You are on page 1of 133

Master of Science Thesis

The aerodynamics of an
impulsively-started insect wing
An experimental and numerical investigation
F.J. Venneman
June 26, 2009
The aerodynamics of an
impulsively-started insect wing
An experimental and numerical investigation
Master of Science Thesis
For obtaining the degree of Master of Science in Applied Physics
at Delft University of Technology
F.J. Venneman
June 26, 2009
Faculty of Aerospace Engineering Faculty of Applied Sciences
Delft University of Technology
Copyright c Aerospace Engineering & Applied Physics, Delft University of Technology
All rights reserved.
DELFT UNIVERSITY OF TECHNOLOGY
DEPARTMENT OF AERODYNAMICS & APPLIED PHYSICS
The undersigned hereby certify that they have read and recommend to the Faculty of Ap-
plied Sciences for acceptance the thesis entitled The aerodynamics of an impulsively-
started insect wing by F.J. Venneman in fulllment of the requirements for the degree
of Master of Science.
Dated: June 26, 2009
Supervisors:
Prof.dr. R.F. Mudde
Dr.eng. L.M. Portela
Dr.ir. M. Tummers
Prof.dr.ir.drs. H. Bijl
Dr.ir. B.W. van Oudheusden
Preface
In the process of choosing a subject for my graduation project I knew for sure that I wanted
something that I was really passionate for. I also wanted to do something that I could
basically explain on a birthday party. Many elds of applied physics seemed attractive, so
it was a hard choice. Until a morning in the summer of 2008, under the shower, I realized I
had to study the uid dynamics of insects or birds. Flight of insects and specically birds
has astonished me for years and uid dynamics has always been the part of physics that I
felt most condent with.
So I went shopping at various research groups (Multi-Scale Physics, the Laboratory for
Aero & Hydrodynamics and the Aerodynamics group) and ended in the position that they
all were part of my graduation project. At the start it was dicult to get a grip of the
Aerodynamics vocabulary, but once used to it, you realize all the terms relate to basic
physical transport phenomena. Besides the language, the challenge of graduation for me
was to keep focused on the insects while debugging my corrupt scripts. Luckily once in a
month a bug landed on one of my screens to tell me that he/she was the subject of my
graduation.
I would like to end this preface with some words of gratitude. First of all my girlfriend,
Saskia, for understanding when I was in a dip and innite support. Then I would like to
thank, Christian Poelma for providing his piv data, Frank Bos for helping me out when
I got segmentation errors, Bas van Oudheusden for advice and reading my report over
and over again, Rob Mudde for coaching me through my graduation project and of course
Hester Bijl for her enthusiasm and keeping me on schedule.
Freek Venneman
Delft, June 15, 2009
MSc. Thesis F.J. Venneman
vi Preface
F.J. Venneman MSc. Thesis
Abstract
The design of Micro Air Vehicles (mavs) is currently an area of rapid growth. To improve
mavs, a better understanding of bird and insect aerodynamics is very helpful. In this
study the time-dependent three-dimensional ow around an impulsively-started fruit-y
wing is investigated at a Reynolds number of 256 and a constant angle of attack ( = 50

).
Both the forces on the wing and the ow-eld structure are studied from an experimental
and a computational point of view.
On the experimental part, the work presented in this thesis is a rst application of
a new tool to determine uid dynamic forces in three dimensions. Experimentally
obtained velocity-elds around a dynamically-scaled robotic wing, with the previously
mentioned kinematics, are provided for the purpose of this study. Pressure-elds are
deduced from these velocity-elds, by means of a new tool, called the planar Poisson
approach. The forces on the wing are obtained by the momentum approach, using
both the velocity and the pressure-elds. These forces are in good agreement with the
forces that were directly measured by a sensor mounted on the robotic wing. Excellent
agreement is found in the steady part of the stroke, where the relative dierence with the
measurements is within 13%.
For the computational element of this study, the Navier-Stokes equations around
the insect wing are solved using OpenFOAM as a framework. The incompressible ow
is solved on a dynamically moving mesh, which deforms based on Radial Basis Function
interpolation, which is a new mesh deformation tool. The forces obtained from the
computations are also in good agreement with the measurements, in the steady part of
the stroke, the relative dierence with the measurements in the steady part of the stroke
is within 22%.
MSc. Thesis F.J. Venneman
Table of Contents
Preface v
Abstract vii
List of Figures xiii
List of Tables xv
1 Introduction 1
1.1 Goal and context of present study . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 MAVs and sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Crop inspection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Disaster detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Environmental monitoring . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Brief background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Objectives and approach of present study . . . . . . . . . . . . . . . . . . . . 5
2 Flapping insect wing aerodynamics and problem denition 7
2.1 Physics of apping insect wing aerodynamics . . . . . . . . . . . . . . . . . . 7
2.2 Experimental investigation of insect ight . . . . . . . . . . . . . . . . . . . . 10
2.3 Numerical investigation of insect ight . . . . . . . . . . . . . . . . . . . . . 11
2.3.1 2D Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.2 3D Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.4 Comparison of experiments and computations . . . . . . . . . . . . . . . . . . 15
2.5 Problem denition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
MSc. Thesis F.J. Venneman
x Table of Contents
3 Experimental test case description 17
3.1 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Kinematics of the wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3 PIV measurement technique . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.4 Floweld around the wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4 Determining forces on insect wings experimentally 23
4.1 The Blasius approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2 The momentum approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.2.1 Scientic literature about the momentum approach . . . . . . . . . . . 26
4.2.2 Determining pressure from two dimensional cross-sections . . . . . . . 27
Numerical scheme and boundary conditions . . . . . . . . . . . . . . . 28
Masking unreliable data regions . . . . . . . . . . . . . . . . . . . . . 29
Transforming velocity to a rotating frame of reference . . . . . . . . . 30
Details Poisson scheme . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2.3 Determining forces from two dimensional pressure-elds . . . . . . . . 31
4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3.1 Results Blasius approach . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3.2 Results momentum approach . . . . . . . . . . . . . . . . . . . . . . 33
Results momentum approach: Pressure elds . . . . . . . . . . . . . . 33
Results momentum approach: Forces . . . . . . . . . . . . . . . . . . 35
4.4 Conclusion: Good estimation of forces on steady stroke . . . . . . . . . . . . 42
4.4.1 Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.4.2 Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5 Determining forces on insect wings numerically 45
5.1 Theory for a numerical study of insect wing aerodynamics . . . . . . . . . . . 45
5.2 2D numerical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.2.1 From three-dimensional to two-dimensional motion . . . . . . . . . . . 47
5.2.2 Building the grid in two dimensions . . . . . . . . . . . . . . . . . . . 48
5.2.3 Solving the system in two dimensions . . . . . . . . . . . . . . . . . . 48
5.2.4 Setup cases in two dimensions . . . . . . . . . . . . . . . . . . . . . . 50
Variation of grid density and Courant number . . . . . . . . . . . . . . 52
Validation impulsively-started wing . . . . . . . . . . . . . . . . . . . 52
Reynolds-Strouhal relation . . . . . . . . . . . . . . . . . . . . . . . . 53
5.2.5 Discussion of the two-dimensional results . . . . . . . . . . . . . . . . 54
Variation of grid density and Courant number . . . . . . . . . . . . . . 54
F.J. Venneman MSc. Thesis
Table of Contents xi
Validation impulsively-started wing . . . . . . . . . . . . . . . . . . . 54
Reynolds-Strouhal relation . . . . . . . . . . . . . . . . . . . . . . . . 54
5.3 3D numerical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.3.1 Building the grid in three dimensions . . . . . . . . . . . . . . . . . . 58
5.3.2 Solving the system in three dimensions . . . . . . . . . . . . . . . . . 60
5.3.3 Setup cases in three dimensions . . . . . . . . . . . . . . . . . . . . . 61
Variation of Courant number . . . . . . . . . . . . . . . . . . . . . . 62
Variation of grid density . . . . . . . . . . . . . . . . . . . . . . . . . 62
Variation of wing planform . . . . . . . . . . . . . . . . . . . . . . . . 62
Variation of the aspect ratio of the wing . . . . . . . . . . . . . . . . 63
Variation of kinematics of the start . . . . . . . . . . . . . . . . . . . 64
5.3.4 Discussion of the three-dimensional results . . . . . . . . . . . . . . . 64
Variation of Courant number . . . . . . . . . . . . . . . . . . . . . . 66
Variation of grid density . . . . . . . . . . . . . . . . . . . . . . . . . 67
Variation of wing planform . . . . . . . . . . . . . . . . . . . . . . . . 67
Variation of the aspect ratio of the wing . . . . . . . . . . . . . . . . 67
Variation of kinematics of the start . . . . . . . . . . . . . . . . . . . 67
5.3.5 Qualitative images of the ow . . . . . . . . . . . . . . . . . . . . . . 69
5.4 Conclusion: Good estimation of force-coecients on steady stroke . . . . . . . 72
5.4.1 Two-dimensional simulations . . . . . . . . . . . . . . . . . . . . . . 72
5.4.2 Three-dimensional simulations . . . . . . . . . . . . . . . . . . . . . . 72
6 Conclusions and recommendations 75
6.1 Detailed conclusions experimental force determination . . . . . . . . . . . . . 76
Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.2 Detailed conclusions numerical force determination . . . . . . . . . . . . . . . 77
Two-dimensional simulations . . . . . . . . . . . . . . . . . . . . . . 78
Three-dimensional grid . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.3 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Bibliography 80
A Derivations of approaches to determine force on an immersed body 85
A.1 Derivation of the momentum approach . . . . . . . . . . . . . . . . . . . . . 85
A.2 Derivation of the Blasius approach . . . . . . . . . . . . . . . . . . . . . . . 89
A.2.1 Velocity integrals in large bounded regions . . . . . . . . . . . . . . . 90
A.2.2 Force acting on a large body . . . . . . . . . . . . . . . . . . . . . . 92
A.2.3 Aerodynamic Force . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
MSc. Thesis F.J. Venneman
xii Table of Contents
B Rotational motion 95
B.1 Transforming velocity to a rotating frame of reference . . . . . . . . . . . . . 95
B.2 Rotational frame of reference . . . . . . . . . . . . . . . . . . . . . . . . . . 97
B.3 Exponential t for rotational motion . . . . . . . . . . . . . . . . . . . . . . 98
C Moving Grids 101
C.1 Radial Basis Function interpolation . . . . . . . . . . . . . . . . . . . . . . . 101
C.2 Space Conservation Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Mass conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
Momentum conservation . . . . . . . . . . . . . . . . . . . . . . . . . 104
D Grid quality 105
D.1 Grid quality measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Aspect Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Non-orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Cell Skewness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
D.2 2D grid quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
D.3 3D grid quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
E OpenFOAM settings 111
E.1 Solution and algorithm control . . . . . . . . . . . . . . . . . . . . . . . . . . 111
E.2 2D OpenFOAM Settings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
E.3 3D OpenFOAM Settings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
F Fourier analysis of a force signal 115
F.J. Venneman MSc. Thesis
List of Figures
1.1 Three dierent types of mavs . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1 Schematic illustration of the leading-edge vortex on fruit-y wings. . . . . . . 8
2.2 Inertial frame and rotating frame of reference . . . . . . . . . . . . . . . . . . 9
2.3 Smoke visualization of the ow around a tethered hawkmoth . . . . . . . . . . 11
2.4 Time courses of the vertical lift forces over a apping cycle. . . . . . . . . . . 14
2.5 Comparison of time-history of computed lift with experiment . . . . . . . . . . 14
2.6 Lift and drag coecients vs. sweeping angle for model wings . . . . . . . . . 15
2.7 Time-histories of model of plate nite elements . . . . . . . . . . . . . . . . . 16
3.1 Experimental setup and important angles. . . . . . . . . . . . . . . . . . . . . 18
3.2 Rotation, angular velocity and angular acceleration . . . . . . . . . . . . . . . 19
3.3 Top view of the experimental setup . . . . . . . . . . . . . . . . . . . . . . . 21
3.4 Isovorticity in six consecutive visualizations . . . . . . . . . . . . . . . . . . . 22
4.1 Control volume determining integral aerodynamic forces. . . . . . . . . . . . . 24
4.2 Forces at the beginning of the stroke of an impulsively-started wing . . . . . . 25
4.3 Force histories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.4 Simplied overview of the grid used to calculate the pressure-eld. . . . . . . . 29
4.5 Control volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.6 Results for the Blasius approach . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.7 Timewise cross-sections of the pressure, t = 0.1 0.8 s . . . . . . . . . . . . 36
4.8 Timewise cross-sections of the pressure, t = 0.9 2.5 s . . . . . . . . . . . . 37
4.9 Planewise cross-sections of the pressure, 0.38R 0.66R . . . . . . . . . . . . 38
4.10 Planewise cross-sections of the pressure, 0.70R 0.98R . . . . . . . . . . . . 39
4.11 Contributions to the force of dierent terms Navier-Stokes equations . . . . . 40
4.12 Forces for dierent integration contours . . . . . . . . . . . . . . . . . . . . . 41
4.13 Result of the force determination on the impulsively-started wing. . . . . . . . 42
MSc. Thesis F.J. Venneman
xiv List of Figures
5.1 Most important aspects of the two-dimensional grid of 46k cells . . . . . . . . 49
5.2 Boundary conditions of the two-dimensional case . . . . . . . . . . . . . . . . 51
5.3 Transient force and a phase for Re = 192, Co
max
= 1 and 46k cells . . . . . . 52
5.4 C
L
and C
D
trajectories in time for 8 chord lengths of travel . . . . . . . . . . 55
5.5 Vorticity plots for the two-dimensional impulsively-started wing. . . . . . . . . 55
5.6 The force-coecients plotted as a function of angle of attack. . . . . . . . . . 56
5.7 Vorticity plot and phase plot for ow around a wing at = 50

. . . . . . . . 57
5.8 Re St relation for a translating wing at = 50

. . . . . . . . . . . . . . . 58
5.9 Most important aspects of the grid of 280k cells . . . . . . . . . . . . . . . . 59
5.10 Boundary conditions of the three-dimensional case . . . . . . . . . . . . . . . 61
5.11 Force-coecients of the measurement . . . . . . . . . . . . . . . . . . . . . . 62
5.12 Planforms of the dierent wings used for the simulations. . . . . . . . . . . . 63
5.13 Rotation, angular velocity and angular acceleration of the alternative kinematics 65
5.14 Comparison of experimental and computational force-coecients . . . . . . . 68
5.15 Iso-Q-surfaces of the ow around the cad-wing (530k mesh) . . . . . . . . . 70
5.16 Comparison of the pressure-elds . . . . . . . . . . . . . . . . . . . . . . . . 71
A.1 Control volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
A.2 Figure showing what forces act on what parts of the volume. . . . . . . . . . 93
B.1 Transforming velocity to a rotating frame of reference . . . . . . . . . . . . . 96
B.2 Roboy setup and angular velocity . . . . . . . . . . . . . . . . . . . . . . . 97
B.3 Angular velocity and angular velocity of the impulsively-started wing. . . . . . 99
C.1 A rectangular control volume whose size increases in time. . . . . . . . . . . . 104
D.1 Illustration to help explain the aspect ratio and non-orthogonality . . . . . . . 106
D.2 Skewness error on the face . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
D.3 Figures indicating the quality of the two-dimensional grid of 46k cells . . . . . 108
D.4 Histograms showing non-orthogonality and the skewness of 2D cells . . . . . . 109
F.1 Time-history and amplitude spectrum of C
L
. . . . . . . . . . . . . . . . . . 115
F.J. Venneman MSc. Thesis
List of Tables
5.1 C
D
for a quiescent wing in a steady ow. Re = 192 . . . . . . . . . . . . . . 54
5.2 C
L
for a quiescent wing in a steady ow. Re = 192 . . . . . . . . . . . . . . 54
5.3 Aspect ratios for dierent wings . . . . . . . . . . . . . . . . . . . . . . . . . 64
A.1 Integration conventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
B.1 Coecients of exponential t with 95% condence bands. . . . . . . . . . . . 98
C.1 Radial Basis Functions with global support . . . . . . . . . . . . . . . . . . . 102
C.2 Most important parameters for the RBF mesh deformation tool. . . . . . . . . 103
D.1 Indicators of the quality of the two dimensional mesh. . . . . . . . . . . . . . 107
D.2 Indicators of the quality of the 3D meshes at a maximum deformation. . . . . 110
E.1 Linear solvers and their tolerance used in the two-dimensional case . . . . . . 112
E.2 Numerical schemes used in the two-dimensional and three-dimensional case. . . 113
E.3 Linear solvers and their tolerance used in the three-dimensional case . . . . . . 113
MSc. Thesis F.J. Venneman
xvi List of Tables
F.J. Venneman MSc. Thesis
Chapter 1
Introduction
1.1 Goal and context of present study
The design of Micro Air Vehicles (mavs) is currently an area of rapid growth. The term
mav generally refers to a new type of remotely controlled aircraft with a target dimension
of approximately 15 cm. Development of insect-size aircraft is expected in the near future.
Use in hazardous environment and enabling indoor ight are the driving factors of its
development. Ultimately, mavs should be able to y independently, with sensors and
ight control instrumentation on board to y around obstacles. Three types of mavs are
under investigation. Airplane-like xed wing models (Figure 1.1(a)), helicopter-like rotary
wing models (Figure 1.1(b)) and bird- or insect- like ornithopter (apping wing) models
(Figure 1.1(c)).
In this study we focus on the insect wing model. The physics of apping insect
wings has been studied extensively, in order to understand biological ight as well as to
improve mav design. Ellington et al. (1996) and Dickinson et al. (1999) showed that in
insect ight, exotic lift generating mechanisms, originating from vorticity generation, are
dominating.
The Delft University of Technology is an acknowledged player in the development of
mavs. Recently the DelFly (Figure 1.1(c)) and the Roboswift (in corporation with
Wageningen University) were developed with great success. Nevertheless, the physical
phenomena that produce the forces that keep these mavs aloft are still not understood
thoroughly. Both experiments and numerical simulations have been performed to see
which parameters inuence force production, but only incidentally the results were
compared quantitatively. Therefore the goal of this study is to determine forces on
a specic test case (an impulsively-started revolving wing for which an experimental
oweld database is provided) and compare the results of the experimental and numerical
work in a quantitative manner.
MSc. Thesis F.J. Venneman
2 Introduction
(a) Fixed wing model (b) Rotary wing model (c) Flapping insect
wing model
Figure 1.1: Three dierent types of mavs. Sources: (a) AeroVironment Inc. (b) Interactive
Toy Concepts (c) DelFly team
1.2 MAVs and sustainability
In order to assess the sustainable character of mav development, a clear denition of
sustainability is needed. Multiple of these denitions are provided, but the best known is
the one stated by the Brundtland Commission, convened by the United Nations in 1983.
It states:
Sustainable development is development that meets the needs of the present
without compromising the ability of future generations to meet their own
needs.
With background knowledge about both mavs and sustainable development, it is possible
to show that the development of mavs supports sustainability. In the next paragraphs a
closer look will be taken at the following areas of application of mavs:
Crop inspection
Disaster detection
Wildre detection
Weather monitoring
Environmental monitoring
Crop inspection
mavs can be very useful for farmers. They can inspect crops and detect insect epidemics
in an early stage. Flying through the trees or the bushes, they can easily detect locusts
or harm-causing beetles. As populations of these insects often grow exponentially it is
F.J. Venneman MSc. Thesis
1.2 MAVs and sustainability 3
advantageous to get rid of them in an early stage. This can be either by a very small
amount of insecticides or a predator that feeds on the harmful insects. So the mavs can
both minimize the damage as well as the amounts of anti-products. This would certainly
result in higher yields of the elds. Besides ying above crops outside, they can also
perform inspection in greenhouses (by using their unique ability to y slowly or hover
inside buildings).
Disaster detection
mavs are extremely applicable in hazardous environments. This makes them the ideal
machines to detect disasters and send an early warning to the authorities who can try to
minimize the results of the disaster. Two examples will be given below: wildre detection
and weather monitoring.
Wildre detection mavs can easily survey a forest with an infra-red camera on
board in order to detect heat. Not only by ying over the woods, but especially by
ying through the woods (remember that they will be able to y independently).
This would permit it to detect wildres in a very early stage. Once a re is de-
tected, they can immediately warn the local re brigade which is able to extinguish
the re when the re is still on a small scale. This application is in line with the
Brundtland denition, because future generations can benet from the forests, both
for recreational purpose and storage of carbon dioxide.
Weather monitoring In general mavs can easily monitor the weather. Besides
sensing and storing information, they can be used as an early warner in areas with
high chance of sudden change in atmospheric conditions. For instance heavy rains
(with risk of ooding) or tornados (with risk of damage on buildings or crops) can be
detected well in advance. Knowing that a sudden atmospheric change is coming up,
people can protect there buildings or crops, in order to minimize disastrous eects.
A lot of weather monitoring equipment already exists. However mavs will combine
their low costs, maneuverability and light-weight properties to be very competitive
in this eld of application.
Environmental monitoring
One of the fears is that in future wars, biological and chemical weapons will be used.
Employing mavs to track down traces means that contaminated areas can be explored
without risking human lives. The minuscule size of the mav would allow it to be transported
quickly to areas to detect the presence of gases and test the safety of the environment for
human operations.
MSc. Thesis F.J. Venneman
4 Introduction
1.3 Brief background
Flapping ight has inspired people for ages. Leonardo da Vinci was the rst person who
considered to mimic biological wing motion, which led to the well-known Ornithopter
design. In the 20
th
century people started developing analytical models as a rst
approximation of the forces that enable ight. It soon became clear that these classical
wing theory models were insucient for the description of complex, apping ight, as they
did not account for unsteady eects, so experiments were performed to visualize the ow
around wings. Once computers were able to solve the Navier-Stokes equation in a time
ecient manner, ow around wings could also be simulated numerically. In this section a
brief history of the analytical, experimental and numerical work will be presented. For a
more detailed description about the experimental and numerical see sections 2.2 and 2.3,
respectively.
Analytical
It has proven to be almost impossible to capture ight performance of insects in an
analytical way. Up to today there exists no satisfactory equation that predicts ight
performance. The rst well known attempt to relate forces with parameters like instan-
taneous velocity, wing geometry and angle of attack was done by Weis-Fogh and Jensen
(1956). In this study the quasi-steady theory was constructed, which greatly simplies
the time-dependent problem by converting it to a sequence of independent, steady-state
problems. Therefore it neglects wing motion and ow history and predicts unreliable
forces. Sane and Dickinson (2002) tried to improve this theory by including some
rotational eects but even then the results remain questionable.
Experimental
To fully understand what is happening in insect ight, many experiments have been
performed in the past. For example hawkmoths were tethered and smoke was used
to visualize the ow pattern (Srygley and Thomas, 2002). Besides that Particle Image
Velocimetry (piv) was used to visualize the ow around a dynamically-scaled robotic
wing that mimics insect ight (Dickinson et al., 1999). Although piv is becoming more
and more accurate, it is still very dicult to capture all the relevant aspects, notably
small-scale ow phenomena as well as the ow very close to solid surfaces.
Numerical
Since computers become powerful enough to solve the Navier-Stokes equation on grids
with sucient resolution, researchers have tried to simulate insect wing performance.
Wang (2000b) for example investigated simple single translational motion with low
amplitude. Bos (2005) studied the inuence of dierent wing kinematic models (including
rotational eects) of hovering insects. These two investigations were both performed in
two dimensions, in the last decade numerical simulations were also performed in three
dimensions, which require a substantially greater computational eort.
F.J. Venneman MSc. Thesis
1.4 Objectives and approach of present study 5
1.4 Objectives and approach of present study
As already mentioned in the previous section, researchers have tried to study insect wing
aerodynamics in various ways. Only a few have succeeded to compare experimental with
numerical work in a transparent manner. Therefore the objective of the present study is
to compare forces deduced from experiments and computations on a specic kinematic
test case, the impulsive start.
Experimentally obtained velocity-elds around this wing are provided by Dr. Ir.
C. Poelma. He performed time-resolved, stereoscopic piv measurements around a
dynamically-scaled wing, known as the Roboy (Poelma et al., 2006).
Besides force acquisition from the force sensor mounted on the Roboy-wing, the
present study applies a new tool to obtain forces, through a combination of the momen-
tum approach and an algorithm to obtain pressure-elds from velocity-elds, called the
planar Poisson approach. This method deduces the pressure-elds from the velocity-elds
by making use of the ow constitutive equations.
For the computational simulations, the kinematics of the Roboy-wing are mim-
icked as accurately as possible. The Navier-Stokes equations are solved using OpenFOAM
as a framework. The incompressible ow equations are solved on a dynamically deforming
mesh. This mesh deforms based on Radial Basis Function interpolation, which is a new
mesh deformation tool developed at the Aerodynamics group of the Delft University of
Technology.
MSc. Thesis F.J. Venneman
6 Introduction
F.J. Venneman MSc. Thesis
Chapter 2
Flapping insect wing aerodynamics and
problem denition
The dazzling physics of ow around bird and insect wings has inspired scientists for ages.
In this thesis the ow around the wing of a common insect is studied, namely the fruit-
y (or Drosophila melanogaster). In this chapter a small overview of recent insect wing
studies will be provided. But rst the basic physics of apping wing aerodynamics will be
discussed in section 2.1. Since we are especially interested in forces on insect wing in this
study, section 2.2 describes dierent methods to obtain forces from velocity-elds around
insect wings. This is followed by an overview of the computational work on insect wing
aerodynamics, provided in section 2.3. In section 2.4 some scientic articles are described
in which computations and experimental work are compared. In the last section (2.5) the
problem denition of this study is stated.
2.1 Physics of apping insect wing aerodynamics
The ow around apping fruit-y wings can be considered incompressible since the Mach
number is extremely low; M 0.0015 < 0.03 (Bos, 2005). In physical transport phenomena
is dealt with conservation of especially mass and momentum. The conservation of mass is
stated in the incompressible continuity equation:
u = 0 , (2.1)
with u the ow velocity. The conservation of momentum is represented by the incompress-
ible Navier-Stokes equations (neglecting gravity):

u
t
+ (u ) u = p + . (2.2)
With t the time, the uid density, p the static ow pressure and the stress tensor.
The fact that insects are able to y, highly depends on their ability to create vortices
MSc. Thesis F.J. Venneman
8 Flapping insect wing aerodynamics and problem denition
around their wings. The kinematics of insect wings is applied in such a way, that they
make use of these vortices in a very ecient way to create lift and thrust. The leading-edge
vortex (or lev) is an excellent example of a lift creating vortex (a schematic illustration is
provided in Figure 2.1). This vortex (with its main component in the spanwise direction)
stays attached to the top of the wing during an upstroke or a downstroke.
Two dimensionless numbers are important in apping insect wing aerodynamics. The
Figure 2.1: Schematic illustration of the leading-edge vortex on the wings of a Drosophila
melanogaster (top view). The shaded area represents the area swept by the wing.
(Poelma et al., 2006)
rst one is the Reynolds number, which can be interpreted as the ratio between the inertial
force and the viscous force. In an unsteady case where a body is excited it is also possible
to describe the Reynolds number as the relation between two time scales, the time-scale to
dissipate a vortex viscously and the time scale needed for convective transport. For ow
around airfoils the Reynolds number is dened as:
Re =
cU

, (2.3)
where c is the typical chord length, U the typical velocity and the kinematic viscosity. In
two dimensions the chord length, c, is evident and the velocity U is the incoming velocity
(for a quiescent wing). In three dimensions the chord length c is the maximum chord
length over the wing and U is the velocity of the wing tip U
tip
. In scientic literature the
typical chord length and velocity are not consistently dened in this way, sometimes the
average c and U are taken. Be aware of this in comparing results obtained in this thesis
with other results. In general, insect ight takes place in the intermediate Reynolds regime
(10 < Re < 1000).
Another important dimensionless parameter is the Strouhal number, which describes the
oscillating frequency of a ow. It is dened as:
St =
fc
U
, (2.4)
F.J. Venneman MSc. Thesis
2.1 Physics of apping insect wing aerodynamics 9
where f is the frequency of the ow. In two-dimensional ow this is the vortex shedding
frequency.
The majority of ow problems are dened and solved in a xed, or inertial, coordinate
system. However, in this thesis, a case arises where we wish to use non-inertial coordinates,
moving with a rotational system. Therefore we must modify the governing equations, in
which the acceleration term is valid only if the absolute acceleration of the particles is
relative to the inertial coordinates. In Figure 2.2 the XY Z system is xed in an inertial
frame and the xyz system rotates relative to it with angular velocity .
Y
Z
X
y
x
z
r
P
Figure 2.2: Inertial frame XY Z and rotating frame xyz. In the present study the rotating
frame is also rotating around the y-axis, just as depicted in this gure.
The expression for the absolute acceleration of point P in Figure 2.2 is (Greenwood, 1988):

2
r
P
t
2
=
u
P
t
=
u
P,r
t
+

t
r
P,r
+

r
P,r
_
+ 2

u
P,r
, (2.5)
where the subindex r refers to the coordinates in the rotational frame and

is the angular
velocity. The rst term on the right hand side of this equation is the acceleration relative
to the xyz frame (viewed by the rotating observer). The second and third term together
represent the acceleration of P relative to the origin as viewed by the non-rotating observer.
The second term represents the tangential acceleration and the third term represents the
centripetal acceleration (it points outward from P). The fourth term is known as the
Coriolis term, which is due to the changing direction in space of the velocity of P relative
to the moving system.
Once we substitute the absolute acceleration (2.5) to the Navier-Stokes equations (2.2) we
obtain:

_
u
t
+ (u ) u
_
= p +
_

t
r +

r
_
+ 2

u
_
. (2.6)
MSc. Thesis F.J. Venneman
10 Flapping insect wing aerodynamics and problem denition
Mind that the velocity u and displacement term r are now relative to the rotating frame
of reference (u = u
P,r
, r = r
P,r
). Special notice should be taken of the term

t
, which is
often neglected in scientic literature because the angular velocity is constant. Nevertheless
it will turn out to be very important in the test case of this thesis. According to the
constitutive equation for an incompressible Newtonian uid (i.e., =
2
u), equation
(2.6) is equal to:

_
u
t
+ (u ) u
_
= p +
2
u
_

t
r +

r
_
+ 2

u
_
. (2.7)
2.2 Experimental investigation of insect ight
To fully understand what is happening in insect ight, many experiments have been per-
formed in the past. For example hawkmoths were tethered and smoke was used to visualize
the ow pattern (Ellington et al., 1996; Srygley and Thomas, 2002) (see Figure 2.3). The
most important feature considering apping insect ight found from these studies, was the
appearance of a leading-edge vortex (lev). Apart from direct observations of live animal
species, many experiments have been carried out that simulate (insect) ight with me-
chanical means using simplied congurations. Many researchers have tried to mimic the
kinematics and planform of insect wings and measure the forces on these wings. In the next
two paragraphs, examples of measurements on insect wings in two and three dimensions
are shown.
Dickinson and Gotz (1993) studied the aerodynamic forces on a two-dimensional
impulsively-started wing. That study couples measurements of lift and drag on a two-
dimensional at plate with simultaneous ow visualizations. The Reynolds number, angle
of attack, camber and roughness of the wing are varied in an independent manner. The
main conclusion is that the unsteady process of vortex generation at large angles of attack
may contribute to the production of aerodynamic forces in insect ight. Especially inter-
esting in the scope of this thesis is that impulsive movement resulted in the production of
a leading-edge vortex that stayed attached to the wing the rst 2 chord lengths of travel,
resulting in 80% increase in lift compared to the performance measured 5 chord lengths
later.
By Dickinson et al. (1999), the aerodynamic forces on a three-dimensional Drosophila wing
are studied. This is done by dynamically-scaled robotic wings, known as the Roboy (see
further information in 3.1. Three interacting mechanisms are distinguished: delayed stall,
rotational circulation and wake capture. Delayed stall functions during the translational
portions of the stroke, when the wing sweeps through the air with a large angle of attack.
It causes the formation of a leading-edge vortex that reduces pressure over the wing. Ro-
tational circulation and wake capture generate aerodynamic forces during stroke reversals,
when the wings rapidly rotate and change in direction. Rotational circulation causes lift
when the insect rotates the angle of attack, increasing the speed on the top side of the
wing relative to the bottom, creating a lower pressure on the top. Wake capture refers
F.J. Venneman MSc. Thesis
2.3 Numerical investigation of insect ight 11
Figure 2.3: Smoke visualization of the ow around a tethered hawkmoth late in the down-
stroke (smoke injection at half the span of the wing, airspeed is 3.7 m s
1
). The ow
separates at the leading-edge and reattaches to the upper surface in the posterior half of the
wing, enclosing a leading-edge vortex (lev) (Ellington et al., 1996).
to the principal that a wing benets from the shed vorticity of the previous stroke. The
ow generated by one stroke can increase the eective uid velocity at the start of the
next stroke and thereby increase force productions above that which could be explained
by translation alone.
2.3 Numerical investigation of insect ight
Simulation of low Reynolds number ow around airfoil structures has been going on for
years. In this research we are focusing on insect wings. Because of the computational eort
required for realistic (3D) simulations, most simulations were performed for a simplied
geometry in two dimensions (see section 2.3.1), however, attempts for simulations in three
dimensions were also performed in the last decade (see section 2.3.2).
2.3.1 2D Simulation
In the two-dimensional case, ow around a cross-section of a single wing is simulated. In
relevance to insect ight two dierent classes of wing simulations may be distinguished: a
apping rigid wing and a exible non-apping wing.
Fixed exible wing
Insect wings have a high aspect ratio (large spanwise length with respect to the
surface area), which are easy to twist and turn. In order to simulate such a wing, it may
be argued that the wing behavior caused by interaction with the airow (Fluid-Structure
Interactions) must be analyzed. In this paragraph a study on exible membrane wing
airfoils is shown, performed by Gordnier (2008). A well-validated, robust Navier-Stokes
MSc. Thesis F.J. Venneman
12 Flapping insect wing aerodynamics and problem denition
solver is employed coupled with a membrane structural model suitable for the highly
nonlinear structural response of the membrane. A low Reynolds number, Re = 2500,
consistent with mav ight is chosen for the majority of the calculations. The most notable
eect of the membrane exibility is the introduction of a mean camber to the membrane
airfoil. A close coupling between unsteady vortex shedding and the dynamic structural
response is demonstrated. The dynamic motion of the membrane surface is also shown to
signicantly alter the unsteady ow over the membrane airfoil at high angles of attack.
The coupling of this dynamic eect and the mean camber results in a delay in stall with
enhanced lift and reduced drag for higher angles of attack. Exploratory computations
investigating the eects of angle of attack, membrane rigidity, membrane pretension and
Reynolds number on the membrane airfoil response are also presented.
Rigid apping wing
In (Bos et al., 2008), the inuence of the dierent wing kinematics models on the
aerodynamic performance of a hovering insect is investigated by means of time-dependent
Navier-Stokes simulations. With increasing complexity, a harmonic model, a Roboy
model and two more realistic fruit-y models are considered, all dynamically scaled at
Re = 110. Details of the vortex dynamics, as well as the resulting lift and drag forces,
were studied. The simulation results reveal that the fruit-y wing kinematics result in
forces that dier signicantly from those resulting from the simplied wing kinematic
models.
In Lentink (2003) a apping wing is modeled with a sinusoidal plunging airfoil. The
corresponding unsteady, incompressible Navier-Stokes equations are solved with respect
to a body-xed coordinate system. This non-inertial coordinate system introduces a body
force. The equations are solved with a validated and veried Navier-Stokes solver for
Reynolds number 150 and Strouhal number 0.25. Two airfoils with dierent shapes are
compared. It was found that the sub-critical airfoils outperform the supercritical airfoils
and that aft camber is highly important for good performance in insect ight.
Wang (2000a) showed that a two-dimensional hovering motion is able to generate enough
lift to support a typical insect weight. That computation reveals a mechanism of creating
a downward dipole jet of counteracting vortices, which are from leading and trailing-edge
vortices. The vortex dynamics further elucidates the role of the phase relation between
the wing translation and rotation in lift generation and explains why the instantaneous
forces can reach a periodic state after only a few strokes.
In (Wang, 2000b) a computational tool is devised to solve the Navier-Stokes equations
around a moving wing, which mimics biological locomotion. The focus of the work is
frequency selection in forward apping ight. Besides that time scales associated with the
shedding of the trailing-edge vortex and leading-edge vortex, as well as the corresponding
time-dependent forces are investigated.
F.J. Venneman MSc. Thesis
2.3 Numerical investigation of insect ight 13
2.3.2 3D Simulation
Pure insect ight is inherently three-dimensional, so a pertinent question is the pertinence
of two-dimensional simulations and the impact of three-dimensional geometry on the
ow eects. This section describes simulations of three-dimensional apping insect ight.
Since a 3D simulation (in contrast to a 2D simulation) also accounts for span wise ow,
comparison with measurements is more relevant. The rst two articles compare their
ndings with measured forces on a dynamically scaled robotic model (Fry et al., 2003,
2005).
In Aono et al. (2008) an integrative computational uid dynamics study of near- and
far-eld aerodynamics in insect hovering ight using a biology-inspired, dynamic ight
simulator is presented. This simulator, which has been built to encompass multiple
mechanisms and principles related to insect ight, is capable of ying an insect on the
basis of realistic wing-body morphologies and kinematics. The cfd study integrates
near- and far-eld wake dynamics and shows the detailed three-dimensional near- and
far-eld vortex ows: a horseshoe-shaped vortex (just as in Poelma et al. (2006)) is
generated and wraps around the wing in the early down- and upstroke; subsequently,
the horseshoe-shaped vortex grows into a doughnut-shaped vortex ring, with an intense
jet-stream present in its core, forming the downwash; and eventually, the doughnut-shaped
vortex rings of the wing pair break up into two circular vortex rings in the wake.
The computed aerodynamic forces show reasonable agreement with experimental results
(measured by Fry et al. (2005)) in terms of both the mean force (vertical, horizontal and
sideslip forces) and the time course over one stroke cycle (lift and drag forces). A large
amount of lift force (approximately 62% of total lift force generated over a full wingbeat
cycle) is generated during the upstroke, most likely due to the presence of intensive and
stable, leading-edge vortices and wing tip vortices. See Figure 2.4 for an example of the
obtained forces and a comparison with measurements.
In Ramamurti and Sandberg (2007), three-dimensional unsteady computations of
the ow past a fruit-y under hovering and free ight conditions are computed. A nite
element ow solver was employed to compute unsteady ow past a fruit-y body. To carry
out computations of the ow about oscillating geometries, the moving surface is coupled to
the volume grid. The volume grid in the proximity of the moving surface is then remeshed
every time step to eliminate badly distorted elements (Ramamurti and Sandberg, 2002).
The kinematics of the wings and the body of the fruit-y are prescribed from experimental
observations. The computed unsteady lift and thrust forces are validated with experimen-
tal results and are in good agreement (see Figure 2.5). The unsteady aerodynamic origin
of the time-varying yaw moment is identied. The dierences in the kinematics between
the right and the left wings show that a subtle change in the stroke angle and deviation
angle can result in the yaw moment for the turning maneuver. This investigation leads to
the conclusion that it is the forward force and a component of the lift force that combine
to produce the turning moment while the side force alone produces the restoring torque
during the maneuver.
MSc. Thesis F.J. Venneman
14 Flapping insect wing aerodynamics and problem denition
Figure 2.4: Time courses of the vertical lift forces over a apping cycle. Blue, red and yellow
lines represent the measurements of the upper (Exp-u), average (Exp-a) and lower (Exp-l)
values obtained by Fry et al. (2005), respectively; the broken line is the computed result
(Com-a). T is the dimensionless period of one apping cycle. (Aono et al., 2008)
In Luo and Sun (2005) the eects of wing planform (shape and aspect ratio) on the
Figure 2.5: Comparison of time-history of lift (L) forces computed in
Ramamurti and Sandberg (2007) and experiments described in Fry et al. (2003). Gray and
white bars indicate downstroke and upstroke, respectively.
aerodynamic force production of model insect wings in impulsively-started sweeping
motion at Reynolds number 200 at angle of attack of 40

are investigated. Wing-shape


and aspect ratio of ten representative insect wings are considered, amongst which the
fruit-y. A time-history of obtained force-coecients is depicted in Figure 2.6, in which
typical peaks are shown followed by a constant force production for both lift and drag. It is
concluded that the variation in wing shape has only minor eects on the force-coecients.
This also counts for the aspect ratio. Increasing the aspect ratio, on one hand, increases
F.J. Venneman MSc. Thesis
2.4 Comparison of experiments and computations 15
the force due to reduction of three-dimensional ows; on the other hand, they will
be decreased due to shedding of part of the leading-edge-vortex. Luo and Sun (2005)
conclude that these two eects approximately cancel each other, resulting in only minor
changes of the force-coecients.
Figure 2.6: Lift and drag coecients vs. sweeping angle for model wings with various shapes,
but the same aspect ratio ( = 40

; Re 200). (Luo and Sun, 2005)


2.4 Comparison of experiments and computations
In the previous section, research was described in which numerical modeling of an insect
wing was performed. In most cases, the results were validated with experimental work
((Fry et al., 2003, 2005), (Dickinson et al., 1999)). Some researchers also combined
experiment and computation. Two articles describing a comparison of experiments and
computation are described below.
Wang et al. (2004) compared 2D computational, 3D experimental and quasi-steady
forces in a hovering wing undergoing sinusoidal motion along a horizontal stroke plane for
Re 100, unsteady eects of this motion are investigated. In all cases the drag compares
well, but the lift only compares for distinct kinds of rotation.
The rst conclusion of that investigation is a weak dependence of the stroke angle on the
force-coecients. Secondly the forces are very sensitive to the phase between the stroke
angle and the angle of attack. It was also found that the main dierence between a 3D
revolving wing and a 2D translating wing is the absence of vortex shedding by a revolving
wing over a distance much longer than the typical stroke length of insects.
Singh and Chopra (2008) validate a model of plate nite elements with 3D measurements
on the Roboy-wing. The predicted vertical force was compared throughout one apping
MSc. Thesis F.J. Venneman
16 Flapping insect wing aerodynamics and problem denition
cycle (see Figure 2.7). As can be seen the two signals do not correspond very well, i.e. the
thrust near the end of the downstroke (t 0.4) and upstroke (t 0.9) is overpredicted to
a large extent.
Figure 2.7: Time histories of the vertical force from Roboy experimental data and results
of the model of plate nite elements. (Singh and Chopra, 2008).
2.5 Problem denition
In this chapter the basics of the aerodynamics of insect ight have been shown. A major
interest in studies on insect ight is to study the eect of kinematic modeling on the
time-histories of the force. Dynamically scaled robotic insect wings were used in order
to experimentally test this. Besides experimental work, a lot of numerical investigations
concerning (apping) insect wings have been performed.
The Aerodynamics group at the Delft University of Technology has developed two tools
that are very useful in the analysis of insect wing aerodynamics:
1. A procedure to obtain forces from velocity-eld data, intended to be applied under
experimental conditions (see detailed information in section 4.2, (Gurka et al., 1999)).
2. A mesh deformation tool based on Radial Basis Function interpolation, to eciently
calculate ows around a moving body (see detailed information in Appendix C,
(Boer et al., 2007))
These tools can be combined to put experimental and computational work on insect wing
aerodynamics together in a quantitative comparison. A piv dataset of velocity-elds
around a dynamically scaled robotic wing (Roboy) was available from the research of
Poelma et al. (2006) carried out at CalTech Laboratories. The procedure to derive forces
from velocity will be applied to this dataset. These forces will then be validated with the
results of a numerical investigation using the same kinematics. A ow solver written within
the open-source framework of OpenFOAM will be used for this purpose.
F.J. Venneman MSc. Thesis
Chapter 3
Experimental test case description
In this thesis the forces on insect wings are investigated experimentally and numerically.
It was decided to emphasize on a specic motion, the impulsive start, for which an ex-
perimental data base was available. These data were obtained from a dynamically-scaled
robotic wing moving in mineral oil. This robotic wing is better known as the Roboy,
of which the experimental setup is described in section 3.1. The impulsive motion of the
Roboy is based on the kinematics derived from the real Drosophila Melanogaster (fruit-
y), as described in section 3.2. Measurements of the velocity-eld around this scaled wing
were performed with stereoscopic piv, this technique is explained in section 3.3 followed
by the resulting ow eld, shown in section 3.4. As already mentioned, the original ex-
periments were performed by Dr. ir. C. Poelma. For a more thorough explanation of the
experimental setup therefore see Poelma et al. (2006).
3.1 Experimental setup
As allready described in the introduction of this chapter, a dynamically-scaled robotic
wing, known as the Roboy, has been used to mimic the wing motion of a fruit-y. This
wing was suspended in a large (11.53 m
3
) rectangular tank. Using computer-controlled
servo motors, the wing position and motion were accurately controlled with respect to four
degrees of freedom. For this thesis only the angle of attack () and the rotation angle ()
are important (see Figure 3.1 for the denition of these angles).
The wing was cut in the planform shape of a Drosophila wing from a 2.25-mm thick acrylic
sheet (see Figure 3.1). The maximum chord length (c) of the wing is 10 cm. The distance
(L) from the point-of-rotation (o) to the wing tip is 25 cm, but the rst 7 cm were taken
up by a gear box and a force sensor. This force sensor could measure forces perpendicular
and parallel to the wing during the uid velocity measurements. The mineral oil that was
used, had a density () of 880 kg m
3
and a kinematic viscosity () of 115 10
6
m
2
s
1
.
After acceleration the wing moved at a typical angular velocity () of (3/8) rad s
1
,
equivalent to a wing-tip velocity (U
tip
) of 0.29 m s
1
. This leads to a maximum chord
MSc. Thesis F.J. Venneman
18 Experimental test case description
Figure 3.1: Picture with the experimental setup and the important angles in this study.
Source: Christian Poelma
length based Reynolds number similar to that of a real Drosophila:
Re =
c
max
U
tip

= 256 . (3.1)
3.2 Kinematics of the wing
In Figure 3.2 the kinematics of the impulsively-started wing are shown. In Figure 3.2(a)
the elapsed rotation angle, , is plotted versus time, these data were provided by Dr. Ir.
C. Poelma. This is followed by Figure 3.2(b), in which the angular velocity versus time is
plotted. The blue crosses indicate the results obtained from numerical dierentiation of the
data in Figure 3.2(a). Since the time resolution of the data is low ( 0.1 s, see for further
information section 3.3 and Figure 3.3), the derivative is not very accurate. Therefore an
exponential t was used for the angular velocity (red lines in 3.2(b)). The exponential t
is of a form prescribed by Poelma et al. (2006, eqn. 1):
(t) =
max
_
1 exp
_

t
t
c
__
, (3.2)
with the maximum angular velocity,
max
= 1.2 rad s
1
and the characteristic time t
c
=
0.17 s (i.e., the wing reached 63% of its maximum velocity in the rst 0.17 s). The time
derivative of this t was used for the the angular acceleration (see Figure 3.2(c)). This
angular acceleration was needed for later pressure and force calculation, as described in
section 4.2. An explanation how the t was obtained is provided in Appendix B.3.
F.J. Venneman MSc. Thesis
3.2 Kinematics of the wing 19
time (s)

(
r
a
d
)
0 0.5 1 1.5 2 2.5 3
0
0.5
1
1.5
2
2.5
3
3.5
(a) Rotation
time (s)

(
r
a
d
s

1
)
0 0.5 1 1.5 2 2.5 3
0
0.2
0.4
0.6
0.8
1
1.2
(b) Angular velocity
time (s)

t
(
r
a
d
s

2
)
0 0.5 1 1.5 2 2.5 3
-1
0
1
2
3
4
5
6
7
8
(c) Angular acceleration
Figure 3.2: Rotation, angular velocity and angular acceleration of the impulsively-started
wing. Blue crosses indicate the (derivatives of the) data provided by Dr. Ir. C. Poelma, red
lines indicate actually used values derived from the angular velocity (that was tted, see
Appendix B.3).
MSc. Thesis F.J. Venneman
20 Experimental test case description
3.3 PIV measurement technique
Particle Image Velocimetry (piv) was used to obtain velocity-elds surrounding the robotic
wing. piv is an optical method in which the uid is seeded with tracer particles which are
generally assumed to follow the ow dynamics. It is the motion of these seeding particles
that is used to calculate velocity information of the ow being studied. Since this part of
the thesis is about the post-processing of the velocity-elds, only a very brief overview of
the measurement technique will be given here. For further information see Poelma et al.
(2006). The local uid velocity was measured by means of stereoscopic piv. Stereoscopic
means that two cameras are used in order to extract the velocity information in three
dimensions. A 25 25 cm
2
eld-of-view (fov) was recorded using two Imager Intense
cameras (1, 376 1, 024 pixel resolution, LaVision).
Silver-coated hollow glass spheres with a mean diameter of 13 m were used as tracer
material. Due to the high viscosity of the uid, the particles could accurately follow all
motions of the uid (their settling velocity was negligible 0.7 m s
1
).
As a light source, an Nd:YAG laser (120 mJ pulse
1
at 532 nm) was used; the light sheet
was approximately 300 mm high and 2 mm thick. The ow was captured by means of
a phase-averaging approach at various phases and at dierent spanwise sections of the
wing. This implies that the light sheet was kept on the same position and always parallel
to the chord of the wing at the time of measurement. Image pairs were recorded with a
laser pulse delay time of 5, 000 or 7, 500 s, depending on the ow pattern; the maximum
tracer displacement was always kept below 8 pixels; a higher delay time led to an increased
error due to out-of-plane pair loss. For the processing, a three-pass cross-correlation was
chosen (one pass at 64 64, two at 32 32 pixels) using a straightforward fft-based
cross-correlation algorithm (DaVis 7.0, LaVision GmbH).
In some images reections of the wing or the gear box were visible, which led to spurious
vectors due to parallax eects. These spurious vectors were very distinct, as they were
often nearly an order of magnitude larger than the surrounding vectors. The reections
were removed by subtracting an averaged image using 32 recordings. Ideally, an image
with only the reections and no tracer particles should have been used for this. This was
nearly impossible, since every image also contains the light scattered by the tracer parti-
cles. The averaged image contained the reections, yet also weakened images of the tracer
(approximately 1/32 of their mean intensity). This meant that subtracting the average
image reduces the image quality somewhat.
Spurious vectors were detected in the post-processing step of the nal ensemble of averaged
vector elds, by means of a local median test; a threshold value of 1.5 times the local stan-
dard deviation was chosen, and spurious vectors were replaced using bilinear interpolation.
Due to the ensemble averaging, typically less than 2% of the vectors needed to be replaced
in the ensemble result. For comparison, if individual image pairs were processed (i.e., no
ensemble averaging), less than 5% of the vectors needed to be replaced.
The velocity-eld around the impulsively-started wing was measured in 16 time steps.
These time steps are depicted in Figure 3.3. At t = 1.5 s, the wing is shown with 18
cross-sections (measurement planes), all 1 cm apart from each other. 2 other measurement
F.J. Venneman MSc. Thesis
3.4 Floweld around the wing 21
Figure 3.3: Top view of the experimental setup. The 16 time steps are shown with their times
(in seconds) at the outside of the semi-circle. These time/angle combinations correspond
with Figure 3.2(a). At t = 1.5 s, the wing is shown with 18 cross-sections (measurement
planes). The 2 other measurement planes are taken behind the wing tip.
planes were taken behind the wing tip. The total fov of one measurement plane is 3324
cm in size. There are 81 (N
i
) datapoints in the x-direction and 60 (N
j
) datapoints in the
y-direction, so the total eld consists of 4860 datapoints in total per fov. The datapoints
in the fov are 4.08 mm apart from each other.
3.4 Floweld around the wing
Although our main focus is on the force development, some information on the ow eld
structure is presented in this section to illustrate its character and evolution in time. In
Figure 3.4, six consecutive panels are shown during the impulsive start. The view point
is from the center of rotation. Each panel shows isosurfaces of spanwise vorticity: blue
and red represent clock-wise (
z
= 15 s
1
) and counter-clockwise motion (
z
= 10 s
1
),
respectively. The vector in the rst frame indicates the direction of the displacement of
the wing. The outline of the wing is shown by the black line. Also indicated is a schematic
representation of the position in the impulsive start (red wing prole). In the rst frame
it is clearly shown that both a leading-edge vortex and a trailing-edge vortex develop. A
part of the trailing-edge vortex is shed in the third and fourth frame. In the stationairy
part of the stroke (frame ve and six) both a leading and a trailing-edge remain on the
surface of the wing.
MSc. Thesis F.J. Venneman
22 Experimental test case description
Figure 3.4: Composite of six consecutive visualizations (isovorticity in the spanwise direction)
of an impulsively-started wing. The vector in the rst frame indicates the direction of the
wing. See the text in section 3.4 for a description and denition of the labels.
F.J. Venneman MSc. Thesis
Chapter 4
Determining forces on insect wings
experimentally
The previous chapter described how the velocity-eld around the dynamically-scaled
robotic wing is obtained. In this chapter we are interested in obtaining the forces on
the wing. Besides measuring forces directly from force sensors mounted on the wing, it is
very challenging to indirectly obtain forces from the ow eld information. A number of
methods exist to obtain forces from velocity data, such as provided by piv. In scientic
literature, multiple approaches of evaluating forces from piv data are presented. Out of
these three approaches, two are applied in this thesis:
The Blasius approach. (section 4.1, for results see section 4.3.1)
The momentum approach (section 4.2, for results see section 4.3.2)
The results of the tested approaches are shown in section 4.3. The conclusions of this
chapter are stated in section 4.4.
These methods have in common that the force on an object is calculated using the mo-
mentum approach. In this approach a xed control volume V , boundary S and outward
pointing normal n is considered, enclosing an object as in Figure 4.1.
MSc. Thesis F.J. Venneman
24 Determining forces on insect wings experimentally
Figure 4.1: Control volume approach for determining integral aerodynamic forces in a two-
dimensional ow conguration. The black body indicates a cross-section of a wing. The
dashed line is the integration path.
4.1 The Blasius approach
According to the Blasius theorem, instantaneous uid dynamical forces can be related to
the time rate of change of the the rst moment of vorticity in an innite domain. In
Poelma et al. (2006), the following straight forward and exact equation, derived by Wu
(1981), is used to calculate the instantaneous force (an extensive derivation can be found
in Appendix A.2):

F(t) =
1
2

d
dt
+
d
dt
___
V
S
udV . (4.1)
The rst term in equation (4.1) accounts for the eect of the ow eld under consideration
on an immersed object. Where is the rst moment of vorticity and described by:
=
___
S
r dV , (4.2)
with , the vorticity and r the position vector enclosed in volume V .
The second term in equation (4.1) accounts for the change in momentum of the displaced
uid by the object (enclosed in volume V
S
, object velocity is u) itself. Since the wing
is thin and acceleration is negligible for the case under consideration, the second term in
equation (4.1) can be neglected in computing the total force.
In (Poelma et al., 2006), force measurements, using strain gauges, were compared to force
predictions extracted from piv data for the rst time. The results, see Figure 4.2, cor-
respond reasonably well for the initial stages of the impulsively started wing. However,
in later stages of the stroke, when vorticity that leaves the eld of view is not taken into
account anymore, the forces are seriously under predicted. Since we have obtained the
raw velocity data from Poelma et al. (2006), we could check this method. The results are
shown in section 4.3.1.
Birch and Dickinson (2003) also use the concept of the rst moment of vorticity to predict
F.J. Venneman MSc. Thesis
4.1 The Blasius approach 25
Figure 4.2: Forces at the beginning of the stroke of an impulsively-started wing (total stroke
takes 3 s). Continuous line and crosses represent the horizontal forces (drag), dashed line
and circles represent the vertical forces (lift). (Poelma et al., 2006)
forces. The same experimental setup as in Poelma et al. (2006) was used, but in that case
the forces are only compared relatively (see Figure 4.3, note that the units on the left axis
are N and on the right axis N m
1
).
Noca et al. (1999) has proposed another formulation that would permit to apply the
vorticity concept on a nite domain, including the eect of vorticity convection across the
domain edges. The method proves to be successful for large normalized force coecients
( 2 3). On the other hand, the authors warn for use of large domains or smaller force
coecients, which may cause convergence problems. Birch and Dickinson (2003) also tried
this approach, but concluded that it was not robust for their case. Since the measurements
used in this study are obtained from the same experimental setup, this method would not
be an obvious choice.
MSc. Thesis F.J. Venneman
26 Determining forces on insect wings experimentally
Figure 4.3: Predictions of equation (4.1) and measured forces give similar time histories at
the start of stroke one. The x-axis covers the rst 16% of stroke one. Blue traces (left-
hand y-axes) show measured lift and drag forces. Red circles (right-hand y-axes) plot values
of sectional lift and drag calculated from Blasius approach calculation. Predictions beyond

t = 0.15 were unreliable because starting vorticity moves out of the visualized frame. note
that the left and right axis have dierent units (N on the left axis and N m
1
on the right
axis), i.e. the total forces are compared with planar forces. (Birch and Dickinson, 2003)
4.2 The momentum approach
A more genereal approach to determine forces from velocity data is the momentum ap-
proach. The scientic literature about this method is shortly described in section 4.2.1.
This is followed by an extensive explanation how the momentum approach is used for this
thesis. Section 4.2.2 shows how to obtain pressure-elds from velocity-elds and section
4.2.3 describes how ultimately the force is obtain from these pressure-elds. The results of
this investigation are presented in section 4.3.1.
4.2.1 Scientic literature about the momentum approach
The momentum approach, as discussed by Unal et al. (1997), relates the instantaneous
value for the force experienced by the object to the ow variables as:

F(t) =
___
V
u
t
dV
. .
i

__
S
(u n) udS
. .
ii

__
S
p ndS
. .
iii
+
__
S
ndS
. .
iv
. (4.3)
with V the control volume, S its contour and n the outward pointing normal (like in Figure
4.1). Flow eld properties are the velocity vector u, the pressure p, the density , and the
viscous stress . Therefore, the principal contributions to the force are due to (i) time
F.J. Venneman MSc. Thesis
4.2 The momentum approach 27
rate of change of momentum within the control volume, (ii) the net momentum ux across
the boundaries of the control surfaces, (iii) the instantaneous pressure force acting on the
control surface, and (iv) the instantaneous shear force on the control surface. Assuming
incompressible ow, the density is constant. The contribution of the viscous stress can be
neglected when the control volume contour is taken suciently far away from the body. A
direct application of the control-volume formulation requires the velocity and acceleration
distribution inside the volume, and the pressure on the outer contour. However, determin-
ing the pressure from the velocity-eld is not a trivial task. First one obtains the pressure
gradient (note: not the pressure itself) from the Navier-Stokes equations (2.2):
p =
_
du
dt
+u u
_
+
2
u . (4.4)
From this equation it is clear that for incompressible ow, the instantaneous pressure gra-
dient can be derived directly from the velocity information, after which the pressure itself
is obtained from spatial integration of the pressure gradient (using a Dirichlet condition,
i.e. setting a reference value for the pressure). As can be seen in equation (4.3), only the
pressure on the contour is required to obtain the integrated loads. However, the pressure
gradient integration may be extended to the entire ow domain of interest, using either
some sort of gradient-integration scheme or through solving the Poisson equation for the
pressure (Gurka et al. (1999), Fujisawa et al. (2005), Oudheusden et al. (2007), Kat et al.
(2008)).
4.2.2 Determining pressure from two dimensional cross-sections
In this thesis the pressure is obtained by calculating the pressure gradient from the Navier-
Stokes equations for incompressible ow in a rotating frame of reference (repeating equation
(2.6)):

_
u
t
+ (u ) u
_
= p +
2
u
_

t
r +

r
_
+ 2

u
_
. (4.5)
In two dimensions the in-plane pressure gradient components look like this (see Kat et al.
(2008) for the non-rotational terms and Appendix B.2 for the rotational terms):
p
x
=
_
u
t
+u
u
x
+v
u
y
+w
u
z
_
+
_

2
u
x
2
+

2
u
y
2
+

2
u
z
2
_

t
z
2
x + 2w
_
, (4.6a)
p
y
=
_
v
t
+u
v
x
+v
v
y
+w
v
z
_
+
_

2
v
x
2
+

2
v
y
2
+

2
v
z
2
_
. (4.6b)
These equations properly describe the pressure gradient in planar ow. There are no
rotational terms for the y-component of the pressure gradient, because the wing is ro-
tating around this axis. Without rotational terms, these equations have been used by
Baur and Kongeter (1999). In this study we used the planar Poisson approach, rst stated
MSc. Thesis F.J. Venneman
28 Determining forces on insect wings experimentally
by Gurka et al. (1999) for this purpose. The planar Poisson approach uses the in-plane
divergence of the pressure gradient.
f =
2
p =

x
_
p
x
_
+

y
_
p
y
_
. (4.7)
Including equations (4.6a) and (4.6b) in equation (4.7) and dierentiating to x and y
respectively yields:
f =
_
(
u
x
+
v
y
)
t
+ u

x
_
u
x
+
v
y
_
+ v

y
_
u
x
+
v
y
_
+ w

z
_
u
x
+
v
y
_
+
+
_
u
x
_
2
+ 2
v
x
u
y
+
_
v
y
_
2
+
w
x
u
z
+
w
y
v
z

2
x + 2
w
x
_
+
+
_

3
u
x
3
+

3
u
xy
2
+

3
u
xz
2
+

3
v
x
2
+

3
v
y
3
+

3
v
y
2
z
_
. (4.8)
After some rewriting, the planar Poisson equation in the presence of three dimensional ow
becomes:
f =
_
divxy
t
+ (u ) div
xy
+
_
u
x
_
2
+ 2
v
x
u
y
+
_
v
y
_
2
+
w
x
u
z
+
w
y
v
z
+

2
x + 2
w
x
_
+
2
div
xy
. (4.9)
Where div
xy
is the in plane divergence:
div
xy
=
u
x
+
v
y
. (4.10)
Numerical scheme and boundary conditions
If we obtain the value of f per datapoint, we can use a simple iterative Poisson solver
scheme to compute the pressure value for each datapoint. See Figure 4.4 for a simplied
overview of the grid used to calculate the pressure-eld. The Poisson scheme is used in
combination with Neumann boundary conditions (pressure gradient), next to the use of
a Dirichlet reference point. More specic, the points p
1
(a typical internal point), p
2
(a
boundary point) and p
3
(a corner point) are calculated in the following manner (mind that
p
2
is not equal to p-squared):
p
1
=
p
i,j+1
+ p
i+1,j
+ p
i,j1
+ p
i1,j
f(p
1
)h
2
4
, (4.11a)
p
2
=
p
i,j+1
+ p
i,j1
+ 2p
i1,j
+ 2h
p
2
x
f(p
2
)h
2
4
, (4.11b)
p
3
=
2p
i,j1
+ 2p
i1,j
+ 2h
p
3
x
+ 2h
p
3
y
f(p
3
)h
2
4
. (4.11c)
F.J. Venneman MSc. Thesis
4.2 The momentum approach 29
Figure 4.4: Simplied overview of the grid used to calculate the pressure-eld. The diagonal
line represents the cross-section of the wing. As discussed in section 3.3 there are 81 (N
i
)
datapoints in the x-direction and 60 (N
j
) datapoints in the y-direction, so the total eld
consists of 4860 datapoints in total. The distance between two datapoints, h is 4.08 mm.
The Neumann conditions are directly obtained from the pressure gradients in (4.6). We set
the reference pressure (Dirichlet condition) in the bottom left to the pressure predicted by
the irrotational Bernoulli equation. This region was chosen since it complies best with the
Bernoulli restrictions (least time dependency, least vorticity). The value for the Dirichlet
point is therefore chosen as:
p = C
1
2

_
u
2
+ v
2
_
, where C = 0 . (4.12)
Since there is only one point set as a Dirichlet condition, this point acts only as a reference
pressure for the specic velocity-eld at consideration, so there is no need to set it exactly
right.
Masking unreliable data regions
As the red datapoints in Figure 4.4 indicate, some regions of the grid are masked. Masking
a region makes sure that the velocity components of this region are not taken into account
for the determination of the pressure-eld. There are two specic regions chosen to be
masked:
1. The region immediately around the wing
The region immediately around the wing was masked because we are not sure of the
reliability of the vectors in that region. This is related to the way the velocity was
determined there. The wing itself is only 2.25 mm thick, but during cross correlation,
64 64 interrogation windows where used (equal to approximately 1.5 1.5 cm),
MSc. Thesis F.J. Venneman
30 Determining forces on insect wings experimentally
which denitely would overlap the wing. So for one interrogation window velocity
components from both sides of the wing where taken into account, which would
certainly lead to unreliable vectors. Therefore it was chosen to mask the vectors
less than 1.2 cm (about three vector distances) around the wing. This is a trade-o
between not losing too much information and using reliable information.
2. The boundary of the fov
Since the fov was tilted a little in the yz-plane (Poelma et al., 2006, Fig. 5), some
vectors at the boundary of the fov where not to be trusted. As well as for the region
around the wing, a band of 1.2 cm was masked.
Transforming velocity to a rotating frame of reference
The cameras did not rotate with the wing, which means that the measured velocity com-
ponents are measured with respect to a static reference frame. Nevertheless we do use
rotating terms in equation 4.5, which is based on the denition of the coordinates system
attached to the wing. Therefore the velocity had to be transformed to a rotating frame
of reference. This is very simple with the knowledge of (t) (see section 3.2). The trans-
formed velocities are obtained with the following relations for respectively the x, y and z
component of the velocity (see the derivation in Appendix B.1):
u
r
=u (t)z
r
, (4.13a)
v
r
=v, (4.13b)
w
r
=w (t)x
r
. (4.13c)
Where symbols with subscript r are in the rotating frame of reference. The velocity in the
y-direction does not change since the wing is rotating around this axis.
Details Poisson scheme
In order to calculate the derivatives in equation 4.5 the second order three-point rule was
used. At the boundary and next to the masked region a simple rst order two point rule
was used.
To speed up the process of convergence a rst guess for the pressure-eld was made using
the Marching scheme. This scheme uses the local gradient to determine the pressure-
eld (by integrating the gradient over space). The scheme marches through the eld
calculating the pressure at the points with the most neighbors. More about this can be
read in (Turella, 2008, page 24-25).
After obtaining the rst guess the Poisson scheme could start. For this iterative process
a convergence criterion is needed. This convergence criterion requires a denition of the
error :
=
1
N
i
N
j

i=1,N
i

j=1,N
j
|p
n
i,j
p
n1
i,j
| , (4.14)
F.J. Venneman MSc. Thesis
4.2 The momentum approach 31
where N
i
is the number of datapoints in the x-direction (81), N
j
is the number of datapoints
in the y-direction (60) and n is the iteration parameter. The convergence criterion for
was set to = 1 10
6
. Fortan 90 was used to calculate the planar pressure-elds,
each measurement plane would approximately take 10 seconds on a Pentium 4, 3.4 GHz
processor. To calculate the pressure-eld for all 20 planes and 16 time steps would therefore
take about one hour of computation time.
A concern for the pressure calculation is the possible lack of resolution. The leading-edge
vortex is only covered within 5 datapoints in width. This number is too low to get a good
reconstruction of the pressure drop. More important, it has a signicant inuence on the
far eld as shown by preliminary calculations. The fact that lack of resolution in one place
inuences the whole reconstructed pressure-eld is is a true disadvantage of the Poisson
scheme.
4.2.3 Determining forces from two dimensional pressure-elds
Having obtained the pressure-eld from the velocity-eld, it is fairly straight forward to
compute the force. As described in section 2.2, the momentum approach is used. The
derivation of this approach is in Appendix A.1, but a brief explanation will be presented
here.
In this thesis the momentum approach is based on the incompressible Navier Stokes equa-
tions including the Coriolis and centripetal acceleration (2.6). Using a momentum approach
the force on the wing is written as the pressure and viscous force on the wing (see Figure
4.5 and equation (A.8)).
F(t) =
_

__
EFG
p ndS +
__
EFG
ndS
_
=
__
ABCI
(u n) u p n + ndS +
+
___
V

u
t

r
_
2

t
rdV . (4.15)
Where V is the volume of the control volume with boundary S. The indices of the
integration path are depicted in Figure 4.5. After some extensive rewriting (see Appendix
MSc. Thesis F.J. Venneman
32 Determining forces on insect wings experimentally
Figure 4.5: Control volume. The black body indicates a cross-section of the wing. The
dashed line is the integration path
A.1), the planar lift and drag are calculated as follows:
F
1
(t) =
__
ABCI
u
2
dy
__
ABCI
uvdx
__
ABCI
2
u
x
dy +
+
__
ABCI

_
u
y
+
v
x
_
dx +
__
ABCI
pdy +
+
___
V

u
t
+
2
x 2w

t
xdV , (4.16a)
F
2
(t) =
__
ABCI
vudy
__
ABCI
v
2
dx +
__
ABCI
2
v
y
dx +
+
__
ABCI

_
u
y
+
v
x
_
dy
__
ABCI
pdx +
+
___
V

v
t
dV . (4.16b)
Where F
1
is the drag force and F
2
is the lift force. The resulting forces have units N m
1
,
because volume V (see Figure 4.5) is a two-dimensional volume. Calculating the force over
each cross-section and multiplying by their distance (1 cm, as described in section 3.3)
gives the force with unit N.
The major advantage of the momentum approach is that one can easily distinguish the
contribution of every physical aspect of the ow, this will be presented in section 4.3.2.
It turned out that integrating close to the wing (like in Figure 4.5), resulted in more
reliable results than taking the whole fov as a contour. This may be the result of lack
resolution (the leading-edge vortex was typically only captured within 5 datapoints). This
lack probably even has a more signicant inuence on the far eld than on the eld near
F.J. Venneman MSc. Thesis
4.3 Results 33
the wing itself, since the Poisson algorithm propagates errors throughout the whole domain.
The fact that relevant ow was masked, could also be a cause of unrealistic pressure-elds.
The integration contour for the force determinations was taken 32 mm from the wing. To
get a measure of the uncertainty, the force on the wing was also determined for contours
in the range of 20 53 mm, which will result in a range of possible forces that will be used
as an indication for the uncertainty of the deduced force.
4.3 Results
In this section, the results of the post-processing of the experimental data are shown.
Section 4.3.1 shows the results for the Blasius approach, followed by section 4.3.2 in which
the results for the momentum approach are shown.
4.3.1 Results Blasius approach
The reproduced results of the force calculations using the Blasius approach are shown in
Figure 4.6(a). Poelmas results are also shown in Figure 4.6(b) to see that both gures
are in good agreement (at least for the rst part of the stroke (t < 1 s) since Poelma only
showed this).
Looking at Figure 4.6(a) one can conclude that the Blasius approach works quite well when
the vorticity is still contained within the eld of view. Ones it is leaving the eld of view
(around t = 0.5), the Blasius approach is not applicable anymore. After t = 1 s, lift and
drag even become negative.
4.3.2 Results momentum approach
This section has been split in two parts. First the results of the pressure deduction are
shown followed by the results of the force deduction.
Results momentum approach: Pressure elds
In this section a timewise and a planewise reconstruction of the pressure-elds are shown.
A timewise reconstruction of the pressure-eld is given in Figures 4.7 and 4.8. In each
gure a cross-section of the wing can be found at a 0.66R (with R = 0.25 m) distance
from the center of rotation (the viewpoint is also from the center of rotation). This is close
to the radius of gyration (for further information see section 5.2). The black line in the
gures represents a cross-section of the wing. The yellow area around the wing is a masked
region, because the velocity vectors there are not trusted (see section 4.2 for detailed
information). In Figures 4.7(a) and 4.8(a) a distance scale (in m) and the pressure-scale
(in Pa) are presented, which are also valid for the other gures. It is a clearly shown that
a local pressure minimum develops near the leading-edge of the wing, while the pressure
in front of the wing remains stable. From t = 1.5 s, a local pressure maximum develops in
MSc. Thesis F.J. Venneman
34 Determining forces on insect wings experimentally
time (s)
F
o
r
c
e
(
N
)
Drag Blasius appr.
Lift Blasius appr.
Drag sensor
Lift sensor
0 0.5 1 1.5 2 2.5 3
-0.5
0
0.5
1
(a) Present Analysis
(b) Analysis Poelma et al. (2006)
Figure 4.6: Results for the Blasius approach, showing forces on the beginning of the stroke
of an impulsively-started wing (total stroke takes 3 s). Continuous line and crosses represent
the horizontal forces, dashed line and circles represent vertical forces.
F.J. Venneman MSc. Thesis
4.3 Results 35
the right-bottom of the Field Of View, this seems to be an unphysical pressure maximum.
The fact that the pressure-eld shows unphysical behaviour can be caused by the lack of
resolution of data points. The leading-edge vortex is only covered within 5 datapoints in
width. This is too low to get a good reconstruction of the pressure drop. More important,
it has a signicant inuence on the far eld as shown by preliminary calculations. The
fact that lack of resolution in one place inuences the whole reconstructed pressure-eld
is is a true disadvantage of the planar Poisson scheme. A planewise reconstruction of the
pressure-eld is given in Figures 4.9 and 4.10. The same pressure scale is used as for
the timewise reconstruction. The pressure-eld near the wing base (Figure 4.9) is rather
smooth whereas the pressure-eld near the wing tip (Figure 4.10) is quite disturbed. The
latter is caused by the tip vortex, which is a vortex near the tip in the direction of the
wake of the wing as described by Poelma et al. (2006, Fig. 10).
Results momentum approach: Forces
The contributions of the dierent terms in equation (4.16) to the total force are depicted in
Figure 4.11. It can easily be seen that the force due to the pressure (Figure 4.11(a)) is the
dominant term in the total force, especially when the wing had stopped accelerating and
the owpattern around the wing is steady. The force due to change of momentum inside
the contour (Figure 4.11(b)) is signicant in the initial phase of the stroke (t < 0.5 s). It is
convincing to see that for t > 0.5 s, the force due to momentum ux through the boundary
(Figure 4.11(c)) of the contour comes into play. Until that time the ux through the
boundary has not been signicant. The force due to viscosity (Figure 4.11(d)) is negligible
(which is as expected because separating ow is generally pressure dominated). The force
due to the centripetal acceleration (Figure 4.11(e)) is also negligible. The force due to the
Coriolis eect (Figure 4.11(f)) has a contribution in the steady part of the stroke. The
force due to the rotational acceleration 4.11(g) has a very big inuence in the beginning.
This term is often neglected in scientic literature, but turns out to be very important in
the case of an accelerating wing.
Figure 4.12 shows the eect of the integration contour placement on the the total force, for
4 dierent time phases of the rotation. Theoretically, the total force should be independent
at all distances away from the wing, so the variation is an indication of the uncertainty of
the force determination. In the rst 12 mm away from the wing, the pressure was masked,
since the velocity vectors there are unreliable (see section 4.2). As can be seen there is a
transient region in which the force has to develop from 12 to 20 mm away from the wing.
The extrema of the shaded regions are taken as the uncertainty per time-step for the total
force depicted in Figure 4.13. Figure 4.12(a) (t = 0.1 s) attracts the attention because the
uncertainty band is so large. This is probably due to the large acceleration there, whereas
Poelma et al. (2006) mentions a t
c
(the wing reached 63% of its maximum velocity within
time t
c
) of 0.2 s, the exponential t yielded 0.17 s, the numerical part of the present study
even suggests that the acceleration was not purely exponential. The other gures in Figure
4.12 show a reasonably constant trend for the domain between 20 and 53 mm. This is
an indication that the pressure and force calculations performed, are heading in the right
MSc. Thesis F.J. Venneman
36 Determining forces on insect wings experimentally
y
(
m
)
x (m)
Pa
-0.1 -0.05 0 0.05 0.1 0.15
-80
-60
-40
-20
0
20
40
-0.1
-0.05
0
0.05
0.1
(a) t = 0.1 s (b) t = 0.2 s
(c) t = 0.3 s (d) t = 0.4 s
(e) t = 0.5 s (f) t = 0.6 s
(g) t = 0.7 s (h) t = 0.8 s
Figure 4.7: Timewise cross-sections of the pressure around the wing, from t = 0.1 0.8 s
at 0.66R from the center of rotation.
F.J. Venneman MSc. Thesis
4.3 Results 37
y
(
m
)
x (m)
Pa
-0.1 -0.05 0 0.05 0.1 0.15
-80
-60
-40
-20
0
20
40
-0.1
-0.05
0
0.05
0.1
(a) t = 0.9 s (b) t = 1.0 s
(c) t = 1.1 s (d) t = 1.2 s
(e) t = 1.5 s (f) t = 1.7 s
(g) t = 2.0 s (h) t = 2.5 s
Figure 4.8: Timewise cross-sections of the pressure around the wing, from t = 0.9 2.5 s
at 0.66R from the center of rotation.
MSc. Thesis F.J. Venneman
38 Determining forces on insect wings experimentally
y
(
m
)
x (m)
Pa
-0.1 -0.05 0 0.05 0.1 0.15
-80
-60
-40
-20
0
20
40
-0.1
-0.05
0
0.05
0.1
(a) 0.38R (b) 0.42R
(c) 0.46R (d) 0.50R
(e) 0.54R (f) 0.58R
(g) 0.62R (h) 0.66R
Figure 4.9: Planewise cross-sections of the pressure around the wing, at 0.38R0.66R from
the center of rotation at t = 1.0 s ( = 55.7

).
F.J. Venneman MSc. Thesis
4.3 Results 39
y
(
m
)
x (m)
Pa
-0.1 -0.05 0 0.05 0.1 0.15
-80
-60
-40
-20
0
20
40
-0.1
-0.05
0
0.05
0.1
(a) 0.70R (b) 0.74R
(c) 0.78R (d) 0.82R
(e) 0.86R (f) 0.90R
(g) 0.94R (h) 0.98R
Figure 4.10: Planewise cross-sections of the pressure around the wing, at 0.70R 0.98R
from the center of rotation at t = 1.0 s ( = 55.7

).
MSc. Thesis F.J. Venneman
40 Determining forces on insect wings experimentally
time (s)
f
o
r
c
e
(
N
)
Drag
Lift
0 0.5 1 1.5 2
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
(a) Force due to pressure
time (s)
f
o
r
c
e
(
N
)
Drag
Lift
0 0.5 1 1.5 2
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
(b) Force due to change of
momentum inside the con-
tour
time (s)
f
o
r
c
e
(
N
)
Drag
Lift
0 0.5 1 1.5 2
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
(c) Force due to momen-
tum ux
time (s)
f
o
r
c
e
(
N
)
Drag
Lift
0 0.5 1 1.5 2
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
(d) Force due to viscosity
time (s)
f
o
r
c
e
(
N
)
Drag
0 0.5 1 1.5 2
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
(e) Force due to centripetal
acceleration
time (s)
f
o
r
c
e
(
N
)
Drag
0 0.5 1 1.5 2
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
(f) Force due to coriolis ef-
fect
time (s)
f
o
r
c
e
(
N
)
Drag
0 0.5 1 1.5 2
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
(g) Force due to rotational
acceleration
Figure 4.11: The contributions of the dierent terms in equation (4.16) to the total force.
The blue lines indicate a contribution to the drag and the red lines indicate a contribution to
the lift.
F.J. Venneman MSc. Thesis
4.3 Results 41
direction.
Figure 4.13 shows the result of the force determination on the impulsively-started wing.
distance from wing (mm)
f
o
r
c
e
(
N
)
5 10 15 20 25 30 35 40 45 50
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
(a) t = 0.1 s
distance from wing (mm)
f
o
r
c
e
(
N
)
5 10 15 20 25 30 35 40 45 50
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
(b) t = 0.5 s
distance from wing (mm)
f
o
r
c
e
(
N
)
5 10 15 20 25 30 35 40 45 50
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
(c) t = 0.9 s
distance from wing (mm)
f
o
r
c
e
(
N
)
5 10 15 20 25 30 35 40 45 50
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
(d) t = 2.0 s
Figure 4.12: Forces in x (drag, blue) and y (lift, red) direction for dierent sizes of integration
contours. The x-axis displays at what distance from the wing the contour is taken.
One can see that for the steady part (t > 0.7 s) of the stroke, the forces coincide very
well with the measurements obtained with the force sensor. The drag is within 13% of
the measured values, the lift is even within 4% of the measured values. This is certainly
an advantage with respect to the Blasius approach, where this part of the stroke was
mismatched due to vorticity leaving the fov. The initial force peaks in Figure 4.13 however,
are not reached (for both lift and drag). This can be caused by the - already mentioned -
lack of resolution, which causes the extrema of the pressure to be smoothed and thus the
forces to be less accurate. The mismatch in the peak can also be caused by the fact that
MSc. Thesis F.J. Venneman
42 Determining forces on insect wings experimentally
the interesting physics in the start of the stroke is taking place close to the wing, which
unfortunately was masked for the pressure determination. In that case the most important
information is absent. Furthermore it can been seen that in the initial stage of the stroke
lift compares better with the sensor than drag. As discussed in the previous paragraph
this is probably caused by an overestimated acceleration.
time (s)
f
o
r
c
e
(
N
)
Drag Momentum appr.
Lift Momentum appr.
Drag sensor
Lift sensor
0 0.5 1 1.5 2 2.5 3
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Figure 4.13: Result of the force determination on the impulsively-started wing by a combi-
nation of the momentum approach and the planar Poisson approach.
4.4 Conclusion: Good estimation of forces on steady
stroke
In this section the conclusions will be drawn about the experimental part of the present
study. The determined forces (using the momentum approach) are discussed in section
4.4.1 and the determined pressure-elds (using the planar Poisson approach) are discussed
in section 4.4.2. Generally it can be concluded the force history deduced by a combination
of the momentum approach and the planar Poisson approach yield a good estimation of
the forces on the impulsively started wing. However there are two main weaknesses of
the combined system. Firstly the momentum approach heavily depends on the pressure
deduction (planar Poisson approach), since it yields unphysical results, once an unrealistic
pressure-eld is prescribed. Secondly the incorrect assumptions about the initial accelera-
tion had an eect on both the pressure deduction and the force deduction, since it appears
in both algorithms.
F.J. Venneman MSc. Thesis
4.4 Conclusion: Good estimation of forces on steady stroke 43
4.4.1 Forces
The forces on the wing predicted by the momentum approach (Figure 4.13) correspond rea-
sonably well to the forces measured with a sensor mounted on the Roboy (Poelma et al.,
2006). These measured forces are assumed to be realistic, yet dr. ir. C. Poelma informed
me that the force sensors had suered from wear and could well be 10% o. The estimated
forces of the present study in the steady part of the stroke (t > 0.7 s, for which the rota-
tional velocity is approximately constant) match very well to the experimental results. The
drag is within 13% of the measured values, the lift is even within 4% of the measured values.
They match much better than the Blasius approach, that shows unphysical results when
the vorticity has left the eld of view. However, the initial forces (t < 0.7 s), as predicted
by the momentum approach, did not correspond very well to the measurements. One can
see that especially the drag is overpredicted (and varies in a large domain for dierent
integration contours), whereas the lift corresponds rather good to the measured forces.
This is probably caused by an overestimated initial acceleration (which only inuenced the
drag directly, see equation (4.16a)). Figure 4.11(g) shows that the rotational acceleration
inuences the initial force signicantly. As can be seen in Figure 3.2(c), the assumed initial
acceleration seems far o the derived acceleration (however, it was prescribed like this in
Poelma et al. (2006)), so this may be an origin of the mismatch of the initial forces. It
should be noticed that the rotational acceleration not only inuenced the force, but also
inuenced the pressure gradient for the planar Poisson approach. So if the acceleration is
overpredicted, the error is taken into account twice. The estimated pressure-elds used for
the momentum approach, may also have been determined incorrectly, this is discussed in
the next section.
4.4.2 Pressure
The pressure-elds around the wing look quite as expected, i.e. there is a pressure minimum
near the leading-edge and a small pressure maximum in front of the wing. However, during
force determination, it turned out that only integrating the pressure-eld in the vicinity of
the wing delivered reliable results. The far eld of the pressure seems rather unreliable, for
example the pressure increases on both the left and right side of the far eld. Besides that
an unexpected pressure maximum shows up in the right-bottom of the wing for t 1.5
s (see Figures 4.8(e)-4.8(h)). The fact that the pressure-eld shows unreliable behaviour
can be caused by multiple aspects, from which the lack of resolution of data points and
the masking of important information near the wing seem most presumable.
MSc. Thesis F.J. Venneman
44 Determining forces on insect wings experimentally
F.J. Venneman MSc. Thesis
Chapter 5
Determining forces on insect wings numerically
In this chapter a numerical study on insect wings is performed. The theory for this
numerical study is provided in section 5.1. After that the forces on a two-dimensional wing
are calculated in section 5.2, concerning an impulsively-started wing in both transient and
steady parts of the stroke. The two-dimensional study is a preliminary study for the work
presented in section 5.3, in which the forces on an impulsively-started three-dimensional
wing are calculated. The kinematics described in Chapter 3 are mimicked as accurate as
possible. The force-coecients measured by the sensor of the experiment will be compared
to the computational results. The conclusions of this Chapter are discussed in section 5.4
Throughout the present research, OpenFOAM was used as a framework for solving
the ow. OpenFOAM is a freely available and open source solver of continuum mechanics
(including Computational Fluid Dynamics), licensed under the gnu General Public
Licence.
5.1 Theory for a numerical study of insect wing aerody-
namics
In this section the background of some important parameters is discussed, which are used
in the next sections. As already mentioned in the introduction of this chapter, the force
was used in order to compare dierent cases with each other an with experiments. This
force consists of the viscous and the pressure force on the surface of the wing (just as in
equation A.8 of the experimental part of this study).

F =
__
ndS +
__
p ndS, (5.1)
with S the surface of the wing, p the pressure and n the vector normal to the wing, as
can be seen, the gravitational body force has neglected. For Newtonian uids the viscous
MSc. Thesis F.J. Venneman
46 Determining forces on insect wings numerically
stress tensor in equation (5.1) can be rewritten in Einstein notation as:

ij
=
_
u
i
x
j
+
u
j
x
i
_
. (5.2)
The experimental and computational results could be compared by their force-coecients,
these are dened as:.
C
D
(t) =
D(t)
qA
, C
L
(t) =
L(t)
qA
. (5.3)
Here the drag, D, describes the force in the horizontal (x) direction and the lift, L, is the
force in the vertical y direction. A is the wing surface area (equal to c in two dimensions)
and the dynamic pressure q is given by:
q =
1
2
U
2
ref
, (5.4)
where U
ref
is the steady velocity of the wing; in two dimensions simply the nal velocity
is obtained. In three dimensions the nal wing-tip velocity, U
tip
is chosen for reasons of
consistency in this report, because it was also used in Poelma et al. (2006) (mind that this
is not a trivial choice, one can also take the average velocity over the wing span).
A dimensionless geometric property of wings, which will be used in this part of the thesis
is the wing aspect ratio, this ratio is dened as:
AR
wing
=
b
2
A
, (5.5)
with b the span of the wing. Another important dimensionless number that will appear in
the context of this chapter is the Courant number. This dimensionless number is dened
for one cell as:
Co =
t |u|
x
, (5.6)
where t is the time step, |u| is the magnitude of the ow velocity through that cell and x
is the cell size in the direction of the velocity. So varying the Courant limit for a specic
grid, basically means varying the time step. It is advantageous to ensure that Co < 1,
so that convective ow does not surpass a cell in one timestep. However, implicit time
stepping (which is used in the present study) is also stable for Co > 1, unlike explicit time
stepping for which the limitation is very severe (Jasak, 1996, pag. 90).
For visualization purposes, iso-surfaces of the second invariant of the uid gradient tensor,
Q, are shown at the end of this chapter. For an incompressible ow, Q is dened as
(Piomelli et al., 2000; Chakraborty et al., 2005):
Q
1
2
(
ij

ij
S
ij
S
ij
) , (5.7)
F.J. Venneman MSc. Thesis
5.2 2D numerical model 47
in which
ij
and S
ij
are the antisymmetric and symmetric components of the gradient
velocity tensor u, respectively:

ij
=
1
2
_
u
i
x
j

u
j
x
i
_
, (5.8a)
S
ij
=
1
2
_
u
i
x
j
+
u
j
x
i
_
. (5.8b)

ij
is the resolved rotation tensor and S
ij
the strain-rate tensor (notice the similarity
between equation (5.8b) and equation (5.2)). The Q-value can be interpreted as a balance
between the local swirling strength () and shear (S). Positive values indicate regions
where the swirling motion is dominant; this means that the local value of Q is a useful
tool to identify vortical structures. On the iso-surfaces of the second invariant of the uid
gradient tensor Q, values of the normalized helicity, H will be shown. The normalized
helicity is dened as (Batchelor, 1967):
H =
u
|u| ||
. (5.9)
The H-value can be interpreted as a balance between the swirling direction and the ow
direction. For |H| 1, the vorticity is aligned with direction of the velocity (think of
the tip vortex). For |H| = 0, the vorticity is the perpendicular to the ow (think of the
leading-edge vortex).
5.2 2D numerical model
In this section the two-dimensional model will be described. In section 5.2.1, the transition
from three-dimensional to two-dimensional motion is explained. This is followed by a
description of the development of the grid in section 5.2.2. Section 5.2.3 explains how the
system of equations is solved. Finally the dierent case studies will be described in section
5.2.4 of which the results are shown in section 5.2.5.
5.2.1 From three-dimensional to two-dimensional motion
Initial computations were performed for a two-dimensional ow geometry for simplicity.
The three-dimensional motion of the Roboywing (as described in 3.1) had to be conned
to two-dimensional motion. A translating wing at a constant angle of attack was chosen
as a research object.
The two-dimensionality is obtained by considering the motion in a plane at a representative
radius of the wing. Birch and Dickinson (2003) suggested that at 0.65R of the Roboy-
wing, the leading-edge vortex was still attached and exhibited near-maximal spanwise
vorticity, so this would be a good cross-section to investigate. Ellington (1984) found that
the radius of gyration (at 0.64R of the Roboy wing (Bos et al., 2008)) was a better option,
MSc. Thesis F.J. Venneman
48 Determining forces on insect wings numerically
because that is the place where the mean lift acts.
From the experimental data, the closest known chord length was known at 0.66R (c =
0.1018), so this cross-section will be investigated. With a nal rotational velocity of =
1.181 rad s
1
(see Appendix B.3) the velocity at this cross-section was 0.66R = 0.19
m s
1
(with R = 0.25 m). The kinematic viscosity of the uid in the experiment was
= 115 10
6
m
2
s
1
. This results in a chord length based Reynolds number of
Re =
cU

= 172 . (5.10)
The mean angle of attack in the experiment was 50.4

, which was also mimicked in the


numerical work.
Translating a wing through a uid is the same as simulating a quiescent wing in a stream.
For numerical work the advantage of the latter option is that no mesh motion is involved.
Besides that the ow pattern after the transient is investigated in this study, which would
require very large grids in a moving mesh setup.
5.2.2 Building the grid in two dimensions
In the present study, a 10% (thickness versus chord-length) percent 2
nd
order ellipsoid
1
shape represents the wing. Figure 5.1 shows the most important aspects of the structured
O-grid around this wing (46k cells, 200 cells along one chord length, 10000 cells within 4
chord lengths from the center of the wing). In Figure 5.1(a) the schematic setup of the
domain of the grid is depicted. It can be seen that the height of the grid takes 20 chord
lengths, in order to obtain a reasonably unperturbed ow around the wing. The setup is
tube like to be able to visualize the wake as well. The part in front of the wing could have
been reduced, since the ow is still very uniform there. For that reason the cell density
was very low in this part of the domain, so not much of the computational eort was put
in it. In order to obtain a mesh with good quality an elliptical mesh around the wing is
used following Bos et al. (2008). Such a grid is suitable for our problem since the vorticity
is strongest near the tip of the ellipse and is weaker away from the body. Figure 5.1(b)
shows the details of the mesh around the wing. Figure 5.1(c) shows a detailed structure of
the mesh around the leading-edge. GridPro was used to generate the grids. The quality
(by means of the characteristic parameters as the aspect ratio, non-orthogonality and the
skewness) of the mesh is described in Appendix D.2.
5.2.3 Solving the system in two dimensions
Having obtained the grid, the system of equations could be solved in order to acquire a
pressure and a local velocity vector per cell. This system is solved using backward, Euler
(implicit, second order time stepping) for temporal discretization and central dierencing
1 x
a
c
a
1
+
y
a
c
a
2
= 1, with a = 2 (hence 2
nd
order), c
1
and c
2
are the radii in two directions.
F.J. Venneman MSc. Thesis
5.2 2D numerical model 49
(a)
(b) (c)
Figure 5.1: Most important aspects of the grid of 46k cells around the wing. (a) A schematic
setup of the grid-domain. (b) Details of the mesh around the wing, it has 200 cells around
the chord. (c) Details of the mesh around the leading-edge.
MSc. Thesis F.J. Venneman
50 Determining forces on insect wings numerically
for spatial discretization. No turbulence model is used so in fact a Direct Numerical
Simulation (dns) is used. The specic solver used, is applicable for incompressible, laminar
ow of Newtonian uids and a dynamic mesh. In the two-dimensional case there was no
moving mesh, but nevertheless the solver could be used. The boundary conditions in the
domain are depicted in Figure 5.2. At the inlet, a Dirichlet condition is set for the velocity.
The velocity u
i
is time dependent, where the acceleration of the experiment is mimicked
(see equation (B.11)):
u
i
(t) = (t)0.66R =
_
1 exp
_

t
t
c
__
0.66R , (5.11)
where t
c
is the typical time to reach 63% of the maximum velocity (0.1685 s) and R is the
distance from the point-of-rotation to the wing tip (0.25 cm). At the outow there is no
velocity gradient in the tangential direction, so a Neumann boundary condition is set, i.e.
u
x
= 0. The top and bottom of the domain where established as a SymmetryPlane, which
means that there is no convection in the normal direction (v = 0) and no gradient in the
normal direction (
u
x
= [0, 0]). The velocity on the wing itself was zero in all directions.
To obtain a well-posed problem (Wesseling, 2001) the boundary condition for the pressure
inlet was set to a Neumann condition (
p
y
= 0). A Dirichlet condition was used at the
outow (reference pressure set to p = 0). The top and the bottom where again established
as a SymmetryPlane, i.e.
p
y
= 0. The boundary condition on the wing was set to a zero
gradient on the normal of the wing (
p
n
= 0). More details about the settings used in
OpenFOAM are stated in Appendix E.
5.2.4 Setup cases in two dimensions
Since impulsively-started two-dimensional wings shed their vortices (Dickinson and Gotz,
1993), whereas they may stay attached on rotating wings (Poelma et al., 2006), the two-
dimensional case and the three-dimensional case cannot be compared in a quantitative
manner. Nevertheless, a lot of characteristics of the three-dimensional ow will appear in
the two-dimensional case. Three distinct cases were tested in two dimensions:
1. Variation of grid density and Courant number, to check the dependency on spatial
and temporal discretization.
2. Validation of an impulsively-started wing, to see whether our results would make
sense.
3. Reynolds-Strouhal relation, to check the transition of the vorticity in the wake of the
wing.
The setup of the numerical investigations is described in the following paragraphs, the
results are shown in section 5.2.5.
F.J. Venneman MSc. Thesis
5.2 2D numerical model 51
(a)
(b)
Figure 5.2: Boundary conditions of the two-dimensional case. (a) velocity (b) pressure.
MSc. Thesis F.J. Venneman
52 Determining forces on insect wings numerically
t [s]
C
D
[
-
]
0 20 40 60 80 100
1.4
1.5
1.6
1.7
1.8
1.9
2
2.1
2.2
2.3
2.4
(a) Transient of the force
C
D
[-]
C
L
[
-
]
1 1.2 1.4 1.6 1.8 2 2.2
0.7
0.8
0.9
1
1.1
1.2
1.3
1.4
1.5
1.6
1.7
(b) Phaseplot for 300 < t < 500
Figure 5.3: Transient force (a) and a phaseplot, C
L
C
D
(b) for a quiescent wing in a
steady ow. Re = 192, Co
max
= 1 and 46k cells.
Variation of grid density and Courant number
The rst case investigates the characteristics of the steady ow around the wing. In order
to obtain a reproducible result, it checks whether the solution of the calculation converges
for increasing grid density and decreasing Courant number. The maximum Courant num-
ber, Co
max
, is decreased from 2, 1 to
1
2
. Decreasing Co implies that the time step t will
be adjusted to the cell with the highest value of
|u|
x
.
The number of cells is approximately doubled twice, from 22k, 46k to 97k. For every num-
ber of cells all Courant numbers have been tested, this means that in total 9 calculations
have been run. The kinematics described at the beginning of this section (see 5.2.1) have
been used for this case (impulsive start until the ow is steady at Re = 172).
The values that were compared are the average force-coecients (C
L
and C
D
) for an in-
teger number of periods. To obtain a steady solution the wing had to travel many chord
lengths, because the impulsive start initiated a transient in the ow pattern. This is shown
in Figure 5.3(a), where C
D
is plotted for the rst 100 s. It can be seen that the periodic
force convergences to a steady pattern. For sake of good convergence it was chosen to
obtain C
L
and C
D
for 300 < t < 500. To show that there is good periodicity for this time
frame, a phase plot of C
D
and C
L
is plotted in Figure 5.3(b). The results of this case are
presented in section 5.2.5.
Validation impulsively-started wing
The second case is a small validation case. Dickinson and Gotz (1993) performed measure-
ments on a two-dimensional impulsively-started wing. This wing was of 2% thickness and
translated at a constant angle of attack (for various values of ) at Re = 192 based on
the nal velocity. It was linearly accelerated at 62.5 cm s
2
to a nal velocity of 10 cm
F.J. Venneman MSc. Thesis
5.2 2D numerical model 53
s
1
. For this validation only one angle of attack of 50

was used, to compare to the results


of Dickinson and Gotz (1993). The kinematic viscosity was adapted to obtain the right
Reynolds number of 192. The force-coecients where obtained for 46k and 97k cells and
for a wing thickness of 2%. The results of this case are discussed in section 5.2.5.
Reynolds-Strouhal relation
The last case invokes the relation between two dimensionless numbers, i.e. the Reynolds
number and the Strouhal number. The Strouhal number (2.4) will be repeated here:
St =
fc
U
, (5.12)
where f is the dominant frequency in the ow (as obtained by analysing an ampli-
tude spectrum of the frequency as described in Appendix F), represented by the vor-
tex shedding frequency in two-dimensional ow. Simulations were made for Re =
[120, 160, 200, 260, 300, 330] at 46k cells and Co
max
= 1 by varying the nal velocity.
A lot of experimental research has been performed on Re St relations, especially for the
ow around circular cylinders, as extensively described in Zdravkovich (1997). In his book
three dierent ow regimes are distinguished for the range of 120 < Re < 330:
Periodic laminar regime (30 to 48) < Re < (180 to 200), a fully developed Karman
street arises.
Transition in Wake 1 (180 to 200) < Re < (220 to 250), transition of laminar eddies in
the wake.
Transition in Wake 2 (220 to 250) < Re < (350 to 400), transition of an irregular eddy
during its formation.
Although a cylinder is not a wing, many of the aspects described above are similar for a
two-dimensional wing at a large angle of attack. Three vorticity gures will be presented
in the results (next section), to show that ow around a wing agrees with the classication
as described above.
The transition in the wake eects the eddy shedding (and thus St) to a large extent.
Zdravkovich (1997, page 82) refers to a study of Roshko (1954), who performed measure-
ments on the shedding frequency of ow past circular cylinders in a large low-turbulence
wind tunnel. He noted that for the range of 150 < Re < 300 there are irregular bursts in
the ow, which makes the ow unstable, leading to aperiodic time-histories of the force.
Since the research performed in this part of the thesis has a domain of 120 < Re < 330,
an unstable St regime is expected.
MSc. Thesis F.J. Venneman
54 Determining forces on insect wings numerically
5.2.5 Discussion of the two-dimensional results
Variation of grid density and Courant number
The values of the average drag coecient, C
D
, for dierent grid densities and dierent
Co
max
are stated in Table 5.1. The values are averaged over 40 complete periods. It can
be seen that the value of C
D
converges for decreasing Co
max
. Besides that, C
D
converges
for increasing grid density. One can also notice that C
D
converges when Co
max
is doubled
while the number of cells is also doubled (roughly), so convergence is more sensitive to
doubling the cell density than to doubling the amount of time steps. To see this, look from
the bottom left to the top right of Table 5.1.
The values of the average lift coecient, C
L
, for dierent grid densities and dierent
number of cells
22k 46k 97k
Co
max
2 1.7135 1.6993 1.6782
1 1.7050 1.6935 1.6736
1
2
1.6983 1.6875 1.6703
Table 5.1: C
D
for a quiescent wing in a
steady ow. Re = 192
number of cells
22k 46k 97k
Co
max
2 1.1524 1.1323 1.1151
1 1.1421 1.1288 1.1125
1
2
1.1421 1.1251 1.1104
Table 5.2: C
L
for a quiescent wing in a
steady ow. Re = 192
Co
max
are stated in Table 5.2. The same trend is shown as for C
D
while varying the
number of cells and Co
max
. So, one can state that the solution converges for increasing
cell density and decreasing Co
max
.
Validation impulsively-started wing
The values for C
D
and C
L
for 8 chord lengths of travel are depicted in Figure 5.4 for two
dierent grid sizes. The size of the grid does not alter the solution to more than 2%, so
the grid is of reasonable density. The alternating pattern for both C
D
and C
L
is due to
the vortex streets that occurs as in Figure 5.5, where the vorticity is shown for 2 and 7
chord lengths of travel, respectively. The force-coecient of the 97k grid are compared
to the results obtained by Dickinson and Gotz (1993) in Figure 5.6. They measured the
lift and drag after 2 and 7 chord lengths of travel. The computed force-coecients are in
the vicinity of the measured lines for 2 (open circles) and 7 (lled circles) chord lengths of
travel. This shows that the computed lift and drag are realistic.
Reynolds-Strouhal relation
Figure 5.7 shows three isovorticty plots and matching phaseplots for Re = 120 (a), Re =
200 (c) and Re = 330 (e), respectively. In Figure 5.7(a) (Re = 120) a periodic Karman
street arises. In Figure 5.7(c) (Re = 200), the wake starts to show some aperiodic behaviour
and Figure 5.7(e) (Re = 330) depicts irregular eddies. From the matching phaseplots it can
be concluded that the forces become higher for increasing the Reynolds number. What
should be remarked as well, is that the line in the phaseplot in Figure 5.7(f) is much
F.J. Venneman MSc. Thesis
5.2 2D numerical model 55
Chord lengths
C
L
46k
97k
1 2 3 4 5 6 7 8
0
0.5
1
1.5
2
2.5
3
(a)
Chord lengths
C
D
46k
97k
1 2 3 4 5 6 7 8
0
0.5
1
1.5
2
2.5
3
(b)
Figure 5.4: This gure show the C
L
(a) and C
D
(b) trajectories in time for 8 chord lengths of
travel for two dierent grid sizes (46k abnd 97k cells). The Reynolds number was 192, based
on the steady-state velocity of 10 cm s
1
, achieved through an initial constant acceleration
of 62.5 cm s
2
.
(a) (b)
Figure 5.5: Vorticity plots for the two-dimensional impulsively-started wing. A vortex street
starts to appear. (a) 2 chord lengths travelled (b) 7 chord lengths travelled.
MSc. Thesis F.J. Venneman
56 Determining forces on insect wings numerically
Figure 5.6: The force-coecients plotted as a function of angle of attack. Open circles
and lled circles show the data for early (2 chords) and late (7 chords) points, respectively.
The colored points are obtained from the simulation and correspond to the cross-sections in
Figure 5.4. The original Figure is from Dickinson and Gotz (1993, Fig. 4).
thicker than in the other phase plots, indicating that the force production is not stable
anymore for this Reynolds number. Increasing Reynolds even further, likely results in a-
periodic behavior. But then turbulence needs to be modelled, which is not accounted for
in the current solver. The Reynolds-Strouhal relation for a translating wing at = 50

is depicted in Figure 5.8. As expected, St is unstable for this regime. The blue points
(acquired by varying the velocity) show a remarkable peak near Re = 160. To verify this
peak, a Re = 172-case was included from the numerical work previously obtained in this
section (to check the variation of grid density and Co
max
). The St number for Re = 172
(red point) corresponds well to the St number for Re = 160.
F.J. Venneman MSc. Thesis
5.2 2D numerical model 57
(a) Re = 120
C
D
[-]
C
L
[
-
]
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6
0.8
1
1.2
1.4
1.6
1.8
2
(b) Re = 120
(c) Re = 200
C
D
[-]
C
L
[
-
]
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6
0.8
1
1.2
1.4
1.6
1.8
2
(d) Re = 200
(e) Re = 330
C
D
[-]
C
L
[
-
]
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6
0.8
1
1.2
1.4
1.6
1.8
2
(f) Re = 330
Figure 5.7: Vorticity plot and phase plot for ow around a wing at = 50

for Re = 120,
Re = 200 and Re = 330.
MSc. Thesis F.J. Venneman
58 Determining forces on insect wings numerically
Re
S
t
100 150 200 250 300 350
0.185
0.19
0.195
0.2
0.205
0.21
0.215
0.22
0.225
0.23
Figure 5.8: Re St relation for a translating wing at an angle of attack 50

. The blue
points are acquired by varying the velocity, the red point is acquired with another kinematic
viscosity .
5.3 3D numerical model
For the three-dimensional case, the Roboy kinematics as described in chapter 3 were used.
An impulsively-started wing with a maximum chord-length-based Reynolds number of 256
was rotated over 90

(not 180

, since the mesh deformation did not properly work in that


case). The geometric planform of the wing was a representation of the Roboy planform
(which in itself a representation of the Drosophila planform) (Dickinson et al., 1999).
Section 5.3.1 provides information on the three-dimensional grid of the wing. This is
followed by section 5.3.2, which explains how the system of equations is solved. Finally
the dierent case studies will be described in section 5.3.3 and their results are discussed
in section 5.3.4. A qualitative impression and a comparison with the experiments of the
ow around the wing is provided in section 5.3.5.
5.3.1 Building the grid in three dimensions
A structured three-dimensional mesh was generated around a wing that was a representa-
tion of the Roboy-wing (Dickinson et al., 1999). In order to develop a mesh around this
wing, a cad-drawing of the Roboy-wing was made, therefore the wing will be referred to
as the cad-wing in the remainder of this thesis. The cad-wing was developed with the
known data points on the wing surface acquired from the piv measurements and had a
constant thickness of 2% (with respect to the maximum chord length). Some angles had
to be rounded o, to make sure that a mesh with low skew cells could wrap around the
wing. A schematic overview of the size of the mesh around the wing is shown in Figure
5.9(a). The planform is shown in Figure 5.9(b), followed by an enlarged image of two
perpendicular cross-sections of the mesh around the wing in Figure 5.9(c).
Two meshes were created around the cad-wing, a 280k and a 530k cell mesh. The gener-
ation of ner meshes resulted in highly skewed cells, so these were not used. The quality
F.J. Venneman MSc. Thesis
5.3 3D numerical model 59
(a) (b)
(c)
Figure 5.9: Most important aspects of the grid of 280k cells. (a) Schematic setup of the
grid-domain. (b) Overview of the planform of the wing, where the characteristic bend in a
Drosophila wing is visible on the left. (c) Details of the mesh around the wing. Two perpen-
dicular cross-sections are provided, the non-ellipsoidal lines near the wing tip are artifacts of
the cross-sections (the cells are not perfectly aligned with the cross-sections).
MSc. Thesis F.J. Venneman
60 Determining forces on insect wings numerically
of the three-dimensional mesh is described in Appendix D.3.
5.3.2 Solving the system in three dimensions
Having obtained the grid, a system of equations could be solved in order to calculate
the local velocity and the pressure on the grid-points. This system of equations is solved
using backward Euler (implicit, second order time stepping) for temporal discretization
and central dierencing for spatial discretization. No turbulence model is used so in fact a
Direct Numerical Simulation (dns) is used. Just as in the two-dimensional case, a solver
for incompressible, laminar ow of Newtonian uids and a dynamic mesh was used.
As already pointed out in the introduction of this section, the conditions of the simulation
were that of an impulsively-started wing with a maximum chord length based Reynolds
number of 256. In order to obtain this exact Reynolds number the kinematic viscosity, ,
was set to 7.38 10
3
m s
2
, with a maximum chord length c = 0.64 m and U
tip
= R
tip
=
2.95 m s
1
. The location of the rotation axis was taken as in the experiment (see chapter
3. The angle of rotation, , was obtained by integrating the rotational velocity, (t) (as
described in section 3.2), with respect to time and setting it = 0 rad for t = 0 s:
(t) =
max
_
t + t
c
exp
_
t
t
c
__

max
t
c
(5.13)
The calculations were performed on a moving mesh, this means that the time dependency
of moving control volumes has to be taken into account. This is done by using the Space
Conservation Law, which is described in Appendix C.2. For a structured mesh there are
ecient deformation techniques available, the one used in this thesis is based on Radial
Basis Function (rbf) interpolation (Boer et al., 2007). To understand RBF interpolation
and the parameters used for the present work, one is referred to read Appendix C.1.
Deformation of a mesh has its limits, as large deformation decreases the quality of a mesh.
For the present work, the wing was rotated over 90

, instead of the 180

rotation of the
experiment. Nevertheless, the interesting part of the motion (the impulsive start and a
part of the steady stroke) is contained within the rst 90

, so one only loses the rest of the


steady rotation of the wing.
The boundary conditions of the three-dimensional case are depicted in Figure 5.10. Con-
cerning the velocity (Figure 5.10(a)), a Dirichlet boundary condition was set on the top
of the domain and a Neumann boundary condition was set at the bottom of the compu-
tational domain. For the pressure (Figure 5.10(b)) exactly the opposite was done in order
to obtain a well posed problem (Wesseling, 2001). On the wing the velocity, u was set to
the wing velocity u
wing
and the normal pressure gradient was set to zero. The front, back,
left and right side of the domain were set as a SymmetryPlane. For a SymmetryPlane the
normal velocity and velocity gradient are zero and the normal pressure gradient is zero.
The settings used in OpenFOAM (numerical schemes and the linear solvers including tol-
erances) are stated in Appendix E.3.
F.J. Venneman MSc. Thesis
5.3 3D numerical model 61
(a) (b)
Figure 5.10: Boundary conditions of the three-dimensional case. (a) velocity, (b) pressure.
5.3.3 Setup cases in three dimensions
A couple of dierent cases were tested for the three-dimensional simulation. Five param-
eters were varied in order to see whether the time-history of the force on the wing was
sensitive to this variation. The following parameters were varied:
1. The Courant number, to check the sensitivity to the numerical time-step.
2. The grid density, to check the sensitivity to the cell sizes.
3. The wing planform, to check whether an ellipsoidal wing would yield another time-
history of the force than a cad-wing.
4. The aspect ratio of the wing, to check the sensitivity to the chord length.
5. The kinematics of the start, to check the sensitivity to the initial acceleration.
An inventory of the ve investigated cases is given in the paragraphs below and the results
are shown in section 5.3.4. The values that are compared, are the time-histories of the
force-coecients, C
L
(t) and C
D
(t). In order to obtain force-coecients of the experiment,
the forces measured by a sensor on the Roboy-wing are non-dimensionalized with the
following values:
= 880 kg m
3
,
R = 0.25 m ,

max
= 1.181 rad s
1
,
U
2
tip
= (
max
R)
2
= 0.087 m
2
s
2
,
A = 0.0167 m
2
.
MSc. Thesis F.J. Venneman
62 Determining forces on insect wings numerically
time [s]
C
D
/
C
L
[
-
]
CD
CL
0 0.5 1 1.5 2
0
0.5
1
1.5
Figure 5.11: Force-coecients of the measurement. Only
2
3
of the stroke is shown because
the three-dimensional simulation was limited to a 90

rotation (till t 1.5 s).


Where the value for A is obtained from Bos (2005, pag.32). It should be noted that the
force-coecients presented in this part of the thesis are in the frame of reference rotating
with the wing (similar to the results presented in the experimental part of the thesis). To
give an impression of the size of the force-coecients as a function of time, they are shown
in Figure 5.11.
Variation of Courant number
The rst case compares the force-coecients in time for various Courant numbers. Just as
in the two-dimensional case, the maximum Courant number is decreased from 2, 1 to
1
2
.
The tests were performed on a single grid of 280k cells, so in fact the time-step was varied.
Variation of grid density
The second case compares the force-coecients resulting from two dierent grid densities.
Grids consisting of 280k and 530k cells, were compared. The maximum Courant number
in this computation was 2.5.
Variation of wing planform
The third case investigates the inuence of the planform of the wing on the forces. This case
was executed to investigate whether an ellipsoidal wing would yield another time-history
of the force than the cad-wing. An ellipsoidal wing of order 2.25 was considered on a 480k
mesh. The force-coecients were compared with the force-coecients on the cad-wing
(278k cells, Co
max
= 0.5). The ellipsoidal wing possessed approximately the same aspect
ratio as the cad-wing, i.e. the aspect ratio of the cad-wing was 3.1, whereas the aspect
ratio of the ellipsoidal wing was 3.4 (see Figure 5.12(b)). A study on the eect of the wing
planform on a rotating impulsively-started insect wing was performed by Luo and Sun
F.J. Venneman MSc. Thesis
5.3 3D numerical model 63
(2005). While keeping the aspect ratio and the radius constant they found that varying
the planform could result in reasonably dierent time-histories of the force (up to 33%).
Variation of the aspect ratio of the wing
A drawback of the cad-wing (as shown in Figure 5.9(b)), was its small chord length. This
caused its aspect ratio to be smaller than the Roboy-wing. Whereas the wing aspect ratio
of the Roboy-wing is 1.9 (with b = 0.18 m), the aspect ratio of the cad-wing was 3.1 (see
Figure 5.12(a)). This is a non-negligible dierence, of which the inuence is investigated
in this section. Two ellipsoidal model wings (of order 2.25) were compared with aspect
ratios comparable to the cad-wing and the Roboy-wing (the chord length was changed,
the span remained constant). The cad-like ellipsoidal wing possessed an aspect ratio of
3.4 (the same wing as tested in the paragraph above, see Figure 5.12(b)) and the Roboy-
like ellipsoidal wing possessed an aspect ratio of 2.2 (see Figure 5.12(c)). To make this
paragraph a little clearer, the aspect ratios are summarized in Table 5.3.3.
Luo and Sun (2005) studied the eect of varying the aspect ratio while keeping the shape
of the wing constant (for an impulsively-started insect wing). They found that when the
aspect ratio is increased, on one hand, the aerodynamic force coecients are increased
due to the reduction of 3D ow eects, and on the other hand, they are decreased due to
the shedding of the part of the lev at the outer part of the wing; these two eects may
approximately cancel each other, resulting in only a minor change of the force-coecients.
So it is expected that variation of the aspect ratio of the wing, leads to negligible changes
in the time-history of the force.
(a) cad-wing (b) ellipsoidal
wing, aspect
ratio 3.4
(c) ellipsoidal
wing, aspect
ratio 2.2
Figure 5.12: Planforms of the dierent wings used for the simulations.
MSc. Thesis F.J. Venneman
64 Determining forces on insect wings numerically
Roboy-wing cad-wing
real 1.9 3.1
ellipsoidal copy 2.2 3.4
Table 5.3: This table shows the aspect ratios of the Roboy-wing and cad-wing. Their
ellipsoidal copies used in the computational comparison of dierent aspect ratios are stated
on the bottom of the table.
Variation of kinematics of the start
The kinematics of the beginning of the stroke of the wing were varied to check whether
a change in kinematics would be of any inuence to the force-coecients. Two kinematic
models were investigated, in which the rotational velocity was altered. The rst one is a
small correction of the original kinematics, but with less acceleration at the beginning,
and is therefore denoted as smoothed. The assigned rotational velocity, as described in
section 3.2 was corrected as follows:
(t) =
max
_
1 exp
_

t
t
c
__
. .
original (t)
2 sin(t) exp
_
2t
t
c
_
. .
correction
, (5.15)
The nal rotational velocity in (5.15) is equal to the original velocity, so the nal Reynolds
number does not change. Dierentiating the new rotational velocity with respect to time
yielded the rotational acceleration,

t
, integrating it (and setting = 0 radat t = 0 s)
yielded the angle of rotation (t). These are shown in Figure 5.3.3 (denoted as smoothed
kinematics). As can be seen, the smoothed acceleration (black dash-dotted line) starts
from approximately 0 rad s
2
, whereas the original acceleration (red line) starts from 7
rad s
2
. The second kinematic model is an adapted version of the kinematics described in
Luo and Sun (2005). This model consists of a delayed (but large) acceleration for the rst

a
seconds and is therefore referred to as delayed. The rotational velocity is a piecewise
continuously dierentiable function, described as:
(t) =
_
t
a
max
a
_
t
a

sin
_
t
a
__
,
t >
a

max
,
(5.16)
where
a
is set to 0.3 s in order to let the total angle of displacement be approximately
equal to the original kinematics for the whole stroke. The rotation, rotational velocity
and rotational acceleration are also shown in Figure 5.3.3 (green dashed line denoted as
delayed acceleration).
The new kinematics was tested on the cad-wing (280k mesh, Co = [0.5, 1, 2]).
5.3.4 Discussion of the three-dimensional results
In Figure 5.14 the force-coecients are shown for the dierent cases. The initial part of
the stroke (rst 0.1 s) is not shown, because the computational time-step was too large to
F.J. Venneman MSc. Thesis
5.3 3D numerical model 65
time (s)

(
r
a
d
)
derived from measurements
original kinematics
smoothed kinematics
delayed acceleration
0 0.5 1 1.5
0
0.5
1
1.5
(a) Rotation
time (s)

(
r
a
d
s

1
)
derived from measurements
original kinematics
smoothed kinematics
delayed acceleration
0 0.5 1 1.5
0
0.2
0.4
0.6
0.8
1
1.2
(b) Angular velocity
time (s)

t
(
r
a
d
s

2
)
derived from measurements
original kinematics
smoothed kinematics
delayed acceleration
0 0.5 1 1.5
0
1
2
3
4
5
6
7
8
(c) Angular acceleration
Figure 5.13: Rotation (a), angular velocity (b) and angular acceleration (c) of the dierent
kinematics of the impulsively-started wing.
MSc. Thesis F.J. Venneman
66 Determining forces on insect wings numerically
properly resolve the impulsive start for the nite acceleration of 7 rad s
2
and therefore the
force-coecients exhibited non-physical behavior. In general the force-coecients in the
steady part of the stroke (0.7 < t < 1.5) correspond reasonably well to the experimentally
obtained coecients. The drag-coecient and the lift-coecient are within 11% and 22%
of the experimentally measured values, respectively (for the solution where the Courant
number is varied). The force peak near t = 0.3 however, does not appear in the computed
coecients, although a small bump is visible. This force peak is expected though, since it
appears in the sensor measurements. The fact that the force is absent in the present study
can be the result of several aspects that were not taken into account (suciently):
1. The kinematics of the wing were not exactly known. An equation for the rotational
velocity was described in Poelma et al. (2006), but provided angles of rotation by
Dr. Ir. Poelma himself did not exactly agree with this it. An exponential t was
performed on the provided data (see Appendix B.3), to mimic the kinematics as good
as possible, but these kinematics do not yield a force peak at the start of the stroke.
A sensitivity analysis is performed to the inuence of the initial kinematics of the
wing (see later in this section), which shows that the inuence can be signicant (a
force peak arises for delayed initial acceleration). This makes it presumable that the
initial acceleration is the main reason for the absence of a force peak.
2. Flexible characteristics of the wing are not taken into account in the simulation, this
could be a cause of the absence of the force peak. The piv measurements showed
that the wing was a little exible. Taking Fluid Structure Interactions (fsi) into
account would make the simulation much more realistic.
3. The mesh is not ne enough to resolve the vortical structures around the wing (see
Figure 5.9(c)). Although a small mesh renement study is performed, which indicates
that the solution does not vary signicantly while increasing the mesh density (see
later in this section), rening the mesh could lead to better compliance with the
measured force peak. However, ner meshes lead to highly skew cells around the
sharp corners of the wing, so they were not used in the present study.
The results of the dierent cases (as described in the previous section) are listed in the
remainder of this section.
Variation of Courant number
First the Courant number is varied from 2, 1 to
1
2
. The results for Co = 1 are depicted
in Figure 5.14(a), in which a comparison of the experimental and computational force-
coecients is shown.
The variations of the force-coecients for various Courant numbers cannot be visualized on
the scale of Figure 5.14(a), being of the order of 0.1%. n general the force-coecients in the
steady part of the stroke (0.7 < t < 1.5) correspond reasonably well to the experimentally
obtained coecients. The drag is within 11% of the measured values and the lift is within
22% of the measured values.
F.J. Venneman MSc. Thesis
5.3 3D numerical model 67
Variation of grid density
The force-coecients for the two grids of 280k and 530k cells are shown in Figure 5.14(b)
in which no signicant dierence is shown. The lift and drag are overall a little lower on
the ne grid.
Variation of wing planform
The force-coecients for the cad-wing and the ellipsoidal wing are compared to the ex-
perimental force measurements in Figure 5.14(c). In general, the forces on the cad-wing
are slightly larger than on the ellipsoidal wing, but the planform of the wing does not
seem to alter the force-coecients to a large extent. This is not totally as expected, since
Luo and Sun (2005) showed that the time-histories of the force could change up to 33%,
but that is for a very wide variety of planforms. Besides that their kinematics is a little
dierent and the point of rotation is on the base of the wing, so we are not worried about
this dierence.
Variation of the aspect ratio of the wing
The force-coecients for the two ellipsoidal wings with aspect ratios of 2.2 and 3.4 are com-
pared to the experimental force measurements in Figure 5.14(d). The dierence between
the force-coecients of the AR = 2.2 and AR = 3.4 wing is approximately 10%, the forces
being higher on the wing with a lower aspect ratio. The wing with the aspect ratio almost
similar to the Roboy compares best to the measured force-coecients, especially in the
steady part of the stroke. According to Luo and Sun (2005) the aspect ratio should not
change signicantly for reasons explained in 5.3.3, 10% is not signicant nor insignicant.
However, the computational uid dynamics diered with the present study for at least two
reasons mentioned in the previous paragraph, so a true comparison is dicult.
Variation of kinematics of the start
The force-coecients for the dierent kinematics are compared with the original kinematics
and the experimental force measurements in Figure 5.14(e). For dierent Courant numbers
no signicant variations appeared (maximum dierence around 1%).
The drag of the wing with smoothed kinematics starts a little lower and is a little higher
at t = 0.4 s (where the experimental data show a peak as well), but no signicant dierence
with the original data appears. The delayed kinematics, however, do show a signicant
dierence with the original kinematics. Clearly a force peak is shown for both lift and
drag at t = 0.3 s (this is actually the moment that the acceleration stops, see Figure
5.13(c)). In the steady part of the stroke the original and delayed kinematics converge to
the same force-coecients again. This is a promising results, since we now can state that
the force-coecients are sensitive to initial the kinematics.
MSc. Thesis F.J. Venneman
68 Determining forces on insect wings numerically
time [s]
C
D
/
C
L
[
-
]
CD experiment
CL experiment
CD simulation
CL simulation
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
0
0.5
1
1.5
(a) Variation of Courant number and
kinematics. Co = 1
time [s]
C
D
/
C
L
[
-
]
CD experiment
CL experiment
CD 280k
CL 280k
CD 530k
CL 530k
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
0
0.5
1
1.5
(b) Variation of grid density. 280k
and 530k cells.
time [s]
C
D
/
C
L
[
-
]
CD experiment
CL experiment
CD CAD-wing
CL CAD-wing
CD ellipsoidal wing
CL ellipsoidal wing
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
0
0.5
1
1.5
(c) Variation of planform. cad-wing
versus ellipsoidal wing.
time [s]
C
D
/
C
L
[
-
]
CD experiment
CL experiment
CD ellips. wing AR = 2.2
CL ellips. wing AR = 2.2
CD ellips. wing AR = 3.4
CL ellips. wing AR = 3.4
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
0
0.5
1
1.5
(d) Variation of aspect ratio.
time [s]
C
D
/
C
L
[
-
]
CD experiment
CL experiment
CD original kinematics
CL original kinematics
CD smoothed kinematics
CL smoothed kinematics
CD delayed kinematics
CL delayed kinematics
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
(e) Variation of kinematics
Figure 5.14: Comparison of experimental and computational force-coecients of an
impulsively-started wing for a rotation of 90

.
F.J. Venneman MSc. Thesis
5.3 3D numerical model 69
5.3.5 Qualitative images of the ow
To get an idea of the ow eld around the cad-wing, iso-surfaces of the second invariant of
the gradient tensor, Q (for an explanation see section 5.1), are shown in Figure 5.15. From
this gure it can be concluded that although the force-coecients on the wing are rather
stable, the ow eld around the wing shows an appreciable development. For a detailed
description of the ow-phenomena, the reader is referred to the caption of Figure 5.15.
To compare the pressure-elds obtained from the simulations and the planar Poisson ap-
proach, Figure 5.16 shows consecutive cross-sections of the pressure-elds of both. The
cross-sections are provided for three dierent time steps (t = 0.3 s, t = 0.8 s, t = 1.5 s) at
0.66R of the wing (with R the spanwidth of the wing). For details about the cross-sections
from the planar Poisson approach, one is referred to section 4.3.2. It should be noted that
the pressure-elds in Figure 5.16 do not have the same scale, one can only distinguish the
locations of the extrema. What is most striking in Figure 5.16, is the relative dierence
of pressure in the far-eld (left and right of the wing). For the planar Poisson approach,
the maximum pressure is in the far eld on both sides of the wing, whereas the pressure
maximum in case of the simulations is in front of the wing. The computed pressure-elds
comply better with what one would expect, since there is a clear pressure minimum near
the leading-edge and a pressure maximum in front of the wing.
MSc. Thesis F.J. Venneman
70 Determining forces on insect wings numerically
(a) t = 0.4 s (b) t = 0.6 s
(c) t = 0.8 s (d) t = 1.1 s
(e) t = 1.3 s (f) t = 1.5 s
Figure 5.15: Iso-Q-surfaces of the ow around the cad-wing (530k mesh). The value of Q
is +10, so the surfaces are regions where the swirling motion is dominant. The colors on the
surfaces denote the helicity-value which shows whether the local vorticity is aligned with the
local velocity (red or blue) or perpendicular to it (green). The arrow points in the direction
of the velocity of the wing. (a. t = 0.4 s) The leading-edge vortex (lev) and trailing-edge
vortex (tev) are developed, a tip vortex connects the lev with the tev. (b. t = 0.6 s) The
tev sheds from the wing, whereas the lev remains on the wing. (c. t = 0.8 s) The tev
separates from the tip vortex, the lev at the tip of the wing separates from the wing. (d.
t = 1.1 s) A larger part of the lev sheds, the tev is now almost perpendicular to the wing
(e. t = 1.3 s) A new tip vortex on the bottom of the tip is formed. (f. t = 1.5 s) End of the
stroke, fairly stable Q surfaces with respect to the previous time step.
F.J. Venneman MSc. Thesis
5.3 3D numerical model 71
(a) simulation t = 0.3 s (b) planar Poisson approach
t = 0.3 s
(c) simulation t = 0.8 s (d) planar Poisson approach
t = 0.8 s
(e) simulation t = 1.5 s (f) planar Poisson approach
t = 1.5 s
Figure 5.16: Comparison of the pressure-elds of the simulations (left) and the planar Poisson
approach (right). Cross-sections are obtained for dierent time steps (t = 0.3 s, t = 0.8
s, t = 1.5 s) at 0.66R of the wing (with R the spanwidth). It should be noted that the
pressure-elds do not have the same scale, one can only distinguish the locations of the
extrema. Most remarkable is the relative dierence in the far eld of the pressure.
MSc. Thesis F.J. Venneman
72 Determining forces on insect wings numerically
5.4 Conclusion: Good estimation of force-coecients on
steady stroke
In this section the conclusions will be drawn about the numerical part of the present study.
The most important conclusions are drawn in section 5.4.2, in which the three-dimensional
results are discussed. The preliminary calculations in two dimensions are rst discussed in
section 5.4.1.
5.4.1 Two-dimensional simulations
The preliminary calculations in two dimensions showed time-histories of the force and
vorticity patterns that were useful for the calculations in three dimensions. The most
important conclusion of the two-dimensional computations are:
1. The vortex pattern in the wake of quiescent wing at constant angle of attack (50.4

)
in a steady translating ow (Re = 192) causes the force-coecients to oscillate.
The average lift and drag on this wing do converge for decreasing the time step or
increasing grid density. Convergence is more sensitive to doubling the cell density
than to doubling the amount of time steps.
2. The computed lift and drag on an impulsively-started wing, at constant angle of
attack and Re = 192, show reasonable agreement with measurements performed by
Dickinson and Gotz (1993).
3. It is conrmed that increasing the Reynolds number (from 120 to 330) of the ow
around a wing with a constant angle of attack, yields approximately the same tran-
sition of vorticity in the wake, as for a circular cylinder (for the same Reynolds
domain).
5.4.2 Three-dimensional simulations
The time-history of the force-coecients on the simulated three-dimensional rotating wing
corresponded reasonably well to experimentally obtained force-coecients. Especially for
the steady part of the stroke, in which the rotational velocity is constant. The drag and lift
are maximally within 11% and 22% of the experimentally measured values, respectively.
The peak in the force-coecients at the start of the stroke, however, did not appear in
the force-histories of the simulations. Three aspects were mentioned at the beginning of
section 5.3.4 that can be the cause of this absence, which will briey be repeated here:
1. Kinematics of the wing could have varied, this is accepted as the most presumable
cause.
2. No exibility and uid structure interactions taken into account.
F.J. Venneman MSc. Thesis
5.4 Conclusion: Good estimation of force-coecients on steady stroke 73
3. Mesh density not ne enough.
Some conclusions can also be drawn according to the dierent cases that were presented
in section 5.3.3.
1. In the region of 0.1 < t < 1.5 s, the force-coecients were not sensitive to a change
in Courant number at the same grid density (so in changing the time step), for the
range of
1
2
< Co < 2.
2. A grid density of 280k cells seems enough for a proper simulation of the forces around
an impulsively-started insectwing.
3. Altering the planform (while keeping the wing aspect ratio approximately constant) of
the wing does not change the forces to a large extent for the present wing kinematics.
4. For the present study, the force-coecients are more sensitive to the aspect ratio of
the wing than to the planform of the wing. This can be concluded from the fact
that the time-history of the force on an ellipsoidal wing with the same aspect ratio
as the Roboy-wing, corresponded better to the experimentally obtained forces than
the cad-wing (with a corresponding planform, but a dierent aspect ratio).
5. The force-coecients are very sensitive to the initial kinematics of the wing of the
present study. Since the initial kinematics of the experimental wing are not known
exactly, the initial force-coecients can not truly be compared. It is shown that
alternative initial kinematics lead to a force peak. This suggests that the initial
kinematics of the present test case could have been dierent.
MSc. Thesis F.J. Venneman
74 Determining forces on insect wings numerically
F.J. Venneman MSc. Thesis
Chapter 6
Conclusions and recommendations
A study is performed on an impulsively-started wing of a fruit-y (Drosophila
Melanogaster). Both the forces on the wing and the ow-eld are studied from an
experimental and a computational point of view. The main challenge of the present
study is to compare forces, deduced from experiments and computations, with each other.
The forces in the steady part of the stroke, obtained using both techniques, are in good
agreement with each other and the experiment.
The present research concerns an impulsively-started, rotating wing at a Reynolds
number of 256 at a constant angle of attack ( = 50.4

). Experimentally obtained
velocity-elds around this wing are provided by Dr. Ir. C. Poelma. He performed
time-resolved, stereoscopic piv measurements around a dynamically-scaled wing, known
as the Roboy (Poelma et al., 2006).
Besides force acquisition from the force sensor mounted on the Roboy-wing, the
present study applies a new tool to obtain forces, through a combination of the momen-
tum approach and an algorithm to obtain pressure-elds from velocity-elds, called the
planar Poisson approach. This method deduces the pressure-elds from the velocity-elds
by making use of the ow constitutive equations. The planar Poisson approach has
successfully been tested on two-dimensional ows, but never on three-dimensional ows.
For the computational simulations, the kinematics of the Roboy-wing are mim-
icked as accurately as possible. The Navier-Stokes equations are solved using OpenFOAM
as a framework. The incompressible ow equations are solved on a dynamically deforming
mesh. This mesh deforms based on Radial Basis Function interpolation, which is a new
mesh deformation tool developed at the Aerodynamics group of the Delft University of
Technology.
After investigating the experimental (E) and numerical (N) results, the following
main conclusions can be drawn.
MSc. Thesis F.J. Venneman
76 Conclusions and recommendations
E.1 The forces on the impulsively-started wing predicted by the combination of the planar
Poisson approach and the momentum approach correspond reasonably well to the
forces measured by the sensor mounted on the wing. For the steady part of the
stroke, the drag-coecient and the lift-coecient are within 11% and 22% of the
experimentally measured values, respectively
E.2 The estimated initial drag using the planar Poisson approach and the momentum
approach is overestimated. This is probably caused by an initial acceleration, that
was assumed to be too large.
E.3 The pressure-elds in the vicinity of the wing, determined by the planar Poisson
approach, seem physical. The pressure-elds in the far eld, do not seem physical,
this is probably caused by the lack of resolution of the velocity data.
N.1 The time-history of the force-coecients on the simulated three-dimensional rotat-
ing wing corresponds reasonably well to experimentally obtained force-coecients.
Especially the force-coecients of the steady part of the stroke agree well, the drag-
coecient and the lift-coecient are within 13% and 4% of the experimentally mea-
sured values, respectively
N.2 In the simulations a force peak in the initial stage of the stroke is almost absent.
This is most likely the result of the fact that the initial kinematics prescribed for
the motion of the wing were not equal to those in the actual kinematics of the
measurement. It is shown that the force-coecients are very sensitive to varying the
initial acceleration.
Conclusions E.2 and N.2 agree with each other, since they both indicate that the initial
acceleration used in the present study, does not comply with the acceleration of the exper-
imental work.
Details about the experimental conclusions can be found in section 6.1. This is followed by
a detailed description of the numerical conclusions in section 6.2. At the end of this chapter
some recommendations for future research on insect wings will be provided in section 6.3.
6.1 Detailed conclusions experimental force determina-
tion
The conclusions about the experimental force determination can be divided in two groups;
the deduction of the forces and the deduction of the pressure. First some general conclu-
sions are drawn:
1. Especially the deduced forces in the steady part are in good agreement with the force
obtained from the sensor.
F.J. Venneman MSc. Thesis
6.2 Detailed conclusions numerical force determination 77
2. The momentum approach heavily depends on the pressure deduction (planar Poisson
approach), since unphysical forces are predicted, once an unrealistic pressure-eld is
prescribed.
3. The force is highly sensitive to the initial assumed acceleration, it has an eect on
both the pressure deduction and the force deduction, since the term appears in both
algorithms.
Forces
1. The forces on the impulsively-started wing predicted by the momentum approach
correspond reasonably well to the forces measured with a sensor mounted on the
Roboy. Especially in the steady part of the stroke of the wing (t > 0.7 s), where the
drag-coecient and the lift-coecient are within 13% and 4% of the experimentally
measured values, respectively. These measured forces are assumed to be realistic, yet
Dr. ir. C. Poelma informed me that the force sensors had suered from wear and
could well be 10% o. Especially the forces in the steady part of the stroke match
very well to the experimentally obtained forces.
2. The momentum approach is a better predictor of forces than the Blasius approach,
specically when vortices have left the eld of view.
3. The initial drag is overpredicted (and varies in a large domain for dierent integration
contours), whereas the initial lift corresponds rather good to the measured forces.
This is probably caused by an overestimated initial acceleration (which only inuences
the drag directly).
Pressure
1. The pressure-elds around the wing look quite as expected, i.e. there is a pressure
minimum near the leading-edge and a small pressure maximum in front of the wing.
2. However, during force determination, it turned out that only integrating the pressure-
eld in the vicinity of the wing delivered reliable results. The far eld of the pressure
seems rather unreliable. This may be caused by a lack of resolution of data points.
The assumption that the pressure-elds in the far eld of the wing are unreliable is
conrmed by the pressure-elds obtained, using computational dynamics.
6.2 Detailed conclusions numerical force determination
The conclusions about the numerical force determination are also divided into two groups;
the two-dimensional simulations and the three-dimensional simulations.
MSc. Thesis F.J. Venneman
78 Conclusions and recommendations
Two-dimensional simulations
In two dimensions, a quiescent wing in an impulsively-started translating ow is simulated.
A number of conclusions can be drawn according to these simulations.
1. The vortex pattern in the wake of quiescent wing at constant angle of attack (50.4

)
in a steady translating ow (Re = 192) causes the force-coecients to oscillate. The
average lift and drag on this wing do converge for a decreasing time step and an
increasing grid density. Convergence is more sensitive to doubling the cell density
than to doubling the amount of time steps.
2. The computed lift and drag on an impulsively started wing at constant angle of attack
at Re = 192, show reasonable agreement with earlier measurements from literature.
3. It is conrmed that increasing the Reynolds number (from 120 to 330) of the ow
around a wing with a constant angle of attack, yields approximately the same tran-
sition of vorticity in the wake, as for a circular cylinder (for the same Reynolds
domain).
Three-dimensional grid
The time-history of the force-coecients on the simulated three-dimensional rotating wing,
correspond reasonably well to experimentally obtained force-coecients. Especially for the
steady part of the stroke, in which the rotational velocity is constant (t > 0.7 s), where
the drag-coecient and the lift-coecient are within 11% and 22% of the experimentally
measured values, respectively. The peak in the force-coecients at the start of the stroke,
however, does not appear in the force-histories of any of the simulations. This is most likely
the result of the use of kinematics which are slightly dierent compared to the experiments.
Some conclusions can also be drawn according to the dierent cases that were presented
in section 5.3.3.
1. The force-coecients were not sensitive to a change in Courant number at the same
grid density (so in fact, changing the time step), for the range of
1
2
< Co < 2.
2. A grid density of 280k cells has proven to be sucient for a proper simulation of the
forces around the impulsively-started wing.
3. Altering the planform (while keeping the wing aspect ratio approximately constant) of
the wing does not change the forces to a large extent for the present wing kinematics.
4. For the present study, the force-coecients are more sensitive to the aspect ratio of
the wing than to the planform of the wing.
5. The force-coecients are very sensitive to the distribution of the acceleration of the
rotating wing of the present study.
F.J. Venneman MSc. Thesis
6.3 Recommendations 79
6.3 Recommendations
Like in every investigation, solving one questions, yields many more questions. For future
work, the following items could be investigated:
1. Dierent kinematics could be investigated in the same manner. It would, for ex-
ample, be interesting to see how the forces compare on a hovering wing. In order
to validate computational uid dynamics with experiments it is very important to
exactly prescribe the same kinematics.
2. Insect ight can be modelled less simplied. One can model two wings instead of
one, model the body or investigate the inuence of uid structure interactions (fsi).
3. Once changing shapes and kinematics, it would be very challenging to maintain a
mesh of good quality for a wide range of rotation angles using mesh deformation
based on Radial Basis Function interpolation.
4. It would be interesting to see what would change in the pressure-elds deduced by
the planar Poisson approach, using velocity-elds with better resolution. One can
obtain these data by piv or even by cfd (and see whether the pressure-eld agree
with the earlier computed pressure-elds).
MSc. Thesis F.J. Venneman
80 Conclusions and recommendations
F.J. Venneman MSc. Thesis
Bibliography
H. Aono, F. Liang, and H. Liu. Near- and far-eld aerodynamics in insect hovering ight:
an integrated computational study. The Journal of Experimental Biology, 211:239257,
2008. URL http://dx.doi.org/10.1242/jeb.008649. [13, 14]
G. K. Batchelor. An Introduction to Fluid Dynamics. Cambrige University Press, 1967.
[47, 90]
T. Baur and J. Kongeter. Piv with high temporal resolution for the determination of
local pressure reductions from coherent turbulence phenomena. pages 671676, Santa
Barbara, Sep 1999. [27]
J. M. Birch and M. H. Dickinson. The inuence of wingwake interactions on the production
of aerodynamic forces in apping ight. The Journal of Experimental Biology, 206:2257
2272, March 2003. URL http://dx.doi.org/10.1242/jeb.00381. [24, 25, 26, 47]
A. de Boer, M. S. van der Schoot, and H. Bijl. Mesh deformation based on radial basis
function interpolation. Computers and Structures, 85(11-14):784795, June 2007. URL
http://dx.doi.org/10.1016/j.compstruc.2007.01.013. [16, 60, 102]
F. M. Bos. Inuence of wing kinematics on performance in insect ight. Masters thesis,
Delft University of Technology, 2005. [4, 7, 62]
F. M. Bos, D. Lentink, B. W. van Oudheusden, and H. Bijl. Inuence of wing kinematics
on aerodynamic performance in hovering insect ight. Journal of Fluid Mechanics, 594:
341368, 2008. URL http://dx.doi.org/10.1017/s0022112007009172. [12, 47, 48]
P. Chakraborty, S. Balachandar, and R. J. Adrian. On the relationships between local
vortex identication schemes. Journal of Fluid Mechanics, 535:189214, 2005. URL
http://dx.doi.org/10.1017/S0022112005004726. [46]
M. H. Dickinson and K. G. Gotz. Unsteady aerodynamic performance of model wings
at low reynolds numbers. The Journal of Experimental Biology, 174(1):4564, January
1993. URL http://jeb.biologists.org/cgi/content/abstract/174/1/45. [10, 50,
52, 53, 54, 56, 72]
MSc. Thesis F.J. Venneman
82 BIBLIOGRAPHY
M. H. Dickinson, F. O. Lehmann, and S. P. Sane. Wing rotation and the
aerodynamic basis of insect ight. Science, 284(5422):19541960, 1999. URL
http://dx.doi.org/10.1126/science.284.5422.1954. [1, 4, 10, 15, 58]
C. P. Ellington. The aerodynamics of hovering insect ight. Phil. Trans. R. Soc. Lond.,
1984. [47]
C. P. Ellington, C. van den Berg, A. P. Willmott, and A. L. R. Thomas. Leading-edge
vortices in insect ight. Nature, 384:626630, 1996. [1, 10, 11]
J. H. Ferziger and M. Peric. Computational Methods for Fluid Dynamics. Springer, 2
nd
edition, 1999. [103, 104, 112]
S. N. Fry, R. Sayaman, and M. H. Dickinson. The aerodynamics of free-
ight maneuvers in drosophila. Science, 300(5618):495498, 2003. URL
http://dx.doi.org/10.1126/science.1081944. [13, 14, 15]
S. N. Fry, R. Sayaman, and M. H. Dickinson. The aerodynamics of hovering ight
in drosophila. The Journal of Experimental Biology, 208:23032318, 2005. URL
http://dx.doi.org/10.1242/jeb.01612. [13, 14, 15]
N. Fujisawa, S. Tanahashi, and K. Srinivas. Evaluation of pressure eld and uid forces on
a circular cylinder with and without rotational oscillation using velocity data from piv
measurement. Measurement Science and Technology, 16(4):989996, April 2005. URL
http://dx.doi.org/10.1088/0957-0233/16/4/011. [27]
R. E. Gordnier. High delity computational simulation of a membrane wing airfoil. Amer-
ican Institute of Aeronautics and Astronautics, Jan 2008. [11]
D. T. Greenwood. Principles of dynamics. Prentice-Hall, Inc., 2
nd
edition, 1988. [9]
R. Gurka, A. Liberzon, D. Hefetz, D. Rubinstein, and U. Shavit. Computation of pressure
distribution using piv velocity data. pages 671676. 3rd International Workshop PIV
(Santa Barbara), Jan 1999. [16, 27, 28]
H. Jasak. Error Analysis and Estimation for the Finite Volume Method with Applications
to Fluid Flows. PhD thesis, Imperial College, 1996. [46, 106]
R. de Kat, B. W. van Oudheusden, and F. Scarano. Instantaneous planar pressure eld
determination around a square-section cylinder based on time-resolved stereo-piv. Sym-
posium on Applications of Laser Techniques to Fluid Mechanics (Lisbon), July 2008.
[27]
P. K. Kundu and I. M. Cohen. Fluid Mechanics. Academic Press, 2
nd
edition, 2002. [92]
D. Lentink. Inuence of airfoil shape on performance in insect ight. American Insitute of
Aeronautics and Astronautics, Jun 2003. [12]
F.J. Venneman MSc. Thesis
BIBLIOGRAPHY 83
G. Luo and M. Sun. The eects of corrugation and wing planform on the aerodynamic
force production of sweeping model insect wings. Acta Mechanica Sinica, 21:531541,
2005. URL http://dx.doi.org/10.1007/s10409-005-0072-4. [14, 15, 62, 63, 64, 67]
F. Noca, D. Shiels, and D. Jeon. A comparison of methods for evaluating time-
dependent uid dynamic forces on bodies, using only velocity elds and their
derivatives. Journal of Fluids and Structures, 13(5):551578, July 1999. URL
http://dx.doi.org/10.1006/jfls.1999.0219. [25]
B. W. van Oudheusden, F. Scarano, E. W. M. Roosenboom, E. W. F. Casimiri, and L. J.
Souverein. Evaluation of integral forces and pressure elds from planar velocimetry data
for incompressible and compressible ows. Experiments in Fluids, 43:153162, 2007.
URL http://dx.doi.org/10.1007/s00348-008-0546-9. [27]
U. Piomelli, E. Balaras, and A. Pascarelli. Turbulent structures in accel-
erating boundary layers. Journal of Turbulence, 1(1), May 2000. URL
http://dx.doi.org/10.1088/1468-5248/1/1/001. [46]
C. Poelma, W. B. Dickson, and M. H. Dickinson. Time-resolved reconstruction of the full
velocity eld around a dynamically-scaled apping wing. Experiments in Fluids, 41(2):
213225, August 2006. URL http://dx.doi.org/10.1007/s00348-006-0172-3. [5, 8,
13, 16, 17, 18, 20, 24, 25, 30, 34, 35, 43, 46, 50, 66, 75, 98]
R. Ramamurti and W. C. Sandberg. A three-dimensional computational study of the aero-
dynamic mechanisms of insect ight. The Journal of Experimental Biology, 205:1507
1518, March 2002. URL http://jeb.biologists.org/cgi/reprint/205/10/1507.
[13]
R. Ramamurti and W. C. Sandberg. A computational investigation of the
three-dimensional unsteady aerodynamics of drosophila hovering and maneu-
vering. The Journal of Experimental Biology, 12:881896, 2007. URL
http://dx.doi.org/10.1242/jeb.02704. [13, 14]
A. Roshko. On the development of turbulent wakes from vortex streets. National Advisory
Committee for Aeronautics, 1954. [53]
S. P. Sane and M. H. Dickinson. The aerodynamic eects of wing rotation and a revised
quasi-steady model of apping ight. The Journal of Experimental Biology, 205(8):
10871096, 2002. [4]
B. Singh and I. Chopra. Insect-based hover-capable apping wings for micro air ve-
hicles: experimental and analysis. AIAA Journal, 46(6):21152135, 2008. URL
http://dx.doi.org/10.2514/1.28192. [15, 16]
R. B. Srygley and A. L. R. Thomas. Unconventional lift-generating mech-
anisms in free-ying butteries. Nature, 420:660664, Dec 2002. URL
http://dx.doi.org/10.1038/nature01223. [4, 10]
MSc. Thesis F.J. Venneman
84 BIBLIOGRAPHY
A. Turella. An accuracy assessment of instantaneous planar pressure imaging based on
tr-piv. Masters thesis, Delft University of Technology, 2008. [30]
M. F. Unal, J. C. Lin, and D. Rockwell. Force prediction by piv imaging: a
momentum-based approach. Journal of Fluids and Structures, 11:965971, 1997. URL
http://dx.doi.org/10.1006/jfls.1997.0111. [26]
J. Z. Wang. Two dimensional mechanism for insect hovering. Physical Review Letters, 85
(10):22162219, Sep 2000a. [12]
J. Z. Wang. Vortex shedding and frequency selection in apping ight. Journal of Fluid
Mechanics, 410:323341, 2000b. [4, 12]
J. Z. Wang, J. M. Birch, and M. H. Dickinson. Unsteady forces and ows in low
reynolds number hovering ight: two-dimensional computations vs robotic wing ex-
periments. The Journal of Experimental Biology, 207(3):449460, January 2004. URL
http://dx.doi.org/10.1242/jeb.00739. [15]
T. Weis-Fogh and M. Jensen. Biology and physics of locust ight. i. basic principles in
insect ight. a critical review. Royal Society of London Philosophical Transactions Series
B, 239:415458, jul 1956. [4]
P. Wesseling. Elements of computational uid dynamics. Applied Mathematics, Delft
University of Technology, 2001. [50, 60]
J. C. Wu. Theory for aerodynamic force and moment in viscous ows. AIAA Journal, 19
(4):432441, 1981. [24, 89]
M. M. Zdravkovich. Flow around circular cylinders, volume 1. Oxford University Press,
1997. [53]
F.J. Venneman MSc. Thesis
Appendix A
Derivations of approaches to determine force
on an immersed body
In this appendix two ways of determining forces on an immersed body in a ow eld are
derived. In section A.1 the momentum approach is derived and in section A.2 the Blasius
approach is derived.
A.1 Derivation of the momentum approach
In this section the momentum approach will be explained, explicitly in deriving equation
(4.16). We begin with restating the incompressible Navier Stokes equation (2.6) including
the Coriolios and centripetal acceleration:

_
u
t
+ (u ) u
_
= p +
_

t
r +

r
_
+ 2

u
_
. (A.1)
With the density (assumed constant), u the ow velocity, t the time, p the static ow
pressure, the stress tensor,

the angular velocity and r the displacement vector in the
rotating frame of reference. Applying a volume integral to (A.1) the Navier Stokes equation
in integral form is obtained:

___
V
u
t
dV +
___
V
(u ) udV =

___
V
pdV +
___
V
dV +
___
V

r
_
2

t
rdV . (A.2)
MSc. Thesis F.J. Venneman
86 Derivations of approaches to determine force on an immersed body
Applying the divergence theorem to some of the terms yields:
___
V
(u ) udV =
__
S
(u n) udS , (A.3)
___
V
pdV =
__
S
p ndS , (A.4)
___
V
dV =
__
S
ndS , (A.5)
with n the outward pointing normal. This enables us to write (A.2) as follows:

___
V
u
t
dV +
__
S
(u n) udS =

__
S
p ndS +
__
S
ndS +
___
V

r
_
2

t
rdV . (A.6)
Now consider the control volume as in Figure A.1, this is a spanwise cross section of the
A
E
D
B
C
G
F
I H
V
S
Figure A.1: Control volume. The black body indicates a cross section of the wing. The
dashed line is the integration path (with the arrows indicating the direction of the integration).
wing. Our integration will take place in volume V . The dashed line in the gure indicates
the surface integration path S. All surface integrations on the cuts DE and GH are equal
and opposite to each other, so they cancel each other. Now we divide the surface integrals
in (A.6) into an integral over the outer boundary ABCI and an integral over the wing
F.J. Venneman MSc. Thesis
A.1 Derivation of the momentum approach 87
EFG:

___
V
u
t
dV +
__
ABCI
(u n) udS +
__
EFG
(u n)
. .
=0
udS =

__
ABCI
p ndS
__
EFG
p ndS +
__
ABCI
ndS +
__
EFG
ndS +
+
___
V

r
_
2

t
rdV . (A.7)
Where we assume the velocity normal to the solid non-porous wing is zero, as indicated
in the third term of (A.7). We are interested in the instantaneous force on the wing due
to the uid, which is equal in size, but opposite to the force on the uid due to the wing
(Newtons third law). If we write this out for equation (A.7) and split the equation in a
surface integral and a volume integral we obtain the following:

F(t) =
_

__
EFG
p ndS +
__
EFG
ndS
_
=
__
ABCI
(u n) u p n + ndS +
+
___
V

u
t

r
_
2

t
rdV . (A.8)
Which is the same equation as (4.3). To write the integral more explicit way, we now
switch to Einstein notation. Rewriting the right hand side of equation (A.8) yields:
F
i
(t) =
__
ABCI
u
i
u
j
n
j
pn
i
+
_
u
i
x
j
+
u
j
x
i
_
n
j
dS +
+
___
V

u
i
t

j
_

j
r
k

ijk
_

ijk
2
j
u
k

ijk


j
t
r
k

ijk
dV , (A.9)
where we have used the Levi-Civita symbol
ijk
with

ijk
=
_

_
+1 if (i, j, k) is (1, 2, 3), (3, 1, 2) or (2, 3, 1) ,
1 if (i, j, k) is (3, 2, 1), (1, 3, 2) or (2, 1, 3) ,
0 otherwise: i = j or j = k or k = i .
(A.10)
and the characteristic for Newtonian uids (see equation (5.2)):

ij
=
_
u
i
x
j
+
u
j
x
i
_
. (A.11)
Mind that the following relations are valid in this case (integrating the outside integral as
MSc. Thesis F.J. Venneman
88 Derivations of approaches to determine force on an immersed body
side dx dy
Top > 0 =0
Right =0 < 0
Bottom < 0 =0
Left =0 > 0
Table A.1: Integration conventions
depicted in Figure A.1 clockwise):
n
i
dS = dS
i
,
n
j
dS = dS
j
. (A.12)
We obtain the segments determined by the rotational frame from Appendix B.3. Rewriting
equation (A.9) for i = 1, 2 then yields:
(i = 1) F
1
(t) =
__
ABCI
uu
j
+
_
u
x
j
+
u
j
x
_
dS
j

__
ABCI
pdS
x
+
+
___
V

u
t
+
2
x 2w

t
xdV , (A.13a)
(i = 2) F
2
(t) =
__
ABCI
vu
j
+
_
v
x
j
+
u
j
y
_
dS
j

__
ABCI
pdS
y
+
+
___
V

v
t
dV . (A.13b)
Filling in j = 1, 2 yields:
F
1
(t) =
__
ABCI
u
2
dS
x

__
ABCI
uvdS
y
+
__
ABCI
2
u
x
dS
x
+
+
__
ABCI

_
u
y
+
v
x
_
dS
y

__
ABCI
pdS
y
+
+
___
V

u
t
+
2
x 2w

t
xdV , (A.14a)
F
2
(t) =
__
ABCI
vudS
x

__
ABCI
v
2
dS
y
+
__
ABCI
2
v
y
dS
y
+
+
__
ABCI

_
u
y
+
v
x
_
dS
x

__
ABCI
pdS
y
+
+
___
V

v
t
dV . (A.14b)
F.J. Venneman MSc. Thesis
A.2 Derivation of the Blasius approach 89
where the following relations are valid:
dS
x
= dy ,
dS
y
= dx . (A.15)
This gives:
F
1
(t) =
__
ABCI
u
2
dy
__
ABCI
uvdx
__
ABCI
2
u
x
dy +
+
__
ABCI

_
u
y
+
v
x
_
dx +
__
ABCI
pdy +
+
___
V

u
t
+
2
x 2w

t
xdV , (A.16a)
F
2
(t) =
__
ABCI
vudy
__
ABCI
v
2
dx +
__
ABCI
2
v
y
dx +
+
__
ABCI

_
u
y
+
v
x
_
dy
__
ABCI
pdx +
+
___
V

v
t
dV . (A.16b)
The equations in (A.16) represent the formulations used to compute the lift (A.16b) and
drag (A.16a) over the control volume V. Where lift and drag are dened in respectively
the positive y and x direction. The resulting forces have units N m
1
, because volume
V (see Figure A.1) is a 2D volume. Calculating the force over each cross section and
multiplying by their distance (1 cm, as described in section 3.3) gives the force with unit
N. Equation (A.16) does not take surfaces into account with the normal aligned with the
span of the wing. However the Coriolis terms and the centripetal terms do take spanwise
movement into account. Besides that the results in section 4.3.2 also suggest that the terms
on the surfaces with a normal in the spanwise direction do not play a big role, because the
obtained forces are quite close to the measured forces.
A.2 Derivation of the Blasius approach
In this section the Blasius approach will be explained, explicitly in deriving equation (4.1).
This derivation is very briey performed in (Wu, 1981). We start of with the denition of
the total vorticity:
___
V
dV , (A.17)
and the rst moment of the vorticity eld (or total vorticity moment):
=
___
V
r dV . (A.18)
MSc. Thesis F.J. Venneman
90 Derivations of approaches to determine force on an immersed body
A.2.1 Velocity integrals in large bounded regions
From the continuity and vorticity denition equations, one obtains the following expression
for the velocity vector in three-dimensional ows:
u =
1
2
[r + (ru) (r u)] . (A.19)
Let V
L
be a spherical region with r L. Let L be suciently large so that V
L
contains the
wing and that the total vorticity and the total vorticity moment outside V
L
are negligible.
Integrating (A.19) over V
L
, one obtains:
___
V
L
udV =
1
2
___
V
L
(r ) dV +
1
2
___
V
L
(ru) (r u) dV . (A.20)
Using (xy) = ( x) y + ( y) x and mass conservation the latter integral of (A.20)
can be rewritten:
___
V
L
(ru) (r u) dV =
___
V
L
( r) u + ( u)
. .
=0
r (r u) dV ,
=
___
V
L
( r) u (r u) dV . (A.21)
Applying the divergence theorem and its corollary to the individual terms in (A.21) yields:
___
V
L
( r) udV =
__
S
L
(r n) udS , (A.22)
___
V
L
(r u) dV =
__
S
L
(r u) ndS . (A.23)
Where S
L
is the boundary of volume V
L
. Using the above and x (y z) = y (x z)
z (x y) (A.20) therefore becomes
___
V
L
udV =
1
2
___
V
L
(r ) dV
1
2

__
S
L
(r u) n (r n) udS ,
=
1
2
___
V
L
(r ) dV
1
2

__
S
L
r ( n u) dS . (A.24)
Since the total moment of vorticity is negligible outside S
L
, the rst therm on the right-
hand side of (A.24) is recognized to be , dened by (A.18). In Batchelor (1967, Chapter
2.9) it is shown that the principle of conservation of total vorticity leads to an equation
for the velocity-eld at large distance from an object (wing in our case):
u(r, t) =
1
8

1
r

_
+ terms of order r
n
, n 4 . (A.25)
F.J. Venneman MSc. Thesis
A.2 Derivation of the Blasius approach 91
If L is suciently large, then only the rst term on the right-hand side of (A.25) has a
non-negligible contribution to the last integral in (A.24). Taking r =
_
x
2
+ y
2
+ z
2
on S
L
yields:

1
r
=
x
(x
2
+ y
2
+ z
2
)
3/2

x
+
y
(x
2
+ y
2
+ z
2
)
3/2

y
+
z
(x
2
+ y
2
+ z
2
)
3/2

z
=
r
r
3
. (A.26)
This yields the velocity vector on S
L
(where r = L):
u(r, t) =
1
8

1
r

_
=

8L
3
+
3(r )r
8L
5
. (A.27)
Now we place (A.27) in the second part of (A.24). Noting that r = L n on S
L
, the second
term of (A.27) drops out ( n n = 0).
r ( n u) =
n (r )
8L
3
. (A.28)
Using Stokes theorem, one then obtains

__
S
L
r ( n u) dS =
1
8L
3

__
S
L
n (r ) dS =
1
8L
3
___
V
L
(r ) dV . (A.29)
The last integrand in equation (A.29) can be rewritten in the following manner:
(r ) =r ( ) ( r) + ( )r (r ) . (A.30)
Where is constant in space so = 0. r = 3 because we are working in three
dimension. The third term on the right side of equation (A.30) can be written as follows:
( )r =
x
r
x
x
+
y
r
y
y
+
z
r
z
z

k = . (A.31)
The last term in (A.30) is equal to zero because r = 0. Now it can simply be shown
that equation (A.30) is equal to 2:
(r ) = 0 3 + 0 = 2 . (A.32)
Placing equation (A.32) in equation (A.29) yields:

__
S
L
r ( n u) dS =
1
8L
3
___
V
L
2dV =

4L
3
___
V
L
dV =

3
. (A.33)
Equation (A.24) therefore simply becomes:
___
V
L
udV =
1
2
___
V
L
(r ) dV
1
2

__
S
L
r ( n u) dS =
1
2

1
2
1
3
=
1
3
. (A.34)
It is important to note that equation (A.34) is valid for any suciently large but nite value
of L. Although equation (A.34) is independent of the value of L, one may not consider L
to be innitely large.
MSc. Thesis F.J. Venneman
92 Derivations of approaches to determine force on an immersed body
A.2.2 Force acting on a large body
Outside the vortical regions, the vorticity is zero and the viscous stress is absent. The
absence of vorticity implies the existence of a scalar potential in the inviscid region such
that the gradient of gives u. The unsteady inviscid momentum equation then yields
(similar like the Bernoulli equation in Kundu and Cohen (2002, equation 4.81)):
p =

t

v
2
2
+ f(t) . (A.35)
Where f(t) is is an integrating function independent of location. For three-dimensional
ows past nite bodies, the region in which exists is singly connected and therefore is
independent of path.
The force

F
L
acting on the boundary S
L
is expressible in terms of the pressure p. Since
the vorticity decays exponentially with r, the shear stress on V
L
is negligible for suciently
large value of L. Using equation (A.35), one obtains:

F
L
=
__
S
L
_

t

v
2
2
+
f(t)

_
ndS . (A.36)
The asymptotic behavior of u shows that the term v
2
/2 does not contribute to (A.36). The
function f(t), being independent of position, also does not contribute to (A.36). Using the
expression for u in (A.25), one obtains for three dimensional ows:

F
L
=
__
S
L

t
ndS =

8
d
dt

__
S
L

1
r
ndS . (A.37)
The integrand on the right-hand side of equation (A.37) can be rewritten as
r
r
3
(see
equation (A.26)). Using Stokess theorem again, one then obtains:

F
L
=

8
d
dt

__
S
L
r
L
3
ndS =

8L
3
d
dt
___
V
L
(r ) dV . (A.38)
It is simple to show that (r ) = . Filling this in equation (A.38) yields the result of
this subsection:

F
L
=

8L
3
d
dt
___
V
L
(r ) dV =

8L
3
d
dt
___
V
L
dV =

8L
3
d
dt
4L
3
3
=

6
d
dt
.
(A.39)
A.2.3 Aerodynamic Force
Consider again the control volume V
L
bounded externally by S
L
and containing the solid
body occupying the region V
S
. The momentum theorem gives for the total force

F
t
,S
F.J. Venneman MSc. Thesis
A.2 Derivation of the Blasius approach 93
acting on the system:

F
t
=
d
dt
___
V
L
udV +
__
S
L
u(u n) dS , (A.40)
where is either the uid density or the solid density
s
, depending on the region within
V
L
.
On account of the asymptotic behavior of u, the last integral in equation (A.40) is negligible
for suciently large values of L. The integral over S
L
can be written as a sum of integrals
as follows:

F
t
=
d
dt
___
V
L
udV
d
dt
___
V
S
udV +
d
dt
___
V
S

s
udV . (A.41)
The total force

F
t
on the system in V
L
consists of the force

F
L
acting on the boundary V
L
V
L
F
e+
F
F
t
=F
e+
F
L
V
S
S
L
Figure A.2: Figure showing what forces act on what parts of the volume.
and the external force

F
e
which is exerted from outside the system and acts on the solid
bodies (see Figure A.2). The total force acting on the solid bodies is

F
e
+

F, where

F is
the aerodynamic force exerted by the uid on the solid bodies. Newtons second law of
motion states that the last term of equation (A.41) is equal to

F
e
+

F. In consequence,
one obtains::

F =

F
e

d
dt
___
V
S

s
udV ,

F
e
=

F
t

F
L
,
___
V
S

s
vdV =

F
t

d
dt
___
V
L
udV +
d
dt
___
V
S
udV ,

F =

F
L

d
dt
___
V
L
udV +
d
dt
___
V
S
udV . (A.42)
MSc. Thesis F.J. Venneman
94 Derivations of approaches to determine force on an immersed body
For three-dimensional incompressible ows, placing equations (A.34) and (A.39) in equa-
tion (A.42) yields:

F =

6
d
dt


3
d
dt
+
d
dt
___
V
S
udV =
1
2

d
dt
. .
i
+
d
dt
___
V
S
udV
. .
ii
. (A.43)
Equation (A.43) is equal to equation (4.1). It shows that the aerodynamic force exerted by
a uid on a solid body immersed in and moving relative to the uid is equal to the inertia
force due to the mass of uid displaced by the solid body (ii) plus a term proportional to
the time rate of change of total rst moment of the vorticity eld in the solid body and
the uid (i).
F.J. Venneman MSc. Thesis
Appendix B
Rotational motion
When objects start to rotate, it is always dicult to perform a consequent book-keeping
of the various terms that come into play. The aim of this Appendix to help the reader
understand how the rotating terms are dealt with in the current study. In Appendix B.1
a transformation is derived to write an expression for the velocity in the rotating frame of
reference. In Appendix B.2 the dierent rotational terms in the Navier-Stokes equations
are explicitly written out. Appendix B.3 shows the details about an exponential t through
the obtained values for the rotational velocity.
B.1 Transforming velocity to a rotating frame of reference
The aim of this section is to get an expression for the velocity in the rotating frame of
reference in terms of the velocity in the non-rotational frame of reference and u(u, v, w),
the angular velocity and the position in the rotating frame of reference r(x
r
, y
r
, z
r
).
We start this derivation with the point O in Figure B.1, where the coordinates x
r
and
z
r
are in the rotating frame of reference. This frame rotates anti-clockwise (negative x
r
-
direction) with angular velocity . It is easy to see that the velocity of point O in the
rotational frame of reference in respectively the x
r
, y
r
and z
r
direction are:
u
r
=u U = u R , (B.1a)
v
r
=v , (B.1b)
w
r
=w . (B.1c)
where u, v and w are in the non-rotational frame of reference (obtained with piv). There
is no dierence between v
r
and v because the frame rotates around this axis. There is no
dierence between w
r
and w because the z
r
-axis is aligned with point O.
This was the simple case, now take a look at point P, which is also in the fov (dashed
line). This point is rotating with U
P
= R
P
, where U
P
> U, because R
P
> R. We see
that U
P
has a component in both the x
r
and z
r
direction. We dene the angle positive
MSc. Thesis F.J. Venneman
96 Rotational motion
R
R
x
z
P U
u w
U
P
P P
P
O
F
O
V
r
r
Figure B.1: Transforming velocity to a rotating frame of reference. The dashed line in this
gure indicates the fov from a top view.
from R to R
P
. It is easy to see that the angle between u
P
and U
P
is also equal to (the
sum of the angles in a triangle and along a straight line is both equal to , also make use
of the right angles). The components of U
P
can be written as follows:
u
P
=U
P
cos = R
P
cos , (B.2a)
w
P
=U
P
sin = R
P
sin . (B.2b)
can also be written as:
=cos
1
_
R
R
P
_
, or (B.3a)
=sin
1
_
x
r,P
R
P
_
. (B.3b)
Substituting B.3a in B.2a and B.3a in B.2a yields:
u
P
=U
P
cos = R , (B.4a)
w
P
=U
P
sin = x
r,P
. (B.4b)
Surprisingly the x
r
-component of the velocity is always corrected with R (just like in
B.1a). It should be noted here that R = z
r,O
, where z
r,O
is a negative number. Now
we have an expression for the velocity in the rotating frame of reference in terms of the
velocity in the non-rotational frame of reference and u(u, v, w), the angular velocity and
the position in the rotating frame of reference r(x
r
, y
r
, z
r
):
u
r
=u z
r
, (B.5a)
v
r
=v , (B.5b)
w
r
=w x
r
, (B.5c)
which are equal to equation 4.13.
F.J. Venneman MSc. Thesis
B.2 Rotational frame of reference 97
B.2 Rotational frame of reference
In a rotating right handed Cartesian coordinate system the direction of the angular velocity

is in the positive y-direction (see Figure B.2). This is shown by the denition of the
Figure B.2: Roboy setup with the rotating frame of reference and the direction of the
angular velocity

.
angular velocity:

=
r v
|r|
2
=


k
0 0 r
z
v
x
0 0
=
v
x
r
z
. (B.6)
Here we restart the Navier-Stokes equation from (2.6):

_
u
t
+ (u ) u
_
= p +
_

t
r +

r
_
+ 2

u
_
. (B.7)
In this part of the appendix the dierent rotating terms will be written in three dimensions.
We start with the tangential acceleration term relative to the rotating frame of reference.
Since

is in the y-direction,

t
is also in that direction.

t
r =


k
0

t

y
0
r
x
r
y
r
z
=

t

y
r
z


t

y
r
x

k . (B.8)
For the centripetal acceleration we have to obtain a cross product twice.

r =


k
0
y
0
r
x
r
y
r
z
=
y
r
z

y
r
x

k ,

r
_
=


k
0
y
0

y
r
z
0
y
r
x
=
2
y
r
x

2
y
r
z

k . (B.9)
MSc. Thesis F.J. Venneman
98 Rotational motion
The Coriolis term looks as follows:
2

u = 2


k
0

t

y
0
u
x
u
y
u
z
= 2
y
u
z
2
y
u
x

k . (B.10)
B.3 Exponential t for rotational motion
In this section it will be shown how the exponential ts in Figure 3.2(b) and 3.2(c) were
obtained. These gures are repeated for convenience in Figure B.3.
In (Poelma et al., 2006, equation (1)) the empirical function for the tip velocity is given:
U
tip
(t) = U
max
_
1 exp
_

t
t
c
__
, (B.11)
where t
c
is the typical time to reach 63% of the maximum velocity U
max
. Knowing the
function for the tip velocity, the empirical function for the angular velocity () is:
(t) =
U
max
L
_
1 exp
_

t
t
c
__
=
max
_
1 exp
_

t
t
c
__
, (B.12)
with L the distance from the point-of-rotation to the tip of the wing (wing span). Using
the cftool in MatLab a nonlinear least-squares t of the form (a
_
1 exp{
t
b
}
_
) through
the numerical derivatives was obtained. This resulted in the values with 95% condence
bands stated in Table B.1. These values are indicated by the red line in Figure B.3(a).
This empirical function for the angular velocity was then derived with respect to time in
order to obtain the rotational acceleration (plotted in red in B.3(b)).
variable value 95% minimum 95% maximum
a =
max
1.181 1.157 1.206
b = t
c
0.1685 0.1477 0.1894
Table B.1: Coecients of exponential t with 95% condence bands.
F.J. Venneman MSc. Thesis
B.3 Exponential t for rotational motion 99
time (s)

(
r
a
d
s

1
)
0 0.5 1 1.5 2 2.5 3
0
0.2
0.4
0.6
0.8
1
1.2
(a) Angular velocity
time (s)

t
(
r
a
d
s

2
)
0 0.5 1 1.5 2 2.5 3
-1
0
1
2
3
4
5
6
7
8
(b) Angular acceleration
Figure B.3: Angular velocity and angular velocity of the impulsively-started wing. Blue
crosses indicate the (derivatives of the) data provided by Dr. Ir. C. Poelma, red lines indicate
the exponential ts.
MSc. Thesis F.J. Venneman
100 Rotational motion
F.J. Venneman MSc. Thesis
Appendix C
Moving Grids
The dynamic mesh utility in OpenFOAM updates the grid at each time step according
to Radial Basis Function interpolation. This method is explained in section C.1. When
moving grids are considered, the time-dependent variation of the control volumes has to
be taken into account. This is done by using the Space Conservation Law, which is derived
in section C.2.
C.1 Radial Basis Function interpolation
The generation of a complex grid is a time-consuming and nontrivial task. For structured
meshes there are ecient techniques available to deform a moving mesh, one of them is
Radial Basis Function (rbf) interpolation. To understand rbf interpolation one should
rst understand what a Radial Basis Function is. Basically it is a function whose value
depends on the distance from a certain origin, i.e. (x) = ( x c ), with x a vector
in space and c the vector pointing at the origin. There are various Radial Basis Functions
available in scientic literature, which can be divided into two groups: functions with
compact support and functions with global support. Compact functions have the following
property:
(x) =
_
f(x) 0 x r with f(x) 0 ,
0 x > r ,
(C.1)
an example is f(x) = (r x)
2
. Functions with global support are not equal to zero outside
a certain radius, but cover the whole interpolation space. A number of these functions is
listed in Table C.1.
Now the reader has some understanding of rbfs, we can start explaining the interpolation
technique. The rbf interpolation scheme moves each grid point individually based on
its position in space and is therefore called a point-by-point scheme. The interpolation
function, s, is used to transfer the displacements known at the boundary of the mesh to
all the nodes of the aerodynamic mesh. This interpolation function can be described by a
MSc. Thesis F.J. Venneman
102 Moving Grids
Name Abbreviation f(x)
Gaussian Gaus exp(x
2
)
Thin plate spline TPS x
2
log(x)
Multiquadratic Biharmonics MQB

a
2
+ x
2
Inverse Multiquadratic Biharmonics IMQB
_
1
a
2
+x
2
Table C.1: Radial Basis Functions with global support
sum of Radial Basis Functions (Boer et al., 2007)
s(x) =
n
b

j=1

_
x x
b
j

_
+ p(x) , (C.2)
where x
b
j
=
_
x
b
j
, y
b
j
, z
b
j

are the centers in which the values are known (in this case the
boundary nodes), p a polynomial, n
b
the number of boundary nodes and a given Radial
Basis Function with respect to the Euclidean distance x . The coecients
j
and the
polynomial p are determined by the interpolation conditions.
s(x
b
j
) =

d
b
j
, (C.3)
with

d
b
containing the discrete known values of the displacement at the boundary, and the
additional requirements
n
b

j=1

j
q(x
b
j
) = 0 , (C.4)
for all polynomials q, with a degree less or equal than that of polynomial p. The minimal
degree of polynomial p depends on the choice of the basis function . If the Radial Basis
Function is a conditionally positive denite function of order m 2
1
, a linear polynomial
can be used. The values for the coecients
j
and the linear polynomial can be obtained
by solving the system:
_

d
b
0
_
=
_
M
b,b
P
b
P
T
B
0
_ _

_
, (C.5)
with containing the coecients
j
,

the coecients of the linear polynomial p, M
b,b
an n
b
n
b
matrix containing the evaluation of the Radial Basis Functions
b
i
b
j
= (
x
b
i
x
b
j
) and P
b
an n
b
4 matrix with row j given by
_
1 x
b
j
y
b
j
z
b
j

. The values
of the interior of the ow mesh

d
in
, can then be derived by evaluating the interpolation
function (C.2) in the internal grid points:

d
in
j
= s(x
in
j
) . (C.6)
1
It goes beyond the scope of this thesis to explain what this exactly means, but if you are interested see
Denition 2.1 on http://www.maths.sussex.ac.uk/
~
wendland/Forschung/reconhtml/node2.html.
F.J. Venneman MSc. Thesis
C.2 Space Conservation Law 103
Variable Value Detail
rbf Function IMQB see Table C.1
coarseningRatio 100
innerRadius 3 wing length is 1.8
outerRadius 40 just touched the outside of the total domain
Table C.2: Most important parameters for the Radial Basis Function mesh deformation tool.
The OpenFOAM utility icoDymFoamRBF uses Radial Basis Function interpolation as a
base for mesh deformation. It assigns a value to the number of points, x
b
j
, taken into
account on the moving object (wing in this case), this is equal to the number of cells on
the moving wall divided by the coarseningRatio. Furthermore it uses an innerRadius
and an outerRadius for rigid mesh deformation (inside the innerRadius) and less rigid
mesh deformation (decreasing from the innerRadius to the outerRadius). Outside the
outerRadius, the mesh deformation is zero. The parameters used for the three-dimensional
case, described in section 5.3, are summarized in Table C.2.
C.2 Space Conservation Law
In this part of the Appendix the Space Conservation Law (scl) for moving grids is derived.
The extended version of this derivation is is in Ferziger and Peric (1999). First consider
the 1D continuity equation:

t
+
u
x
= 0 . (C.7)
Integrating this equation over a control volume with moving boundaries the following
equation is obtained:
x
2
(t)
_
x
1
(t)

t
dx +
x
2
(t)
_
x
1
(t)
u
x
dx = 0. (C.8)
Using Leibniz rule for dierentiating under the integral sign pulls the time-derivative out-
side the integral. This results in an extra term concerning the movement of the boundary
of the control volume u
b
:
d
dt
x
2
(t)
_
x
1
(t)
dx +
x
2
(t)
_
x
1
(t)

x
[ (u u
b
)] dx = 0 . (C.9)
When the boundary moves with uid velocity, i.e. u = u
b
, the second integral becomes
zero and we have the Lagrangian mass conservation equation,
dm
dt
.
Using a similar analysis the 3D version of the integral mass equation can be obtained.
MSc. Thesis F.J. Venneman
104 Moving Grids
Mass conservation
d
dt
_
V (t)
dV +
_
S(t)
(u u
b
) ndS = 0 . (C.10)
The same principles apply to any transport equation, so using the same analysis the integral
momentum equation for moving grids can be derived from 2.2:
Momentum conservation
d
dt
_
V (t)
udV +
_
S(t)
u(u u
b
) ndS =
_
S(t)
p n + ndS . (C.11)
According to Ferziger and Peric (1999) articial mass sources may occur due to the mesh
motion resulting in insucient mass conservation. The articial mass sources are due to
the dashed volume in Figure C.1.
Figure C.1: A rectangular control volume whose size increases in time due to a dierence in
grid velocities at the boundaries. (Ferziger and Peric, 1999)
These articial mass sources may accumulate with time and cause serious problems. There-
fore mass conservation is enforced by the Space Conservation Law (scl):
d
dt
_
V (t)
dV
_
S(t)
u
b
ndS = 0 . (C.12)
This equation describes the conservation of space when the CV changes its shape and/or
position of the control volume.
F.J. Venneman MSc. Thesis
Appendix D
Grid quality
Mesh quality is very important in Computational Fluid Dynamics. The correctness of
results is directly linked to the quality of the mesh. Mesh-induced errors show up at
insucient mesh resolution. If the shape of the wing changes rapidly in some regions of
the domain and the local number of computational points is not sucient to describe the
change, the quality of the solution will be lost.
D.1 Grid quality measures
OpenFoam uses several indicators to determine the quality of a mesh. In the present study,
three important indicators are reviewed for the dierent meshes.
1. The Aspect Ratio
2. Non-orthogonality
3. Skewness
These indicators will be described below.
Aspect Ratio
The aspect ratio is a measure for the ratio between the length and the width of a cell. It
can best be explained by Figure D.1(a). With e
1
= (a+c)/2 and e
2
= (b+d)/2, the aspect
ratio AR is:
AR =
max (e
1
, e
2
)
min(e
1
, e
2
)
. (D.1)
A utility in OpenFOAM to check the quality of the mesh, uses a threshold for the aspect
ration of 1000.
MSc. Thesis F.J. Venneman
106 Grid quality
(a) (b)
Figure D.1: Illustration to help explain the aspect ratio (a) and non-orthogonality (b).
Non-orthogonality
Figure D.1(b) shows two cells that are not orthogonal with respect to each other. P and
N are the centroids of the two cells. The vector e points from P to N. The vector

f is
perpendicular to the face between points P and N. The angle is the angle between e
and

f and is a measure of the non-orthogonality. In two dimensions each cell possesses
four neighboring cells. The maximum angle of the four sides of the cell is taken as the
non-orthogonality of that cell. The threshold OpenFOAM uses, is a non-orthogonality
angle of 70

.
Cell Skewness
The skewness of a cell is a measure of the skew shape of the two neighboring cells. The
accuracy of a face integral between two very skew is reduced, so this is an unwanted
situation. Figure D.2 shows a typical setup of two neighboring cells causing a skewness
error. P and N are the centroids of the two cells. The vector

d points from P to N, the
Figure D.2: Skewness error on the face. Redrawn from Jasak (1996)
vector m points from the intersection of

d with the cell face to middle of the cell face. The
skewness is then measured by:
SK =
| m|

(D.2)
The threshold OpenFOAM uses, is a cell skewness of 4.
F.J. Venneman MSc. Thesis
D.2 2D grid quality 107
22k 46k 97k
max mean max mean max mean
aspect ratio 59.98 92.23 149.43
non-orthogonality 39.98 3.28 33.33 3.28 28.30 3.11
skewness 0.64 0.0066 0.55 0.0051 0.47 0.0039
Table D.1: Indicators of the quality of the two dimensional mesh.
D.2 2D grid quality
Three dierent meshes are tested in two dimensions. First we focus on the 46k mesh,
which has the same setup as the other meshes (described in section 5.2.2). Figure D.3
gives an indication of the quality of the two-dimensional grid of 46k cells. It is shown that
is especially large at the leading and trailing-edge of the wing, but is much smaller than
the threshold of 4 (Figure D.3(a)). Figure D.3(b) shows that the skewness in the fareld
of the wing is high where the ellipsoid grid meets surrounding rectangular blocks. At these
four connection points, ve cells are connected. Figure D.3(c) shows the non-orthogonality
of the cells in the fareld of the wing. Just like in Figure D.3(b), the connection points
can be interpreted as sources of non-orthogonality. Still the maximum non-orthogonality
angle is far lower than the threshold of 70

.
Figure D.4 shows histograms of the non-orthogonality and the skewness of all the two-
dimensional meshes (22k, 46k and 97k from top to bottom). As can be seen the distribution
of less orthogonal or skew cells is fairly constant for the dierent meshes.
Table D.1 shows the maximum and mean values for the three indicators of the mesh quality
for all the three two-dimensional grids.
MSc. Thesis F.J. Venneman
108 Grid quality
(a) Cell skewness around the wing (b) Cell skewness in the fareld of
the wing
(c) Non-orthogonality in the far eld
of the wing
Figure D.3: Figures indicating the quality of the grid of 46k cells. (a) The skewness is
especially large at the leading and trailing-edge of the wing, but is much smaller than the
threshold of 4. (b) The skewness in the fareld of the wing is high where the ellipsoid
grid meets surrounding rectangular blocks. At these four connection points, ve cells are
connected. (c) Non-orthogonality of the cells in the fareld of the wing. Just like in (b),
the connection points can be interpreted as sources of non-orthogonality. Still the maximum
non-orthogonality angle is far lower than the threshold of 70

.
F.J. Venneman MSc. Thesis
D.2 2D grid quality 109
[degrees]
n
u
m
b
e
r
o
f
c
e
l
l
s
0 5 10 15 20 25 30 35 40
0
1000
2000
3000
4000
5000
6000
7000
8000
(a) non-orthogonality, 22k
SK [-]
n
u
m
b
e
r
o
f
c
e
l
l
s
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
0
1000
2000
3000
4000
5000
6000
7000
(b) skewness, 22k
[degrees]
n
u
m
b
e
r
o
f
c
e
l
l
s
0 5 10 15 20 25 30 35 40
0
5000
10000
15000
(c) non-orthogonality, 46k
SK [-]
n
u
m
b
e
r
o
f
c
e
l
l
s
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
0
2000
4000
6000
8000
10000
12000
14000
16000
(d) skewness, 46k
[degrees]
n
u
m
b
e
r
o
f
c
e
l
l
s
0 5 10 15 20 25 30 35 40
0.0e+00
5.0e+03
1.0e+04
1.5e+04
2.0e+04
2.5e+04
3.0e+04
3.5e+04
(e) non-orthogonality, 97k.
SK [-]
n
u
m
b
e
r
o
f
c
e
l
l
s
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
0.0e+00
5.0e+03
1.0e+04
1.5e+04
2.0e+04
2.5e+04
3.0e+04
3.5e+04
(f) skewness, 97k.
Figure D.4: Histograms showing the non-orthogonality (left) and the skewness (right) of the
dierent meshes (22k, 46k and 97k from top to bottom). As can be seen the distribution of
less orthogonal or skew cells is fairly constant for the dierent meshes.
MSc. Thesis F.J. Venneman
110 Grid quality
D.3 3D grid quality
Four dierent grids were used for the three dimensional computations. Two CAD-wings
with 280k cells and 530k cells, respectively and two ellipsoidal wings with 480k cells and
aspect ratios of 2.2 and 3.4 respectively. The maximum aspect ratio, non-orthogonality
and skewness of these grids, at maximum deformation, are shown in Table D.2.
CAD-wing ellipsoidal wing
280k 530k AR = 2.2 AR = 3.4
maximum aspect ratio 9.68 4.80 10.79 16.80
maximum non-orthogonality 53.69 52.60 60.08 60.10
maximum skewness 2.02 3.98 0.94 0.96
Table D.2: Indicators of the quality of the three-dimensional meshes at a maximum defor-
mation.
F.J. Venneman MSc. Thesis
Appendix E
OpenFOAM settings
OpenFOAM-1.5-dev was used as a framework for solving the ow around an impulsively-
started wing. OpenFOAM (Open Field Operation And Manipulation) is primarily a C++
toolbox for the customization and extension of numerical solvers for continuum mechanics
problems, including Computational Fluid Dynamics. It is produced by OpenCFD Ltd.
and is freely available and open source, licensed under the GNU General Public License.
In this Appendix the settings used in OpenFOAM are discussed. Section E.2 shows the
settings used in two dimensions, section E.3 shows the settings used in three dimensions.
But rst a little background will be given about the solvers and preconditioners that are
used in the present study
E.1 Solution and algorithm control
icoDyMFoam was used as a solver for incompressible, laminar ow of Newtonian uids and
a dynamic mesh. For solving the equations for pressure and velocity PISO is used, this is
an iterative method. Every time-step piso evaluates an initial u and p, corrects them twice
(in this thesis) until a convergence criterium is met. An iterative method like PISO can
be accelerated by Krylov subspace methods. This means that the matrix A of the system
Ay = b to be solved is split as: A = M N (with N small). M is used to precondition
the problem, which means that Ay = b is replaced by its preconditioned version:
M
1
Ay = M
1
b. (E.1)
In this thesis Incomplete Cholesky decomposition is used to precondition the system as
follows:
L
1
AL
T
y = L
1
b, y = L
T
y , (E.2)
where LL
T
is an Incomplete Cholesky decomposition of A and L
T
= (L
T
)
1
Another Krylov subspace method is the Conjugate Gradient method for linear systems
Ay = b, with a symmetric self-adjoint
1
positive denite matrix A. It goes beyond the
1
A matrix A is self adjoint if A
H
= A
T
= A
MSc. Thesis F.J. Venneman
112 OpenFOAM settings
scope of this thesis to explain it, more information can be found in Ferziger and Peric
(1999, chap 5.3.6). One can use preconditioning together with Conjugate Gradient meth-
ods, in OpenFOAM this is referred to as pcg.
Combining ic and cg results in a new method called iccg. If the matrix A is not self
adjoint, one can apply BiConjugate Gradient method (bcg). Using this in combination
with IC results in biccg, the combination with cg results in pbicg.
The preconditioners used are dic and dilu. dic is an abbreviation for Diagonal Incomplete
Cholesky. Incomplete Cholesky decomposition has been explained, dic is only applicable
for symmetric (or Diagonal, hence the D) matrices. dilu is an abbreviation for Diagonal
Incomplete lu decomposition. The idea of ilu is to determine lower and upper tridiagonal
matrices L and U such that LU A, and to choose M = LU (where M is the precondi-
tioner matrix, as explained before). The elements of L and U are prescribed to be zero,
except if (i, j) Q, where Q is the set of index pairs where nonzeros are allowed.
E.2 2D OpenFOAM Settings
For every variable calculated the linear solver methods and their solution tolerances are
listed in Table E.1. The numerical schemes used for determining gradients and other
operators are listed in Table E.2.
1
Gaussian integration is based on summing values on cell faces, which must be interpolated at the cell
centers
Variable Solver Tolerance
p ICCG 1 10
6
pcorr ICCG 1 10
6
pFinal ICCG 1 10
6
U BICCG 1 10
5
Table E.1: Linear solvers and their tolerance used in the two-dimensional case
F.J. Venneman MSc. Thesis
E.3 3D OpenFOAM Settings 113
Keyword Category of mathe-
matical terms
Scheme Details
ddtSchemes First order time
derivatives /t
backward second order implicit
gradSchemes Gradient Gauss
linear
1
divSchemes Divergence Gauss
linear
1
laplacianSchemes Laplacian
2
Gauss
linear
corrected
Unbounded, second
order, conservative
1
interpolationSchemes Point-to-point inter-
polation of values
linear central dierencing
snGradSchemes Component of gradi-
ent normal to a cell
face
corrected explicit non-
orthogonal correction
Table E.2: Numerical schemes used in the two-dimensional and three-dimensional case.
E.3 3D OpenFOAM Settings
For every variable calculated the linear solver methods, solution tolerances and precondi-
tioners are listed in Table E.3. The numerical schemes used for determining gradients and
other operators are exactly the same as for the two-dimensional case, which are described
in Table E.2.
Variable Solver Tolerance Preconditioner
p PCG 1 10
2
DIC
pcorr PCG 1 10
6
DIC
pFinal PCG 1 10
6
DIC
U PBiCG 1 10
6
DILU
cellMotion PCG 1 10
12
DIC
cellDisplacement PCG 1 10
8
DIC
Table E.3: Linear solvers and their tolerance used in the three-dimensional case
MSc. Thesis F.J. Venneman
114 OpenFOAM settings
F.J. Venneman MSc. Thesis
Appendix F
Fourier analysis of a force signal
The Fourier spectrum of the time-history of C
L
is depicted in Figure F.1(b). A sucient
amount of periods (approximately 40) was taken for 300 < t < 500 in order to ensure the
right frequency was obtained. A power spectrum
1
was used to acquire this frequency (see
Figure F.1(b)). The highest peak in the Figure shows a frequency of f = 0.20871, which
roughly corresponds to one period per 5 s (which agrees with Figure F.1(a)).
t [s]
C
L
[
-
]
300 350 400 450 500
0.8
0.9
1
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
(a) Time-history
f [Hz]
|
C
L
(
f
)
|
f= 0.20871
0 0.5 1 1.5 2
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
(b) Amplitude spectrum
Figure F.1: Time-history and amplitude spectrum of C
L
for Re = 192, Co
max
= 1 and 46k
cells.
1
See for detailed information http://en.wikipedia.org/wiki/Spectral_density
MSc. Thesis F.J. Venneman

You might also like