)
analysis techniques for aerodynamic loads prediction. The wing was designed for
flight testing to investigate the use of active control systems to alleviate manoeuvre and gust loads as well as for flutter suppression. A series of windtunnel tests
were carried out at NASA Langley Research Centre using the Langley Transonic
Dynamics Tunnel (TDT) using heavy gas R-12 (also known as Dichlorodifluoromethane [107]) in the late 1980s and early 1990s. After the first set of tests
were completed the planned flight test program was cancelled. As a result a
second windtunnel test was performed to explore the region of large-amplitude
response that had been uncovered by the first set of windtunnel tests.
The wing planform and dimensional data are presented in Figure 6.1 (a). The
wing has an aspect ratio of 10.3 and a leading edge sweep angle of 28.8. The
wing was designed with supercritical aerofoil sections for a lift coefficient of 0.53
at Mach 0.8 with a cruise angle of attack of 1.53 at 46,800ft and dynamic pressure of 126.4psf. The wing has hydraulically driven trailing edge control surfaces,
two inboard and one outboard, their locations are given in Figure 6.1 (b). For
the windtunnel test the wing was instrumented with 182 pressure transducers, 10
accelerometers and strain gauges at the wing root to measure bending moments.
The instrumentation locations are also shown in Figure 6.1 (b). The wing structure was designed iteratively taking into account the load and stiffness reduction
benefits associated with the use of an active control system. This led to a more
flexible wing than traditional designs. Due to this flexibility a jig shape was defined, which is the shape that the wing is manufactured to, so that at cruise the
117
wing will deform into the desired aerodynamic shape. The first four modes in
still air were first-bending (8.1Hz), second-bending (29.7Hz), in-plane (39.9Hz)
and first-torsion (62.6Hz). The wing geometry is given in Sandford et al. [14]
with the coordinates for the jig shape. This report also included data for the
structural model that was generated with the mode shapes given on the upper
surface of the wing.
tip acceleration reached a peak of 32g. The remainder of the chapter discusses
the previous work, the aerodynamic grid and structural model used in this work,
followed by results of aeroelastic CFD calculations.
6.2
Previous Work
In 1987 Seidel et al. [108] presented experimental results from the first windtunnel
test of the ARW-2 with the objective of providing an early assessment of the wing
aeroelastic stability over a wide range of angles of attack. The experiment varied
the Mach number between 0.6 to 0.9 at dynamic pressures of 50 to 300psf. The
angle of attack was also varied in the range 0 to 2 . They found an unusual
wing instability at a Mach number of 0.9 with a dynamic pressure of 100psf.
Zero damping points where avoided to prevent damage to the wing, since it was
expected to be flight tested. The instability had a frequency of 8.6Hz, which is
very close to the first-bending mode frequency. The frequency of the instability
was found to increase with dynamic pressure to 13Hz at the highest dynamic
pressure tested and was found to be sensitive to angle of attack with minimum
damping near zero degrees. A sustained limit amplitude oscillation at Mach 0.895
with the lowest dynamic pressure was discovered, however the wing became stable
again at Mach 0.9.
Following the first windtunnel test the flight test program was cancelled, so
a second windtunnel test was completed to firmly establish the existence of the
instability that had been found and to gather data to help understand the mechanism forcing the wing oscillations. The results from this second test were published in a number of papers [23, 15, 109]. The second test did not find the
instability predicted in the first windtunnel test, however a region of high dynamic wing response was found instead between Mach numbers 0.92 and 0.94.
Eckstrom et al. [15] published a large amount of data from the second windtunnel
test. This is a brief summary of the findings of the report. At Mach 0.80 there
was little to no response, however there was a high frequency (40Hz) unsteady
flow on the upper surface. At Mach 0.85 there was little response, but a strong
shock was starting to form on the upper surface. There was also flow unsteadiness on both the upper and lower surfaces. At Mach 0.88 there was a moderate
increase in motion for the medium dynamic pressure and a significant increase
in motion for the high dynamic pressure. At this Mach number there was now
a strong oscillating shock with a frequency of about 15Hz. At Mach 0.90 there
was a significant increase in wing motion for all dynamic pressures. There was a
strong shock on both the upper and lower surfaces and for some of the dynamic
pressures the lower surface pressure was lower than the upper surface. This in119
dicated that there was a strong upwards bending and, due to the wing sweep, a
nose-down twist. There were also regions of separation on the lower surface. At
Mach 0.92 the near maximum motion was achieved for all dynamic pressures and
for the high dynamic pressure, at an angle of attack of 1 , wing-tip acceleration
peaked at 32g. Some combinations of angle of attack and dynamic pressure were
not tested due to concerns for the structural safety of the wing. There was a large
region of separation on the lower surface and another region at the trailing edge
of the upper surface. At Mach 0.94 there was a significant decrease in motion for
the high dynamic pressure, whereas for the other dynamic pressures the motion
remained at the same level. At Mach 0.96 for all dynamic pressures the wing
motion decreased substantially. This is summarised in Figure 6.2, which shows
the relative peak wing-tip response as the Mach number is increased. The mech-
Figure 6.2: Maximum PSD peak response from front spar wing-tip accelerometer
(from ref. [15])
anism for the response is related to the chordwise shock movement in conjunction
with flow separation and reattachment on both the upper and lower surfaces of
the wing. After a Mach number of 0.94 the wing response reduced, which was
attributed to full flow separation downstream of the shockwaves with no reattachment. The wing-tip accelerometers show that the frequency of the response
between 8 to 10Hz, which is near the first-bending mode. The amplitude of the
dynamic wing response was found to increase with increasing dynamic pressure.
It was also found to be angle of attack dependant with the maximum motion occurring near an angle of attack of zero degrees and as the angle of attack moved
120
from zero degrees, the maximum motion was shifted to Mach 0.94 at angles of
attack of 2 . Seidel et al. [23] tested two methods to reduce the buffet, the first
was a mean control surface deflection of 6 and the second was a lower surface
spanwise fence. Both methods were found to significantly reduce the maximum
wing response and reduce the Mach number of the peak wing response slightly.
In 1996 Farhangnia et al. [110] published CFD results and compared them
with the experimental data. The calculations were done using ENSAERO which
used the strong conservation-law form of the thin-layer Reynolds-averaged NavierStokes equations with the Baldwin-Lomax turbulence model and a modal approach for the structural solver. Pressure coefficient comparisons were presented
for rigid, static and dynamic cases. An improvement in the results was shown
as the aeroelasticity was taken into account. This paper says that the structural
response is due to the upper shock oscillation.
6.3
6.3.1
The geometry was built from the design coordinates from Sandford et al. [14]
and a C-H type blocking was used to block the wing. The grid size is 2,333,940
nodes and is shown in Figure 6.3.
6.3.2
Structural Model
The structure and the mode shapes are given in Sandford et al. [14] in the form
of coordinates and displacements at 100 points on the upper surface of their finite
element model. A shell model (referred to as the shell model) was generated to
replicate the shell model presented in Sandford et al. [14] (referred to as the
report model). The resulting model is compared with the model [14] in both
frequencies and mode shapes, with the frequencies given in Table 6.1.
Table 6.1: Comparison of the structural model frequencies
Mode Model [14] (Hz) Shell Model (Hz)
1
8.0935
8.0272
2
29.2299
29.0527
3
32.7980
32.6560
4
60.8920
61.4565
The frequencies are all within 1% of the model [14]. The mode shape comparison of the upper surfaces is shown in Figure 6.4. It can be seen that the shell
121
122
model mode 4 has less torsion compared to the report model, however all the
other modes are captured well.
(a) Mode 1
(b) Mode 2
(c) Mode 3
(d) Mode 4
Figure 6.4: Comparison of the mode shapes. Grey - Undeformed shell model,
Red - Report model mode shapes, Green - Shell model mode shapes
The shell model was converted into a beam model using the leading and
trailing edge deflections. The mode shapes are almost exactly the same as the
original data, except for the torsion mode. The torsion mode from the shell
model has the section stretching which cannot be recreated using a beam model,
however the wing is unlikely to stretch to the degree of the shell model, so this
is believed to be an artifact of the finite element model that is presented in the
reference and recreated by the shell model. This is backed up by the paper by
Byrdsong et al. [111] that states that there was sufficient chordwise rigidity to
prevent chordwise bending. The beam model along with a rigid direction that
is parallel with the majority of the ribs was used to generate the mode shapes
123
on the aerodynamic surface using the method discussed in Chapter 3. The mode
shapes defined on the surface mesh are shown in Figure 6.5.
(a) Mode 1
(b) Mode 2
(c) Mode 3
(d) Mode 4
Figure 6.5: Comparison of the mode shapes with the beam model. Grey - Undeformed shell model, Red - Report model mode shapes, Green - Shell model mode
shapes, Pale Blue - Undeformed CFD mesh, Blue - Beam model mode shapes on
the CFD mesh
6.4
Results
gas R-12, the ratio of specific heats was set to 1.14 as per the data given in
Cole and Rivera [107].
Case
1
2
3
4
6.4.1
Mach
number
0.80
0.85
0.92
0.92
6
0.0
2.41 10
123.6psf
91
82
1.0
2.41 106 134.6psf
243
914
0.0
2.60 106 152.5psf
98
6
2.0
2.60 10
154.8psf
111
-
The experiments by Eckstrom et al. [15] showed that this test condition was
free from buffet and there is experimental data for the static deflections from
Byrdsong et al. [111, 112]. Byrdsong et al. did not record any results if the
deflections were not static or quasi-static, implying that this test condition is
static.
The pressure distributions are shown in Figure 6.6 for both the static and
rigid cases. The effect of not taking the structural flexibility into account is considerable with a strong shockwave generated in the outboard sections. By taking
into account the structural flexibility the pressure distributions match the experimental data better. The noticeable discrepancy is a large suction peak in the
static results that are not shown in the experimental results. This improvement
appears to be due a nose-down twist, however the lower surface is not matched
as well as expected. This is likely to be due to an insufficient amount of twist
being generated.
The Mach number contours for the rigid and static solutions are shown in
Figure 6.7 side by side. The sections at 10% span are very similar, since the
effect of the static deflections will have a minimal effect so close to the root of
the wing. The section at 56% span has the static effects starting to show as
the static section has risen slightly. There are also small differences appearing
in the Mach contours. At 80% span the deflection of the static section is more
pronounced and the shockwave seen in the rigid section has weakened significantly
and shifted to the leading edge in the static solution. Finally in the 95% sections,
again there is significant deflection in the static case. The Mach contours have a
strong shock in the rigid case that has shifted forwards to the leading edge and
weakens considerably in the static case. This implies that the large suction peak
is possibly a weak shockwave.
125
The streamlines are shown in Figure 6.8 for both the rigid and static cases.
For both the rigid and static cases, the flow over the entire span is attached. This
is backed up by the experimental observations by Eckstrom et al. [15].
Figure 6.9 shows the 3D representation of the wing in its rigid and deflected
states. The plots showing the front and rear spar deflections show a very good
match with the experimental data of Byrdsong et al. [111, 112]. The twist angle
is shown in Figure 6.10 along with the change in twist between the rigid and
static cases. Although the deflections match the experimental values well, the
twist does not match as well. The change in twist for the CFD follows a different
profile to the experimental result as seen in Figure 6.10(b). This observation
matches with the pressure distribution results, which appears to have insufficient
nose down twist. This could be due to the errors introduced due to the modelling
of the structure, both by Sandford et al. [14] and subsequently by the author.
In the paper by Sandford et al. the frequencies of the structural model did not
match those obtained by the experiment exactly. Then the shell model generated
by the author was within 1% of the Sandford et al. model and mode 4 did not
recreate the same level of twist as the Sandford et al. model. This could be the
reason for the lack of agreement for the twist.
6.4.2
The experimental results of Eckstrom et al. [15] report no large scale motion
at these test conditions. There are also static deflection results available from
Byrdsong et al. [111, 112], implying it is a static case.
The sectional results are shown in Figure 6.11. As seen before the rigid results
produce a strong shockwave and much higher pre-shock levels. The static results
are a significant improvement over the rigid results and match the experimental
results well. Again this improvement is indicative of a nose-down twist.
The Mach number contours are shown in Figure 6.12 and a strong shockwave
can be seen to have formed on the upper surface throughout the entire span for
both rigid and static cases. However the shock weakens as the tip is approached
for the static case, due to the deflection of the wing and the slight nose down
twist.
The streamline plots are given in Figure 6.13. The rigid case does have regions
of separated flow caused by the shock wave especially at the midspan. The static
case shows no regions of separation for any spanwise station. The experiments
by Eckstrom et al. [15] did not report any regions of separated flow, but did
comment that there were regions of unsteady flow on both the upper and lower
surfaces. It is not known whether this unsteadiness is due to shock oscillation or
separated flow.
126
127
129
(a) Twist
The spar deflections in Figure 6.14 do not match the experimental data of
Byrdsong et al. [111, 112] as well as the previous case. This is mainly due to the
dynamic pressure used that matches the experimental data by Eckstrom et al.
[15], but is about halfway between two of the dynamic pressures used to record
the static deflections by Byrdsong et al.. As such it would be expected that the
static deflections do not match.
The twist results are shown in Figure 6.15 again showing a discrepancy between the CFD and the experimental results. This will be due to the dynamic
pressure used and the failure of the structural model to fully capture the amount
of twist in mode 4.
131
133
(a) Twist
6.4.3
This case did not produce static deflections in the experiments and had near
maximum dynamic wingtip motions [15]. As such there are no static results to
compare against by Byrdsong et al. [111, 112], since they did not record any
dynamic deflections. The static calculation lift and residual histories are shown
in Figure 6.16. These show that the calculation has not converged to a steady
solution and is therefore is likely to be an unsteady case. The lift history shows
oscillations that indicate a possible shock-oscillation.
6.4.4
This case like the last also has large dynamic wingtip motion in the experiment
[15] and therefore has no static deflection results to compare with from Byrdsong
et al. [111, 112]. This experimental observation is backed up by the calculation
lift and residual histories in Figure 6.17. The lift history has an oscillation,
which indicates a possible shock-oscillation. The residual history shows that the
static calculation converges to about -4 and then converges no further, which also
indicates that this case is unsteady.
6.4.5
Buffet Search
A similar search to the one performed for the BGK number one aerofoil was
performed for the ARW-2 to find the most likely buffet cases using steady static
calculations. The calculations were performed at angle of attack 0.0 , Reynolds
number 2.5 106 , dynamic pressure 150 psf and the Mach number was varied
135
136
6.5
Summary
It was shown that for the static case the AMB solver matched the experiments
well. However when the flow conditions where expected to lead to an unsteady
result the solver was unable to match the experiment using steady static calculations. This is unsurprising and can be rectified by running computationally
expensive unsteady calculations in the future. A buffet search was completed to
find the most promising case for buffet calculations, leading to Mach 0.92 and 0.94
being the most likely candidates. Cases 3 and 4 are also promising candidates for
buffet calculations as these also failed to converge and hint at shock-oscillation.
Although all the preparation was completed for buffet calculations on the ARW
wing, the actual unsteady calculation was not ran. This was because the calculation was unable to be ran in a reasonable time period on the computational
resources available. To run this calculation in a reasonable time period more computational nodes are required and the grid needs to be split into more domains
to allow it to be well load balanced on the increased number of nodes.
137
138
Chapter 7
Conclusions
7.1
Conclusions
beam stick models by specifying a direction for extrapolation and it was shown
to perform as well as IIM. It was also shown that when the rigid direction was
not close to perpendicular to the beam, the section shapes were worse.
The effect of the transformation methods on the flutter boundary of the
Goland and MDO wings was also investigated in Chapter 4. It was shown that
the local methods had less spread in their prediction of the flutter point as the
structural model was varied than the global methods. More importantly there
was more spread in the flutter point when the transformation methods extrapolated. This adds weight to the previous conclusion that the transfer methods
should not extrapolate and that extrapolation should be done as a pre-processing
step. It is important to stick to the structural model assumptions and extrapolate perpendicular to the beam, as this was shown to reduce the spread in the
predicted flutter point. Since it was shown that the choice of transfer method can
have a significant effect on the predicted flutter point, it is therefore advisable to
use multiple transfer methods to build confidence in the results obtained.
To facilitate buffet calculations the AMB solver was verified for shock-induced
buffet. This was completed using the BGK No. 1 aerofoil and a number of convergence studies in Chapter 5. A buffet search using steady state calculations
was performed to estimate the buffet boundary. Two methods were used, the
first was to use the lift history and the second was to use the calculation residual.
The lift history method matched the experimental shock oscillation region very
well, but performed less well against the experimental buffet onset. The residual
method performed better at predicting buffet onset, but predicted a slightly larger
region than the experiment. This did provide evidence that steady state calculations could be used to predict shock-oscillation cases and potentially buffet onset
boundary. A number of unsteady calculations were used to check the results
of the steady state predictions and covered the pre, post and shock-oscillation
regimes. The results from these calculations matched the steady state predictions and the experimental data. For the two cases that proved to be unsteady,
the mechanism behind the shock-oscillation was investigated. Through the use of
cross-covariance and probes inside and outside the boundary-layer, pressure waves
could be detected travelling downstream inside the boundary-layer and upstream
outside the boundary-layer. The time lags of these pressure waves could be calculated and they correlated to the shock-oscillation period for the 4.905 angle of
attack case.
Finally the static aeroelasticity of the ARW-2 wing was calculated to verify
the AMB solver for static aeroelasticity in Chapter 6. For cases that had been
shown to be static experimentally the solver produced good results providing reasonable agreement with experimental pressure distributions and wing deflections.
140
The results did not match as well as expected for wing twist, this is thought to
be due to the structural model failing to capture the torsion mode well. The
experimentally unsteady cases failed to converge when computed using steady
state calculations. This led to a similar buffet search to the BGK No. 1 aerofoil,
which predicted the Mach numbers 0.92 and 0.94 as the most likely candidates for
unsteady cases for an angle of attack of zero degrees, Reynolds number 2.5 106
and dynamic pressure 150psf. This was consistent with the experimental data.
7.2
Future Work
Buffet calculations can now be performed on the ARW-2 wing. This would be a
combination of the experience gained from the 2D aerofoil buffet calculations and
the static aeroelastic work already preformed on the ARW-2 wing. It has already
been commented that the literature indicates the buffet is due to shock-oscillation
on the upper surface, but the mechanism behind it has not been studied. As such
it would be interesting to investigate the buffet mechanism on the wing and see
if it is the same as the aerofoil case with propagating pressure waves.
There has been a lot of work in recent years on the RBF transfer method
for aeroelasticity and grid deformation. The most interesting development is
the pointwise partition of unity and has been shown to localise the method. It
would be an interesting extension of the current work to see how this would effect
the extrapolation characteristics of RBF and if it prevents the oscillation in the
section shapes that were be generated. The effect of the pointwise partition of
unity on flutter prediction variability would also be an interesting study.
Although the rigid ribs transformation model was useful in testing how varying
the direction of the ribs affects the flutter prediction, it requires extending for
use with more complex geometries. This can be achieved in a similar way to that
used for CVT, by using a hierarchy of components and blending functions.
An extension of the aerofoil buffet work would be to model the turbulence
using DES or LES to investigate whether significant improvements could be made
by using these more expensive options over URANS.
141
142
Bibliography
[1] Spalart P.R. Strategies for Turbulence Modelling and Simulations. International Journal of Heat and Fluid Flow, 21:252263, 2000.
[2] Rendall T.C.S. and Allen C.B. Unified Fluid-Structure Interpolation and
Mesh Motion using Radial Basis Functions. International Journal for Numerical Methods in Engineering, 74:15191559, 2008.
[3] Xiao Q., Tsai H.M., and Liu F. Numerical Study of Transonic Buffet on a
Supercritical Airfoil. AIAA Journal, 44(3):620628, March 2006.
[4] Canfield R.A., Morgenstern S.D., and Kunz D.L. Alleviation of BuffetInduced Vibration using Piezoelectric Actuators. Computers and Structures, 86:281291, 2008.
[5] NASA. Website, 1989. http://www.dfrc.nasa.gov/Gallery/Photo/F-18HARV/HTML/EC89[6] ESDU International plc. An Introduction to Aircraft Buffet and Buffeting.
Technical Report 87012, ESDU International plc, 2006.
[7] Woodgate M.A., Badcock K.J., Rampurawala A.M., Richards B.E., Nardini
D., and Henshaw M.J. de C. Aeroelastic Calculations for the Hawk Aircraft
using the Euler Equations. Journal of Aircraft, 42(4):10051012, 2005.
[8] Fukushima
C.
and
Westerweel
J.
http://en.wikipedia.org/wiki/File:Jet.jpg.
Website,
2007.
[9] Wilcox D.C. Turbulence Modelling for CFD. DCW Industries, Inc., 3
edition, 2006.
[10] Rampurawala A.M. Aeroelastic Analysis of Aircraft with Control Surfaces
using CFD. PhD thesis, University of Glasgow, 2006.
[11] Samareh J.A. and Bhatia K.G. A Unified Approach to Modelling Multidisciplinary Interactions. AIAA Paper AIAA-2000-4704, 2000.
143
[38] Farhat C., Geuzaine P., and Brown G. Application of a Three-Field Nonlinear Fluid-Structure Formulation to the Prediction of the Aeroelastic Parameters of an F-16 Fighter. Computers and Fluids, 32:329, 2003.
[39] Badcock K.J., Rampurawala A.M., and Richards B.E. Intergrid Transformation for Aircraft Aeroelastic Simulations. AIAA Journal, 42(9):1936
1939, 2004.
[40] Badcock K. J., Woodgate M. A., and Richards B. E. Direct Aeroelastic
Bifurcation Analysis of a Symmetric Wing Based on the Euler Equations.
Journal of Aircraft, 42(3):731737, 2005.
[41] Badcock K.J. and Woodgate M.A. Prediction of Bifurcation Onset of Large
Order Aeroelastic Models. In AIAA-2008-1820 in the proceedings of the
49th Structural Dynamics and Materials Conference, AIAA, Schaumberg,
Illinois, 2008.
[42] Shang J.S. Computational Fluid Dynamics Application to Aerospace Science. The Aeronautical Journal, 113(1148):619632, 2009.
[43] Batina J.T. Unsteady Euler Airfoil Solutions Using Unstructured Dynamic
Meshes. AIAA Journal, 28(8):13811388, August 1990.
[44] Farhat C., Degand C., Koobus B., and Lesoinne M. Torsional Springs
for Two-Dimensional Dynamic Unstructured Fluid Meshes. Computational
Methods in Applied Mechanics and Engineering, 163:231245, 1998.
[45] Blom F.J. Considerations on the Spring Analogy. International Journal
for Numerical Methods in Fluids, 32:647668, 2000.
[46] Dubuc L., Cantariti F., Woodgate M., Gribben B., Badcock K.J., and
Richards B.E. A Grid Deformation Technique for Unsteady Flow Computations. International Journal for Numerical Methods in Fluids, 38:285311,
2000.
[47] Chen P.C. and Jadic I. Interfacing of Fluid and Structural Models via Innovative Structural Boundary Element Method. AIAA Journal, 36(2):282
287, 1998.
[48] Liu X., Qin N., and Xia H. Fast Dynamic Grid Deformation based on
Delaunay Graph Mapping. Journal of Computational Physics, 211:405
423, 2006.
146
[49] de Boer A., van der Schoot M.S., and Bijl H. Mesh Deformation based on
Radial Basis Function Interpolation. Computers and Structures, 85:784
795, 2007.
[50] Rendall T.C.S. and Allen C.B. Parallel Efficient Mesh Motion using Radial
Basis Functions with Application to Multi-Bladed Rotors. International
Journal for Numerical Methods in Engineering, 81:89105, 2009.
[51] Farhat C. and Lesoinne M. Two Efficient Staggered Algorithms for the Serial and Parallel Solution of Three-Dimensional Nonlinear Transient Aeroelastic Problems. Computer Methods in Applied Mechanics and Engineering,
182:499515, 2000.
[52] Harder R.L. and Desmarais R.N. Interpolation Using Surface Splines. Engineering Notes, pages 189191, February 1972.
[53] Menter F.R. Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications. AIAA Journal, 32(8):15981605, 1994.
[54] Menter F.R., Kuntz M., and Langtry R. Ten Years of Industrial Experience
with the SST Turbulence Model. In Hanjalic K., Nagano Y., and Tummers
M., editors, Turbulence, Heat and Mass Transfer 4. Begell House Inc., 2003.
[55] Badcock K.J., Richards B.E., and Woodgate M.A. Elements of Computational Fluid Dynamics on Block Structured Grids using Implicit Solvers.
Progress in Aerospace Sciences, 36:351392, 2000.
[56] Bekas C. and Saad Y. Computation of Smallest Eigenvalues using Spectral
Schur Complements. SIAM Journal on Scientific Computing, 27(2):458
481, 2005.
[57] Guruswamy G.P. and Byun C. Direct Coupling of Euler Flow Equations
with Plate Finite Element Structures. AIAA Journal, 33(2):375377, 1995.
[58] Felker F.F. Direct Solution of Two-Dimensional Navier-Stokes Equations
for Static Aeroelasticity Problems. AIAA Journal, 31(1):148153, 1993.
[59] H
ubner B., Walhorn E., and Dinkler D. A Monolithic Approach to FluidStructure Interaction using Space-Time Finite Elements. Computer Methods in Applied Mechanics and Engineering, 193:20872104, 2004.
[60] Smith M.J., Hodges D.H., and Cesnik C.E.S. An Evaluation Of Computational Algorithms to Interface Between CFD and CSD Methodologies.
Technical Report WL-TR-96-3055, U.S. Air Force Research Lab., November 1995.
147
[61] Kamakoti R. and Shyy W. Fluid-Structure Interaction for Aeroelastic Applications. Progress in Aerospace Science, 40(8):535558, 2004.
[62] Sadeghi M., Lui F., Lai K.L., and Tsai H.M. Application of ThreeDimensional Interfaces for Data Transfer in Aeroelastic Computations. In
22nd Applied Aerodynamics Conference and Exhibit, number 2004-5376,
August 2004.
[63] Done G.T.S. Interpolation of Mode Shapes: A Matrix Scheme using TwoWay Spline Curves. Aeronautical Quarterly, XVI:333349, 1965.
Website.
[74] Cebral J.R. and Lohner R. Conservative Load Projection and Tracking for
Fluid-Structure Problems. AIAA Journal, 35(4):687692, April 1997.
[75] Beckert A. and Wendland H. Multivariate Interpolation for Fluid-StructureInteraction Problems using Radial Basis Functions. Aerospace Science and
Technology, 5:125131, 2001.
[76] Ahrem R., Beckert A., and Wendland H. A Meshless Spatial Coupling
Scheme for Large-Scale Fluid-Structure-Interaction Problems. Computer
Modelling in Engineering and Sciences, 12:121136, 2006.
[77] Ahrem R., Beckert A., and Wendland H. Recovering Rotations in Aeroelasticity. Journal of Fluids and Structures, 23:874884, 2007.
[78] Rendall T.C.S. and Allen C.B. Improved Radial Basis Function FluidStructure Coupling via Efficient Localised Implementation. International
Journal for Numerical Methods in Engineering, 78:11881208, 2009.
[79] Goura G.S.L. Time Marching Analysis of Flutter using Computational
Fluid Dynamics. PhD thesis, University of Glasgow, 2001.
[80] Burnett D.S. Finite Element Analysis from Concepts to Applications.
Addison-Wesley Publishing Company, 1987.
[81] Smith M.J., Hodges D.H., and Cesnik C.E.S. Evaluation of Computational
Algorithms Suitable for Fluid-Structure Interactions. Journal of Aircraft,
37(2):282294, 2000.
[82] Samareh J.A. Aeroelastic Deflection of NURBS Geometry. In 6th International Conference on Numerical Grid Generation in Computational Field
Simulation, July 1998.
[83] Wendland H. Fast Evaluation of Radial Basis Functions: Method Based
on Partition of Unity. Approximation Theory X: Wavelets, Splines and
Applications, pages 473483, 2002.
[84] Beran P.S., Khot N.S., Eastep F.E., Snyder R.D., and Zweber J.V. Numerical Analysis of Store-Induced Limit-Cycle Oscillation. Journal of Aircraft,
41(6):13151326, 2004.
[85] Girodroux-Lavigne P., Grisval J.P., Guillemot S., Henshaw M., Karlsson
A., Selmin V., Smith J., Teupootahiti E., and Winzell B. Comparison of
Static and Dynamic Fluid-Structure Interaction Solutions in the Case of a
Highly Flexible Modern Transport Aircraft Wing. Aerospace Science and
Technology, 7:121133, 2003.
149
[86] Lee B.H.K. and Ohman L.H. Unsteady Pressure and Forces during Transonic Buffeting of a Supercritical Airfoil. Journal of Aircraft, 21(6):439441,
1984.
[87] Lee B.H.K., Ellis F.A., and Bureau J. Investigation of the Buffet Characteristics of Two Supercritical Airfoils. Journal of Aircraft, 26(8):731736,
1989.
[88] Lee B.H.K. Investigation of Flow Separation on a Supercritical Airfoil.
Journal of Aircraft, 26(11):10321037, 1989.
[89] Lee B.H.K. Transonic Buffet on a Supercritical Aerofoil. Aeronautical
Journal, pages 143152, May 1990.
[90] Huang R.F. and Lee H.W. Turbulence Effect on Frequency Characteristics
of Unsteady Motions in Wake of Wing. AIAA Journal, 38(1):8794, 2000.
[91] Xiao Q., Tsai H.M., and Liu F. A Numerical Study of Transonic Buffet
on a Supercritical Airfoil. In 42nd AIAA Aerospace Sciences Meeting and
Exhibit, 5-8 January 2004, Reno, Nevada, AIAA 2004-1056.
[92] Lee B.H.K. and Tang F.C. Characteristics of the Surface Pressures on a
F/A-18 Vertical Fin due to Buffet. Journal of Aircraft, 31(1):228235, 1994.
[93] Lee B.H.K. Self-Sustained Shock Oscillations on Airfoil at Transonic
Speeds. Progress in Aerospace Sciences, 37:147196, 2001.
[94] Levy Jr. L.L. Experimental and Computational Steady and Unsteady Transonic Flows about a Thick Airfoil. AIAA Journal, 16(6):564572, 1978.
[95] McDevitt J.B. Supercritical Flow about a Thick Circular-Arc Airfoil. Technical Report 78549, NASA, 1979.
[96] Rumsey C.L., Sanetrik M.D., Biedrom R.T., Melson N.D., and Parlette
E.B. Efficient and Accuracy of Time-Accurate Turbulent Navier-Stokes
Computations. Computations and Fluids, 25(2):217236, 1996.
[97] Prananta B.B. Two-Dimensional Transonic Aeroelastic Analysis using
Thin-Layer Navier-Stokes Method. Journal of Fluids and Structures,
12:665676, 1998.
[98] Raghunathan S., Gillan M.A., Cooper R.K., Mitchell R.D., and Cole J.S.
Shock Oscillations on Biconvex Aerofoils. Aerospace Science and Technology, 1:19, 1999.
150
[99] Chen L.W., Xu C.Y, and Lu X.Y. Numerical Investigation of the Compressible Flow Past an Aerofoil. Journal of Fluid Mechanics, 643:97126,
2010.
[100] Jacquin L., Molton P., Deck S., Maury B., and Soulevant D. Experimental
Study of Shock Oscillation over a Transonic Supercritical Profile. AIAA
Journal, 47(9):19851994, September 2009.
[101] Deck S. Zonal-Detached-Eddy Simulation of the Flow Around a High-Lift
Configuration. AIAA Journal, 43(11):23722384, 2005.
[102] Garnier E. and Deck S. Large-Eddy Simulation of Transonic Buffet over a
Supercritical Airfoil. In Direct and Large-Eddy Simulation VII, volume 13
of ERCOFTAC, pages 549554. Springer Netherlands.
[103] Raghunathan S., Mitchell R.D., and Gillan M.A. Transonic shock oscillations on naca 0012 aerofoil. Shock Waves, 8:191202, 1998.
[104] Iovnovich M. and Raveh D.E.
Reynolds-Averaged Navier-Stokes
Study of the Shock-Buffet Instability Mechanism.
In 52nd
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and
Materials Conference, number AIAA 2011-2078, 2011.
[105] Singh J.P. Estimation of Shock-Induced Buffet Onset. Acta Mechanica,
151:245253, 2001.
[106] Murrow H.N. and Eckstrom C.V. Drones for Aerodynamic and Structural
Testing (DAST) - A Status Report. Journal of Aircraft, 16(8):521526,
1979.
[107] Cole S.R. and Rivera Jr. J.A. The New Heavy Gas Testing Capability in
the NASA Langley Transonic Dynamics Tunnel. Technical Report NASATM-112702, NASA Langley Research Center, 1997.
[108] Seidel D.A., Sandford M.C., and Eckstrom C.V. Measured Unsteady Transonic Aerodynamic Characteristics of an Elastic Supercritical Wing. Journal of Aircraft, 24(4):225230, April 1987.
[109] Eckstrom C.V., Seidel D.A., and Sandford M.C. Unsteady Pressure and
Structural Response Measurements on an Elastic Supercritical Wing. Journal of Aircraft, 27(1):7580, 1990.
[110] Farhangnia M., Guruswamy G., and Biringen S., editors. Transonic-Buffet
Associated Aeroelasticity of a Supercritical Wing, number AIAA 96-0286 in
34th Aerospace Sciences Meeting and Exhibit, Reno, NV, January 1996.
151
[111] Byrdsong T.A., Adams R.R., and Sandford M.C. Close-Range Photogrammetric Measurement of Static Deflections for an Aeroelastic Supercritical
Wing. Technical Report NASA TM 4194, NASA, Langley Research Center,
Hampton, Virginia, 1990.
[112] Byrdsong T.A., Adams R.R., and Sandford M.C. Close-Range Photogrammetric Measurement of Static Deflections for an Aeroelastic Supercritical
Wing. Technical Report Supplement to NASA TM 4194, NASA, Langley
Research Center, Hampton, Virginia, 1990.
152
Appendix A
Infinite Plate Spline
This appendix will provide the derivation of the transformation matrix for infinite
plate splines. The deflections normal to the plate surface due to N point forces
Fi at given locations (xi , yi ) on the plate can be written as
dz(x, y) = a0 + a1 x + a2 y +
N
X
Fi ri2 ln ri2
(A.1)
i=1
where ri2 = x2i + yi2 . The unknowns, ak and Fi are obtained from the equilibrium
conditions
N
N
N
X
X
X
Fi =
xi Fi =
yi Fi = 0
(A.2)
i=1
i=1
i=1
N
X
Fi rij2 ln rij2
(A.3)
i=1
where rij2 = (xi xj )2 +(yi yj )2 . By applying Equations A.2 and A.3 to both the
structural and aerodynamic points the following equations are obtained, expressed
in matrix form as
dxs = Rs a + Zs Fs
(A.4)
dxa = Ra a + Zas Fs
(A.5)
Rs =
1 xs1 ys1
1 xs2 ys2
..
..
..
.
.
.
1 xsN ysN
153
(A.6)
and
Ra =
1 xa1 ya1
1 xa2 ya2
..
..
..
.
.
.
1 xaN yaN
xai xsj
2
+ yai ysj
2 i
ln
xai xsj
2
+ yai ysj
2 i
which are effectively matrices of the basis function (rij2 ln rij2 ). Given that
dxa = T dxs
(A.7)
where T is the transformation matrix and from the principle of virtual work
Fs = T T Fa
(A.8)
154
(A.9)
Appendix B
Inverse Isoparametric Mapping
This appendix will provide the derivation of the transformation matrix for Inverse
Isoparametric Mapping. The same shape functions Ni are used for both the
aerodynamic grid points and the structural deformation
xa =
n
X
Ni (, )xesi
(B.1)
Ni (, )ysei
(B.2)
i=1
ya =
n
X
i=1
where n is the number of nodes for the element. The element is assumed to lie in
the x y plane and the superscript e indicates the structural node is a member
of that element. The shape function for a quadrilateral element[80] is given by
N1 (, ) = (1 )(1 )/4
N2 (, ) = (1 + )(1 )/4
N3 (, ) = (1 + )(1 + )/4
N4 (, ) = (1 )(1 + )/4
, [1, 1]
(B.3)
where
(B.4)
(B.5)
(B.6)
155
where
(B.7)
ya ay4 ay2
ay1 + ay3
(B.8)
This can be substituted into Equation B.4 and rearranged into the form
where
A 2 + B + C = 0
(B.9)
(B.10)
xa ax4 ax2
ax1 + ax3
to substitute into Equation B.6, this leads to
=
(B.11)
(B.12)
Equation B.9 can be solved as a normal quadratic, there will be one solution
of the equation inside the bounds [1, 1]. Then can be solved for using either
Equation B.8 or B.11. Once the local coordinates for the aerodynamic grid points
have been calculated the shape functions can be calculated using Equation B.3.
Then the displacements can be calculated using
dxa =
n
X
Ni (, )dxesi
(B.13)
i=1
(B.14)
(B.15)
where T is the local transformation matrix. This then allows the forces to be
calculated using
Fes = T T Fa
(B.16)
from the principle of virtual work and therefore satisfies conservation of energy.
156
Appendix C
Radial Basis Functions
This appendix will provide the derivation of the transformation matrix for radial
basis functions. The radial basis function interpolant [2] has the form
s(x) =
N
X
i=1
i (kx xi k) + p(x)
(C.1)
where s(x) is the function to be evaluated at x, is the basis function, the index i
identifies the centres for the RBF and xi is the location of that centre. The linear
polynomial p(x) is used to ensure that translations and rotations are recovered.
The coefficients i are found by requiring the exact recovery of the structural
node positions. The polynomial in the x-direction is given by
px (x) = x0 + xx x + xy y + xz z
(C.2)
When the polynomials are included in the system, the additional requirement is
also required
N
X
i q(x) = 0
(C.3)
i=1
for all polynomials q(x) with degree less than or equal to p(x). For ease of
computation the exact recovery of centres gives using up to linear polynomial
terms
Xs = Css ax
(C.4)
Ys = Css ay
(C.5)
Zs = Css az
(C.6)
157
where
Xs =
0
0
0
0
xs
xs1
.
, xs =
..
xsns
, ax =
0 0
0
0
1
1
0 0
0
0
xs1
xs2
0 0
0
0
y
ys2
s1
0 0
0
0
zs1
zs2
Css =
1 xs
ys1 zs1 s1 s1 s1 s2
1
..
..
..
..
..
..
.
.
.
.
.
.
1 xsns ysns zsns sns s1 sns s2
x0
xx
xy
xz
xs1
..
.
xsns
...
...
...
...
...
..
.
xsns
ysns
zsns
s1 sns
..
.
. . . sns sns
(C.7)
(C.8)
..
..
..
..
..
..
..
Aas = ...
(C.9)
.
.
.
.
.
.
.
1 xana yana zana ana s1 ana s2 . . . ana sns
The positions of the aerodynamic surface points given by the the vectors xa ,
ya and za analogous to xs in Equation C.7 is given by
xa
T 0 0
Xs
ya = 0 T 0 = Ys
(C.10)
za
0 0 T
Zs
1
where T = Aas Css
. Since the matrix Css has a zero block, the matrix can be
subdivided as shown below
0 P
(C.11)
Css =
PT M
The vectors ax , ay and az are split into the parts which multiply the RBF (superscript RBF) and polynomial (superscript PLY)terms,
aPx LY = Mp P M 1 xs
(C.12)
= M 1 M 1 P T Mp P M 1 xs
(C.13)
aRBF
x
1
where Mp = P M 1 P T
. Forming the transformation matrix is now straight
forward
Mp P M 1
(C.14)
T = Aas
M 1 M 1 P T Mp P M 1
158