You are on page 1of 75

OPTIMAL AIRFOIL DESIGN USING 3D

COMPUTATIONS
A Thesis Submitted
in Partial Fulllment of the Requirements
for the Degree of
Master of Technology
by
Devashish Sharma
to the
Department of Aerospace Engineering
INDIAN INSTITUTE OF TECHNOLOGY
KANPUR
INDIA
May, 2012
To my family
i
Abstract
A section of wing is optimized for a three dimensional unsteady ow at Re = 110
4
and = 4
o
for maximization of lift. Before starting the optimization process in 3D,
the eects of span and mesh resolution on ow are studied, as these parameters play
a vital role in 3D computations. The study is done on an airfoil geometry, which is
optimal for 2D unsteady ows, at Re = 1 10
4
and = 4
o
. The airfoil surface is
parameterized by 13 control points using the 4
th
order NURBS curve. Four cases
have been studied using 13, 26, 39 and 52 slices in span-wise direction for 0.25, 0.5,
0.5 and 1.0 as the span to chord ratios, respectively. It is shown that 26 slices along
with a span to chord ratio of 0.5 are optimum to do calculations for a 3D ow.
Gradient-based continuous adjoint method for shape optimization, proposed by Sri-
nath and Mittal[1, 2], is applied for shape optimization of airfoils in three dimensions.
The Limited memory-Broyden-Fletcher-Goldfarb-Shanno(L-BFGS) algorithm has
been used as an optimizer. A nite element method, based on the streamline upwind
Petrov/Galerkin(SUPG) and pressure stabilized Petrov/Galerkin(PSPG) stabiliza-
tion techniques, is used to solve the adjoint equations.
Further, shape optimization of airfoils in an unsteady ow is done. A shape that is
optimal for 2D unsteady ow is used as an initial guess in 3D optimization. This
optimal shape was obtained after carrying out optimization in 2D unsteady ow
with NACA-0012 as the initial guess. Maximization of lift is used as the objective
function for 3D optimization at Re = 1 10
4
. Due to the large amount of time
required for a complete unsteady problem and owing to limitations of the present
ii
iii
optimization code, such as huge disk space needed to save ow and adjoint les,
this unsteady shape optimization problem could not be completed. Three iteration
cycles of optimization have been carried out and the results obtained after the third
iteration are shown and analyzed. It is also shown that values of the aerodynamic
coecients diminish from 2D to 3D computations for the same airfoil shape. In
3D optimization, the time averaged lift coecient is found to increase from 0.86 to
0.90 and the drag coecient is found to reduce from 0.089 to 0.075 after the third
iteration.This leads to the conclusion that a shape which is optimal for 2D ows
may not necessarily be optimal for 3D ows. Also, shape optimization is carried out
in 3D using NACA-0012 directly as an initial guess, at Re = 1 10
4
. Maximization
of lift is used as the objective function for 3D optimization. In this 3D optimization,
the time averaged lift coecient is found to increase from 0.15 to 0.85 after the rst
iteration, yielding an increase of 471% in the value of lift coecient.
Acknowledgments
I would like to express my sincere gratitude towards my thesis supervisor, Prof.
Sanjay Mittal, who has been an immense source of inspiration for the last two
years. His knowledge, expertise and work ethic has enlightened and motivated me
throughout my research work. He has been a source of constant support in my
tough times during these two years.
I would like to thank Naveen and Nikhil who were my mentors and whose research
I have carried forward. My thanks, to my other lab mates Suresh, Murali, Navrose,
Sandeep and Mohsin for their help and making the lab environment enjoyable. I
will cherish the moments I have spent with them in lab.
I would also like to thank my friends Akshay, Manoj, Pathanjali and Karan for
supporting and encouraging me in my dicult times at IIT Kanpur.
Finally, I would like to thank my family for their unconditional love, support and
motivation.
iv
Contents
Certicate ii
Abstract ii
Acknowledgment iv
Contents v
List of Figures viii
List of Tables xi
Nomenclature xii
1 Introduction 1
1.1 Work done earlier on Aerodynamic Shape optimization . . . . . . . . 2
1.1.1 Adjoint Calculations: Continuous and discrete approaches . . 4
1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Thesis organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
v
CONTENTS vi
2 Governing Flow Equations and Finite Element Formulation 7
2.1 The Navier-Stokes equations . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Finite element formulation . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Stabilization parameter . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3 The Adjoint Method for Shape Optimization 12
3.1 Problem set-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 The adjoint approach . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2.1 The continuous adjoint equations . . . . . . . . . . . . . . . . 16
3.2.2 Adjoint boundary and terminal conditions . . . . . . . . . . . 16
3.2.2.1 Adjoint boundary conditions for steady ows . . . . 17
3.2.2.2 Adjoint boundary conditions for unsteady ows . . . 17
3.2.3 Finite element formulation for adjoint equations . . . . . . . . 18
3.2.4 The gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3 The optimizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4 Study of eect of span and mesh resolution in 3D ows at Re =
1 10
4
21
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2 Problem description and mesh . . . . . . . . . . . . . . . . . . . . . 22
4.2.1 Parameterization using NURBS curves: . . . . . . . . . . . . . 23
4.2.2 Various cases studied and the parameters used in those cases: 24
4.2.3 Time averaging of aerodynamic coecients: . . . . . . . . . . 24
4.2.4 Optimal shape for 2D unsteady ows, used as initial guess: . . 27
4.3 Eect of span to chord ratio in 3D ow computations for Re = 110
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4 Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
CONTENTS vii
5 Shape optimization of 2D geometry for 3D unsteady ows 40
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.2 Problem description and mesh . . . . . . . . . . . . . . . . . . . . . 41
5.2.1 Mesh movement strategy . . . . . . . . . . . . . . . . . . . . . 42
5.3 Flow Chart of an optimization code used for 3D optimization in un-
steady ows at Re=1 10
4
. . . . . . . . . . . . . . . . . . . . . . . 43
5.4 Shape optimization of 2D airfoil geometry in a 3D unsteady ow at
Re=1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . . 44
5.5 Shape optimization in an unsteady 3D ow, taking NACA-0012 as
initial guess at Re=1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . 46
6 Conclusion 53
6.1 Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
References 60
List of Figures
2.1 Schematic of the problem set-up: boundary conditions.
U
and
D
are the upstream and downstream boundaries, respectively.
S
and

W
are the lateral boundaries.
B
represent the surface body . . . . . 9
4.1 Parameterization of airfoil, used as initial guess for 3D computations,
using 4
th
order of NURBS curve with 13 control points. The y-co-
ordinates of the control points 2-6 and 8-12 are used for optimizing
the airfoil geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2 Close-up view of nite element mesh for the 2D optimized shape at
4
o
angle of attack. The mesh contains 44804 nodes and 89304 elements. 25
4.3 A 3D nite element mesh used for 3D calculations. Mesh consists of
wedge element. 26 slices of 2D mesh 5.5 has been used to generate
this mesh. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.4 Time history of C
l
along with averaging and rms from reverse direc-
tion for: 13 slices and 0.25 span case . . . . . . . . . . . . . . . . . . 32
4.5 Time history of C
l
along with averaging and rms from reverse direc-
tion for: 26 slices and 0.5 span case . . . . . . . . . . . . . . . . . . . 32
4.6 Time history of C
l
along with averaging and rms from reverse direc-
tion for: 39 slices and 0.5 span case . . . . . . . . . . . . . . . . . . . 33
4.7 Time history of C
l
along with averaging and rms from reverse direc-
tion for: 52 slices and 1.0 span case . . . . . . . . . . . . . . . . . . . 33
4.8 Time history of C
d
along with averaging and rms from reverse direc-
tion for: 13 slices and 0.25 span case . . . . . . . . . . . . . . . . . . 34
viii
LIST OF FIGURES ix
4.9 Time history of C
d
along with averaging and rms from reverse direc-
tion for: 26 slices and 0.5 span case . . . . . . . . . . . . . . . . . . . 34
4.10 Time history of C
d
along with averaging and rms from reverse direc-
tion for: 39 slices and 0.5 span case . . . . . . . . . . . . . . . . . . . 35
4.11 Time history of C
d
along with averaging and rms from reverse direc-
tion for: 52 slices and 1.0 span case . . . . . . . . . . . . . . . . . . . 35
4.12 Comparison of time averaged lift coecients for all the cases . . . . . 36
4.13 Comparison of rms lift coecients for all the cases . . . . . . . . . . . 36
4.14 Comparison of time averaged drag coecients for all the cases . . . . 37
4.15 Comparison of rms drag coecients for all the cases . . . . . . . . . . 37
4.16 Averaged uv component of Reynolds Stress is compared for the cen-
ter slice: (a) 7th slice for 13 slices case, (b) 13th slice for 26 slices
case, (c) 20th slice for 39 slices case, (d) 26th slice for 52 slices case . 38
4.17 Averaged uw component of Reynolds Stress is compared for the
center slice: (a) 7th slice for 13 slices case, (b) 13th slice for 26 slices
case, (c) 20th slice for 39 slices case, (d) 26th slice for 52 slices case . 38
4.18 Averaged w component of velocity is compared for the center slice:
(a) 7th slice for 13 slices case, (b) 13th slice for 26 slices case, (c) 20th
slice for 39 slices case, (d) 26th slice for 52 slices case . . . . . . . . . 39
4.19 Averaged pressure coecient is compared for the center slice: (a) 7th
slice for 13 slices case, (b) 13th slice for 26 slices case, (c) 20th slice
for 39 slices case, (d) 26th slice for 52 slices case . . . . . . . . . . . . 39
5.1 A ow-chart detailing the steps of the optimization process. . . . . . 45
5.2 Further optimization of shape, optimal for 2D ow, in a 3D unsteady
ow at Re=110
4
and = 4
o
: initial and optimal shapes after three
iterations of optimization cycle. . . . . . . . . . . . . . . . . . . . . . 46
5.3 Time histories of lift coecient for both initial and optimal shapes
at Re=1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . 47
LIST OF FIGURES x
5.4 Time histories of drag coecient for both initial and optimal shapes
at Re=1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . 47
5.5 Close-up view of nite element mesh for the NACA-0012 at 4
o
angle
of attack. The mesh contains 44804 nodes and 89304 elements. . . . . 49
5.6 Further optimization of shape, optimal for 2D ow, in a 3D unsteady
ow at Re=110
4
and = 4
o
: initial and optimal shapes after three
iterations of optimization cycle. . . . . . . . . . . . . . . . . . . . . . 49
5.7 Time histories of lift coecient for both initial and optimal shapes
at Re=1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . 50
5.8 Time histories of drag coecient for both initial and optimal shapes
at Re=1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . 50
5.9 Unsteady optimization of NACA-0012 at Re=1 10
4
and = 4
o
:
magnitude of instantaneous pressure eld for (a) Initial Shape i.e.
NACA-0012, (b) Optimal Shape, after one iteration of optimization
cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.10 Unsteady optimization of NACA-0012 at Re=1 10
4
and = 4
o
:
instantaneous pictures of isosurfaces of stream-wise vorticity(
x
=
0.05) for (a) NACA-0012, (b) Optimal Shape. Instantaneous pic-
tures of isosurfaces of span-wise vorticity(
z
= 5) for (c) NACA-
0012, (d) Optimal Shape . . . . . . . . . . . . . . . . . . . . . . . . . 52
List of Tables
4.1 The number of slices of 2D mesh and span to chord ratio for various
cases, computed at Re = 1 10
4
and = 4
o
. . . . . . . . . . . . . . 25
4.2 Comparison of averaged and rms aerodynamic coecients of 13, 26
and 52 slices for span to chord ratio of 0.25, 0.5 and 1.0 respectively,
to show the eect of span to chord ratio . . . . . . . . . . . . . . . . 28
4.3 The number of nodes and elements in 3D mesh of all the four cases
computed at Re = 1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . 30
4.4 Comparison of averaged and rms aerodynamic coecients of 13, 26
and 39 slices for span to chord ratio of 0.25, 0.5 and 0.5 respectively,
to show the eect of mesh resolution . . . . . . . . . . . . . . . . . . 31
5.1 The time averaged lift and drag coecients of the initial and the
optimal airfoil shapes for maximization of lift coecient at Re =
1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.2 The time averaged lift and drag coecients of NACA-0012 and the
optimal airfoil shapes for maximization of lift coecient at Re =
1 10
4
and = 4
o
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
xi
Nomenclature
C
d
coecient of drag
C
l
coecient of lift
C
p
coecient of pressure
f body force
h element length
I identity tensor
I
c
objecive function
I augmented objective function
N B-spline basis functions
n Unit normal vector
P coodinates of control points
p pressure
q weight function
Re Reynolds number
s adjoint stress tensor
T deviatric stress tensor
T extent of time integration
t non-dimensional time
U Velocity eld
U

free stream velocity


u, v inline and transverse components of velocity
w weight function
x spatial coordiantes
xii
NOMENCLATURE xiii
Greek letters
density
stress tensor
spatial domain
boundary of spatial domain

U
,
D
,
S
,
B
upstream, downstream, lateral and surface boundaries
dynamic viscosity
kinematic viscosity

SUPG
coecient of SUPG stabilization terms

PSPG
coecient of PSPG stabilization terms

LSIC
coecient of least square stabilization terms
design parameters
adjoint eld

u1
,
u2
inline and transverse components of adjoint velocity

p
adjoint pressure
Chapter 1
Introduction
Engineers strive to continuously improve design, both, to increase operational ef-
fectiveness and reduce cost. While designing any machine or device, the biggest
problem for an engineer or a designer is to optimize the various trade-os. These
trade-os may be its operational eciency, its cost, its structural strength etc. Tra-
ditionally, important decisions during a design process that could aect the eciency
of, both, the product and the process depended on the judgement and experience,
of designers and engineers, based on hit and trial methods. But, since this method-
ology may not only be time consuming but is also expensive and may not lead to
the proper optimal designs. So, various optimization methods came into play which
ensure improved operational eciency at low cost. They are less time consuming
also.
Optimization is a quantitative study of obtaining best results under given conditions.
There are three main elements of an optimization problem. The rst element is to
identify the quantitative measures of the performance of the sytem, which is known
as objective function. The second is to identify the parameters which governs the
performance of the system, which are called as design variables. The last element
1
1.1. Work done earlier on Aerodynamic Shape optimization 2
is selection of an algorithm that can be used to optimize the objective function by
nding suitable design variables.
1.1 Work done earlier on Aerodynamic Shape op-
timization
Real ows and behavior of various aerodynamic bodies in real ows have become one
of the active research areas in recent times. Signicant work has been done in past to
design airfoils operating at low Reynolds number. There utilization is in variety of
applications such as micro-aerial vehicles (MAV), unmanned aerial vehicles (UAV)
and remotely piloted vehicles (RPV). [3, 4, 5, 6]
In most of the aerospace engineering applications, Computational Fluid Dynam-
ics(CFD) is widely used as a tool in design process. But CFD has very rarely been
integrated with optimization tools to improve the performance of existing designs.
First Lighthill [7] used conformal mapping to design airfoils with desired pressure
distribution assuming ow to be incompressible. Further, McFadden [8] extended
this method to compressible ows. Hicks and Henne [9] used conjugate gradient
methods to design a wing in potential ows. In gradient based methods, gradi-
ents are computed by nite dierence methods in which each design parameter is
independently perturbed by a small amount and each perturbed design requires a
complete ow solution. In this method the cost is proportional to the number of de-
sign variables, making it very expensive for problems with a large number of design
parameters.
Optimization can be broadly classied into gradient based and non-gradient based
methods. Two types of optimization methods are studied by Foster and Dulikravich
[10]. They performed a 3D shape optimization using these two optimization meth-
ods. Non-gradient based methods only use the cost function to obtain an optima.
1.1. Work done earlier on Aerodynamic Shape optimization 3
Simplex [11], simulated annealing [12] and genetic algorithms [13] are a few such
methods that have been widely used for engineering applications. One of these non-
gradient based optimization method is Genetic Algorithm. This is a global approach
to optimization and is very time consuming and ultimately expensive also. The other
method is a Gradient based method. It is a local search method of optimization
based on gradient of objective function with respect to identied design parameters.
Interest in the design and the development of the Unmanned Aerial Vehicles (UAVs)
has increased dramatically in the last two and a half decades. Shape optimization of
conforming airfoils is studied by Gano et al. [14]. The Buckle-Wing UAV concept
being developed in this study is designed to morph in a way which facilitates
variations in wing loading, aspect ratio and wing section shapes. In the subsonic
ows aerodynamic wing optimization is carried out by Mialon et al. [15]. They
have presented the description of the analysis and the design process, using CFD
tools, and have presented the results regarding plan-form optimization. Further the
airfoil shape optimization was extended to transonic ows by Congedo [16]. He
constructed high performance airfoils for transonic ows using a robust and ecient
Euler ow solver coupled with a multi-objective genetic algorithm. Huyse and Lewis
[17] studied the aerodynamic shape optimization of two-dimensional airfoils under
uncertain operating conditions. They observed that a deterministic optimization for
discrete operating conditions may result in dramatically inferior performance when
the actual conditions are dierent from these, somewhat arbitrarily chosen, design
values. Lions [18] formulated a framework in which the design problem can be con-
sidered as a problem of optimal control governed by partial dierential equations.
In this approach which is known as adjoint approach, the constraint equations are
augmented to the cost function via Lagrange multipliers. The main benet of this
method over nite dierence method is that the cost of computing the gradients is
independent of number of design parameters. In spite of the fact that this method
1.1. Work done earlier on Aerodynamic Shape optimization 4
requires solution of an additional system of equations, it is clearly much cheaper
than nite dierence method. Jameson [19, 20, 21] rst applied the adjoint method
in the design of airfoils in transonic ows where the uid was assumed to be inviscid.
He carried out the redesign of the RAE2822 airfoil to reduce its shock induced pres-
sure drag at Mach number of 0.73 while the lift is held constant. Further, Jameson
et al. [22, 23, 24] extended the method for design using Navier-Stokes equations
and applied the method to get optimal wing section shapes, where the plan-form
of the wing was kept xed. Okumura and Kawahara [25] rst carried out shape
optimization in unsteady ows and obtained a streamlined body from a circular
cylinder while minimizing the drag.
1.1.1 Adjoint Calculations: Continuous and discrete ap-
proaches
There are two methods by which adjoint calculations can be done. First is discrete
adjoint approach and second is continuous adjoint approach. In discrete adjoint
approach, rst the ow equations are discretized and then the set of discrete ow
equations are dierentiated with respect to design parameters. The gradients ob-
tained from discrete adjoint approach are theoretically closer to those obtained from
nite dierence methods. In continuous adjoint approach, rst objective function
and constraint equations are dierentiated with respect to design parameters to ob-
tain the continuous adjoint equations and then they are discretized using the similar
scheme, which are used to discretize ow equations. In general both the approaches
lead to dierent approximations but converge to the same answer as size of element
tends to zero.
RAE 2822 airfoil was optimized by Jameson [22] using the continuous adjoint ap-
proach. Reuther et al. [26] also used this method for design of complete aircraft
1.2. Objectives 5
conguration. Anderson and venkatakrishnan [27] studied the restrictions on the
various types of cost functions could be used with continuous adjoint approach.
Recently, Srinath and Mittal [1] formulated a continuous adjoint based method for
shape-optimization and implemented it for steady low Reynolds number ows. The
method was applied to nd optimal airfoil shapes at low Re ows [2] and was also
implemented for multi-point shape optimization of airfoils at low Re [28]. Also,
shape-optimization in unsteady ows [29] was carried out for Re = 110
4
ow [30].
1.2 Objectives
There are two main objectives of this work. First, is to investigate the eect of
span to chord ratio and mesh resolution for 3D unsteady ows at Re = 1 10
4
and
= 4
o
. Secondly, to obtain an optimal 2D airfoil geometry for 3D unsteady ows
at Re = 1 10
4
and = 4
o
for maximization of averaged lift coecient.
1.3 Thesis organization
Chapter 2 describes the ow equations and the nite element method used to solve
them. It also describes the boundary conditions used in the computations. Chapter
3 describes the adjoint method for optimal shape design. A complete description
about the adjoint boundary conditions for steady and unsteady ows is given. A
stabilized nite element method is presented to solve the adjoint equations. A brief
description of the L-BFGS algorithm, which is used as the optimizer. Chapter 4
deals with the investigation of eect of span to chord ratio and mesh resolution on 3D
computations. Four cases have been studied at Re = 1 10
4
and = 4
o
. Optimal
span length and number of slices of 2D mesh to be used in span-wise direction are
found out in this chapter by studying various cases. Chapter 5 deals with the shape
optimization of 2D airfoil geometry in 3D unsteady ows at Re = 1 10
4
. The
1.3. Thesis organization 6
2D airfoil surface is parameterized by 13 control points using 4
th
order NURBS
curve. The initial guess taken in this 3D optimization is optimal shape for a 2D
unsteady ow. Further, shape optimization is carried out in a 3D unsteady ow with
NACA-0012 as initial guess at Re = 110
4
and = 4
o
. For the case of NACA-0012
optimization, the aerodynamic characteristics of airfoil geometry obtained after rst
complete iteration of optimization cycle is shown and three dimensionality of ow for
optimal airfoil is discussed. Concluding remarks and possible directions for future
research are presented in chapter 6.
Chapter 2
Governing Flow Equations and
Finite Element Formulation
The formulation of governing equations for the viscous, incompressible and unsteady
ows is discussed in this chapter. Also, the nite element discretization which is to
solve these equations is presented.
2.1 The Navier-Stokes equations
Let R
n
sd
and (0, T) be the spatial and temporal domains, respectively, where n
sd
is the number of space dimensions, and let denote the boundary of . The spatial
and temporal co-ordinates are denoted by x and t. The Navier-Stokes equations
governing an incompressible ow are

_
u
t
+ u.u
_
. = 0 on (0, T) (2.1)
.u = 0 on (0, T) (2.2)
7
2.1. The Navier-Stokes equations 8
where, , u and are the density, velocity and stress tensor, respectively. The
stress tensor is written as the sum of its isotropic and deviatric parts:
= pI + T, T = 2(u), (u) =
1
2
(u + (u)
T
)
where, p, I and are the pressure, identity tensor and dynamic viscosity, respec-
tively. The boundary conditions are either on the ow velocity or stress. Both,
Dirichlet and Neumann type boundary conditions are accounted for:
u = g on
g
(2.3)
n. = h on
h
(2.4)
where, n is the unit normal vector on the boundary . Here,
g
and
h
are subsets
of the boundary . More details on the boundary conditions are given in Figure
2.1.
U
and
D
represent the upstream and downstream boundaries, respectively.

S
and
W
represent the lateral boundaries. The surface of the body is represented
by
B
.
The initial condition on the velocity is specied on :
u(x, 0) = u
0
on (2.5)
where, u
0
is divergence free.
The drag and lift force coecients, (C
d
, C
l
), on the body are calculated using the
following expression:
(C
d
, C
l
) =
2
U
2
S
_

B
nd (2.6)
2.2. Finite element formulation 9
Figure 2.1: Schematic of the problem set-up: boundary conditions.
U
and
D
are
the upstream and downstream boundaries, respectively.
S
and
W
are the lateral
boundaries.
B
represent the surface body
.
The time-averaged coecients are calculated as follows:
C
d
=
1
T
_
t
0
+T
t
0
C
d
(t)dt (2.7)
C
l
=
1
T
_
t
0
+T
t
0
C
l
(t)dt (2.8)
The time-averaging begins at t = t
0
to leave out the transient eect of the initial
condition on the fully developed ow.
2.2 Finite element formulation
Consider a nite element discretization of into sub-domain
e
,e = 1, . . . , n
el
,
where n
el
is the number of elements. We dene:
H
1h
() = {
h
|
h
C
0
(),
h
|

e P
1
, e = 1, 2, . . . , n
el
} (2.9)
2.2. Finite element formulation 10
with P
1
representing rst-order polynomials. The trial and test function spaces are
dened as:
S
h
u
= {u
h
|u
h
(H
1h
)
n
sd
, u
h
= g
h
on
g
} (2.10)
V
h
u
= {w
h
|w
h
(H
1h
)
n
sd
, w
h
= 0 on
g
} (2.11)
S
h
p
= V
h
p
= {q
h
|q
h
H
1h
} (2.12)
where, n
sd
is the number of space dimensions.
The stabilized nite element formulation of Equations (2.1) and (2.2) is written as
follows: nd u
h
S
h
u
and p
h
S
h
p
such that w
h
V
h
u
, q
h
V
h
p
_

w
h
.
_
u
h
t
+ u
h
.u
h
_
d +
_

(w
h
) : (p
h
, u
h
) d
+
_

q
h
.u
h
d +
n
el

e=1
_

e
1

(
SUPG
u
h
.w
h
+
PSPG
q
h
).
_

_
u
h
t
+ u.u
_
.
_
d
e
+
n
el

e=1
_

LSIC
.w
h
.u
h
d
e
=
_

h
w
h
.h
h
d. (2.13)
The rst three terms and the right-hand side of Equation (2.13) constitute the
Galerkin formulation of the problem. The terms involving the element level integrals
are the stabilization terms added to the basic Galerkin formulation to enhance its
numerical stability. The term with coecient
LSIC
is also a stabilization term
based on the least squares of the incompressibility constraint and is found to be
useful for large Reynolds number ows. The generalized trapezoidal rule is used for
time discretization of u
h
. This is given by:
u
h
= u
i
+ (1 )u
i+1
, (2.14)
u
h
t
=
u
i+1
u
i
t
(2.15)
2.3. Stabilization parameter 11
where 0 1, and u
i
represents the velocity eld at time step i.
2.3 Stabilization parameter
The coecients
SUPG
and
PSPG
in Equation (2.13) are given by:

SUPG
=
PSPG
=
_
_
2||u
h
||
h
_
2
+
_
12
h
2
_
2
_
(2.16)
where h is the element length. In the present work, h is taken as the length of the
shortest edge in an element. The coecient
LSIC
is dened as

LSIC
=
_
_
2
h||u
h
||
2
_
2
+
_
12
h
2
||u
h
||
2
_
_
1/2
(2.17)
More details on the formulation can be found in Tezduyar et al. [31].
Chapter 3
The Adjoint Method for Shape
Optimization
In this chapter the problem set-up and the solution strategy used for optimal shape
design is presented. The adjoint method for design in uid ows is presented in
detail in section 3.2. The adjoint equations along with the boundary conditions are
derived this section. Section 3.3 briey describes the optimizer used in this work.
3.1 Problem set-up
The objective of optimal shape design is to alter either a part or whole of a surface
to achieve a desired objective. The objective could be to achieve a shape having
the maximum lift, least drag or highest aerodynamic eciency. The objectives are
posed in the form of cost functions. The cost function I
c
is a function of and
U. is the set of shape parameters that are used to represent the shape and U
represents the ow parameters. An optimal shape problem is formulated as:
Obtain , that minimizes or maximizes the cost function I
c
(U, ) subject to satis-
fying the following constraints:
12
3.1. Problem set-up 13
PDE constraints (U) = 0
Geometric constraints g() = 0
Optimization can be broadly classied as non-gradient based and gradient based
methods. Non-gradient based methods use only the cost function to obtain a op-
tima. Simplex [11], simulated annealing [12] and genetic algorithms [13] are a few
such methods that have been used for engineering applications. These methods are
simple to code and are very robust. They are well suited for problems in which the
gradient information is dicult to obtain. The drawback of these methods is the
high computational cost involved. For N design variables, the number of function
evaluations required is O(N
2
) while using genetic algorithms.
Gradient based methods use both the value of cost function and the gradient infor-
mation while seeking an optima. From a Taylor series expansion
I
c
+ I
c
I
c
+
_
I
c

_
T
The gradient vector (I
c
/) may be used to obtain a search direction that mini-
mizes the cost function. If one chooses
=
_
I
c

_
then, for suciently small , the cost function has to diminish.
I
c
+ I
c
= I
c

_
I
c

__
I
c

_
T
< I
c
Further improvements can be obtained by using quasi-Newton methods which use
the Hessian (
2
I
c
/
2
) to obtain new search direction. The simplest way to ob-
tain gradients (I
c
/
i
) is by using nite dierences. For each design variable, its
3.2. The adjoint approach 14
value is varied by a small quantity
i
, the ow and objective function recomputed
for the modied shape and the derivative of I
c
with respect to
i
calculated. An
optimization problem with N design variable requires N + 1 ow computations if
rst order forward dierences are used. Although the approach is simple, it involves
large computational costs if the number of design variables is large. Another disad-
vantage while using this approach is that the appropriate step size
i
is problem
specic and is not known apriori.
The gradients can be obtained in a fast and more ecient manner if the optimiza-
tion problem is formulated in the framework of theory of optimal control of systems
governed by partial dierential equations [18]. For example, shape parameters repre-
senting an airfoil can be considered as controls and the ow equations as constraints.
This leads to the so called adjoint based methods.
3.2 The adjoint approach
Let
B
be the segment of the boundary whose shape is to be determined. Let =
(
1
, . . . ,
m
) be the set of shape parameters that determine
B
. The optimization
problem involves nding the shape parameters that minimize (or maximize) the
objective function, I
c
(U, ) subject to satisfying (U, ) = 0, where represents
the ow equations. Depending on the accuracy needed various approximations to the
ow equations could be used. The ow equations are treated as constraint conditions
on the objective function. An augmented objective function is constructed to convert
the constrained problem to an unconstrained one. The ow equations are augmented
to the objective function by introducing a set of Lagrange multipliers or adjoint
variables .
I = I
c
+
_
T
0
_

.ddt (3.1)
3.2. The adjoint approach 15
The augmented objective function (3.1) degenerates to the original one if the ow
variables, U, exactly satisfy the ow equations. The rst variation of the augmented
objective function is given by:
I =
I
U
U +
I

+
I

(3.2)
The optimal solution is achieved when the variation of the augmented objective
function vanishes, i.e., I = 0. This requires that the variation of I with respect to
the ow variables U, design parameters and the adjoint variables should go to
zero, independently. These variations are given as:
I

=
_
T
0
_

(U, ) ddt (3.3)


I
U
=
_
I
c
U
+
_
T
0
_

T

U
ddt
_
(3.4)
I

=
_
I
c

+
_
T
0
_

ddt
_
(3.5)
Satisfying the ow equations sets the variation of I with respect to to zero .
The variation of I with respect to U, given by Equation (3.4), when set to zero,
leads to a set of equations and boundary conditions that are used to evaluate the
adjoint variables. The gradient,
I

, given by Equation (3.5) is utilized to nd the


optimal shape parameters. When the ow and adjoint equations are satised, the
augmented objective function is seen to be a function only of the design parameters.
This implies that the gradient can be determined without the need for additional
ow computations, i.e. the cost of evaluating the gradients is independent of the
number of design parameters.
3.2. The adjoint approach 16
3.2.1 The continuous adjoint equations
The equations and boundary conditions for the adjoint variables are obtained by
setting the variation of I with respect to the ow variables U, given in equation
(3.4), to zero. The equations governing the adjoint variables are:

u
t
+ (u)
u
(u.)
u
_
.

= 0 on (0, T) (3.6)
.
u
= 0 on (0, T) (3.7)
where,

is similar to the stress tensor and is given by

=
p
I + [
u
+
(
u
)
T
].
The adjoint equations (3.6) and (3.7) are a set of coupled linear partial dierential
equations. Unlike the ow equations (2.1) and (2.2), the equations for the adjoint
variables are posed backward in time.
3.2.2 Adjoint boundary and terminal conditions
The boundary conditions on the adjoint variables are:

u
= 0 on
U
(3.8)
s = 0 on
D
(3.9)
s
1
= 0, s
3
= 0,
u
2
= 0 on
S
(3.10)
s
1
= 0, s
2
= 0,
u
3
= 0 on
W
(3.11)

_
T
0
_

B
(.n).
u
ddt +
I
c
u
u +
I
c
p
p = 0 on
B
(3.12)
where, s = {
u
uI
p
+[
u
+(
u
)
T
]}.n.
U
and
D
represent the upstream
and downstream boundaries, respectively.
S
and
W
represent the lateral bound-
aries. The surface of the body is represented by
B
. The terminal condition on the
3.2. The adjoint approach 17
adjoint velocity is given by:

u
(u, T) = 0 on (3.13)
The conditions on the boundary, given by equation (3.12), depend on the denition
of the objective function. This is illustrated with a few examples for steady and
unsteady ows.
3.2.2.1 Adjoint boundary conditions for steady ows
Let I
c
=
1
2
C
2
d
, where, C
d
is the coecient of drag acting on the body. This objective
function is useful in designing a body that yields minimum drag coecient. In this
case, it can be shown that, the boundary conditions for the adjoint variables on the
body surface given by Equation (3.12) can be simplied to:

u
=
_

2C
d
U
2
S
, 0, 0
_
on
B
3.2.2.2 Adjoint boundary conditions for unsteady ows
The objective function to minimize the time-averaged drag coecient is given by
I
c
=
1
2
C
d
2
. In this situation the boundary condition is given by

u
=
_

2C
d
TU
2
S
, 0, 0
_
on
B
(0, T)
Here, T is the time duration for which the aerodynamic coecients are averaged
and adjoints computed. Similarly, the objective function to minimize the ratio of
the time-averaged drag coecient to the time-averaged lift coecient is given by
I
c
=
1
2
_
C
d
C
l
_
2
. In this case the boundary condition on
B
is given by

u
=
_

2
TU
2
S
C
d
C
l
2
,
2
TU
2
S
C
d
2
C
l
3
_
on
B
(0, T)
3.2. The adjoint approach 18
3.2.3 Finite element formulation for adjoint equations
A stabilized SUPG/PSPG nite element method is proposed to solve the adjoint
equations (3.6) and (3.7). The trial and test function spaces are dened as:
S
h

u
= {
h
u
|
h
u
(H
1h
)
n
sd
,
h
u
.
= g
h
on
g
}
V
h

u
= {w
h

u
|w
h

u
(H
1h
)
n
sd
, w
h

u
.
= 0 on
g
}
S
h

p
= V
h

p
= {q
h

p
|q
h

p
H
1h
}
The stabilized nite element formulation of equations (3.6) and (3.7) is as follows:
given u
h
and p
h
satisfying equations (2.1) and (2.2), nd
h
u
S
h

u
and
h
p
S
h

p
such that w
h

u
V
h

u
and q
h

p
V
h

p
,
_

w
h

u
.
_

h
u
t
+ (u
h
)
h
u
u.
u
_
d
+
_

(w
h

u
) :

(
h
p
,
h
u
) d +
_

q
h

p
.
h
u
d
+
n
el

e=1
_

e
1

SUPG
(
_
u
h
)w
h

u
u
h
.w
h

u
_
+
PSPG
q
h

p
_
.
_

h
u
t
+ (u
h
)
h
u
u.
u
_
.

(
h
p
,
h
u
)
_
d
e
+
n
el

e=1
_

LSIC
.w
h

u
.
h
u
d
e
= 0 (3.14)
The stabilization coecients
SUPG
,
PSPG
and
LSIC
in the formulation proposed
in equation (3.14) are computed based on the ow variables (u, p) and are given in
equations (2.16) and (2.17), respectively.
3.3. The optimizer 19
3.2.4 The gradient
The variation of the augmented objective I function with respect to the design
parameters gives the gradient. The gradient is given by:
I

=
I
c

_
T
0
_

p
.uddt
+

_
T
0
_

u
.
_
(
u
t
+ u.u) .
_
ddt (3.15)
It quanties the sensitivity of the objective function to the design parameters. It is
used by the optimizer to rene the search direction. The vanishing of the gradient
reects the attainment of optimal shape.
3.3 The optimizer
The optimization algorithm used in the present work is the L-BFGS (Limited
memory-Broyden-Fletcher-Goldfarb-Shanno) procedure. This method is derived
from the Newton-Raphson method that seeks to nd a stationary point of a function
f(x) where x could be a vector. In the Newton-Raphson method, marching from
iterate n to n + 1 is is done by the following:
x
n+1
= x
n

n
[Hf(x
n
)]
1
f(x
n
) (3.16)
where
n
is the step length, f(x
n
) and Hf(x
n
) are the gradient and the Hessian
matrix, respectively. In general, computing the Hessian and its inverse are expensive
operations and need to be approximated. In the BFGS approach, the following is
undertaken to march from iterate n to n + 1:
x
n+1
= x
n

n
B
n
f(x
n
) (3.17)
3.3. The optimizer 20
where the inverse Hessian approximation B
n
is updated at every iteration by means
of the formula
B
n+1
= V
T
n
B
n
V
n
+
n
s
n
s
T
n
(3.18)
where

n
=
1
y
T
n
s
n
, V
n
= I
n
y
n
s
T
n
and
s
n
= x
n+1
x
n
, y
n
= f
n+1
f
n
The inverse Hessian approximation B
n
is generally very dense. For a problem with
a large number of design variables the cost of storing B
n
becomes prohibitive. To
circumvent this a certain number of vector pairs {s
i
, y
i
} are stored. The product
B
n
f(x
n
) is then obtained by performing a series of inner products and vector
summations involving f(x
n
) and the pairs {s
i
, y
i
}. More details about this method
can be obtained in the book by Nocedal and Wright [32] and in the article by Byrd
et al. [33]. This algorithm is capable of handling upper and lower bounds on the
design variables.
Chapter 4
Study of eect of span and mesh
resolution in 3D ows at
Re = 1 10
4
4.1 Introduction
This chapter mainly deals with the study of the parameters which play a vital role in
calculations in 3D ows. Signicant research has been carried out so bar for airfoils
at low Re. Carmichael [34] has provided a good deal of data of airfoils at low Re.
Srinath and Mittal [2] have also studied the shape optimization of airfoil in low Re.
Also, Srinath and Mittal [29, 30] have studied the shape optimization of airfoils in
high Re in 2D. But, all these above mentioned optimal airfoils may not be optimal in
3D unsteady ows. So, before studying the optimization in 3D unsteady ows, the
parameters used in a 3D ow are investigated. There are two main objectives of this
study. Firstly, to check the eect of span in three dimensional ows i.e. how does
the variation in span eects the values of aerodynamic coecients and what span
21
4.2. Problem description and mesh 22
is optimal for computations in three dimensional ows. Secondly, to investigate the
eect of mesh resolution for three dimensional unsteady ows. This study gives the
idea about the number of slices which are optimum for study of three dimensional
eects. These studies are carried out at = 4
o
and Re = 1 10
4
for 13 control
points. The time averaged and root mean squared aerodynamic coecients have
been compared at dierent time steps.
4.2 Problem description and mesh
Srinath and Mittal [29, 30] have formulated a continuous adjoint method for shape
optimization of airfoils and implemented it in unsteady optimization problems in 2D.
An optimal shape of airfoil was obtained after optimizing NACA-0012 in 2D. This
optimal airfoil of unit chord length is placed in a rectangular domain. The upstream
and downstream boundaries are located 20 chord lengths from the trailing edge of
airfoils. The lateral boundaries are at a distance of 10 chord length each, from the
airfoil. The Reynolds number is based on chord of airfoil. A typical nite element
mesh, used for computations is shown in Figure 4.2. It consists of 44804 nodes and
89304 elements with 200 points on surface. A structured mesh is employed close to
the boundary to resolve boundary layer ow and the rest of domain is discretized
using an unstructured mesh which is employed by using Delaunay Triangulation
method [35]. Further four cases have been studied in which mesh is stacked to 13,
26, 39 and 52 slices with span to chord equal to 0.25, 0.5, 0.5 and 1.0 respectively,
for 3D calculations. The main objective of the study done in this chapter is to
investigate the eect of span to chord ratio and mesh resolution on the computations
done for a 3D ow. A 3D mesh consisting of 26 slices with a span to chord ratio
equal to 0.5 is shown in Figure 4.3.
4.2. Problem description and mesh 23
4.2.1 Parameterization using NURBS curves:
To carry out shape optimization, parameterization should be the one which oers
a rich design space. A number of methods have been used to parameterize airfoil.
Castonguay and Nadarajah [36] have discussed various types of techniques of pa-
rameterization and their eect on shape optimization. In this work NURBS(Non
uniform rational B-splines) has been used to parameterize the airfoil geometry [37].
A NURBS curve is dened as:
C(t) =

n
i=0
N
i,p
(t)w
i
P
i

n
i=0
N
i,p
(t)w
i
, t T (4.1)
where, T = { t
0
, t
1
, t
2
, ......., t
k
} t
i
( 0, 1 ), k = p + n
Here, T is the knot vector, n is the number of control points, p is the order of the
curve, t is the curve parameter, w
i
is the weight associated with control point P
i
and N
i,p
are the B-Spline basis functions, dened on T, given by:
N
i,0
=
_

_
1 if t
i
t t
i+1
0 otherwise
(4.2)
N
i,p
=
t t
i
t
i+p
t
i
N
i,p1
(t) +
t
i+p+1
t
t
i+p+1
t
i+1
N
i+p,p1
(t) (4.3)
The basis functions can be calculated by equations (4.2) and (4.3). A variety of
curves can be obtained by changing the location and weight of the control points or
by changing the knot sequence. In the present work 4
th
order NURBS curve is used
with 13 control points in sections 4.3 and 4.4 to model the surface of airfoil. Figure
4.1 shows an attempt to model an airfoil, which is optimal for 2D unsteady ow. As
seen from the gure control point 1 and 13 are identical and forms a closed curve
and sharp trailing edge. Leading and trailing edges are xed to achieve the desired
chord length and angle of attack.
4.2. Problem description and mesh 24
Figure 4.1: Parameterization of airfoil, used as initial guess for 3D computations,
using 4
th
order of NURBS curve with 13 control points. The y-co-ordinates of the
control points 2-6 and 8-12 are used for optimizing the airfoil geometry
4.2.2 Various cases studied and the parameters used in those
cases:
Basically, this study is done to validate the parameters to be used in 3D optimization
in further part of this work. Span of wing and Mesh resolution play a vital role in
3D computations. So, their eect on ow calculations, is investigated by doing the
various cases of ow computations. The cases which are done are shown in Table
4.1. One of the stacked mesh is shown in Figure 4.3. This mesh shows the 2D mesh
stacked to 26 slices in span-wise direction for a span to chord ratio of 0.5. Number
of nodes and elements used in 2D mesh are 44804 and 89304 respectively. These
computations are carried out at = 4
o
and Re = 1 10
4
for 13 control points.
The ow averaging is done for 13500 time steps for each case.
4.2.3 Time averaging of aerodynamic coecients:
To investigate the time duration that is necessary for the time averaging of aerody-
namic coecients to achieve a stationary state was of sheer importance. In various
4.2. Problem description and mesh 25
X
Y
Z
Figure 4.2: Close-up view of nite element mesh for the 2D optimized shape at 4
o
angle of attack. The mesh contains 44804 nodes and 89304 elements.
Table 4.1: The number of slices of 2D mesh and span to chord ratio for various
cases, computed at Re = 1 10
4
and = 4
o
Case No. of slices along span Span to Chord ratio
a 13 0.25
b 26 0.5
c 39 0.5
d 52 1.0
4.2. Problem description and mesh 26
Figure 4.3: A 3D nite element mesh used for 3D calculations. Mesh consists of
wedge element. 26 slices of 2D mesh 5.5 has been used to generate this mesh.
cases, done at Re = 1 10
4
, it is observed from the time history of C
l
that varia-
tion of C
l
is non periodic. This is due to the fact that the ow obtained for higher
Reynolds number is non periodic. Therefore, if the ow is not averaged for sucient
time, the value of time averaged aerodynamic coecient is not very reliable. One
way of averaging the ow is to start the averaging in forward direction i.e in the
direction of increasing time, from certain time unit t
0
. The data for t < t
0
are not
utilized for time averaging as it consists of the initial transience which might be
4.3. Eect of span to chord ratio in 3D ow computations for Re = 1 10
4
and
= 4
o
27
aected by the initial condition with which the computations begin. The other way
of averaging the ow is to start the averaging in reverse direction i.e. back in time.
This method also eliminates the eect of initial transience. To have large amount
of data for averaging, ow calculations were done for 135 time units. The analysis
of data has been done at 30, 40, 50 and 60 time units to ensure that maximum of
data has been taken for averaging and also the transience region is excluded.
4.2.4 Optimal shape for 2D unsteady ows, used as initial
guess:
In the earlier work done by Srinath and Mittal [29, 30] in 2D, the initial guess
was taken as NACA-0012 and this airfoil was further optimized. But in this study
initial guess during optimization is taken as shape obtained after optimization in
2D ow, which was obtained by Srinath and Mittal [29, 30]. This optimized shape
obtained in 2D ow is used as initial guess in 3D optimization, rather than starting
the optimization cycle from NACA-0012, to avoid the convergence at some local
optima. The 3D optimization was to be carried out on this shape, optimal for 2D
unsteady ow, as initial guess. Therefore all the ow cases i.e. to obtain an optimal
span and mesh resolution have been done on this shape shown in Figure 4.1
4.3 Eect of span to chord ratio in 3D ow com-
putations for Re = 1 10
4
and = 4
o
In this study, a 2D optimized shape is taken for ow calculations. This 2D optimized
airfoil shape is parameterized by 13 design variables using NURBS curves. Further,
a typical nite element mesh is generated around this shape. This 2D mesh consists
of 44804 nodes and 89304 elements with 200 points on surface. A structured mesh
is employed close to the boundary to resolve boundary layer ow and the rest of
4.3. Eect of span to chord ratio in 3D ow computations for Re = 1 10
4
and
= 4
o
28
Table 4.2: Comparison of averaged and rms aerodynamic coecients of 13, 26 and
52 slices for span to chord ratio of 0.25, 0.5 and 1.0 respectively, to show the eect
of span to chord ratio
Coecients Time units 13 slices % di 26 slices % di 52 slices
C
l
avg 30 0.889 0.43 0.893 0.11 0.892
40 0.894 0.42 0.890 0.22 0.892
50 0.891 0.34 0.887 0.23 0.889
60 0.888 0.79 0.881 0.57 0.886
C
l
rms 30 0.0914 7.85 0.0840 0.71 0.0834
40 0.0925 10.35 0.0838 1.90 0.0822
50 0.0916 9.82 0.0834 1.92 0.0818
60 0.0933 13.67 0.0821 1.09 0.0812
C
d
avg 30 0.0868 1.71 0.0884 0.90 0.0892
40 0.0869 1.70 0.0884 0.68 0.0890
50 0.0868 1.65 0.0882 0.91 0.0890
60 0.0868 1.38 0.0881 0.91 0.0889
C
d
rms 30 0.0114 3.08 0.0112 4.46 0.0117
40 0.0114 1.22 0.0114 2.63 0.0117
50 0.0112 0.58 0.0112 3.57 0.0116
60 0.0110 0.41 0.0110 3.63 0.0114
domain is discretized using an unstructured mesh which is employed by using De-
launay Triangulation method. Three cases are studied to investigate the eect of
span to chord ratio in 3D computations. In these cases, the above mentioned 2D
mesh is stacked to 13, 26 and 52 slices, and span to chord ratio being 0.25, 0.5 and
1.0 respectively. The ow is computed for 13500 time steps for all the three cases.
It is also observed that ow is supposed to run for at least this much time to get
the convergence in the time averaged lift and drag coecients, as can be seen from
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
29
the Figures 4.4 to 4.7 and Figures 4.8 to 4.11. The average and rms of aerodynamic
coecients are calculated in reverse direction. Averaged and rms aerodynamic co-
ecients are calculated at 30, 40, 50 and 60 time units and are compared in Figures
4.12 to 4.15, and the numerical values are compared in Table 4.2. Since, the percent-
age dierence in time averaged and rms aerodynamic coecients is larger between
13 and 26 slices case as compared to between 26 and 52 slices case. Therefore, it
is concluded that span to chord ratio of 0.5 is enough to obtain the complete 3D
eects. Since ow at Re = 1 10
4
is found to be non periodic. Hence, rather
than plotting instantaneous quantities various averaged quantities such as Reynolds
stresses and velocity components are plotted, as shown in Figures 4.16 to 4.18 at
the center slice in all the cases and are compared. Unsteadiness, three dimensional
eects can be studied from these pictures, which look reasonably comparable for all
other cases except for 13 slices case. C
p
plots have been plotted, as shown in Figure
4.19 at the center slice for all the cases and are found to be very much comparable.
These computations conclude the optimal span to chord ratio which must be used
for 3D computations. But, still nothing can be concluded about mesh resolution
from these calculations. So, for mesh resolution further one more case is studied.
4.4 Eect of Mesh Resolution in 3D ow compu-
tations for Re = 1 10
4
and = 4
o
In this study also the same 2D optimized shape is taken for ow calculations and
parameterization is done in the same manner. Also the 2D mesh formed around the
airfoil is same as in previous case. Three cases are studied to investigate the eect of
Mesh resolution in 3D computations. In these cases, the above mentioned 2D mesh
is stacked to 13, 26 and 39 slices, and span to chord ratio being 0.25, 0.5 and 0.5 re-
spectively. The number of nodes and elements in 3D meshes for all the cases studied
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
30
Table 4.3: The number of nodes and elements in 3D mesh of all the four cases
computed at Re = 1 10
4
and = 4
o
Slices Span to Chord ratio No. of nodes No. of elements
13 0.25 582452 1071648
26 0.50 1164904 2232600
39 0.50 1747356 3393552
52 1.00 2329808 4554504
are given in Table 4.3. The ow is computed for13500 time steps for all the three
cases. The average and rms of aerodynamic coecients are calculated in reverse di-
rection as shown in Figures 4.4 to 4.11. Averaged and rms aerodynamic coecients
are calculated at 30, 40, 50 and 60 time units and are compared in Figures 4.12 to
4.15, and the numerical values are compared in Table 4.4. Since, the percentage
dierence in time averaged and rms aerodynamic coecients is larger between 13
and 26 slices case as compared to between 26 and 39 slices case. Therefore, it is
concluded that stacking the 2D mesh to 26 slices gives the reliable results for a 3D
unsteady ow. Since ow at Re = 110
4
is found to be non periodic. Hence, rather
than plotting instantaneous quantities various averaged quantities such as Reynolds
stresses and velocity components are plotted, as shown in Figures 4.16 to 4.18 at the
center slice in all the cases and are compared. These computations nally conclude
the mesh resolution which should be used in 3D computations, which ensures that
full 3D eects are studied and optimal amount of disk space is used.
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
31
Table 4.4: Comparison of averaged and rms aerodynamic coecients of 13, 26 and
39 slices for span to chord ratio of 0.25, 0.5 and 0.5 respectively, to show the eect
of mesh resolution
Coecients Time units 13 slices % di 26 slices % di 39 slices
C
l
avg 30 0.889 0.43 0.893 0.89 0.884
40 0.894 0.42 0.890 1.12 0.879
50 0.891 0.34 0.887 0.51 0.882
60 0.888 0.79 0.881 0.45 0.885
C
l
rms 30 0.0914 7.85 0.0840 1.45 0.0828
40 0.0925 10.35 0.0838 3.46 0.0809
50 0.0916 9.82 0.0834 0.36 0.0827
60 0.0933 13.67 0.0821 1.82 0.0836
C
d
avg 30 0.0868 1.71 0.0884 0.03 0.0884
40 0.0869 1.70 0.0884 0.02 0.0883
50 0.0868 1.65 0.0882 0.29 0.0885
60 0.0868 1.38 0.0881 0.51 0.0885
C
d
rms 30 0.0114 3.08 0.0112 3.91 0.0107
40 0.0114 1.22 0.0114 6.31 0.0107
50 0.0112 0.58 0.0112 2.51 0.0109
60 0.0110 0.41 0.0110 1.72 0.0109
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
32
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0 20 40 60 80 100 120 140
C
l
time units
Cl
Cl avg
Cl rms
Figure 4.4: Time history of C
l
along with averaging and rms from reverse direction
for: 13 slices and 0.25 span case
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0 20 40 60 80 100 120 140
C
l
time units
Cl
Cl avg
Cl rms
Figure 4.5: Time history of C
l
along with averaging and rms from reverse direction
for: 26 slices and 0.5 span case
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
33
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0 20 40 60 80 100 120 140
C
l
time units
Cl
Cl avg
Cl rms
Figure 4.6: Time history of C
l
along with averaging and rms from reverse direction
for: 39 slices and 0.5 span case
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0 20 40 60 80 100 120 140
C
l
time units
Cl
Cl avg
Cl rms
Figure 4.7: Time history of C
l
along with averaging and rms from reverse direction
for: 52 slices and 1.0 span case
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
34
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0 20 40 60 80 100 120 140
C
d
time units
Cd
Cd avg
Cd rms
Figure 4.8: Time history of C
d
along with averaging and rms from reverse direction
for: 13 slices and 0.25 span case
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0 20 40 60 80 100 120 140
C
d
time units
Cd
Cd avg
Cd rms
Figure 4.9: Time history of C
d
along with averaging and rms from reverse direction
for: 26 slices and 0.5 span case
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
35
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0 20 40 60 80 100 120 140
C
d
time units
Cd
Cd avg
Cd rms
Figure 4.10: Time history of C
d
along with averaging and rms from reverse direction
for: 39 slices and 0.5 span case
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0 20 40 60 80 100 120 140
C
d
time units
Cd
Cd avg
Cd rms
Figure 4.11: Time history of C
d
along with averaging and rms from reverse direction
for: 52 slices and 1.0 span case
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
36
0.70
0.75
0.80
0.85
0.90
0.95
1.00
1.05
0 20 40 60 80 100 120 140
C
l
time units
Cl avg 13-slices
Cl avg 26-slices
Cl avg 39-slices
Cl avg 52-slices
Figure 4.12: Comparison of time averaged lift coecients for all the cases
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0 20 40 60 80 100 120 140
C
l
time units
Cl rms 13-slices
Cl rms 26-slices
Cl rms 39-slices
Cl rms 52-slices
Figure 4.13: Comparison of rms lift coecients for all the cases
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
37
0.065
0.070
0.075
0.080
0.085
0.090
0.095
0.100
0.105
0 20 40 60 80 100 120 140
C
d
time units
Cd avg 13-slices
Cd avg 26-slices
Cd avg 39-slices
Cd avg 52-slices
Figure 4.14: Comparison of time averaged drag coecients for all the cases
0.000
0.002
0.004
0.006
0.008
0.010
0.012
0.014
0.016
0 20 40 60 80 100 120 140
C
d
time units
Cd rms 13-slices
Cd rms 26-slices
Cd rms 39-slices
Cd rms 52-slices
Figure 4.15: Comparison of rms drag coecients for all the cases
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
38
Figure 4.16: Averaged uv component of Reynolds Stress is compared for the center
slice: (a) 7th slice for 13 slices case, (b) 13th slice for 26 slices case, (c) 20th slice
for 39 slices case, (d) 26th slice for 52 slices case
Figure 4.17: Averaged uw component of Reynolds Stress is compared for the center
slice: (a) 7th slice for 13 slices case, (b) 13th slice for 26 slices case, (c) 20th slice
for 39 slices case, (d) 26th slice for 52 slices case
4.4. Eect of Mesh Resolution in 3D ow computations for Re = 1 10
4
and
= 4
o
39
Figure 4.18: Averaged w component of velocity is compared for the center slice: (a)
7th slice for 13 slices case, (b) 13th slice for 26 slices case, (c) 20th slice for 39 slices
case, (d) 26th slice for 52 slices case
Figure 4.19: Averaged pressure coecient is compared for the center slice: (a) 7th
slice for 13 slices case, (b) 13th slice for 26 slices case, (c) 20th slice for 39 slices
case, (d) 26th slice for 52 slices case
Chapter 5
Shape optimization of 2D
geometry for 3D unsteady ows
The previous chapter dealt with validation of input parameters such as span to
chord ratio and mesh resolution to be used in 3D optimization for unsteady ows.
This chapter deals with airfoil shape optimization in unsteady ows. Section 5.1
provides a brief description of airfoil design in unsteady ows attempted by others.
The working of optimization cycle is described in Section 5.3. The whole working of
complete design cycle is shown through a ow chart and each module is described.
Section 5.4 deals with the shape optimization of 2D airfoil geometry for 3D unsteady
ows at = 4
o
and Re = 1 10
4
. Shape optimization in a 3D unsteady ow using
NACA-0012 as initial guess, at = 4
o
and Re = 1 10
4
is discussed in Section 5.5.
5.1 Introduction
A thorough study has been done on airfoils operating at low Reynolds number (Re <
5 10
5
) due to their importance in a variety of applications such as micro-aerial
vehicle(MAV), unmanned aerial vehicles(UAV), remotely piloted vehicle(RPV) etc.
40
5.2. Problem description and mesh 41
The conventional airfoils operating at high Re are found to have a non-optimal
performance at low Re. At low Re viscous eects are relatively high leading to
large drag and low lift to drag ratios. Sunada et al.[38] experimentally measured
the aerodynamic characteristics of fteen airfoils at Re < 4 10
3
including sev-
eral NACA 4 digit airfoils and corrugated airfoils. They observed that the airfoil
with better aerodynamic characteristics should be thinner at Re < 4 10
3
than
those of higher Re and also should have sharp leading edge. These airfoils which
are optimal for low Re 2D ows may not necessarily be optimal for 3D unsteady
real ows. Huyse and Lewis [17] studied the aerodynamic shape optimization of
two-dimensional airfoils under uncertain operating conditions. They observed that
a deterministic optimization for discrete operating conditions may result in dra-
matically inferior performance when the actual conditions are dierent from these,
somewhat arbitrarily chosen, design values.
Srinath and Mittal [29, 30] formulated a continuous adjoint method for shape op-
timization of airfoils and implemented it in unsteady optimization problems in 2D.
There are two main objectives of this study. Firstly, to further optimize a shape,
which is optimal for 2D unsteady ows, in a 3D unsteady ow at = 4
o
and
Re = 110
4
. Secondly, the shape optimization in a 3D unsteady ow using NACA-
0012 as initial guess and studying the transition of ow from 2-dimensional to 3-
dimensional.
5.2 Problem description and mesh
Srinath and Mittal [29, 30] formulated a continuous adjoint method for shape opti-
mization of airfoils and implemented it in unsteady optimization problems in 2D. An
optimal shape of airfoil was obtained after optimizing NACA-0012 in a 2D unsteady
ow at Re = 1 10
4
. This airfoil of unit chord length is placed in a rectangular
5.2. Problem description and mesh 42
domain. The upstream and downstream boundaries are located 20 chord lengths
from the trailing edge of airfoils. The lateral boundaries are located at 10 chord
length each, from the airfoil. The Reynolds number is based on chord of airfoil.
A typical nite element mesh, used for computations is shown in Figure 4.2. It
consists of 44804 nodes and 89304 elements with 200 points on surface. A struc-
tured mesh is employed close to the boundary to resolve boundary layer ow and
the rest of domain is discretized using an unstructured mesh which is employed by
using Delaunay Triangulation method [35]. Since, it has already been shown in the
previous chapter that 26 slices along with a span to chord ratio of 0.5 are optimal
for 3D calculations, so these parameters are used for stacking the 2D mesh in span
wise direction for the all the 3D optimizations done. 3D mesh for 26 slices and 0.5
span to chord ratio is shown in Figure 4.3.
5.2.1 Mesh movement strategy
After each optimization cycle, a new geometry of airfoil is obtained. Therefore, a
moved mesh for new obtained geometry is required after each iteration for computing
the ow. In the present work, a mesh moving scheme is employed. A nite element
mesh is generated and the nodes are subsequently relocated as per the new shapes
while maintaining the connectivity between them. The computational domain is
modeled as a linearly elastic solid. Smaller elements are made stier to reduce
the distortion of mesh. The modied equations of linear elasticity are solved for
the internal nodal displacements based on the given shape deformation of the solid
boundary. Further details about this technique can be found in article by Tezduyar
et al. [39]. After each iteration new mesh is obtained by above mentioned procedure
and then 3D mesh for new shape is obtained by stacking the 2D mesh, keeping the
2D mesh same at all slices.
5.3. Flow Chart of an optimization code used for 3D optimization in unsteady
ows at Re=1 10
4
43
5.3 Flow Chart of an optimization code used for
3D optimization in unsteady ows at Re=1
10
4
In this section details of the working of an optimization code are described step by
step. Working of the optimization code used can also be seen in owchart 5.1.
The details of the working of an optimization code are listed below:
1. Starting with an initial guess having certain design parameters and generate
initial mesh for that shape.
The shape parameters, , that represent the surface to be optimized, are
identied. In the present work, surface is parameterized using NURBS (Non-
Uniform Rational Bi-cubic Spline) with 13 control points as shown in Figure
4.1. The mesh consists of 44804 nodes and 89304 triangular elements with
200 nodes on the surface. A structured mesh is employed close to the airfoil
surface and in the near wake to resolve the ow structures adequately. The
remaining domain is lled with an unstructured mesh that is generated via
Delaunay triangulation [35]. Further this 2D mesh 4.2 is stacked with 26 slices
and span length is taken as half of the chord length.
2. Computation of unsteady ow, u and p and calculate the objective function I
c
.
The nite element formulation given by equation (2.13) is utilized to compute
the unsteady ow. The computed ow is used to compute I
c
. The ow solution
is stored on the hard disk after every 100 time steps.
3. Compute the adjoint variables,
u
and
p
.
5.4. Shape optimization of 2D airfoil geometry in a 3D unsteady ow at
Re=1 10
4
and = 4
o
44
The nite element formulation given by equation (3.14) and the unsteady ow
computed in previous step is utilized to compute the adjoint variables.
4. Compute the gradient given by equation (3.15).
A nite dierence approach, as described by Soto and Lohner [40] is used to
calculate the gradient. Mathematically:
I

=
I(+) I()

.
5. If the convergence criteria, either on the objective function or the gradient, is
satised, then stop.
6. Update the shape parameters, , in the direction of the gradient.
The L-BFGS (Limited memory-Broyden-Fletcher-Goldfarb-Shanno) [33] is used
as the optimizer in the present work.
7. Modify mesh to accommodate new shape. A mesh moving scheme is employed
to modify the mesh which is described in Section 5.2.1.
5.4 Shape optimization of 2D airfoil geometry in
a 3D unsteady ow at Re=1 10
4
and = 4
o
In this case, an airfoil geometry is optimized in 2D, whereas, the ow is kept 3D.
An airfoil geometry obtained after optimizing NACA-0012 in a 2D unsteady ow is
taken as initial guess. This airfoil shape is parametrized by 13 control points using
4th order NURBS curve as shown in Figure 4.1. The behavior of this 2D geometry is
studied in 3D unsteady ows. The lift coecient for airfoil geometry, obtained after
2D optimization, is found to reduce from 1.03 to 0.86 from 2D to 3D unsteady ows,
respectively. This optimal geometry of 2D ows is further allowed to optimize with
same bounds as used earlier. Bounds are the range which we provide to the control
points in the y-direction for movement. The objective function is maximization of
5.4. Shape optimization of 2D airfoil geometry in a 3D unsteady ow at
Re=1 10
4
and = 4
o
45
Figure 5.1: A ow-chart detailing the steps of the optimization process.
lift. The same 2D nite element mesh is used as mentioned in previous chapter as
shown in Figure 4.2. 26 slices of 2D mesh are used to obtain a typical 3D nite
element mesh and half of a chord length is taken as span length.This 3D nite
element mesh stacked for 26 slices is shown in Figure 4.3. The optimization is
carried out at Re=1 10
4
and = 4
o
. In one optimization cycle, ow is solved
for 80 time units and time averaging of coecients is done for 10 time units. Three
iteration cycles have been carried out and time averaged lift coecient is found
to increase from 0.86 to 0.90, whereas, time averaged drag coecient is found to
reduce from 0.089 to 0.075 after third iteration. Figure 5.2 shows the initial shape
and the optimal shape obtained after third iteration of optimization cycle. Table
5.1 shows the time averaged aerodynamic coecients obtained as a result, after
three iterations of optimization cycle. Time histories of lift and drag coecients are
also shown for initial and optimal shapes in Figure 5.3 and 5.4. This optimization
also shows that, the shape which is optimal for 2D unsteady ows may not be the
5.5. Shape optimization in an unsteady 3D ow, taking NACA-0012 as initial
guess at Re=1 10
4
and = 4
o
46
Table 5.1: The time averaged lift and drag coecients of the initial and the optimal
airfoil shapes for maximization of lift coecient at Re = 1 10
4
and = 4
o
.
Airfoils C
l
C
d
(
C
l
C
d
)
Initial shape 0.86 0.089 9.66
Optimal shape 0.90 0.075 12.00
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
-1 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0
Y
X
initial shape
3rd iter-shape
Figure 5.2: Further optimization of shape, optimal for 2D ow, in a 3D unsteady
ow at Re=1 10
4
and = 4
o
: initial and optimal shapes after three iterations of
optimization cycle.
optimal shape for 3D unsteady ows.
5.5 Shape optimization in an unsteady 3D ow,
taking NACA-0012 as initial guess at Re=1
10
4
and = 4
o
In this study, NACA-0012 is used as initial guess for optimization in a 3D unsteady
ow. NACA-0012 is parametrized by 13 control points using 4th order NURBS
curve. The 2D nite element mesh generated around NACA-0012 is shown in Figure
5.5. Shape optimization in an unsteady 3D ow, taking NACA-0012 as initial
guess at Re=1 10
4
and = 4
o
47
0.60
0.70
0.80
0.90
1.00
1.10
1.20
1.30
1.40
0 1000 2000 3000 4000 5000 6000 7000 8000
C
l
time units
initial shape
3rd iter-shape
Figure 5.3: Time histories of lift coecient for both initial and optimal shapes at
Re=1 10
4
and = 4
o
0.03
0.04
0.05
0.06
0.07
0.08
0.09
0.10
0.11
0.12
0 1000 2000 3000 4000 5000 6000 7000 8000
C
d
time units
initial shape
3rd iter-shape
Figure 5.4: Time histories of drag coecient for both initial and optimal shapes at
Re=1 10
4
and = 4
o
5.5. This 2D mesh consists of 44804 nodes and 89304 elements. Further, this 2D
mesh is stacked to 26 slices in span-wise direction to generate a 3D mesh. Half of
chord length is taken as span length for 3D ow computations. The optimization
is carried out at Re=1 10
4
and = 4
o
. In one optimization cycle, ow is solved
5.5. Shape optimization in an unsteady 3D ow, taking NACA-0012 as initial
guess at Re=1 10
4
and = 4
o
48
Table 5.2: The time averaged lift and drag coecients of NACA-0012 and the
optimal airfoil shapes for maximization of lift coecient at Re = 110
4
and = 4
o
.
Airfoils C
l
C
d
(
C
l
C
d
)
Initial shape 0.1493 0.049 3.046
Optimal shape 0.8520 0.070 12.17
for 80 time units and time averaging of coecients is done for 10 time units. One
iteration cycle has been carried out and time averaged lift coecient is found to
increase from 0.15 to 0.85. Figure 5.6 shows the initial and optimal airfoil shapes
after one iteration. Table 5.2 shows the time averaged aerodynamic coecients
obtained as a result, after one iteration of optimization cycle. Time histories of lift
and drag coecients are also shown for NACA-0012 and optimal shape in Figure
5.7 and 5.8. Figure 5.9 shows the magnitude of pressure eld for NACA-0012 and
optimal airfoil. It can be seen clearly that for optimal airfoil there is huge suction
at upper surface and high pressure at lower surface as compared to NACA-0012 and
due to this pressure dierence optimal airfoil has a 471% increase in lift coecient
over NACA-0012. Figure 5.10 shows the isosurfaces of the span-wise and stream-
wise component of vorticity. Flow over NACA-0012 airfoil is completely 2D at this
Reynolds Number, which can be seen from Figure 5.10(a). But, in case of optimal
airfoil, the stream-wise vorticity shows the transition of ow from 2D to 3D, which
is shown in Figure 5.10(b).
5.5. Shape optimization in an unsteady 3D ow, taking NACA-0012 as initial
guess at Re=1 10
4
and = 4
o
49
Figure 5.5: Close-up view of nite element mesh for the NACA-0012 at 4
o
angle of
attack. The mesh contains 44804 nodes and 89304 elements.
-0.02
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
-1 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0
Y
X
initial shape
1st iter-shape
Figure 5.6: Further optimization of shape, optimal for 2D ow, in a 3D unsteady
ow at Re=1 10
4
and = 4
o
: initial and optimal shapes after three iterations of
optimization cycle.
5.5. Shape optimization in an unsteady 3D ow, taking NACA-0012 as initial
guess at Re=1 10
4
and = 4
o
50
0.00
0.20
0.40
0.60
0.80
1.00
1.20
1.40
0 1000 2000 3000 4000 5000 6000 7000 8000
C
l
t
NACA-0012
optimal shape
Figure 5.7: Time histories of lift coecient for both initial and optimal shapes at
Re=1 10
4
and = 4
o
0.00
0.05
0.10
0.15
0.20
0 1000 2000 3000 4000 5000 6000 7000 8000
C
d
t
NACA-0012
optimal shape
Figure 5.8: Time histories of drag coecient for both initial and optimal shapes at
Re=1 10
4
and = 4
o
5.5. Shape optimization in an unsteady 3D ow, taking NACA-0012 as initial
guess at Re=1 10
4
and = 4
o
51
Figure 5.9: Unsteady optimization of NACA-0012 at Re=1 10
4
and = 4
o
:
magnitude of instantaneous pressure eld for (a) Initial Shape i.e. NACA-0012, (b)
Optimal Shape, after one iteration of optimization cycle
5.5. Shape optimization in an unsteady 3D ow, taking NACA-0012 as initial
guess at Re=1 10
4
and = 4
o
52
Figure 5.10: Unsteady optimization of NACA-0012 at Re=1 10
4
and = 4
o
:
instantaneous pictures of isosurfaces of stream-wise vorticity(
x
= 0.05) for (a)
NACA-0012, (b) Optimal Shape. Instantaneous pictures of isosurfaces of span-wise
vorticity(
z
= 5) for (c) NACA-0012, (d) Optimal Shape
.
Chapter 6
Conclusion
Prior to starting the optimization in a 3D ow, the parameters to be used in 3D
optimization have been investigated. These parameters are: span to chord ratio,
mesh resolution, and the time for which the ow calculations should be done to get
converged time-averaged aerodynamic coecients. Results in 3D ow solely depend
on these parameters, so it was essential to nd the optimal parameters to be used.
Four cases have been studied, taking 13, 26, 39 and 52 slices of 2D mesh in span-
wise direction with span to chord ratios of 0.25, 0.5, 0.5 and 1.0, respectively. The
airfoil used for these ow calculations is an optimal geometry for 2D unsteady ow
at Re = 110
4
and = 4
o
. The airfoil has been parameterized by 13 control points
via the 4th order NURBS curve. All cases are done at Re = 1 10
4
and = 4
o
.
Cases with 13, 26 and 52 slices have been used to investigate the eect of span to
chord ratio in 3D unsteady ows. It is shown that the span to chord ratio of 0.5
is optimal to get the complete 3D eects. Further, cases with 13, 26 and 39 slices
have been studied to check the eect of mesh resolution on ow calculations. It is
concluded that a 2D mesh stacked to 26 slices in span-wise direction is optimal to
study the eects in a 3D unsteady ow. Time-average and rms of the aerodynamic
coecients have been calculated in the reverse direction and compared.
53
6 Conclusion 54
Further, the gradient-based continuous adjoint method for shape optimization, pro-
posed by Srinath and Mittal [1, 2], is applied for shape optimization in 3D. The
Limited memory-Broyden-Fletcher-Goldfarb-Shanno(L-BFGS) algorithm has been
used as an optimizer. A nite element method, based on the streamline upwind
Petrov/Galerkin(SUPG) and pressure stabilized Petrov/Galerkin(PSPG) stabiliza-
tion techniques, is used to solve the adjoint equations.
After validating the parameters for 3D unsteady ows at Re = 1 10
4
, a 2D
airfoil geometry has been optimized for a 3D unsteady ow. A shape, optimal
for 2D unsteady ow, is used as an initial guess in 3D optimization. This shape
was obtained by Srinath and Mittal [29, 30] after optimizing NACA-0012 in a 2D
unsteady ow at Re = 110
4
and the lift coecient was found to increase from 0.15
to 1.03. The same optimal airfoil is found to have a lift coecient of 0.86 in a 3D
unsteady ow. This shows that values of the aerodynamic coecients diminish from
2D to 3D computations for the same airfoil shape. This shape is further optimized in
a 3D unsteady ow. The 2D mesh, consisting of 44804 nodes and 89304 elements,
is stacked to 26 slices in span-wise direction with a span to chord ratio of 0.5.
Three iteration cycles have been completed in this study and the results after three
iteration cycles showed an increase in the lift coecient value from 0.86 to 0.90.
This also leads to the conclusion that a shape optimal for 2D unsteady ows may
not be optimal for 3D unsteady ows. Further, shape optimization is done in a 3D
unsteady ow, with NACA-0012 directly as an initial guess, at Re = 1 10
4
and
= 4
o
. The 2D mesh, consisting of 44804 nodes and 89304 elements, is stacked to
26 slices in span-wise direction and half of chord length is used as span length. One
iteration cycle has been completed in this study and 471% increase in lift coecient
is found. The time-averaged lift coecient increased from 0.15 of NACA-0012 to
0.85 of optimal airfoil. Finally, the three dimensionality of the ow for the optimal
geometry has been discussed.
6.1. Future directions 55
6.1 Future directions
A few directions for future research motivated by the present work are listed here.
These 3D optimizations can be carried out at higher Reynolds number to check
the eect of Laminar Separation Bubble on the optimal shapes.
Convergence is not assured in the optimal shape obtained. So, further opti-
mization cycles can be carried out to achieve convergence.
Optimization can be started with other shapes as initial guesses to ensure that
global optima are obtained.
Geometries can be optimized by implementing dierent objective functions,
such as maintaining an attached ow for entire surface of geometry.
The formulation can be modied to integrate turbulence modeling. Addition-
ally, compressible ows can be taken into account.
Bibliography
[1] D.N. Srinath and S. Mittal. A stabilized nite element method for shape op-
timization in low reynolds number ows. International Journal for Numerical
Methods in Fluids, 54:14511471, 2007.
[2] D.N. Srinath and S. Mittal. Optimal airfoil shapes for low reynolds number
ows. International Journal for Numerical Methods in Fluids, 61:355381, 2008.
[3] T.J. Mueller, J.C. Kellogg, P.G. Ifju, and S. Shkarayev. Introduction to the
design of xedwing micro air vehicles. AIAA Education Series, 16:11901208,
2006.
[4] T.J. Mueller and J.D. DeLaurier. Aerodynamics of small vehicles. Annu. Rev.
Fluid Mech., 35:89111, 2003.
[5] P.B.S. Lissaman. Low-reynolds number airfoils. Annu. Rev. Fluid Mech.,
15:223239, 1983.
[6] G.R. Spedding and P.B.S. Lissaman. Technical aspects of microscale ight
systems. J. Avian Biol., 29:458468, 1998.
[7] M.J. Lighthill. A new method of two-dimensional aerodynamic design. Aero-
nautical Research Councils Reports and Memoranda, Number 2112, 1945.
56
BIBLIOGRAPHY 57
[8] G.B.McFadden. An articial viscosity method for design of supercriticial wings.
PhD dissertation, NewYork University, 1979.
[9] R.M. Hicks and P.A. Henne. Wing design by numerical optimization. Journal
of Aircraft, 15:407412, 1978.
[10] N.F. Foster and G.S. Dulikravich. Three-dimensional aerodynamic shape opti-
mization using genetic and gradient search algorithms. Journal of Spacecrafts
and Rockets, 34, 1997.
[11] D. Landman and C.P. Britcher. Experimental geometry optimization tech-
niques for multi-element airfoils. Journal of Aircraft, 37:707713, 2000.
[12] X. Wang, M. Damodaran, and S. L. Lee. Inverse transonic airfoil design using
parallel simulated annealing and computational uid dynamics. AIAA Journal,
40:791794, 2002.
[13] S. Obyashi. Aerodynamic inverse optimization with genetic algorithms. Journal
of Engineering and Applied Science, pages 421425, 1996.
[14] S.E. Gano, J.E. Renaud, S.M. Batill, and A. Tovar. Shape optimization for
conforming airfoils. AIAA, 2003.
[15] B. Mialon, T. Fol, and C. Bonnaud. Aerodynamic optimization of subsonic
ying wing congurations. AIAA, 2002.
[16] P.M. Congedo, C. Corre, and P. Cinnella. Airfoil shape optimization for tran-
sonic ows. AIAA, 2001.
[17] L. Huyse and R.M. Lewis. Aerodynamic shape optimization of two-dimensional
airfoils under uncertain conditions. NASA/CR-2001-210648, 2001.
BIBLIOGRAPHY 58
[18] J.L. Lions. Optimal control of systems governed by partial dierential equations.
Springer-Verlag, Berlin, 1971.
[19] A. Jameson. Aerodynamic design via control theory. Journal of Scientic
Computing, 59:117128, 1988.
[20] A. Jameson. Computational aerodynamics for aircraft design. Science, 245:361
371, 1989.
[21] A. Jameson. Automatic design of transonic airfoils to reduce the shock induced
pressure drag. In Proceedings of the 31st Israel Annual Conference on Aviation
and Aeronautics, Tel Aviv, Israel, 1990.
[22] A. Jameson. Computational methods for aerodynamic design. In 14th Interna-
tional Conference on Numerical Methods in Fluid Dynamics, Bangalore, India,
1994.
[23] A. Jameson. Optimum aerodynamic design using cfd and control theory. AIAA
Paper 95-1729, 1995.
[24] A. Jameson Juan J. Alonso James J. Reuther L. Martinelli and J.C. Vassberg.
Aerodynamic shape optimization techniques based on control theory. AIAA
Paper 98-2538, 1998.
[25] H. Okumura and M. Kawahara. Shape optimization of a body located in in-
compressible Navier-Stokes ow based on optimal control theory. Computer
Modelling in Engineering and Sciences, 1:7177, 2000.
[26] J. Reuthers, A. Jameson, J. Farmer, L.Martinelli, and D.Saunders. Aerody-
namic shape optimization of complex aircraft congurations via adjoint formu-
lations. AIAA Paper 96-0094, 1996.
BIBLIOGRAPHY 59
[27] W.K. Anderson and V. Venkarakrishnan. Aerodynamic design optimization
on unstructured grids with a continuous adjoint formulations. AIAA Paper
97-0643, 1997.
[28] D.N. Srinath, S. Mittal, and V. Manek. Multi-point shape optimization of
airfoils at low reynolds numbers. Computer Modelling in Engineering and Sci-
ences, 51:169190, 2009.
[29] D.N. Srinath and S. Mittal. An adjoint method for shape optimization in un-
steady viscous ows. Journal of Computational Physics, 229:19942008, 2010.
[30] D.N. Srinath and S. Mittal. Optimal aerodynamic design of airfoils in un-
steady viscous ows. Computer Methods in Applied Mechanics and Engineering,
199:19761991, 2010.
[31] T.E. Tezduyar, S. Mittal, S.E. Ray, and R. Shih. Incompressible ow com-
putations with stabilized bilinear and linear equal-order-interpolation velocity-
pressure elements. Computer methods in applied mechanics and engineering,
95:221242, 1992.
[32] Jorge Nocedal and Stephen J. Wright. Numerical Optimization. Springer Series
in Operations Research, 1999.
[33] R.H. Byrd, P. Lu, J. Nocedal, and C. Zhu. A limited memory algorithm
for bound constrained optimization. SIAM Journal of Scientic Computing,
16:11901208, 1995.
[34] B.H. Carmichael. Low reynolds number airfoil survey, volume-1. NACA-CR-
165803-VOL, 1981.
BIBLIOGRAPHY 60
[35] H. Borouchaki and P.L. George. Aspects of 2d delaunay mesh generation.
International Journal for Numerical Methods in Engineering, 40:19571975,
1997.
[36] Patrice Castonguay and Shiv K. Nadarajah. Eect of shape parameterization
on aerodynamic shape optimisation. In 45th AIAA Aerospace Sciences Meeting
and Exhibit, Reno, Nevada, 2007.
[37] Les Piegl and Wayne Tiller. The NURBS Book. Springer, 1997.
[38] S. Sunada, A. Sakaguchi, and K. Kawachi. Airfoil section characteristics at low
reynolds number. Journal of Fluid Engineering, 119:129135, 1997.
[39] T.E. Tezduyar, M. Behr, S. Mittal, and A.A. Johnson. Computation of un-
steady incompressible ows with the nite element methods Space-Time
Formulations, iterative strategies and massively parallel implementations. In
P. Smolinski, W.K. Liu, G. Hulbert, and K. Tamma, editors, New Methods in
Transient Analysis, AMD-Vol.143, pages 724, New York, 1992. ASME.
[40] O. Soto and R. Lohner. Cfd shape optimization using an incomplete-gradient
adjoint formulation. International Journal for Numerical Methods in Engineer-
ing, 51:735753, 2001.

You might also like