Full Text 01
Full Text 01
Full Text 01
MARTIN GRANLÖF
Abstract
Due to legislative reasons and environmental concerns the automotive and
transport sector are shifting their focus from traditional internal combustion
engine (ICE) vehicles to development of battery electric vehicles (BEVs). This
brings new challenges to design of cooling systems where axial fans are one of
the key components. Axial fans are usually designed with regards to a certain
operating condition and outside this region the efficiency of the fan drops
drastically. Due to difficulty in specifying the exact operational parameters
when placed in a car, post-design optimization may be necessary to ensure
maximized performance.
This thesis focuses on fan blade shape optimization through mesh morphing
using the surrogate based optimization algorithm called Efficient Global
Optimization (EGO). The target fan was a 9 bladed prototype fan by Johnson
Electric with uneven blade spacing. The optimization uses steady state
Reynolds-averaged Navier-Stokes (RANS) simulations to evaluate the fan
designs and a Bezier curve parametrization in order to change the fan blade
shape together with mesh morphing. The simulation setup was evaluated
before preceding with the optimization, and showed good agreement close to
intended operational conditions. Differences in turbulence modeling treatments
were also evaluated in order to have a satisfactory agreement with measurement
data.
The EGO algorithm manages to provide fan designs with higher total-to-
static efficiency at several different operational conditions. Evaluation of
the optimized fan designs was limited to comparison with the provided
measurement data and corresponding simulations.
Acoustic evaluation of selected fan designs is also attempted, but further work
is required in order for the study to result in a quantitative comparison.
Sammanfattning
På grund av lagstifting och miljöpåverkan har bil- och transportidustrin
börjat skifta fokus från traditionella förbränningsfordon till utveckling av
batteridrivna elbilar. Med detta medfäljer nya utmaningar kring kylsystemsdesign
där axiella fläktar är en av huvudkomponenterna hos systemet. Axiella fläktar
är vanligtvis designade kring ett specifikt driftstillstånd och utanför detta har
fläkten avsevärt lägre verkningsgrad. På grund av svårigheter att specificera
detta driftstillstånd med hög precision, speciellt när fläkten monteras i en bil,
kan efterdesigns-optimering vara nödvändigt för att uppnå maximal prestanda.
Acknowledgments
I would like to thank my supervisors Chenyang Weng and Asuka Gabriele
Pietroniro at VCG. Chenyang Weng for his tenacity through the ups and downs
of this project, for sharing his knowledge and for always being positive and
encouraging. Asuka Gabriele Pietroniro for sharing his knowledge regarding
turbo machinery, unsteady RANS simulations and all things CFD.
I would like to thank Magnus Knutsson at VCG for proposing the thesis
subject, allowing me to work on it and for being a great enabler throughout
the project.
I would like to thank my supervisors at KTH Mihai Mihaescu and Mats Åbom
for their constructive criticism and guidance through out the project. I am
also very thankful for being included in the research meetings at KTH by
Mihai. Being able to partake and listen to the discussions was inspiring and
interesting.
I would like to thank Roberto Adorno and Paolo Cavallo from Johnson Electric
for providing the flow measurement data for the fan.
I would like to thank Emil Ljungskog and Martin Lindstroem from Simcenter
STAR-CCM+ for their help solving issues and answering questions throughout
the thesis work.
I would like to thank Brian Fechner from the thermal management team, Tore
Bark, Jacob Vikström, Karthik Narendra Babu, Dalibor Cuturic, Umut Cirik
from the thermal efficiency team, and all the team members from the GasEx,
Cool & Climate team at Volvo Cars for the discussion regarding the simulation.
Contents
1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Original problem and definition . . . . . . . . . . . . 4
1.2.2 Available methods for problem solving . . . . . . . . 4
1.3 Purpose and objectives . . . . . . . . . . . . . . . . . . . . . 7
1.4 Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Structure of the report . . . . . . . . . . . . . . . . . . . . . . 7
2 Theory 9
2.1 Axial flow fans . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.1 Initial design process . . . . . . . . . . . . . . . . . . 11
2.1.2 Blade construction . . . . . . . . . . . . . . . . . . . 14
2.1.3 Fan performance metrics . . . . . . . . . . . . . . . . 15
2.2 Fan noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3 Efficient global optimization . . . . . . . . . . . . . . . . . . 19
3 Methodology 23
3.1 Simulation method . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Mesh morphing . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3 Optimization application . . . . . . . . . . . . . . . . . . . . 25
References 65
List of Figures
4.8 Cross section of the fan region of "Baseline" mesh (a) showing
the blade surface mesh and (b) the prism layer mesh on the
blades. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.9 Cross section of the fan region of Low y + 2 mesh (a) showing
the blade surface mesh and (b) the prism layer mesh on the
blades. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.10 Relative difference in pressure rise in percent for the Baseline
mesh compared with measurement data. . . . . . . . . . . . . 38
4.11 Relative difference in pressure rise in percent for the “Low y +
2” mesh compared with measurement data. . . . . . . . . . . 38
4.12 Pressure rise for the “Low y + 2” mesh compared with measurement
data. Data on the vertical axis are hidden in the public version
of this report. . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.13 Relative difference in pressure rise in percent for the Baseline,
“Low y + ” and “Low y + 2” mesh compared with measurement
data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.14 Total to static efficiency for the Baseline, “Low y + ” and “Low
y + 2” mesh compared with measurement data. Simulation
data is calculated using equation (4.2) and measurement data
using equation (4.3). Data on the vertical axis are hidden in
the public version of this report. . . . . . . . . . . . . . . . . 40
4.15 Control surface and fan blade mesh with (a) being the un-
perturbed and (b) the perturbed control surface and mesh. . . . 42
4.16 Probe placement around the fan. . . . . . . . . . . . . . . . . 44
List of Tables
GA Genetic Algorithm
HM Hybridised method
KRG Kriging
OA Orthogonal Arrays
SA Simulated Annealing
Chapter 1
Introduction
1.1 Background
The start of a fan design process is usually motivated by a need to fulfill some
system requirement. This is usually a requirement of a provided volumetric
flow at a given system resistance. If the fan is placed in a system with complex
flow phenomena, for example on top of a radiator as shown in figure 1.1, the
specification of these two system parameters can be difficult. Still the design
2 | Introduction
1.4
-0.2
-0.4
-0.6
0 0.2 0.4 0.6 0.8 1 1.2
Normalized volume flow
Figure 1.2 – Performance chart for fan at different percentages of the maximum
fan rotational speed (solid lines), and operational curves for different scenarios
(dotted lines). The efficiency (dashed line) is normalized with the maximum
efficiency in the range. Data reproduced from [3].
Apart from meeting criteria with regards to cooling and efficiency there is
also an ever-growing demand for less noisy fans, driven both by legislation
and customer satisfaction. This means that fans also need to be designed
and optimized with regards to acoustic performance, i.e. reduction of noise
at operation.
1.2 Problem
The problem with fan design relates to the aforementioned operational conditions
and requirements, but also the tools available. Firstly there is some difficulty in
accurately estimating the resistance of the system due to the complex geometry
of the parts involved and their interaction with one another. Secondly there are
many vastly different fan designs that can satisfy the same requirements, but
the best design can be hard to quantify. Depending on dimensions, shape of
the blades, and many other parameters, different design methods are also used,
which makes finding the best fan design even harder.
4 | Introduction
The optimization methods usually all start from a similar problem formulation,
which is described in Skinner et al. [4]. That is minimizing (or maximizing)
an objective function with regards to the design variables, which are provided
partly by the geometry parameterization. This function is then minimized
through an iterative process. Common method categories are:
Kim et al. [9] used a B-spline and Bezier curve combination for geometry
representation and manipulation in order to optimize aerodynamic and aeroacoustic
performance of an axial flow fan. They used a weighted average of a
response surface approximation (RSA) model, a Kriging (KRG) model and a
radial basis neural network (RBNN) model for the aerodynamic optimization,
and a combination of RSA, multi-objective evolutionary algorithm (MOEA)
and non-dominated sorting genetic algorithm (NSGA) was used for the
aeroacoustic optimization.
Öksüz et al. [10] used a multi-fidelity multi-objective genetic algorithm
(MOGA) to optimize blade shape of a low speed turbine with regards to torque
and total adiabatic efficiency. They used several Bezier curves to represent
and alter the blade geometry. With this numerical setup they outperformed
a standard MOGA both with regards to overall performance of the optimized
blade shape and computational cost.
Bamberger et al. [11][12] used multiple GAs to maximize the total to static
efficiency of low pressure axial fans given certain requirements for pressure
rise and volumetric flow. They use a multi-layer perception (MLP) artificial
neural network (ANN) that had been trained using data from RANS CFD
simulations to evaluate the performance of the fan designs. A total of 26
geometric parameters were included, varying from blade section parameters
such as chord length, thickness and position of maximum camber, to overall
geometric parameters such number of fan blades and hub-to-tip ratio. Further
CFD analysis and experimental evaluation of the optimized designs was also
carried out and showed good agreement.
Bamberger and Carolus [13] used and modifies an adjoint optimizer
algorithm available in the OpenSource CFD software OpenFoam to maximize
the total-to-static efficiency and pressure rise of a low pressure axial fan,
separately. Using the surface sensitivity of the fan blades to the objective
functions calculated through the adjoint solver, they change the blade CAD
using spline interpolation.
Cho et al. [14] used a GBM to optimize the efficiency of a low pressure
axial fan. They considered 12 design parameters for the fan rotor, such as
stagger, blade angle and sweep, and one for the stator that is part of the fan
shroud. The fan design was altered through the CAD and then re-meshed. The
results were also compared with experiments and showed good agreement at
design operating condition.
Apart from geometry parametrization and optimization scheme/method
the choice of simulation models within CFD softwares is also important for
the outcome of the optimisation.
Introduction | 7
1.4 Scope
The work is limited to optimising a 9 bladed prototype fan made by Johnson
Electric using the Efficient Global Optimisation (EGO) surrogate-model-
based algorithm, outlined by Jones et al. [8]. A Python implementation of the
EGO algorithm by Bouhlel et al. [15] has been adopted in the optimization
framework. Bezier and sinusoidal curves/surfaces are used for geometry
parametrisation and mesh morphing is used to change the geometry. CFD
simulations are performed in Simcenter STAR-CCM+ [16].
Chapter 2
Theory
• Prerotator generate swirl, i.e. azimutal velocity about the fan axis, in
order to control the angle of incidence to the different sections of the fan
rotor.
• Nose and tail fairings aim to reduce the drag created by the hub of the
fan and also the losses due to separation behind the fan unit.
• Rotor is the part that generates the pressure difference that induces
the flow. The rotor can be made out of cambered blades with uniform
thickness or from blades with airfoil shapes.
• Straightener aim to remove the swirl from the flow behind the fan,
making it purely axial again.
The fan setup for simulations and experiments in this thesis are compliant
with Category A.
Theory | 11
A: Free inlet, free outlet B: Free inlet, ducted outlet C: Ducted inlet, free outlet
with partition
Fan Fan
D: Ducted inlet, ducted outlet E: Free inlet and free outlet without partition
equation (2.1).
qv
us = up = (2.1)
2πrdr
where us and up are respectively the axial velocities at the suction and pressure
side of the rotor annulus at position r, qv is the flow rate at the annulus, and
dr is the thickness of the annulus at r.
Secondly, the theoretical total pressure rise of the fan at an annulus of
position r is given by Euler’s equation of turbomachinery
where Ω is the rotor rotational speed in radians per second, and uθ,s and uθ,p
are respectively the circumferential velocities at the suction and pressure side
of the rotor annulus at position r.
The non-dimensional form of this equation reads
2
hth = (s − p ) (2.3)
λ
with h = ∆p/ 12 ρU 2 being the non-dimensional pressure rise, where U is the
mean axial velocity through the fan, λ = U/Ωr the speed ratio, and = uθ /U
the swirl coefficient.
Step 3: Choosing load distribution and calculating velocity triangles in
order to obtain the total pressure rise at the given flow rate
Moving on to the third point, a load distribution along the span of the blades of
the fan rotor needs to be specified. There are several different design methods
for axial fans but two common design approaches are free vortex flow and
arbitrary, also called non-free, vortex flow. Free vortex flow assumes that the
swirl velocity through the fan is inversely proportional to the radial position,
this also has the benefit that the total pressure distribution and axial velocity
is constant with radial position. For arbitrary vortex flow the swirl velocity is
independent of the radial position.
A general expression for the spanwise distribution of the swirl velocity can
be written as
B
(2.4)
X
ruθ,p = ab r b
b=0
In the above equations Qv is the total volumetric flow rate, ∆pth is the total
theoretical pressure rise across the rotor, ∆p(r) is local total pressure rise, and
rh and rt are the hub radius and tip radius respectively.
Given the knowledge of the variables mentioned above the velocity triangles
can be calculated. The angle ϕr of the resulting velocity wr at a radial position
of a fan blade can be calculated as
λ
tan ϕr = 1 (2.8)
1− (
2 s
− p )λ
where the second term in the denominator is the average swirl coefficient
across the rotor.
Step 4: Choosing blade sections (airfoils) and calculating solidity
Given the inflow angle, ϕr , and the axial and swirl velocities together with the
rotational speed of the rotor the blade forces can be calculated in the fourth
step outlined above. The rotor normal and parallel forces for an annulus of
thickness dr, denoted as Y and X in figure 2.3, can be calculated as
Y = s∆pdr (2.9)
2
CL = (s − p ) sin ϕr − CD cot ϕr (2.12)
σ
where σ = c/s is the solidity, and c is the camber length.
Through equations (2.9) to (2.12) we have a connection among the global
quantities, such as the pressure rise and volumetric flow, and the local blade
quantities such as lift, drag and solidity. With these equations one can design
the rotor blades using blade element theory outlined by, for example Wallis,
using the isolated airfoil method for solidities below 0.7 and the cascade
method above one [18]. The free choice of the designer is to choose the
optimum angle of attack, α of the blade section, whose relation to ϕr can
be seen in figure 2.3 and equation (2.13). This will provide the lift and drag
coefficients enabling the calculation for the solidity σ. The angle β in equation
(2.13) is the angle between the rotor plane and the chord line of the blade
section which can be seen in figure 2.3.
L
Ch Y ϕr
ord
lin
α e
β wr X
ϕr
D
β = ϕr + α (2.13)
No sweep No dihedral
Backward Forward
sweep sweep - dihedral + dihedral
Inflow
Hub Hub
A detailed review of the effects of sweep and dihedral on axial fans and
compressors can be found in the article by Vad [21]. Although the effects
of these parameters are complex and depend on the operational conditions of
the fan they can be generalized a bit. Well designed forward sweep increases
efficiency at both intended loading and at partial load. This seems to be mainly
due to the decrease in low energy flow accumulation near the blade tip that is
a result of radial flow along the fan blades. Lower discharge loss and noise
generation have also been observed [22]. Note that forward sweep unloads
the tip region and moves the load towards inboard blade sections. Depending
on initial loading, this region may experience higher losses due to this load
redistribution. Backward sweep seems to decrease performance in almost all
operating conditions and is therefore undesirable [21].
Positive dihedral at the hub and tip seems to be beneficial to unload these
regions where low energy fluid can accumulate. Similar to forward sweep, this
moves the load to the mid-span part of the fan blades and therefore additional
evaluation might be needed when this is used [23].
could be useful. Since the fan in this case is not in a loop the total-to static
efficiency will be used. Apart from these two distinctions of efficiency, the
total input power used in the efficiency formulation can vary. Common choices
for the input power are shaft power at the fan, electrical input power to the fan
motor and total mechanical power. In this report a similar expression to that
used by Kim et. al [9] will be used, which utilizes the mechanical power at the
fan blades as the total input power. This is written as
∆pts Qv
ηts = , (2.14)
Ωτ
where ∆pts = pp − (ps + us ) is the total-to-static pressure difference between
the suction side and the pressure side of the fan, Qv is the volumetric flow
through the fan, Ω is the angular speed of the fan and τ is the torque on the fan
blades in the axial direction.
In the book by William K. Blake [24] the possibility of a more efficient fan
also having better acoustic performance is discussed. If this is the case then
optimizing towards efficiency could also optimize acoustics.
From this one can insert the quantities of equation (2.15) into the mass and
momentum equations and linearize. Assuming that the acoustic propagation
is isentropic one now arrives at the inhomogeneous wave equation, given in
equation (2.16). The first term on the right hand side is the contribution to
the acoustic field due to an unsteady external force, the second term is due to
Theory | 17
1 ∂ 2 p 0 ∂ 2 p0 ∂fi ∂Qm
2 2
− 2
= + (2.16)
c0 ∂t ∂xi ∂xi ∂t
By splitting space into two parts, the source part and the sound part, where the
source field satisfies the inhomogeneous wave equation (2.16) and the sound
field the homogeneous wave equation we arrive at Lighthill’s aeroacoustic
analogy. Making the same assumptions [25] as for the inhomogeneous wave
equation and using the mass and momentum equations without linearization,
one gets the following for the source field
1 ∂ 2 p0 ∂ 2 p0 ∂ 1 ∂ 0 2 0
∂f
i ∂2
− = Qm + (p − c 0 ρ ) − + (ρui uj − τij ).
c20 ∂t2 ∂x2i ∂t c20 ∂t ∂xi ∂xi ∂xj
| {z }| {z } | {z }
s1 s2 s2
(2.17)
In equation (2.17), which is the famous Lighthill’s equation, the three terms on
the right hand side are usually referred to as, from left to right, the monopole,
the dipole and the quadrupole source term, labeled as s1 , s2 and s3 . The
monopole source s1 field is due to unsteady mass injection and deviations
form adiabatic change of state, the dipole s2 source field is due to an unsteady
external force, and the quadrupole source field s3 is due to the viscous stresses
τij and the Reynolds stresses ρui uj . The quadrupole source field is usually
summarized to be the noise generated by turbulence since the viscous stresses
mostly have a dampening effect on the noise production. The dipole source
term is the one that is responsible for the tonal noise related to moving
fan blades and the quadrupole is responsible for the more broadband noise
generated by the fan.
The two equations above are valid for situations with zero mean flow outside
the source region, i.e. u0,i = 0. For most applications involving fans the mean
flow Mach number is small, around 0.1, and the effects of the flow on the
sound propagation can be neglected. A common method used for acoustic
calculations for propellers and fans is the Ffowcs Williams Hawkings FW-H
equation [26]. This utilizes the Heaviside function in order to write the mass
and momentum equations in a way that is valid for the whole fluid domain
18 | Theory
1 ∂ 2 p0 ∂ 2 p0 ∂2
− = (Tij H)
c20 ∂t2 ∂x2i ∂xi ∂xj
∂ ∂H
+ (ρ0 vn + ρ(un − vn )) (2.18)
∂t ∂xn
∂
0 ∂H
− (ρui (un − vn ) + p δij nˆj ) ,
∂xi ∂xn
with H being the Heaviside function, being zero inside the rotating body and
one in the fluid domain, vn being the normal surface velocity of the body,
un the normal velocity of the fluid and nˆj the outwards facing normal of the
surface. The variable inside the first term parenthesis is the Lighthill stress
tensor Tij = p0 δij + ρui uj − c0 ρ0 δij . Often the assumption of an impermeable
surface is used, i.e. un = vn , further neglecting the first term on the right hand
side of equation (2.18), which is a quadrupole term, reduces the equation to
the following form
1 ∂ 2 p0 ∂ 2 p0 ∂ ∂H ∂ 0 ∂H
− = ρ0 vn − p δij nˆj , (2.19)
c20 ∂t2 ∂x2i ∂t ∂xn ∂xi ∂xn
where the right hand side is only non-zero on the surface of the moving body
due to the gradient of the Heaviside function ∂H/∂xn [27]. The removed
quadrupole term can also be modeled instead of directly calculated and then
added back to equation (2.19) in order to capture the contribution. Looking at
the right hand side of equation (2.19) the first term is referred to as thickness
noise and is generated by the fluid displacement of the moving body. The
second term is called loading noise and is related to the pressure distribution
on the propeller or fan blades [25]. Similarly for the second and third term of
equation (2.18), but without the assumption of an impermeable surface.
In CFD software such as STAR-CCM+ the integral form of the FW-H equation
shown in equation (2.18) is used. This is written as
where p0T and p0L are the contributions to the acoustic pressure due to thickness
and loading, as discussed earlier, and are written as
Zρ0 U̇n + Uṅ
4πp0T = dS
r (1 − Mr )2
ret
(2.21)
Z ρ0 Un rṀr + c0 (Mr − M 2 )
+ dS,
r2 (1 − Mr )3
ret
Z " #
Lr − Li Mi
Z
1 L̇r
4πp0L = dS + dS
c0 r(1 − Mr )3 r2 (1 − Mr )2 ret
ret
Z Lr rṀr + c0 (Mr − M 2 )
(2.22)
1 dS.
+
c0 r2 (1 − Mr )3
ret
These integrals are evaluated at the surface of the FW-H surface with
Ui = (1 − (ρ/ρ0 )) + (ρui /ρ0 ), Li = p0 δij n̂j + ρui (un − vn ) and Mi = vi /c0 .
Subscripts r and n indicates that the variable is taken in the radiation direction
and surface normal direction respectively, doted variables are differentiated
with regards to source time, and subscript ret means that the quantities are
evaluated at retarded or source time. The formulation in equation (2.22) is
known as Farassat 1A or just formulation 1A and is given by Farassat et al
[28]. In this the quadropole term is left out because of the low Mach number
[25], but formulations of the term can be found by, for example, Brentner [29].
The acoustic pressure fluctuations are commonly expressed using sound
pressure level SPL as shown in equation (2.23), where p0ref = 2 · 10−5 Pa.
p0 2
SP L = 10 log10 ( ) (2.23)
p0ref
This gives the acoustic pressure fluctuations in dB.
F = µ + ε(χ) (2.24)
where µ is the average of this the aforementioned stochastic process, χ herein
refers to the design parameter(s) of the optimization, and ε(χ) is the error,
which is assumed to belong to a normal distribution with mean 0 and variance
σ 2 (χ) [30]. Unlike linear regression, which assumes ε(χ) to be independent
between two points in design space, EGO assumes correlation between ε(χ)
to be proportional to the distance between points. The distance in this case
is weighted with regards to the influence a parameter has on the objective
function [8]. This correlation method is usually referred to as the DACE
method, from the name of the paper by Sacks et al. [31]. The parameters of
the DACE model are fitted with the initial data using a maximum likelihood
estimate.
After an initial design population, so called design of experiment (DOE), has
been selected, using for example orthogonal arrays (OA) or Latin hypercube
sampling (LHS), the exploration and exploitation of the design space given by
the parameter limits is evaluated using an expected improvement estimate [7].
The algorithm seeks to find the point that maximizes equation (2.25)
This is written as
m
(2.26)
Y
EIp (χ) = EI(χ)F (χ) = EI(χ) P (gi (χ) < 0)
i=1
where F (χ) is the feasibility function and P (gi (χ) < 0) is the probability
of gi (χ) to be feasible, quantified by the Kriging modeling of the ith constraint.
This penalized expected improvement function can be further alternated in
order to fit the problem at hand. This is done by Bagheri et al. [30] in order to
improve performance when handling more than two constraints.
22 | Theory
Methodology | 23
Chapter 3
Methodology
∂ρ ∂ρUi
+ =0 (3.2)
∂t ∂xi
Here Ui and P are the ensemble-averaged velocity and pressure, respectively.
They are represented by lower-case letters, ui and p, in the rest of the thesis.
The turbulence behavior is modeled through the eddy viscosity νt often
involving the turbulent kinetic energy k. Most commercial CFD codes include
turbulence modeling through transport equations for different turbulence
quantities. Some of the common ones are the standard k- model with
constants from the article by Launder and Spalding [33], the realizable k-
model by Shih et al. [34] and the low-Reynolds number version by Launder
and Sharma [35].
Other common turbulence models are the k-ω model, using the specific
turbulence dissipation ω instead of . Common versions of the k-ω model are
the standard k-ω model by Wilcox [36] and the SST k-ω model by Menter [37].
Both of these are two equation models, i.e. they use two equations in order to
24 | Methodology
User inputs
• Upper and lower bounds of design parameters
• Definition of objective function
Job #1
Design of experiment (DOE)
Train
Job #2 EGO
. Surrogate model
Simulation job
Update
.
manager Expected improvement
.
Job #N
Optimum design
Chapter 4
4.1 2D testing
In order to get a better understanding of how the optimization algorithm
behaves, several 2D simulations were performed on airfoil sections. Since
the blades of a fan are usually made from airfoil shapes, the simulation results
might also be extrapolated to the fan simulations.
The 2D simulations presented below are based on flow parameters and airfoil
shapes investigated in the tutorial on DaFoam’s web page [41].
4.1.1 NACA0012
This simulation is based on the NACA0012 airfoil and the parameters presented
for the incompressible NACA0012 airfoil tutorial on the web page of DAFoam
[41]. Since the original simulation is performed in the CFD software OpenFoam,
the default fluid parameters are slightly different, but the chord Reynolds
number, Rec has been matched. In the tutorial simulation the aim is to
reduce the Drag coefficient, CD while keeping the lift coefficient, CL fixed
at 0.375. To do this they use an adjoint optimization method with 40 free
form deformation (FFD) points, with freedom to move in one of the coordinate
axis, and the angle of attack (AoA) as input parameters for the algorithm. The
starting angle of attack is 3.579107°, and then this is free to change in order
to keep CL constant. The important quantities can be seen in table 4.1, with c
being the chord length of the airfoil.
28 | Simulation setup and testing
Table 4.1 – Important flow quantities used in the optimization performed using
OpenFoam [41] and the one using STAR-CCM+.
OpenFoam STAR-CCM+
c [m] 1 1
Uin [m/s] 35 36
ρ [kg/m3 ] 1 1.18415
Rec 2.3e6 2.3e6
(a) Unperturbed
(b) Perturbed
Figure 4.1 – NACA0012 airfoil. (a), unperturbed shape and (b), an example of
perturbed shape. The “dots” in the figure are the point sets used for the control
of mesh morphing.
Using the suggestion of 10 times the design dimensions for the initial
design amount, proposed by Jones et al. [8], the initial designs are selected
using the Latin hypercube method. In this case each control point has two
degrees of spatial freedom and the angle of attack has one degree of freedom.
Setup one therefore has 50 initial designs, setup two has 90 and setup three has
70. The number of designs investigated after the initial designs was manually
set to 30.
The CD , CL , angle of attack α and the ratio between lift and drag for
the original and the optimized airfoils can be seen below in table 4.2, the
performance of the design samples in setup 1 can be seen in figure 4.2, with
the leading edge of the airfoils corresponding to the data point. The doted
line represents the CL value of the baseline airfoil. The optimized shape of
the airfoil in setup 1 can be seen compared with the original airfoil in figure
4.3 and the local pressure coefficient Cp can be seen in figure 4.4 with points
labeled opt being the optimized airfoil shape. Similar plots shown in figure
4.2, 4.3 and 4.4 can be found for setup 2 and 3 in Appendix A.
30 | Simulation setup and testing
Table 4.2 – Results from the optimized airfoils compared with the original
together with the number of input parameters and total design evaluations.
0.7
3.50
0.6 3.25
0.5 3.00
0.4 2.75
AoA
CL
2.50
0.3
2.25
0.2
2.00
0.1
1.75
0.0 1.50
0.0095 0.0100 0.0105 0.0110 0.0115 0.0120 0.0125 0.0130 0.0135 0.0140
CD
Figure 4.2 – EGO optimization of NACA0012 airfoil for setup 1.
Simulation setup and testing | 31
Figure 4.3 – Optimized airfoil from setup 1 with pink outline being the baseline
airfoil shape.
1.5
Suction side opt
Pressure side opt
1.0 Suction side base
Pressure side base
0.5
Cp
0.0
0.5
1.0
0.0 0.2 0.4 0.6 0.8 1.0
x /c
similar in between the blade designs, the improved lift and drag coefficients
would equate to around 2% higher pressure rise at the same flow rate.
The effect of relative size difference between the lift and coefficient could
influence the optimization as discussed by Marler et al. [42] and therefore
the choice of objective function is paramount for the outcome of the process.
Therefore the penalty coefficient νp should both account for the relative
difference between the included parameters and the importance to fulfill the
lift constraint.
Pressure side
Suction side
The choice of physics and schemes for the ISO5801 fan simulations can
be found below in table 4.3. The setup is a three dimensional steady state
simulation using air at constant density and viscosity as the fluid with the
segregated flow solver. Three turbulence modeling treatments were evaluated,
all using the all y + treatment in Star-CCM+. These treatments were the
realizable k − , SST k − ω and the SST k − ω with enhanced stability control
EST.
Table 4.3 – Simulation physics and schemes for ISO5801 fan simulation.
Three meshes where evaluated using the setup described above. These
meshes only differ inside the fan region (the moving reference frame) and some
34 | Simulation setup and testing
overall mesh quantities are can be found in table 4.4. The mesh called Baseline
and the one called “Low y + ” have the same fan blade surface resolution and
the mesh called “Low y + 2” has a lower surface resolution compared with
these two.
Table 4.4 – Cell count, thickness of the first cell on the surface of the fan blades
and growth rate of the prism layer for the three tested meshes in the ISO5801
setup.
The setup were run close to intended operating conditions until convergence
to check the wall y + at the fan blades. The resulting wall y + can be seen for
the Baseline mesh and the “Low y + 2” mesh in figure 4.6 and 4.7 respectively.
The Baseline mesh has a wall y + significantly higher than one and is therefore
not compliant with the recommendations of the SST k − ω turbulence model,
while the “Low y + ” and “Low y + 2” mesh has a y + of one or lower. Figure
4.8 shows the surface mesh on the blades and prism layer mesh of the Baseline
mesh and figure 4.9 shows the same for the “Low y + 2” mesh.
Figure 4.6 – Wall y + on the fan blades for the Baseline mesh in the ISO5801
simulation setup at Qv = 1.01 m3 /s. Specifics of the mesh can be seen in
table 4.4.
Simulation setup and testing | 35
Figure 4.7 – Wall y + on the fan blades for the “Low y + 2” mesh in the ISO5801
simulation setup at Qv = 1.01 m3 /s. Specifics of the mesh can be seen in table
4.4.
36 | Simulation setup and testing
Figure 4.8 – Cross section of the fan region of "Baseline" mesh (a) showing
the blade surface mesh and (b) the prism layer mesh on the blades.
Simulation setup and testing | 37
Figure 4.9 – Cross section of the fan region of Low y + 2 mesh (a) showing the
blade surface mesh and (b) the prism layer mesh on the blades.
with the measurement data provided by the supplier, for the Baseline mesh
and for the “Low y + 2” mesh can be seen in figure 4.10 and 4.11 respectively.
Note here that the sweep is carried out going from high volume flow rate to
low. The actual pressure data from the sweep ran on the “Low y + 2” mesh can
be seen compared with the measurement data in figure 4.12
2.0
Simulation Sweep k −ω
Simulation Sweep k− ω EST
1.5
Simulation Sweep k−
Rel. diff. Pressure rise
1.0
0.5
0.0
0.5
1.0
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
Volume Flow [m3 /s ]
Figure 4.10 – Relative difference in pressure rise in percent for the Baseline
mesh compared with measurement data.
2.0
Simulation Sweep k −ω
Simulation Sweep k− ω EST
1.5
Simulation Sweep k−
Rel. diff. Pressure rise
1.0
0.5
0.0
0.5
1.0
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
Volume Flow [m3 /s ]
Figure 4.11 – Relative difference in pressure rise in percent for the “Low y +
2” mesh compared with measurement data.
Simulation setup and testing | 39
Pressure rise [Pa]
JE measurement 50 Pa
Simulation Sweep k− ω
Simulation Sweep k− ω EST
Simulation Sweep k −
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
Volume Flow [m3 /s ]
Figure 4.12 – Pressure rise for the “Low y + 2” mesh compared with
measurement data. Data on the vertical axis are hidden in the public version
of this report.
The flow rate interval that is most crucial is that close to Qv = 1m3 /s and
this area shows a discrepancy of less than 6% for the Baseline mesh and less
than 5% for the “Low y + 2” mesh, compared to the measurements.
Since simulations on the "Low y + mesh" are considerably more computationally
expensive, a sweep was only performed using the SST k − ω with EST. A
comparison between all meshes using this turbulence model can be seen in
figure 4.13. What can be noted is that the Baseline mesh under-predicts the
pressure rise, for the most part, compared to the “Low y + ” and the “Low y + 2”
mesh.
2.0
Simulation Sweep k− ω EST Baseline
Simulation Sweep k− ω EST Low y +
1.5 Simulation Sweep k− ω EST Low y+ 2
Rel. diff. Pressure rise
1.0
0.5
0.0
0.5
1.0
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
Volume Flow [m3 /s ]
Figure 4.13 – Relative difference in pressure rise in percent for the Baseline,
“Low y + ” and “Low y + 2” mesh compared with measurement data.
40 | Simulation setup and testing
The total to static efficiency from each of the evaluated points during the
sweep can be seen below in figure 4.14. The efficiency in the simulation is
calculated using equation (4.2) and the efficiency from the fan experimental
data is calculated using equation (4.3). The difference between these two
equation is that the simulation uses the fan angular speed Ω times the fan blade
torque τ as the total work per unit time done by the fan, and the experiment
uses the electrical power input for the fan motor, i.e., the voltage U times the
current I. This means that the experimental efficiency accounts for both fan
motor losses and bearing losses, which should explain the slight offset in the
peak efficiency shown in figure 4.14. Similar behavior is presented also by
Cho et al. [14].
Qv ∆pts
ηts,sim = (4.2)
Ωτ
Qv ∆pts
ηts,exp = (4.3)
UI
JE measurement
Simulation Sweep k− ω EST Baseline
Simulation Sweep k− ω EST Low y +
Simulation Sweep k− ω EST Low y + 2
ts
η
0.2
Figure 4.14 – Total to static efficiency for the Baseline, “Low y + ” and “Low
y + 2” mesh compared with measurement data. Simulation data is calculated
using equation (4.2) and measurement data using equation (4.3). Data on the
vertical axis are hidden in the public version of this report.
Three out of the 13 evaluated points from the sweep were also ran to full
convergence, with regards to pressure rise, to check the accuracy of the sweep
data. These points were all ran using the SST k − ω with EST and the results
can be seen in table 4.5. Close to intended operation, at Qv = 1.01m3 /s, all
the meshes are within 1% of the fully converged results
Simulation setup and testing | 41
Table 4.5 – Total-to-static pressure rise at three different Qv for the three
meshes from full convergence simulations and sweep.
(a) Un-perturbed
(b) Perturbed
Figure 4.15 – Control surface and fan blade mesh with (a) being the un-
perturbed and (b) the perturbed control surface and mesh.
Two objective functions will mainly be considered, one targeting the total-
to-static efficiency ηts and the other the fan RPM. These objective function are
written as
static pressure difference for the baseline fan design at the given volume flow
rate and RPM. Through the structure of the objective functions we aim to
punish designs that do not reach the baseline total-to-static pressure rise while
increasing efficiency and reducing RPM.
As seen in the 2D testing in Section 4.1 the required design evaluations
increase with the number of input parameters. Therefore a setup with one
free control point in the Bezier curve in the cord wise direction for each of the
control surfaces will be used. This, together with the RPM, constitutes 5 input
parameters in total, since the control point has two degrees of freedom. This
means that the initial DOE will be 50 designs and then 30 more designs will
be evaluated as was done for the NACA0012 optimization.
Each of the designs are run for 7000 iterations in order to ensure decent
stability and convergence. Promising designs are run until full convergence
and will presented as such.
placed on a circle at a 2.5m distance from the fan origin with 5 degrees in
between and uses the Farassat 1A integral formulation of the FW-H presented
in Section 2.2. Figure 4.16 shows the placement of the acoustic probes around
the fan.
Chapter 5
Table 5.1 – Volume flow rate, control surface points, weighting factors and
input parameter intervals for the different optimization runs.
5.1.1 RP Mopt
The results from the RP Mopt optimization can be seen below in figures 5.1
to 5.3. The best design with regards to efficiency and pressure rise was run
to full convergence and has a total-to-static pressure rise ∆pts = 257.81 Pa
and efficiency ηts = 0.53904 with the same RPM as the baseline design at the
evaluated volume flow rate. This corresponds to 0.5% lower pressure rise, but
a 3.2% higher total-to static efficiency than the baseline. The design itself is
pushed inwards close to the leading edge on the suction side and similarly on
the pressure side but at the leading edge.
Results and discussion | 47
250000 Simulation samples
EGO convergence curve
200000
150000
Fobj
100000
50000
0
0 10 20 30 40 50 60 70 80
Sample No.
(a) Full
3200
3100
3000
Fobj
2900
2800
Simulation samples
EGO convergence curve
2700
0 10 20 30 40 50 60 70 80
Sample No.
(b) Close-up
5.1.2 ET Aopt1.1
The optimization data for this run was sadly lost in a backup error, but luckily
the best performing design was saved and therefore will be presented. This
design has a pressure rise ∆pts = 255.124 Pa and an efficiency ηts = 0.5337
with the same 2914 RPM as the baseline design at the evaluated volume
flow rate. This corresponds to 1.54% lower pressure rise and 2.16% higher
efficiency than the baseline fan design. The optimized design can be seen
below in figure 5.4, where one can note that the fan blade is pushed down close
to the trailing edge on the suction side and pushed up close to the leading edge
on the pressure side.
Results and discussion | 49
(a) Baseline
(b) Optimized
Figure 5.4 – Control surface and fan blade mesh with (a) being the baseline
design and (b) the optimized from the ET Aopt1.1 run.
5.1.3 ET Aopt1.2
The results from the ET Aopt1.2 optimization can be seen below in figure 5.5
to 5.7. The best design in this case had a pressure rise ∆pts = 256.96 Pa
and an efficiency ηts = 0.53411 with the same RPM as the baseline design
at the evaluated volume flow rate. This corresponds to 0.83% lower pressure
rise and 2.24% higher efficiency than the baseline fan design. This design is
almost unchanged on the suction side and on the pressure side the mid chord
section is slightly pushed outwards.
50 | Results and discussion
Target pts
300 Optimization data
250
pts
200
150
Target pts
300 Optimization data
250
pts
200
150
20000 Simulation samples
EGO convergence curve
15000
Fobj
10000
5000
0
0 20 40 60 80
Sample No.
(a) Full
20
20
Fobj
40
60
Simulation samples
EGO convergence curve
80
0 20 40 60 80
Sample No.
(b) Close-up
5.1.4 ET Aopt2
The results from the ET Aopt2 can be seen below in figure 5.8 to 5.10. The best
design had a pressure rise ∆pts = 303.2 Pa and an efficiency ηts = 0.4952
with the same RPM as the baseline design at the evaluated volume flow rate.
This corresponds to 1.1% lower pressure rise and 1.4% higher efficiency than
the baseline fan. The optimized design shares some of the same features to
that of the ET Aopt1.1 fan. The suction side surface is pushed inwards close to
the trailing edge and the pressure side surface is pushed inwards close to the
leading edge.
Note here that this part of the operational map was quite hard to evaluate and
if more time was available then more iterations per design evaluation would
52 | Results and discussion
be allocated. This could also explain why the optimum design is not that much
better than the baseline.
Target pts
350 Optimization data
300
pts
250
200
150
2700 2750 2800 2850 2900 2950 3000
RPM
Figure 5.8 – Performance of the designs in the ET Aopt2 optimization. The
baseline fan at the evaluated flow rate runs at 2935 RPM.
Target pts
350 Optimization data
300
pts
250
200
150
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
ts
Figure 5.9 – Performance of the designs that reach above ∆pts,target in the
ET Aopt2 optimization. The baseline fan at the evaluated flow rate has an
efficiency ηts = 0.4882.
Results and discussion | 53
Simulation samples
25000 EGO convergence curve
20000
15000
Fobj
10000
5000
0
0 10 20 30 40 50 60 70 80
Sample No.
(a) Full
20
25
30
35
Fobj
40
45
50
55 Simulation samples
EGO convergence curve
60
0 10 20 30 40 50 60 70 80
Sample No.
(b) Close-up
5.1.5 ET Aopt3
The results from the ET Aopt3 optimization run can be seen below in figure
5.11 to 5.13. the best design had a pressure rise ∆pts = 98.14 and an efficiency
ηts = 0.2450 with RPM=2986.2 at the evaluated volume flow rate. This is
54.5% higher pressure rise and 13.4% higher efficiency than the baseline at
a lower RPM. The optimized design seems to have increased the camber of
the baseline design by pushing the suction side surface outward close to mid
chord and the pressure side surface inwards around closer to the trailing edge
of the blade. The design can be seen in figure 5.14.
54 | Results and discussion
0
25
50
75
100
2700 2750 2800 2850 2900 2950 3000
RPM
Figure 5.11 – Performance of the designs in the ET Aopt3 optimization. The
baseline fan at the evaluated flow rate runs at 2992 RPM
0
25
50
75
100
0.75 0.50 0.25 0.00 0.25 0.50 0.75
ts
25000 Simulation samples
EGO convergence curve
20000
15000
Fobj
10000
5000
0
0 10 20 30 40 50 60 70 80
Sample No.
(a) Full
20
10
0
Fobj
10
20
30 Simulation samples
EGO convergence curve
40
0 10 20 30 40 50 60 70 80
Sample No.
(b) Close-up
(a) Baseline
(b) Optimized
Figure 5.14 – Control surface and fan blade mesh with (a) being the baseline
design and (b) the optimized from the ET Aopt3 run.
5.1.6 Summary
In table 5.2 the best performing designs from each of the optimization runs
are shown. Total-to-static pressure rise and efficiency together with the input
for the Bezier curve which modifies the blade shape are shown. The negative
value of npert means that the control surface and the blade surface are pushed
inwards. The number of the optimized design is also shown, which indicates
where in the process it was found. The abbreviations "SS" and "PS" stand for
suction side and pressure side respectively.
Results and discussion | 57
Table 5.2 – Summary of the performance of the best designs from the different
optimization runs, the Bezier curve inputs for these designs and the number of
the optimized design.
As can be noted in table 5.2, the design from the ET Aopt1.1 has a number
higher than the 80 design evaluations mentioned in Section 4.3. The maximum
allowed design evaluations was increased to 90 for ET Aopt1.1 and ET Aopt1.2
since it proved harder to find a good design, using this setup, than previously
thought. The ET Aopt3 run ended up at 82 evaluated designs in total due to the
way it was parallelized. The exact data for the ET Aopt1.1 best design is left
out since the data from the optimization was lost, as previously mentioned.
Note also that the best designs are not usually the designs shown in the EGO
convergence plots for the optimization runs shown above. This could also
indicate that the objective functions are not the optimized for the goal, that is
increasing the efficiency.
180
Base
150 210
Opt1.1
Opt1.2
120 240
90 40 20 270
80 60
60 300
30 330
0
Figure 5.15 – Directivity of the OASPL for the baseline, ET Aopt1.1 and
ET Aopt1.2 fan designs.
Baseline
Opt1.1
Single-sided FFT of pressure (dB)
Opt1.2
10 dB
2 3
10 10
Frequency (Hz)
Figure 5.16 – Single sided FFT of the acoustic pressure for the baseline,
ET Aopt1.1 and ET Aopt1.2 fan designs. Data on the vertical axis are hidden
in the public version of this report.
Results and discussion | 59
Baseline
Opt1.1
2 dB
435 435.5 436 436.5 437 437.5 438 438.5 439 439.5
Frequency (Hz)
Figure 5.17 – Single sided FFT of the acoustic pressure at the first blade
passing frequency for the baseline, ET Aopt1.1 and ET Aopt1.2 fan designs. Data
on the vertical axis are hidden in the public version of this report.
Baseline
Opt1.1
Single-sided FFT of pressure (dB)
Opt1.2
1 dB
Figure 5.18 – Single sided FFT of the acoustic pressure at the second blade
passing frequency for the baseline, ET Aopt1.1 and ET Aopt1.2 fan designs. Data
on the vertical axis are hidden in the public version of this report.
Note here that the frequency resolution is quite coarse, around 4.86 Hz,
60 | Results and discussion
which is the first blade passing frequency multiplied by a factor of 1/90. This
is due to lack of time and also some stability issues with the setup. The flow
quantities are converging within the time step, but the residuals are slowly
diverging. Therefore a shorter sampling time was selected to have some data
that can be used for this comparison.
Using this data one can conclude that the ET Aopt1.1 design seems to be
slightly better than the baseline and the ET Aopt1.2 design in the OASPL and
also at the first and second blade passing frequencies. Note that unevenly
spaced fans shift some of the tonal noise to fractions of the blade passing
frequencies, as discussed by Peng et al.[43]. This can also be seen in figure
5.16 where there is a substantial peak at 2/9, around 97 Hz, of the blade passing
frequency.
The lack of time and stability issues is also the reason why the promising
design from ET Aopt3 was not simulated. Looking at scaling laws for the
pressure, RPM and sound power level from Harris [44] there could be a
possible reduction in RPM in-order to bring the pressure rise down to the same
level as the baseline fan design. This would then equate to a reduction in the
sound power level equal to roughly 10 · log10 (1.55/2 ) ≈ 4.4 dB.
Conclusions and Future work | 61
Chapter 6
6.1 Conclusions
The EGO can provide an aerodynamically improved fan blade design at the
intended volume flow rate, i.e. a design with higher total-to-static efficiency
ηts . It can also be concluded that the toolbox created for manipulating the
fan blade geometry is capable of changing the original design in a suitable
way for post design optimization. It is highly versatile both in adhering to the
original geometry and creating new ones with correct user input. It is also
rather straightforward to add functionality to the toolbox if one would like to.
Considering the results from the different objective function use, for the case
with Qv = 1.01 m3 /s one can conclude that the efficiency based objective
functions is not constructed in the best way. The RPM based objective function
does not intend to increase efficiency directly, but it still manages to provide a
better design than the efficiency-based optimization.
For the ET Aopt3 case that is run at Qv = 1.60 m3 /s the performance gain is a
lot higher than for the lower volume flow rates. This is most likely due to the
fact that the fan was designed for lower flow rates.
The two fan designs that were evaluated acoustically and compared with the
baseline seem to perform quite similarly to the baseline fan. This is most likely
due to them not being significantly improved compared to the baseline. To
conclude that improving efficiency also improves acoustic performance the
design from ET Aopt3 should be investigated and compared with the baseline.
62 | Conclusions and Future work
6.2 Limitations
The limiting factor with optimization and simulations is almost always the
time, and that was the case for this work as well. That is also the reasoning
behind the choice of using steady RANS simulations for the optimization
simulations. A large eddy simulation or a direct numerical simulation would
most likely give more accurate results, being setup correctly, the computational
cost, however, is also a lot higher. An unsteady RANS could also have
been used for the optimization, but since we were interested in averaged
quantities for efficiency and pressure rise this would also be a waste of
computational resources since this is also more computationally demanding.
The time consumption for one design evaluation for the acoustic evaluation
simulations is more than 30 times that of the steady state simulation used in
the optimization.
Another aspect that may have limited the maximum performance of the
optimized design is that the focus of the morphing was placed on the mid span
section of the fan blades. This was a decision based on the literature available
and also because it is the furthest away from any other surface, simplifying the
mesh morphing.
The mesh used for the optimization simulations could also be improved if
more time was available. It was made early in the project and with knowledge
gathered throughout the process it could be better when it comes to the prism
layers connecting to the blade surfaces.
The optimization and the acoustic evaluations were not carried out in exactly
the same simulation setups. Therefore there is some ambiguity how to
interpret the results and connections between the efficiency and the acoustic
performance. The setup itself is not completely stable and is slowly diverging.
The evaluation is carried out with a quite coarse frequency resolution.
References
[9] J.-H. Kim, B. Ovgor, K.-H. Cha, J.-H. Kim, S. Lee, and K.-Y. Kim,
“Optimization of the aerodynamic and aeroacoustic performance of an
axial-flow fan,” AIAA Journal, vol. 52, no. 9, pp. 2032–2044, 2014.
[18] A.R. Wallis, Axial flow fans: design and practise. Academic press,
1961. ISBN 1-4832-7422-5
[22] M. B. Wilkinson, “The design of an axial flow fan for air-cooled heat
exchanger applications.” Ph.D. dissertation, 2017.
[29] K. S. Brentner, “An efficient and robust method for predicting helicopter
high-speed impulsive noise,” Journal of Sound and Vibration, vol. 203,
no. 1, pp. 87–100, 1997.
[34] T.-H. Shih, W. W. Liou, A. Shabbir, Z. Yang, and J. Zhu, “A new k- eddy
viscosity model for high reynolds number turbulent flows,” Computers
& fluids, vol. 24, no. 3, pp. 227–238, 1995.
[39] H. Chen and V. Patel, “Near-wall turbulence models for complex flows
including separation,” AIAA journal, vol. 26, no. 6, pp. 641–648, 1988.
[42] R. T. Marler and J. S. Arora, “The weighted sum method for multi-
objective optimization: new insights,” Structural and multidisciplinary
optimization, vol. 41, no. 6, pp. 853–862, 2010.
[43] Z. Peng, H. Ouyang, Y. Wu, and J. Tian, “Tonal noise control of cooling
fan module by using modulation principles on both rotor and stator,” in
REFERENCES | 69
Turbo Expo: Power for Land, Sea, and Air, vol. 50985. American
Society of Mechanical Engineers, 2018, p. V001T09A007.
Appendix A
0.7
3.50
0.6 3.25
0.5 3.00
0.4 2.75
AoA
CL
2.50
0.3
2.25
0.2
2.00
0.1
1.75
0.0 1.50
0.0095 0.0100 0.0105 0.0110 0.0115 0.0120 0.0125 0.0130 0.0135 0.0140
CD
Figure A.1 – EGO optimization of NACA0012 airfoil for setup 2
72 | Appendix A: NACA0012 result plots
0.7
3.50
0.6 3.25
0.5 3.00
0.4 2.75
AoA
CL
2.50
0.3
2.25
0.2
2.00
0.1
1.75
0.0 1.50
0.0095 0.0100 0.0105 0.0110 0.0115 0.0120 0.0125 0.0130 0.0135 0.0140
CD
Figure A.2 – EGO optimization of NACA0012 airfoil for setup 3
Figure A.3 – Optimized shape of the airfoil in setup 2 with original airfoil in
pink.
Appendix A: NACA0012 result plots | 73
Figure A.4 – Optimized shape of the airfoil in setup 3 with original airfoil in
pink.
0.0
0.5
1.0
0.0 0.2 0.4 0.6 0.8 1.0
x /c
0.0
0.5
1.0
0.0 0.2 0.4 0.6 0.8 1.0
x/ c
www.kth.se