14 - Cooling Fan Optimization For Heavy Electric Vehicles

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 93

Cooling Fan Optimization for Heavy

Electrified Vehicles
A study on performance and noise

Kylfläktoptimering för Tunga Elektrifierade Fordon


En studie om prestanda och buller

Master’s thesis in Aerospace Engineering

Amir Khiabani
Daniel Acebo Alanis

School of Engineering Sciences


Department of Aeronautical and Vehicle Engineering
KTH R OYAL I NSTITUTE OF T ECHNOLOGY
Stockholm, Sweden 2020
Master’s thesis 2020

Cooling Fan Optimization for Heavy


Electrified Vehicles

A study on performance and noise

Amir Khiabani Daniel Acebo Alanis

School of Engineering Sciences


Department of Aeronautical and Vehicle Engineering
KTH Royal Institute of Technology
Stockholm, Sweden 2020
Cooling Fan Optimization for Heavy Electrified Vehicles
A study on performance and noise
Amir Khiabani
Daniel Acebo Alanis

Supervisors: Leila Shafiee and Zoltan Kardos


Examiner: Raffaello Mariani, Assistant Professor

Master’s Thesis 2020


Host Company: Scania Group
School of Engineering Sciences
Department of Aeronautical and Vehicle Engineering
KTH Royal Institute of Technology
SE-100 44 STOCKHOLM
Telephone +46 8 790 60 00

Template: David Frisk

Typeset in LATEX, template by David Frisk


Printed by KTH
Stockholm, Sweden 2020

iv
Abstract
Vehicle electrification plays a significant role in the effort to reduce the environmental
impact of the automotive industry. Scania is one of the leading manufacturers of
heavy vehicles which is currently moving towards a sustainable transport system
by manufacturing a new generation of heavy vehicles powered by batteries. One
of the major concerns with these vehicles is related to the noise generated by the
electric axial fans used in the cooling system. This project was conducted with the
purpose of investigating the factors that positively affect both noise and performance
in the electric fans. Based on two different blade design methods and several noise
control techniques, 11 fan models were developed. The fan models created with
design method 1 are equipped with cambered-plate blades, while the models made
with design method 2 consist of airfoil-shaped blades. Moreover, the performance of
these models was analyzed by using theoretical methods and Computational Fluid
Dynamics (CFD). In addition, two empirical approaches were used to estimate the
acoustic energy emitted by the fan models. Furthermore, the developed models
were compared with two commercially available fans. It was found that both design
methods provide similar performance in low pressure differences. On the other hand,
the efficiency and acoustic energy are influenced by the choice of the noise control
methods.

Sammanfattning
Fordonselektrifiering har en väsentlig roll i arbetet med att minska bilindustrins
miljöpåverkan. Scania är en av de ledande tillverkarna av tunga fordon som för när-
varande går mot ett hållbart transportsystem, genom att tillverka en ny generation
tunga fordon drivna med batterier. Ett stort bekymmer med dessa fordon är relat-
erat till det ljud som genereras av de elektriska axialfläktarna som används i kylsys-
temet. Detta projekt genomfördes i syfte till att undersöka de faktorer som positivt
påverkar både buller och prestanda hos de elektriska fläktarna. Baserat på två olika
bladdesignmetoder och flera brusstyrningstekniker, utvecklades 11 fläktmodeller.
Fläktmodellerna som är utformade med konstruktionsmetod 1 är utrustade med
krökformade plattor, medan modellerna som skapades med designmetod 2 består
av vingprofil blad. Dessutom analyserades prestandan för dessa modeller med an-
vändning av teoretiska metoder och strömningsmekaniska beräkningar. Ytterligare
två empiriska tillvägagångssätt användes för att uppskatta den akustiska energin
som släppts ut av fläktmodellerna. Utöver det jämfördes de utvecklade modellerna
med två kommersiellt tillgängliga fläktar. Detta visade att båda konstruktionsme-
toderna resulterar i liknande prestanda vid lågtrycksskillnader, däremot påverkas
verkningsgraden och den akustiska energin av valet av brusstyrningsmetoder.

Keywords: Electric axial fans, Fan design, Blade design, Battery Electric Vehicle,
Fan noise, Fan performance.

v
Acknowledgments
This study was carried out at Scania Group in Södertälje, Sweden. The work was
a collaboration between Vehicle Cooling & Aerodynamics and Electrical Actuators
Design departments.
We would like to take the opportunity to acknowledge everyone who supported
us to accomplish this project. First of all, we would like to extend our deepest
gratitude to Leila Shafiee and Zoltan Kardos who provided guidance and valuable
advice throughout the project. Furthermore, we are grateful to Mattias Chevalier for
taking time out of his schedule to perform all the CFD simulations. Special thanks
to Stina Johansson and Ola Hall for all the support and motivation. We would also
want to thank our examiner, assistant professor Raffaello Mariani, at KTH Royal
Institute of Technology for his help during the project. Last but not least, our deep
and heartfelt gratitude to our families for their relentless and extraordinary patience
and support over the years.
Daniel would also like to acknowledge the monetary support provided by Conacyt
to obtain his master’s degree.

Amir Khiabani and Daniel Acebo Alanis, Stockholm, July 2020

vii
Contents

List of Figures xi

List of Tables xiii

Nomenclature xiv

1 INTRODUCTION 1
1.1 Fan History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 THEORY 4
2.1 Fans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.1 Axial fan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.2 Centrifugal fan . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.3 Fan laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.4 Fan performance . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2.1 Sound measurement units . . . . . . . . . . . . . . . . . . . . 8
2.2.1.1 Octave bands and weighted sound pressure levels . . 9
2.2.2 Noise in axial fans . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.3 Fan sound law . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2.4 Noise control in axial fans . . . . . . . . . . . . . . . . . . . . 12
2.3 Blade Element Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3.1 Blade element forces . . . . . . . . . . . . . . . . . . . . . . . 13
2.3.2 Relative flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 METHODS 16
3.1 Design Method 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.1.1 Mean line calculation . . . . . . . . . . . . . . . . . . . . . . . 16
3.1.2 Outlet blade angle . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.3 Blade shape computation . . . . . . . . . . . . . . . . . . . . . 18
3.1.4 Losses and diffusion factor . . . . . . . . . . . . . . . . . . . . 18
3.2 Design Method 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.1 Isolated airfoil . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2.2 Cascade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

ix
Contents

3.2.3 Modified isolated airfoil . . . . . . . . . . . . . . . . . . . . . 23


3.2.4 Pressure loss components . . . . . . . . . . . . . . . . . . . . . 24
3.3 Fan Performance Estimation . . . . . . . . . . . . . . . . . . . . . . . 25
3.3.1 Theoretical . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3.2 CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.4 Noise Level Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.4.1 Approach 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.4.2 Approach 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

4 DESIGN PROCEDURE 30
4.1 Design Method 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2 Design Method 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

5 RESULTS 38
5.1 Design Method 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.1.1 Impeller dimensions . . . . . . . . . . . . . . . . . . . . . . . . 38
5.1.2 Blade characteristics . . . . . . . . . . . . . . . . . . . . . . . 39
5.1.3 Fan dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.2 Design Method 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.2.1 Impeller dimensions . . . . . . . . . . . . . . . . . . . . . . . . 41
5.2.2 Blade characteristics . . . . . . . . . . . . . . . . . . . . . . . 42
5.2.3 Fan dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3 Fan Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3.1 Theoretical . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3.2 CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.4 Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.4.1 Approach 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.4.2 Approach 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.4.3 Scania’s fans comparison . . . . . . . . . . . . . . . . . . . . . 50

6 DISCUSSION 52

7 CONCLUSION 55
7.1 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

Bibliography 56

Appendix A Design Method 2 Data I

Appendix B Flow Diagrams IV

Appendix C CAD Models VII

Appendix D CFD Results XIII

x
List of Figures

2.1 Axial fan components . . . . . . . . . . . . . . . . . . . . . . . . . . . 4


2.2 Centrifugal fan components . . . . . . . . . . . . . . . . . . . . . . . 5
2.3 Free inlet-ducted outlet installation setup . . . . . . . . . . . . . . . . 7
2.4 Fan performance curves . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.5 System and fan performance curves . . . . . . . . . . . . . . . . . . . 8
2.6 A-weighted sound pressure curve . . . . . . . . . . . . . . . . . . . . 9
2.7 Fan noise mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.8 Fan impeller diagram illustrating an annular blade element of length
dr at radius r . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.9 Forces on a blade section . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.10 Flow over a blade section . . . . . . . . . . . . . . . . . . . . . . . . . 15

3.1 Mean-line calculation of the velocity diagram with axial entry . . . . 16


3.2 View of the calculated blade shape in (x,y) coordinates . . . . . . . . 18
3.3 Wake momentum thickness versus overall diffusion factor DF for
NACA 65 and C4 airfoils at minimum loss incidence . . . . . . . . . . 19
3.4 Chord length and pitch distance . . . . . . . . . . . . . . . . . . . . . 20
3.5 Typical airfoil section nomenclature . . . . . . . . . . . . . . . . . . . 21
3.6 Isolated airfoil and velocity triangles . . . . . . . . . . . . . . . . . . 22
3.7 Cascade of rotor blades . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.8 Interference factor for airfoils with a circular-arc camber line . . . . . 24
3.9 Velocity triangle for axial fans . . . . . . . . . . . . . . . . . . . . . . 25
3.10 A cut plane view of the fan installation, from left to right: blockage,
fan, fan shroud, and flow resistance . . . . . . . . . . . . . . . . . . . 27
3.11 A center cut plane view of the computational domain, the flow direc-
tion is from right to left . . . . . . . . . . . . . . . . . . . . . . . . . 28

4.1 Cordier diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30


4.2 Range of optimum hub-to-tip ratio . . . . . . . . . . . . . . . . . . . 31
4.3 Specific speed vs. mean head and flow coefficients for axial fans . . . 32
4.4 Cordier diagram based on non-dimensional specific speed and diameter 34
4.5 Inlet and outlet velocity triangles for axial fans . . . . . . . . . . . . 35

5.1 Blade cross section shape at two radial position . . . . . . . . . . . . 39


5.2 Blade angle distribution at two radial positions . . . . . . . . . . . . 39
5.3 Diffusion Factor along the blade . . . . . . . . . . . . . . . . . . . . . 40
5.4 Aerodynamic characteristics of NACA 4412 at Re = 3.24 · 105 . . . . . 43

xi
List of Figures

5.5 Aerodynamic characteristics of NACA 6409 at Re = 3.18 · 105 . . . . . 44


5.6 Two-dimensional blade sections . . . . . . . . . . . . . . . . . . . . . 45
5.7 Theoretical performance graph. . . . . . . . . . . . . . . . . . . . . . 46
5.8 Fan performance and efficiency curves at 2000 rpm. . . . . . . . . . . 48

A.1 Recommended incidence angle i based on camber angle θ . . . . . . . I


A.2 Stagger angle γ vs. deviation angle coefficient m . . . . . . . . . . . . I
A.3 Recommended blade solidity σ based on flow coefficient φ . . . . . . . II
A.4 Lift data corresponding to cambered plate . . . . . . . . . . . . . . . II

B.1 Flow diagram for design method 1 . . . . . . . . . . . . . . . . . . . . IV


B.2 Flow diagram for design method 2 . . . . . . . . . . . . . . . . . . . . VI

C.1 Model 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . VII


C.2 Model 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . VII
C.3 Model 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . VIII
C.4 Model 4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . VIII
C.5 Model 5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . IX
C.6 Model 6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . IX
C.7 Model 7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . X
C.8 Model 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . X
C.9 Model 9. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . XI
C.10 Model 10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . XI
C.11 Model 11. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . XII

D.1 Pressure distribution on the suction side of model 1 at 2000 rpm. . . XV


D.2 Pressure distribution on the suction side of model 2 at 2000 rpm. . . XV
D.3 Pressure distribution on the suction side of model 3 at 2000 rpm. . . XVI
D.4 Pressure distribution on the suction side of model 4 at 2000 rpm. . . XVI
D.5 Pressure distribution on the suction side of model 5 at 2000 rpm. . . XVII
D.6 Pressure distribution on the suction side of model 6 at 2000 rpm. . . XVII
D.7 Pressure distribution on the suction side of model 7 at 2000 rpm. . . XVIII
D.8 Pressure distribution on the suction side of model 8 at 2000 rpm. . . XVIII
D.9 Pressure distribution on the suction side of model 9 at 3000 rpm. . . XIX
D.10 Pressure distribution on the suction side of model 10 at 3000 rpm. . . XIX
D.11 Pressure distribution on the suction side of model 11 at 2000 rpm. . . XX

xii
List of Tables

2.1 Fan laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

3.1 Specific sound power levels in eight lowest octave bands for axial fans 29

5.1 Fan performance requirements. . . . . . . . . . . . . . . . . . . . . . 38


5.2 Models’ characteristics based on method 1. . . . . . . . . . . . . . . . 40
5.3 Performance input parameters. . . . . . . . . . . . . . . . . . . . . . 41
5.4 Main design parameters. . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.5 Models’ characteristics based on method 2. . . . . . . . . . . . . . . . 46
5.6 Fan models’ theoretical performance. . . . . . . . . . . . . . . . . . . 47
5.7 CFD results at/near the peak efficiency point. . . . . . . . . . . . . . 47
5.8 Sound pressure level at 2000 rpm using approach 1. . . . . . . . . . . 49
5.9 Overall sound pressure level at 2000 rpm using approach 2. . . . . . . 50
5.10 Performance of the fans at ∆p = 183 Pa. . . . . . . . . . . . . . . . . 50
5.11 Sound pressure level at ∆p = 183 Pa using approach 1. . . . . . . . . 51

A.1 List of airfoils used in the blade design . . . . . . . . . . . . . . . . . III

D.1 CFD results at/near the peak efficiency point at different rotational
speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . XIII
D.2 CFD performance results for model 3. . . . . . . . . . . . . . . . . . . XIV
D.3 CFD performance results for model 7. . . . . . . . . . . . . . . . . . . XIV

xiii
Nomenclature

SUBSCRIPT γ Blade angle


0 Reference Stagger angle

1 Inlet ν Hub-to-tip ratio


Kinematic viscosity
2 Outlet
Ω Blade rotational speed
axial Impeller axial direction
ω Wake rotational speed
d Design Impeller rotational speed
h Hub ωs Non-dimensional specific speed
hyd Hydraulic Re Reynolds number
m Mean φ Flow coefficient
max Maximum Ψ Static pressure coefficient
s Stall ρ Fluid density
shaf t Mechanical shaft σ Blade solidity
t Total Speed number
Tip τ Torque
t−s Total-to-static θ Camber angle
t−t Total-to-total A Area
SYMBOLS a Axial induction factor
α Angle of attack b Angular induction factor
Blade span
β Relative flow angle
Blade angle c Chord length
Absolute velocity
∆p Pressure rise
Cd Total drag coefficient
δ Diameter number
CL Lift coefficient
δr Drag-to-lift ratio
cu Absolute tangential velocity
∆s Non-dimensional specific diam-
eter Cda Annulus drag coefficient

 Deflection angle Cdp Profile drag coefficient


Convergence tolerance Cds Secondary drag coefficient
η Efficiency Cdt Tip clearance drag coefficient

xiv
Nomenclature

CLi Isolated lift coefficient OASP L Overall sound pressure level


CT Tip clearance p Pressure
D Diameter Q Volume airflow rate
Drag force Qθ Directivity factor
DF Diffusion factor r Radius
f (r) Pressure distribution along the R Reaction factor
r
blade
s Pitch distance
Fa Axial force
T Thrust force
Fd Drag force
U Tangential velocity
FL Lift force
u Tangential velocity
Fu Tangential force
V Absolute velocity
g Gravitational acceleration
Va Axial velocity
H Head rise
Specific sound power level Vp Pick velocity

i Incidence angle Vu Tangential component of abso-


lute velocity
K Multiplane interference factor
V∞ Incoming flow velocity
L Lift force
W Power
Lp Sound pressure level Overal sound power
LW Sound power level Relative velocity
m Deviation angle coefficient w Relative velocity
N Rotational speed wu Relative tangential velocity
n Rotational speed Y Specific head
Ns Specific speed z Number of blades

xv
1
INTRODUCTION

Fans are considered to be one of the high-demand turbomachinery devices whose


key role is to create pressure difference in order to transfer a large volume of air
throughout a system using a rotating impeller. Fan blades are responsible for di-
recting airflow from the fan inlet to the outlet. The blades, which are available in
various shapes, sizes, and numbers, are attached to a hub connected to a rotating
shaft. In order to rotate the fan impeller, different types of power sources such as
an electric motor can be employed. One of the crucial advantages of fans compared
to other turbomachinery devices such as compressors is the capability of generating
high volume flow rates at relatively low pressure differences [1]. This is one of the
main reasons why fans are widely used in various applications, such as ventilation,
cooling, heating, material handling, and air conditioning.
Fans are available in different shapes and dimensions with various rotational speeds.
They can be categorized into four different classes based on direction of the airflow,
namely: axial-flow, centrifugal-flow, mixed-flow, and tangential-flow [2]. The aero-
dynamic performance of a fan is highly dependent on its efficiency and the amount
of air it delivers with respect to the pressure rise. Fan performance is one of the
main factors that distinguishes one fan type from another. In order to pick a suit-
able fan for a specific application, some major constraints such as space limitation,
efficiency, and noise level must be taken into account [2].

1.1 Fan History


The origin of fans can be dated back to the 18th century B.C. in ancient Egypt.
Handheld fans in various sizes were used by different nations around the world for
hundreds of years. The main purpose of using this type of fan was to create airflow
in order to cool down the body temperature [3]. One of the very first rotary fans,
known as crude fan, was used in Germany for the purpose of ventilation in metal
mines during the 16th century, according to Georgius Agricola (1494-1555) [3]. This
kind of fan was made out of wood and is considered as one of the first centrifugal
fans due to the existence of radial paddle vanes to generate airflow [3]. Watermills
and animals were usually used to rotate crude fans in those days. During the 19th
century, several attempts were made by various engineers to invent new ventilation
fans. William Brunton (1777-1851) was a Scottish engineer who designed one of
the first steam-driven centrifugal fans which enhanced the ventilation system of
Gelligaer Colliery in Wales [2].
After about a century of using noisy steam engines to drive different types of fans,

1
1. INTRODUCTION

electric motors replaced steam engines in order to enhance the quality of human life.
Upon the advent of electric motors, the use of axial-flow fans flourished. Many fan
manufacturing companies, such as Aerex Ltd, started to develop new generations
of axial fans for mine ventilation [2]. A new type of impeller equipped with airfoil-
shaped blades was employed. Furthermore, inlet and outlet guide vanes were used
to improve the efficiency of the fans. Axial-flow fans play a significant role in today’s
growing industries. Automotive industry is one the major business fields in which
this type of fan is highly demanded.

1.2 Background
At the present time, environmental sustainability issues are of huge concern to all
nations of the world. Vehicle electrification is one of the major solutions to con-
trol the environmental damage, including climate change due to greenhouse gases.
Scania, one of the leading heavy vehicle manufacturers, is currently focusing on
innovative technologies with the purpose of electrifying heavy duty vehicles, such
as trucks and buses, as well as infrastructures [4]. Scania’s Battery Electric Ve-
hicle (BEV) is not only compatible with the urban living environment due to low
noise pollution but also considered as an environmentally friendly product with zero
tailpipe emissions. Despite all the efforts made to reduce the noise emitted from the
vehicle by replacing the combustion engine with an electric powertrain, there are
still some components which can generate high-pitched noise. The electric axial fans
used in the cooling system of first-generation BEVs produce loud noise, especially
when running at high rotational speeds. Heavy electrified vehicles are equipped with
batteries which provide the required power to run the electric machine as well as
other components. In order to dissipate the heat from the radiators connected to dif-
ferent cooling circuits, puller electric fans are employed. These fans are responsible
for generating a high capacity of airflow to maintain proper operating temperatures
when the vehicle is travelling at lower speeds. Although this kind of fan works with
appropriate efficiency according to the supplier, its performance still does not fulfill
the requirements from an acoustic point of view, as it is not specially designed for
use in a BEV. Therefore, the main focus of this project is to explore the possibility
of improving the performance of electric fans along with reducing the associated
noise.

1.3 Objectives
The overall objective of the present work is to study the main factors that influence
the performance of a cooling fan with the purpose of optimizing the capacity of
the airflow as well as mitigating the noise generated by the fan. In order to meet
the required cooling capacity for BEV, several fan models are developed based on
two different blade design approaches. This strategy provides the possibility to
design different impellers equipped with cambered-plate and airfoil-sectioned blades.
Moreover, various efficient noise reduction methods are implemented for the purpose
of minimizing the noise emitted by the fan designs.

2
1. INTRODUCTION

The main part of the project, including the blade design, fan dimension, and noise
evaluation process, is conducted by applying suitable design methods in Matlab.
Furthermore, the theoretical results are employed to create fan models using Solid
Edge, a Computer-Aided Design (CAD) software. Then, the performance of the
models are investigated by conducting a Computational Fluid Dynamics (CFD)
analysis in STAR-CCM+. The outcome is compared with the performance parame-
ters of two different electric fans which are already used in Scania’s trucks and buses.
This study will then be used by engineers at Scania to discuss the possibilities to
optimize the performance of future electric axial fans for the use in BEV’s cooling
system.

1.4 Layout
In this section, the overall layout of the project is described and briefly presents the
upcoming chapters.
Chapter 2 covers the theoretical background related to the fundamental principles
of fan performance and noise. Furthermore, it provides information about useful
theories that can be applied in order to design a fan with high efficiency.
In the third chapter, two different design methods which are used to develop fans
with different blade shapes are introduced. Moreover, various noise and performance
estimation methods are presented in chapter 3 with the purpose of analyzing the
fan models.
The design procedures based on the two fan design methods are explained in detail
in chapter 4. This chapter provides an overview of the calculation steps which are
being applied to develop new fan models.
The results from the performance and noise estimations along with the CFD out-
comes are considered in chapter 5. In addition, the characteristics of two commer-
cially available fans are compared with the results related to the developed models.
Finally, chapters 6 and 7 provide a summary of the key outcomes and findings of
the study, and highlight relevant future work.

3
2
THEORY

2.1 Fans
A fan is a device that uses a rotating impeller with the purpose of producing airflow
by creating a pressure difference [5]. Fans can be placed in two main categories
based on flow direction [5], namely:
1. Axial
2. Centrifugal

2.1.1 Axial fan


Axial-flow fans are used to generate airflow parallel to the shaft axis. The main
parts of the axial fans are the impeller, casing, and motor. This type of fan is
equipped with blades which can be structured in various shapes, such as airfoils and
cambered plates. One of the prominent features of the axial fan is the capability
of generating a high volume of air at comparatively low pressure differences [6].
The dimension of this type of fan can vary based on the installation position and
operational conditions. The general components of axial fans are illustrated in
Figure 2.1.
As mentioned earlier, axial fans are in great demand in operations where high volume
and homogeneous air distribution are required. This sort of fans are extensively used
in industries such as automotive and aviation in order to provide engine cooling
and thrust generation, respectively. Furthermore, domestic extractor, ceiling, and
computer fans are additional examples of this type.

Figure 2.1: Axial fan components [5].

4
2. THEORY

In general, axial-flow fans are categorized as follows:


1. Propeller fans operate in an open air space. They are suitable for low-
pressure, high-volume airflow applications and considered as a low-efficiency
axial fan. Moreover, they commonly consist of two or more impeller blades of
same thickness. Ceiling and pedestal fans belong to this category [5].
2. Tubeaxial fans direct the air from inlet to outlet by constraining the flow in
the radial direction. This category is more efficient than propeller fans and
they are capable of creating more static pressure. Furthermore, tubeaxial fans
have 4 to 8 blades in the shape of airfoils or cambered plates [5].
3. Vaneaxial fans are a type of tubeaxial fan with a set of guide vanes positioned
either after or before the impeller. This type is the most efficient axial fan
which can provide medium-to high-pressure capacity as long as the blades are
appropriately designed [5]. A vehicle cooling fan typically falls into this class.

2.1.2 Centrifugal fan


This kind of fans are designed to boost the inlet air speed in the radial direction
using an impeller. The inlet airflow enters from the center of the fan in the direction
of the drive shaft, and its path to the outlet changes normally by 90 degrees due to
the centrifugal force created by the system [5].
The moving air inside the fan can be controlled using dampers, vanes, and steering
blades. As shown in Figure 2.2, a centrifugal fan model consists of rotating and fixed
parts, such as casing, impeller wheel, drive shaft, bearing, and coupling. This type
of fan is commonly used in cooling and heating cycles as well as air conditioners [7].

Figure 2.2: Centrifugal fan components [5].

Based on the shape of the impeller, centrifugal fans are classified as follows:
1. Forward-curved centrifugal fans are equipped with blades which are bent in
the direction of the impeller’s rotation. They provide low noise level since they
operate at low rotational speeds and are suitable for ventilation and cleaning
service applications [7].
2. Radial blade centrifugal fans are characterised by their radial blades which
expand from the hub center without curvature. They are well suited for low-
volume flow rates, high static pressures and temperatures. This type of fan is

5
2. THEORY

the most appropriate centrifugal model for dust extraction applications [7].
3. Backward-curved centrifugal fans are equipped with blades which are curved
in the opposite direction of the impeller’s rotation. This type of fan is more
efficient compared to radial and forward-curved centrifugal fan models. Fur-
thermore, they are capable of operating with a varying static pressure [7].
4. Airfoil blade centrifugal fans are the most efficient of all centrifugal mod-
els. This type of fan provides a high volume of airflow which makes it an
appropriate candidate for industrial, mining, and biochemical applications [7].

2.1.3 Fan laws


Fan laws, also known as affinity laws, are a set of equations commonly used for
expressing the relation between performance variables for dynamically similar fans
while all flow conditions are similar [5]. Table 2.1 shows the relation between vari-
ables, such as fan pressure difference ∆p, volume airflow rate Q, power W , diameter
D, rotational speed N , and fluid density ρ.

Table 2.1: Fan laws [5].

Dependent Independent
Law No.
Variables Variables
1a Q1 = Q2 × (D1 /D2 )3 (N1 /N2 )
1b ∆p1 = ∆p2 × (D1 /D2 )2 (N1 /N2 )2 (ρ1 /ρ2 )
1c W1 = W2 × (D1 /D2 )5 (N1 /N2 )3 (ρ1 /ρ2 )
2a Q1 = Q2 × (D1 /D2 )2 (∆p1 /∆p2 )1/2 (ρ2 /ρ1 )1/2
2b N1 = N2 × (D2 /D1 )(∆p1 /∆p2 )1/2 (ρ2 /ρ1 )1/2
2c W1 = W2 × (D1 /D2 )2 (∆p1 /∆p2 )3/2 (ρ2 /ρ1 )1/2
3a N1 = N2 × (D2 /D1 )3 (Q1 /Q2 )
3b ∆p1 = ∆p2 × (D2 /D1 )4 (Q1 /Q2 )2 (ρ1 /ρ2 )
3c W1 = W2 × (D2 /D1 )4 (Q1 /Q2 )3 (ρ1 /ρ2 )

In order to determine the performance of any fan using the relations listed in Table
2.1, it is required to gather all data for another fan of the same series.
The first fan law indicates how pressure difference, volume flow rate, and power are
affected due to the change in fan diameter, rotational speed, or density. The second
fan law shows the influence of modifying pressure difference, diameter, or density on
rotational speed, volume flow rate, and power. In addition, it is possible to predict
the effect of changing density, volume flow rate, or diameter on fan power, rotational
speed, and pressure difference using the third fan law illustrated in Table 2.1.

2.1.4 Fan performance


Fan performance is evaluated by conducting special tests under controlled laboratory
conditions. In order to determine the fan performance, four installation types are

6
2. THEORY

suggested by Air Movement and Control Association International, Inc. (AMCA)


[8], namely:
1. Free inlet and outlet
2. Free inlet and ducted outlet
3. Ducted inlet and free outlet
4. Ducted inlet and outlet
Generally, the second installation setup illustrated in Figure 2.3 is selected due to its
advantages in providing uniform airflow into the fan inlet and appropriate evaluation
of static pressure on the outlet [5]. The volume flow rate is accurately determined
using a nozzle which is mounted at the end of the outlet duct. Furthermore, dynamic
and static pressures are measured using a pitot tube and a pressure tap, respectively.
Hence, the fan outlet total pressure and the inlet airflow rate can be calculated [9].

Figure 2.3: Free inlet-ducted outlet installation setup [8].

Normally, performance data consist of information about fan pressure rise ∆p, power
consumption W , volume flow rate Q, and efficiency η, which can be presented in
graphical form, as shown in Figure 2.4.
The initial point on the pressure curve is known as shut-off, and represents the
point of zero volume flow rate. On the other hand, free-delivery point is referred
to the point at which the volume flow rate becomes maximum [10]. The peak of
the pressure curve splits the curve into two regions. The right side of this point is
referred to as the stable section where the operation point can be selected, while the
left part of the peak point represents the stall region of the pressure curve where
the fan is no longer efficient [11].
The performance curve of a given fan depends on the geometry and speed of the
fan, but the operation point depends on the system (installation). Each installation
has a resistance for the required volumetric flow rate [3].

7
2. THEORY

Figure 2.4: Fan performance curves [5].

As performance curves, the resistance of the system can be plotted on a ∆p - Q


graph. For a specific system, the pressure is calculated for a certain number of
volumetric flow rates, and the curve defined by these points is the system charac-
teristic. Then, by having both the system and fan characteristics at rotation N , the
operation point is specified where both curves cross each other. In Figure 2.5 two
operating points, A and B, are identified for two different fan rotation speeds.

Figure 2.5: System and fan performance curves [3].

2.2 Noise
2.2.1 Sound measurement units
Noise can be expressed in terms of Sound Power Level (SWL or LW ) and Sound
Pressure Level (SPL or Lp ). Sound power level is logarithmically compared with a
reference power, the picowatt (10−12 watt), while sound pressure level is logarith-
mically compared with a reference pressure of 2·10−5 Pa. The unit to express both
quantities is the decibel scale, dBW for sound power level and dB for sound pressure

8
2. THEORY

level. The sound power and pressure level can be determined as follows:
W
 
LW = 10 log10 , (2.1)
W0
!
p
Lp = 20 log10 , (2.2)
p0
where W is the sound power of the noise generating equipment, W0 the reference
power, p sound pressure of the noise, and p0 the reference pressure. The main
difference between Equations (2.1) and (2.2) is that the distance is accounted for
when using SPL. Equation 2.3 can be used to relate both measurements,

 
Lp = LW + 10 log10 , (2.3)
4πr2
where r is the distance from the source (m) and Qθ is the directivity factor of the
source in the direction of r. The value of Qθ changes depending on the position of
the fan in the room, as explained in [3].

2.2.1.1 Octave bands and weighted sound pressure levels


Noise analysis are usually represented as a combination of notes of different frequen-
cies, and because these frequencies have different characteristics, the sound pressure
level is separated into octave bands. The most used in engineering acoustics is 1/3
octave bands, with mid-frequencies of 63, 125, 250, 500, 1000, 2000, 4000, and 8000
Hz [3]. Then, in order to obtain a single number or an overall sound pressure level,
different levels are applied to the octave band results. These are A,B,C, and D levels,
where each level has a different value for each octave band that are subtracted to
the pressure levels at every mid-frequency. Figure 2.6 graphically shows the weight-
ings employed for A level. After applying this weighting, the results are then added
logarithmically to give the overall sound pressure level. The resulting noise level is
known as dBA [3]. The reason to apply these filters is based on the ear’s sensitivity
to the range of frequency around 1000 to 4000 Hz, with a poor response to lower
frequency sounds [12]. Hence, the weight of these frequencies are reduced.

Figure 2.6: A-weighted sound pressure curve [3].

9
2. THEORY

2.2.2 Noise in axial fans


There are three types of fan noise sources, these are divided as:
1. Aerodynamic
2. Electromagnetic
3. Mechanical
For fans typically used in the electric automotive industry, electromagnetic and
mechanical contributions are of minor importance when compared to aerodynamic
contributions [3]. On the other hand, the installation also contributes to the total
noise output.
Moreover, the nature of aerodynamic noise can be indicated by its spectrum. This
spectrum is composed by discrete frequency and broadband components. One ex-
ample of discrete harmonic component is Blade Pass Frequency (BPF). If a fan with
z blades, where z is an unspecified number of blades, rotates with constant speed
Ω, the first harmonic presents as a signal with fundamental frequency zΩ, while
the second harmonic has a frequency of twice the fundamental frequency, and so
on. On the other hand, broadband noise, by definition, contains components of all
frequencies with different amplitudes [13].
The sources of this spectrum are usually divided into three categories, such as blade
thickness noise, blade forces noise, and turbulent noise. According to the work of
Lighthill, Ffowcs Williams, Hawkings and Curle, the three categories are idealized
as monopole, dipole and quadrupole radiation sources, respectively [14]. More in-
formation about this can be found in Crocker [12], chapter 3. Therefore, fan noise
mechanisms can be summarized as shown in Figure 2.7. For low speed axial fans,
monopole and quadrupole radiations are not as important as dipole radiation [13].

Figure 2.7: Fan noise mechanisms [14].

As summarized in Figure 2.7, blade forces can be steady or unsteady. The former

10
2. THEORY

are those exerted by the fan on a uniform stationary flow which is known as Gutin-
Noise. While the latter, in contrast, occur when the fan operates in a stationary
but non-uniform flow field. Both forces contribute to the overall sound radiation,
but at low speeds, steady forces noise is negligible compared to noise generated
by unsteady forces [13]. A non-uniform flow can be produced by obstructions at
locations upstream of the fan, such as radiator, rods, guide vanes, duct bends, as
well as an asymmetric position of the intake with respect to adjacent floors and
walls. This means that the installation plays an important role on the overall noise
[13], but the evaluation of these effects is beyond the scope of the present work.
Another important aspect to consider is the optimum operational point of the fan.
The optimum performance, both from an operational and acoustic point of view, is
to the right of the stall point in the fan performance curve, typically at the maximum
static efficiency point. At the stall point, the sound suffers a great change in the
spectrum due to a higher level of the broadband component because of the flow
characteristics [14].

2.2.3 Fan sound law


As previously shown, the fan noise can be expressed by power unit W. Then, it
is possible to say that there is a relation between the noise output and the im-
peller power. Additionally, there is a relation with both the Mach number and the
Reynolds number. For standard air conditions, the following relations are set:
• Impeller power ∝ N 3 D5 ,
• Mach number ∝ Tip speed ∝ N D,
• Reynolds number ∝ Tip speed × diameter ∝ N D2 ,
where N represents impeller velocity and D represents diameter of the rotor. So,
the final relation can be assumed:
Fan SW L ∝ N 3 D5 × (N D)a × (N D2 )b , (2.4)
where a can be 1, 3 or 5, depending on the type of source, i.e. monopole, dipole
or quadrupole, respectively, while b depends on more complex parameters, such
as blade thickness, number of blades, blade angles, etc., but lays in the range of
0 > b > −1 [3].
This relation can be used to estimate the increase or decrease in dBs of a particular
fan design. For example, assuming that dipole sources predominate and b = -0.5,
by converting to a logarithmic decibel scale, the following equation is obtained:
N2 D2
LW = X + 55 log10 + 70 log10 , (2.5)
N1 D1
where
• X = Constant for a particular fan design at a particular point on its char-
acteristics,
• N1 = Original fan speed (rpm),
• N2 = Final fan speed (rpm),
• D1 = Model fan diameter (m or mm),
• D2 = Final fan diameter (m or mm).

11
2. THEORY

Based on Equation (2.5), it can be noted that the sound power level is greatly related
to the velocity and diameter of the fan, two parameters that depend on the desired
performance of the fan.

2.2.4 Noise control in axial fans


Noise control is normally defined as a set of strategies which helps to reduce the
sound pollution in order to improve the quality of human life. Land-based trans-
portation system is one of the main areas in which noise mitigation is of great
importance. In order to minimize the acoustic effect of axial fans used in heavy
electrified vehicles, some fundamental noise reduction methods are discussed.
Proper choice of airfoil profile along the blade span leads to a smooth and balanced
pressure distribution on the suction side of the impeller blades, which in turn reduces
the noise [15].
Aerodynamic losses at low rotational speeds are considered as one of the main noise
sources in low-pressure axial fans. One appropriate way to diminish the losses is to
force the fluid particles in the boundary layer to reach the blade trailing edge by
travelling a shorter distance. In order to do so, a forward-swept blade configuration
can be employed [15]. In addition, forward-swept blades generate an increase in fan
efficiency, and extend the stall limit in low-speed axial fans [16].
Appropriate choice of number of blades is another useful approach which has a
significant effect on reducing the noise. Since the fan rotational speed and number
of blades have a great influence on periodic fluctuating forces, the acoustic resonance
can be mitigated by selecting a prime number for the rotor blades [17]. Moreover,
uneven blade spacing has an advantage over regular spacing from the acoustic point
of view. The main reason is that the acoustic energy generated by the rotating
blades spreads over a wider frequency range which in turn reduces the noise [15].
Rotor blade response is considered as a noise source related to the fan. One efficient
way to diminish the blade response is to design a blade with larger tip chord. The air
flow is constricted upstream of the rotor due to the existence of the hub and forces
the turbulence gusts to move towards the blade tip. Since the blade tip velocity is
much higher compared with the velocity at the hub, it is necessary to have a larger
chord at the tip in order to reduce the noise [13].
Another useful technique for reducing the noise generated by the fan impeller is to
decrease the tip clearance or make use of a rotating shroud attached to the tip of
the rotor blades. This method not only enhances the performance of the fan but
also reduces the tip clearance noise [15].
Distorted inlet airflow has an adverse influence on the fan performance and also
generates noise when passing through the fan impeller. A number of studies have
shown that the inlet airflow distortion is proportional to the change in lift-incidence
distribution along the blade span. In order to mitigate the noise generated due
to distorted inlet flow, it is recommended to choose a suitable blade stagger angle
where the above-mentioned change is small [15].
Another major noise source in axial fans is related to the blade trailing edge. The
interaction of the flows travelling on the pressure and suction sides at the blade
trailing edge generates vortex-shedding noise. There is a robust noise reduction

12
2. THEORY

technique in which the blade trailing edge is in the shape of a sawtooth [15]. The
blade serrations help to diminish the noise generated by the trailing edge in the far
field [18].

2.3 Blade Element Theory


Blade Element Theory (BET) is a method which can be used to analyze the per-
formance of an axial fan by dividing the rotor blades into a series of independent
elements along the span, as illustrated in Figure 2.8. This theory is based on study-
ing torque and thrust force caused by lift and drag on each blade section without
considering the aerodynamic interactions between the elements [19].

Figure 2.8: Fan impeller diagram illustrating an annular blade element of length
dr at radius r [20].

2.3.1 Blade element forces


As illustrated in Figure 2.9, the lift force on a blade element is perpendicular to
the incoming flow, while the drag force is parallel to the flow. Tangential and axial
forces can be determined using the local lift dL, local drag dD, and relative flow
angle β as follows:
dFa = dD cos β + dL sin β, (2.6)
dFu = dL cos β − dD sin β. (2.7)
Furthermore, Lift and drag can be calculated by considering the airfoil properties
of each element. Therefore, dL and dD can be formulated as below:
1
 
dL = CL ρW 2 c dr, (2.8)
2
1
 
dD = Cd ρW 2 c dr, (2.9)
2
where W is the relative velocity, c is the local chord length, and ρ is the air density.
Moreover, the lift and drag coefficients are represented by CL and Cd , respectively.
In order to determine local axial and tangential forces generated by an impeller with

13
2. THEORY

sufficient number of blades z, Equations (2.6)-(2.9) can be employed which result in


the following expressions:

1
 
dFa = z ρW 2 (Cd cos β + CL sin β) c dr, (2.10)
2

1
 
dFu = z ρW 2 (CL cos β − Cd sin β) c dr. (2.11)
2

Figure 2.9: Forces on a blade section [20].

Finally, it can be concluded that the local thrust force dT is identical to the axial
force on each blade section, and the torque applied on each element, dτ , can be
calculated by multiplying the corresponding tangential force by the radius r. Hence:

1
 
dT = dFa = z ρW 2 (Cd cos β + CL sin β) c dr, (2.12)
2

1
 
dτ = dFu r = z ρW 2 (CL cos β − Cd sin β) cr dr. (2.13)
2

2.3.2 Relative flow


Aerodynamic performance of an airfoil is normally analyzed by fixing the airfoil
section in the airstream generated by the wind tunnel. Since the airfoil section is
considered as a rotating part of the fan, it is necessary to make use of the relative
velocity W in order to relate the airflow over the non-stationary section to the one
tested in the wind tunnel [19].

14
2. THEORY

Figure 2.10: Flow over a blade section.

As it is shown in Figure 2.10, the average tangential velocity felt by the blade section
can be defined as Ωr(1 − b), where b = ω/(2Ω) is the angular induction factor.
Furthermore, the wake rotational speed downstream of the blade is represented by
ω, and Ω is the blade rotating speed. In addition, the axial velocity right before the
impeller can be written as a function of the axial induction factor a and the incoming
flow velocity V∞ , as reflected in the velocity diagram in Figure 2.10. Therefore, the
relative flow angle β can be calculated using axial and average tangential velocities

Ωr(1 − b)
!
β = tan −1
. (2.14)
V∞ (1 − a)

Since the relative flow angle for each blade element is dependent on the rotational
speed of the corresponding element, then the value of β increases from the hub to
tip. Moreover, the relative velocity can be expressed as

V∞ (1 − a)
W = . (2.15)
cos β
Finally, torque and thrust can be expressed in form of induction factors by substi-
tuting Equations (2.14) and (2.15) into Equations (2.12) and (2.13).

15
3
METHODS

3.1 Design Method 1


The first design approach is based on the method developed by Maria T. Pascu [21].
This procedure is a design-point method, meaning that the axial fan is designed
for the maximum efficiency point on the system characteristic shown in Figure 2.4,
corresponding to the design volume flow rate Qd .
This technique is a combined inverse–direct design method, which computes the
optimum blade profiles, according to the fan specifications and operational require-
ments, and does not make use of existing shapes, such as airfoils.

3.1.1 Mean line calculation


The method starts with the one-dimensional flow analysis, i.e. the mean-line calcu-
lation. The assumption here is that the fan is composed by a succession of cascades
at different distances from the hub center. Then, the mean-line calculation can be
done with Euler’s turbomachinery equation. For axial fans, it can be assumed that
the flow particles enter and leave the blade at the same section (axial entry and
exit), and therefore u1 = u2 = u. In Figure 3.1, the velocity triangle for axial entry
along with velocity component’s nomenclature are shown. Velocity components w
are for relative velocity while components c refer to absolute velocity.

Figure 3.1: Mean-line calculation of the velocity diagram with axial entry [21].

16
3. METHODS

Euler’s equation, assuming inviscid and incompressible flow can be written as fol-
lows:
ρh 2 i
∆pt = (w1 − w22 ) + (c22 − c21 ) . (3.1)
2
Since axial entry is assumed, the tangential component of the absolute flow velocity
is zero (cu1 = 0) meaning that wu1 = u. Therefore, Equation (3.1) becomes
ρh 2 i ρh 2 i
∆pt = w1 − w22 + c2u2 = 2
u − wu2 + (u − wu2 )2 = ρu(u − wu2 ) = ρucu2 .
2 2
(3.2)
Given that Qd and N are known, the inlet blade angle γ1 and outlet blade angle
γ2 can be obtained from Figure 3.1. The mean component of the relative velocity
wm is constant throughout the cascade, and is equal to the ratio of Qd and the flow
area. The flow area is the ring area between the radial position of the hub rh and
the tip rt . Therefore, wm can be obtained
Qd
wm = . (3.3)
π(rt − rh2 )
2

3.1.2 Outlet blade angle


Most methods for designing axial fans assume a free-vortex flow, this means that
the total pressure difference along the blade is constant. This can be expressed as
∆pt,r
= 1, (3.4)
∆pt,h
where subscripts r and h stand for radial position and hub position, respectively.
As stated in Pascu [21], this assumption is preferred by most designers due to the
good results it gives. Since in reality this is not the case, Pascu [21] developed her
method considering a change in the pressure distribution in the span-wise direction
to obtain better results. This can be expressed by making Equation (3.4) a function
of the radius. Then, by employing the expression for the total pressure difference in
Equation (3.2) and the outlet velocity diagram in Figure 3.1, after some derivation
the following relation of outlet blade angle at different radial positions is obtained
1 2πn h 2 i
= r − f (r)rh2 , (3.5)
tan γ2r rwm
where f (r) is the pressure distribution along the blade. After an iterative process,
where Pascu [21] considered different constraints such as avoiding rapid variations
from one section to the next and negative outlet blade angles, the following expres-
sion was derived:  x
rt
f (r) = x (r − rh )1.35 + 1. (3.6)
rh
In Equation (3.6), x can be changed to have different pressure distributions. For
example, if x = 0 it results in constant pressure (free-vortex flow), for x = 1 there
is a linear pressure distribution, whereas for x = 2 the distribution is parabolic and
so on. Therefore, by combining Equations (3.5) and (3.6), the following expression
to determine the outlet blade angle is obtained:
1 2πn 2
  x
rt
  
= r − x (r − rh )1.35 + 1 rh2 . (3.7)
tan γ2r rwm rh

17
3. METHODS

3.1.3 Blade shape computation


As described by Pascu [21], the shape of the blades is based on the inlet profile angle
γ1 and outlet profile angle γ2 , where the profile angle at any point between them is
γm . The cartesian system is shown in Figure 3.2.

Figure 3.2: View of the calculated blade shape in (x,y) coordinates [21].

In order to compute the distribution of profile angles, y coordinates must be given,


where y1 = 0 and y2 depends on the hub height. At the same time, the hub height,
i.e. y2 , depends on the chord length of the blades, and it must be chosen to the right
height in order to avoid unrealistic and rapid changes in blade profile angle. Then,
the angle distribution γ(y) can be computed, since

∆y
tan γ(y) = . (3.8)
∆x
The increase from γ1 to γ2 is to apply a driving action on the fluid, and it can be
calculated by linear, parabolic or higher degree polynomial distribution. According
to Pascu and Epple [22], a parabolic distribution can increase the static efficiency
of the fan. Therefore, the distribution is γ(y) = Ay 2 + By + C:
• y = y1 → γ = γ2
• y = ym → γ = γm
• y = y2 → γ = γ1
By having the inlet and outlet conditions fixed, there is only one degree of freedom
by (γm , ym ), resulting in the complete distribution of profile angles.

3.1.4 Losses and diffusion factor


Another important parameter to consider in the design phase, according to Pascu,
is the diffusion factor DF . As it will be explained in the next method, there are
many factors that contribute to loss in total pressure, and one of them is boundary
layer separation. Under normal operating conditions, the suction surface will have a
much thicker boundary layer than that on the pressure surface of the blade, therefore,
the latter can be consider negligible. Then, the thickness of the wake, and hence,
the total pressure loss, is determined by that section on the surface over which the
velocity gradient is negative, given that in that section is where most of the boundary
layer growth occurs [21]. So, the diffusion of the fluid from wmax to w2 is a direct

18
3. METHODS

correlation to the total pressure loss. Therefore, the diffusion factor is

wmax − w2
DF = . (3.9)
wmax

According to the research of Lieblein et al. [23], the momentum thickness of the
wake θ was related to the diffusion factor. Several measurements were performed
for two airfoil series, the NACA 65 and British C4 series. It was found out that a
value of DF = 0.6 is the upper limit for the diffusion factor. A higher value than
that, results in a dramatic increase in the diffusion in the boundary layer, as shown
in Figure 3.3.

Figure 3.3: Wake momentum thickness versus overall diffusion factor DF for
NACA 65 and C4 airfoils at minimum loss incidence [21].

3.2 Design Method 2


The second method employed for designing the fan blades is a combination of three
approaches, namely isolated airfoil, cascade, and modified isolated approaches. This
method is applied to two-dimensional flow and results in an appropriate airfoil profile
selection at different predefined span-wise positions. In other words, aerodynamic
characteristics of various airfoils are investigated with the purpose of choosing a
proper airfoil profile at each radial position.
An important factor in selection of a suitable approach among the three approaches
mentioned earlier is the blade solidity σ which can be determined as

c
σ= . (3.10)
s

As expressed in Equation (3.10), the blade solidity is the ratio between the blade
chord length c and the pitch distance s. Blade-to-blade pitch is the distance between
two adjacent blades which is shown in Figure 3.4.

19
3. METHODS

Figure 3.4: Chord length and pitch distance [24].

As already stated, the blade solidity factor is of great importance in designing fan
blades due to the aerodynamic effect of the adjacent blades on each other. The
determination of the solidity factor at different radial positions makes it possible to
select an appropriate approach during the fan blade design process.
The isolated airfoil and cascade approaches are valid for low and high solidity rotors,
respectively. It is required to use isolated airfoil technique for the blade solidity range
less than 0.7 and make use of the cascade approach for solidities larger than 1 in
order to achieve a reasonable blade design [25]. Furthermore, the modified isolated
approach can be applied to fulfill the requirement for the mid-range solidities [25].

3.2.1 Isolated airfoil


A fan blade is basically examined as a series of airfoil sections when using the
isolated airfoil technique to design a fan impeller. This approach is useful when the
aerodynamic interaction between two neighbor blades is negligible [26]. An airfoil
is generally exposed to non-uniform pressure distributions and viscous forces when
submerged in a flow field. As shown in Figure 3.5, the resultant force acting on the
airfoil is determined in term of lift force FL normal to the flow direction and drag
force Fd parallel to the mean flow direction [9]. These two forces can be formulated
as a function of lift coefficient CL and drag coefficient Cd as below:
1
 
FL = CL bc ρWm2 , (3.11)
2
1
 
Fd = Cd bc ρWm2 , (3.12)
2
where Wm is the relative mean flow velocity, b is the blade span and c represents
the chord length. The lift is a resultant force component which provides energy
transformation from the rotor shaft to the fluid. On the other hand, the drag refers
to the friction between the fluid and the surface of the blade [9].

20
3. METHODS

Figure 3.5: Typical airfoil section nomenclature [9].

In order to relate the fan performance parameters with the forces mentioned previ-
ously, a control volume with a width identical to the pitch of a rotor is employed to
surround an isolated airfoil, as illustrated in Figure 3.6. The next step is to derive
the relative mean flow angle βm using

1
tan βm = (tan β1 + tan β2 ), (3.13)
2

where β1 and β2 are the inlet and outlet flow angles, respectively. It is now possible
to derive the net resultant force as a function of the lift and drag forces using the
momentum equation, the velocity triangle shown in Figure 3.6, and Equation (3.13).
Hence, the lift and drag coefficients are

s
 
CL = 2 (tan β1 − tan β2 ) cos βm − Cd cos βm , (3.14)
c

∆p
!
s
 
Cd = . (3.15)
c ρW12 /2

Since the drag coefficient of the airfoil is considerably less than the lift coefficient,
the second term in Equation (3.14) is normally neglected [9]. It can be noted that
the pressure difference through the blades and fan performance are strongly related
to the drag and lift coefficients [9].

21
3. METHODS

Figure 3.6: Isolated airfoil and velocity triangles [9].

As the blade-to-blade pitch distance decreases, the solidity factor increases, meaning
that the blade pressure field can be greatly influenced by other blades. Hence, a
mismatch between the true and calculated lift values occurs [27]. Consequently, the
isolated airfoil technique is considered reliable for an axial fan which is characterized
by low solidity and substantial stagger angles γ [28]. As reflected in Figure 3.5,
stagger is the angle between the chord line and the axis of rotation, and is defined
as
γ = βm − α, (3.16)

where α is the angle of attack which depends on lift coefficient of the isolated airfoil.

3.2.2 Cascade
A fan with a high flow coefficient is distinguished by high solidity and low stagger
angles. In this case, a simple design technique called cascade approach is expected
to be utilized [28]. As illustrated in Figure 3.7, the model includes a row of similar
blades located at equal distance s from each other, which demonstrates the actual
impeller in the axial fan. The cascade technique considers the flow deflection across
the blade instead of focusing on the lift force corresponding to each blade [27]. The
major parameters used in the cascade technique procedure are flow deflection angle
ε, camber angle θ, solidity σ, incidence angle i, and stagger angle γ.
The first step in the cascade approach is to select a proper solidity from a cas-
cade performance database corresponding to a desired airfoil series, and thereafter
calculate the camber angle using Equation (3.17) [28].

ε−i
θ= q (3.17)
1 − m s/c

where the flow deflection angle is defined as ε = β1 − β2 and m is the deviation angle
coefficient which is a function of stagger angle.

22
3. METHODS

Figure 3.7: Cascade of rotor blades [9].

The initial values of incidence angle and deviation angle coefficient can be allocated
to zero and 0.26, respectively, in order to obtain an appropriate preliminary value for
camber angle θ, according to Wallis [28]. It is now possible to find an approximate
value of stagger angle using the following equation:
θ
γ = (β1 − i) − . (3.18)
2
The final step is to select more precise values of incidence angle and of deviation
angle coefficient from the cascade database and recalculate the camber and stagger
angle.

3.2.3 Modified isolated airfoil


As explained earlier, the aerodynamic interference between two adjacent blades in
an axial flow fan is small and can be neglected due to the large blade-to-blade
pitch distance. In order to design a fan impeller with tighter space between the
blades, a multiplane interference factor K for airfoils having a circular camber line
is introduced in Figure 3.8. The interference factor is the ratio of the cascade lift
coefficient of an impeller blade to the lift coefficient of an isolated one [26]. Therefore,
the interference ratio can be obtained by
CL
K= , (3.19)
CLi
where CL and CLi are the lift coefficients of a single cascade and isolated blade,
respectively.
The modified isolated technique is based on the isolated airfoil approach where the
interference factor is taken into account [28]. This method allows the designer to
modify two-dimensional lift and drag coefficients when increasing the blade solidity
which results in more accurate rotor design.

23
3. METHODS

The series of design curves demonstrated in Figure 3.8 shows how the interference
factor is related to the stagger angle and blade solidity. It can be noted that, as the
solidity decreases, the interference coefficient value moves towards unity, whereas it
approaches zero as the blade solidity increases.
Since the isolated airfoil and cascade data are not valid for intermediate solidities, it
is necessary to apply an interference factor in order to have a reasonable transition
between data corresponding to cascade and isolated airfoil.

Figure 3.8: Interference factor for airfoils with a circular-arc camber line [9].

3.2.4 Pressure loss components


Since the early 1940s, many attempts have been made in order to study the fac-
tors which have negative influences on turbomachinery efficiency. Pressure losses
associated with the rotor and duct cause performance deterioration in axial fans.
The effect of these losses should be considered by applying correlation factors when
designing a new fan. The loss sources are categorized based on their mechanisms
and should be analyzed independently. Empirical loss coefficients presented in this
section are developed by Howell et al. [29] .
1. Profile loss is usually referred to the loss created due to the formation of
boundary layer on the pressure and suction sides of a blade. Moreover, this
type of loss arises away from the hub and tip walls [30]. Surface roughness,
Reynolds number, shape of profile, and air turbulence are four main factors
that affect the profile loss [28]. Furthermore, profile drag coefficient Cdp is
obtained from data corresponding to the airfoil which is used to design the
blade.
2. Secondary loss, also known as endwall loss, is induced by the secondary flows
created by the annulus boundary layers between the blades and occurs mainly
at the endwalls [30]. Secondary loss coefficient Cds can be estimated using
Cds = 0.018CL2 . (3.20)
3. Tip clearance loss is generated due to the existence of an empty space
between the impeller tips and fan casing which results in flow leakage over the

24
3. METHODS

rotor tips [30]. Tip clearance drag coefficient Cdt is a function of blade span b
and tip clearance CT . The loss due to the tip clearance leakage is determined
as
CT
 
Cdt = 0.29 CL1.5 . (3.21)
b
4. Annulus loss is referred to the pressure loss due to friction between the fluid
and annular surfaces such as hub and housing walls [9]. In addition, annulus
pressure loss coefficient Cda can be estimated as
0.02s
Cda = , (3.22)
b
where s is the pitch distance.
The pressure loss components explained in this section should be taken into account
in order to predict the fan performance as accurately as possible. Therefore, the
total drag coefficient can be calculated by
Cd = Cdp + Cds + Cdt + Cda . (3.23)

3.3 Fan Performance Estimation


3.3.1 Theoretical
As it was mentioned in Section 2.1.4, the parameters to measure fan performance
are usually volume flow rate Q and pressure difference ∆p, more precisely, the total-
to-static pressure difference ∆pt−s . According to Pascu et al. [22], the total-to-static
pressure difference for axial fans can be expressed as shown in Equation (3.24) by
using the velocity triangle in Figure 3.9 as reference.
 !2 
ρ ρ cm2
∆pt−s = ∆pt − c22 = u2 − (3.24)
2 2 sin β2

Figure 3.9: Velocity triangle for axial fans [22].

25
3. METHODS

Equation (3.24) is valid for one radial position, only. In order to obtain the total-to-
static pressure difference for the whole impeller, Equation (3.24) has to be integrated
for all radii between hub and tip, taking the area average as below:
rt
 !2 
1 Z
ρ 2 cm2
∆pt−s,axial = u −  2πr dr . (3.25)
π(rt2 − rh2 ) r 2 sin β2
h

Then, by setting a zero flow rate, i.e. cm2 = 0, in Equation (3.24), the shut-off point
or maximum pressure difference can be calculated by integrating for all radii and
taking the area average as
ρ
∆pt−s,max = u2 , (3.26)
2
Zrt
1 ρ 2
∆pt−s,max,axial = u 2πr dr . (3.27)
π(rt − rh ) r 2
2 2
h

In the same way, the free flow point or maximum flow rate can be calculated by
setting the pressure difference in Equation (3.24) equal to zero and considering that
cm2 = Q/A which leads to
Qt−s,max = uA sin β2 , (3.28)
rt rt
1Z Z
Qt−s,max,axial = uA sin β2 2πr dr = 4π 2 n r2 sin β2 dr . (3.29)
Ar r
h h

Finally, by having both the shut-off point and the free flow point, it is possible to
describe the complete performance curve of the axial fan, using Equations (3.24),
(3.26) and (3.28):
 !2 
Q
∆pt−s = ∆pt−s,max 1 − . (3.30)
Qt−s,max
The next performance parameter is the efficiency. There are two efficiencies that are
relevant to describe the performance of an axial fan. The total-to-total efficiency
ηt−t of a fan or a pump is defined as the ratio of the total hydraulic power divided
by the shaft power. If no viscous and three-dimensional losses are considered, then
all the mechanical power of the shaft is transferred to the fluid [22]. This can be
expressed as
Whyd,t Q∆pt
ηt−t = = = 1. (3.31)
Wshaf t Q∆pt
Like the pressure difference, the correct efficiency to describe the performance of a
fan is the total-to-static efficiency ηt−s . The total-to-static efficiency is defined as
the ratio of the total-to-static hydraulic power divided by the shaft power
Whyd,t−s Q∆pt−s ∆pt−s
ηt−s = = = . (3.32)
Wshaf t Q∆pt ∆pt
In contrast to the total-to-total efficiency, the efficiency ηt−s will always be less than
one, even if no losses are considered. By rewriting Equation (3.32) with Equation
(3.24), the total-to-static efficiency is
∆pt − (ρ/2)c22 ρ c22
ηt−s = = ηt−t − . (3.33)
∆pt 2 ∆pt

26
3. METHODS

As shown in Equation (3.33), the efficiency ηt−s is less than 1 due to the second
term. This term is independent of the viscosity and it is only a mechanical loss
that has to be accounted for in the performance of fans and pumps [22]. Equation
(3.33) is only valid for one radial section of the blade, and to get the efficiency for
the complete impeller it must be integrated for all radii by taking the area average.

3.3.2 CFD
In order to perform the CFD simulations, 3D models of the 11 designs were created
in Solid Edge 2020. Therefore, the coordinate system of the blade section profiles
obtained with the two design methods was transformed to cylindrical coordinate
system. The CAD models of the 11 fans are shown in Appendix C.
The CFD simulations were performed by Mattias Chevalier from RTGF group in
Scania. In this section, the description of the settings used for the analysis is speci-
fied, as described by Mattias.
The main performance parameters such as volume flow rate, torque, and forces are
measured at three different rotational speeds for each fan model. The simulations
are conducted with STAR-CCM+ v14.06.012, which is developed by Siemens.
The process uses a finite volume approach to discretize the Navier-Stokes equa-
tions which describes the physics of fluids. The flow is assumed to be an ideal gas
considering compressibility effects. In addition, the boundary conditions are set as
stagnation flow inlet and pressure outlet. Furthermore, the airflow is considered to
be turbulent and is modeled using the Reynolds Averaged Navier-Stokes (RANS)
equations. The technique used is the Realizable K-Epsilon All y+ Wall Treatment
turbulence model.
The performance data are measured by installing a shroud around the fan models.
Moreover, each fan assembly is placed between two chambers with a flow resistance
(porous medium) upstream and a blockage downstream, as shown in Figure 3.10.

Figure 3.10: A cut plane view of the fan installation, from left to right: blockage,
fan, fan shroud, and flow resistance.

27
3. METHODS

In order to diminish undesired numerical effects from the boundaries, the computa-
tional domain extends upstream and downstream of the installation. As illustrated
in Figure 3.11, the computational mesh consists of 15 million cells in form of polyhe-
dral inside the domain. In addition, prism layers are used near the walls. Finally, fan
curves are derived by changing the porous medium resistance at a given rotational
speed.

Figure 3.11: A center cut plane view of the computational domain, the flow di-
rection is from right to left.

3.4 Noise Level Estimation


Throughout the years, many attempts have been made in order to estimate or predict
SWL or SPL from a fan. The complexity of these models vary, it can be a complex
mathematical model, such as the work done in Berglund [14] or simple formulas like
the one used in Fukano [31].
Since the physics behind the aerodynamically generated noise by an axial fan is
complex, and due to limitations in time and equipment, the noise level prediction
in this report is done with simple empirical formulas, as explained in this section.

3.4.1 Approach 1
As stated in Cory [3], the first equation used to approximate the sound power level is
the resulting work done by Beranek, Kamperman, and Alien as shown in Equation
(3.34).
LW = 57.3 + log10 (Qd ) + 20 log10 (∆p) − 10 log10 (η) (3.34)
where
• LW = Overall sound power level of noise transmitted along ducts fitted to
the inlet and outlet of fan operating at/near its peak efficiency (dB ref
10−12 W),
• Qd = Design flow rate (or flow rate at or near peak efficiency) (m3 /s),
• ∆p = Pressure difference at or near peak efficiency (Pa),
• η = Peak efficiency (%).

Then, with Equation (2.3) the sound pressure level can be calculated.

28
3. METHODS

3.4.2 Approach 2
The method presented in Section 3.4.1 results in only one number, namely the overall
power level, but it is also possible to estimate the sound power spectrum of the noise
(LW vs frequency). First, it is necessary to introduce the specific sound power level,
which is defined as
W
H(φ) = , (3.35)
∆p2 Q
where W is the overall sound power in watts. As stated in Crocker [12], Table 3.1
gives specific sound power levels in several octave bands for axial fans. The levels
are given in decibels with reference quantities: W0 = 10−12 W, Q0 = 1 m3 /s, and
∆p0 = 1 kPa.

Table 3.1: Specific sound power levels in eight lowest octave bands for axial fans
[12].

Octave Band Center Frequency (Hz)


Rotor
Fan type Diameter 63 125 250 500 1000 2000 4000 8000 BPF
(m)
Vaneaxial
0.3 ≤ ν ≤ 0.4 All 94 88 88 93 92 90 83 79 6
0.4 ≤ ν ≤ 0.6 All 94 88 91 88 86 81 75 73 6
0.6 ≤ ν ≤ 0.8 All 98 97 96 96 94 92 88 85 6
Tubeaxial >1.0 96 91 92 94 92 91 84 82 7
<1.0 93 92 94 98 97 96 88 85 7
Propeller All 93 96 103 101 100 97 91 87 5

It is possible to estimate the noise spectrum of a fan operating at certain Q and ∆p


using Table 3.1. To do this, simply add 10 log10 Q + 20 log10 ∆p to the tabulated
values. In the octave band where the Blade Passing Frequency (BPF) is located, the
tabulated value for BPF is added to that octave band. The blade passing frequency
can be calculated as BPF= nz.
Once the spectrum of the noise is calculated, the values are transformed from SWL
to SPL with Equation (2.3). Then, the A-weighting filter is applied, which is done
by subtracting the values in Figure 2.6 to each octave band value calculated. Lastly,
with Equation (3.36) overall sound pressure level OASPL can be calculated.

OASP L = 10 log10 10Lpi /10 (3.36)


X

29
4
DESIGN PROCEDURE

4.1 Design Method 1


The initial inputs are the design flow rate Qd , the total pressure difference ∆p, and
the speed of the fan N in rpm. It should be noted that the pressure difference for
this step is the total and not the total-to-static. The total-to-static pressure is the
relevant value for fan performance, as shown in Figure 2.4 [32].
Once the requirements are set, it is possible to use the Cordier diagram to find the
optimum dimensions. This diagram is an empirical diagram based on measurements.
It delivers a relation between flow rate, pressure, rotating speed, and diameter. The
Cordier diagram has the speed number σ on the y-axis and the diameter number δ
on the x-axis. Therefore, it is necessary to estimate the speed number σ with the
following equation: √
√ Qd
σ = 2n π , (4.1)
(2Y )0.75
where the specific head Y = ∆p/ρair , and n = N/60. Then, the speed number
is located on the Cordier diagram, shown in Figure 4.1 , to obtain the diameter
number δ. The diameter number must lie on the curve of maximum efficiency for
low-pressure fans.

Figure 4.1: Cordier diagram [21].

30
4. DESIGN PROCEDURE

Next, by solving Equation (4.2) for Diameter Dt , the optimum fan diameter for the
specified performance is found.
√ !0.25
π 2Y
δ= Dt (4.2)
2 Q2d

Then, with Equation (4.3) and Figure 4.2, the hub-to-tip ratio ν can be found:

N Qd
Ns = . (4.3)
∆p0.75

Figure 4.2: Range of optimum hub-to-tip ratio [9].

In Equation (4.3), Qd must be in CFM and ∆p in in.w.g. Given that ν = Dh /Dt ,


the hub diameter can be calculated. Moreover, the number of blades z is found with
Equation (4.4).

z= (4.4)
1−ν
After obtaining the main fan characteristics, the blade profile angles can be calcu-
lated for each blade section at a radial position r.
First, the tangential velocity u and pitch distance s are calculated as

u = 2πrn, (4.5)

2πr
s= . (4.6)
z
To estimate the chord length of the blades, the specific speed Ns along with Figure
4.3 is used.

31
4. DESIGN PROCEDURE

Figure 4.3: Specific speed vs. mean head and flow coefficients for axial fans [9].

Where φ is the flow coefficient and ψ the head coefficient. The flow coefficient φ is
defined as
wm
φ= . (4.7)
u
The mean component of the relative velocity wm was introduced in Equation (3.3).
Finally, chord length c is obtained by solving the value from Figure 4.3. The next
step is to calculate the inlet blade angle γ1 and the outlet blade angle γ2 . From
Figure 3.1:
wm
tan γ1 = . (4.8)
u
As mentioned in the previous chapter, Pascu developed Equation (3.7) to estimate
γ2 . In her work, Pascu concluded that assuming a parabolic distribution, i.e. x = 2,
yields better results [21]. This was confirmed by both CFD and experimental results.
Therefore, Equation (3.7) becomes
1 2πn 2
"  2 # !
rt
= r − 2 (r − rh )
1.35
+1 rh2 . (4.9)
tan γ2 rwm rh
Once both angles are calculated, the blade angle distribution γ(y) is resolved in an
iterative process based on the parabolic distribution mentioned earlier. The iteration
is done with a script in Matlab and stops when chord length constraints are matched
and corrected for abrupt changes in blade angle distribution [21].
Another parameter to consider is the camber angle. The shape obtained by the
previous process can be considered as the camber line of the profile. By adding a
thickness distribution, the typical airfoil shapes are obtained. As stated in Pascu
[21], for low pressure axial fans, such as the fan for cooling purposes, this variable
thickness is unnecessary since both profiles, i.e. cambered plates and airfoils, have
identical performances when operating at low pressures. So, for this method, the
blades have constant thickness across the chord direction. Therefore, the camber
angles are the same as the blade angles in this case. In order to test this statement,
design method 2 considers airfoil cross sections.

32
4. DESIGN PROCEDURE

Last, the diffusion factor DF is calculated with Equation (3.9). From the velocity
triangle in Figure 3.1, the velocity distribution w(y) is defined by
wm
w(y) = . (4.10)
sin γ(y)

By trying to keep the diffusion factor below 0.6 along the blade, some of the losses
will be minimized. This can be done by adjusting γ2 to avoid rapid changes from
one section to another [21].

4.2 Design Method 2


The design process explained in this section is based on the second design method
explained in Chapter 3 and the formulation described by William W. Peng [9],
chapter 5.
After selecting suitable input data including the total pressure difference ∆p, the
design airflow rate Qd , and the impeller rotational velocity ω in rad/s, it is possible
to calculate the non-dimensional specific speed ωs by using

ω Qd
ωs = , (4.11)
(gH)0.75

where g is the acceleration due to gravity and H is the head rise. In addition, the
multiplication of head rise and gravitational acceleration can be reformulated as
∆p
gH = . (4.12)
ρair
The next step is to determine the non-dimensional specific diameter ∆s correspond-
ing to ωs using the Cordier diagram illustrated in Figure 4.4. Therefore, the fan tip
diameter Dt can be found as √
∆s Qd
Dt = . (4.13)
(gH)0.25
The specific speed calculated using Equation (4.3) can be used for the purpose of
determining the hub-to-tip ratio ν. As shown in Figure 4.2, the desired hub-tip ratio
can be chosen based on designer’s preference from the specified domain. Hence, the
fan hub diameter Dh can be calculated by considering the hub-to-tip ratio and tip
diameter,
Dh = νDt . (4.14)
The blade solidity calculated using Equation (3.10) is required to fall within the
domain of 0.4−1.1 for the purpose of preventing stall in axial-flow fans. In case
the solidity does not fulfill the requirement mentioned above, another hub-to-tip
ratio should be selected from the range of 0.3−0.7 based on the calculated specific
speed Ns [9]. Before starting the implementation of the blade design approaches,
it is necessary to determine the number of blades z to be used in Equation (4.4).
Since the number of blades has an influence on both the fan efficiency and the
noise pollution generated by the impeller, it is important to keep the number of

33
4. DESIGN PROCEDURE

blades as low as possible by choosing an appropriate hub-to-tip ratio from Figure


4.2. The main reason is that a suitable number of blades not only enhances the
fan performance, but also diminishes the noise created by the fan [33]. Another
geometry parameter which should be considered at this stage is the blade span
length b. The blade height can be calculated by taking the difference between radii
of the tip rt and the hub rh .

Figure 4.4: Cordier diagram based on non-dimensional specific speed and diameter
[9].

An essential first step in determining the blade chord length c is to calculate the
flow coefficient φ as below:
Va
φ= , (4.15)
U
where Va and U are axial and tangential velocities, respectively. These velocity
components can be defined as

4Q
Va = , (4.16)
π (Dt2− Dh2 )

U = ωr, (4.17)
where r represents the blade radial position. The axial velocity component is identi-
cal to the absolute inlet velocity V1 and remains constant across the fan blades. This
is due to the fact that the two-dimensional flow is assumed to be incompressible,
meaning that the Mach number is lower than 0.3 [28]. The next step is to estimate

34
4. DESIGN PROCEDURE

 
the value of φ sc corresponding to the specific speed from Figure 4.3. Thereafter,
the pitch distance s calculated using Equation (4.6) is employed in order to deter-
mine the chord length of the blade at each radial position.

Figure 4.5: Inlet and outlet velocity triangles for axial fans [9].

The local blade solidity needs to be calculated using Equation (3.10) at each radial
position in order to decide which design approach is more suitable for the analysis.
An iterative process shall be used at this stage with the purpose of checking the
accuracy of the model with respect to the hydraulic efficiency ηhyd , stagger angle
γ, and lift coefficient CL . Hydraulic efficiency, also known as stage efficiency, needs
to be initially defined since the outlet flow angles and velocities are dependent of
this parameter. After selecting an appropriate initial value for hydraulic efficiency,
the next step is to make use of the velocity diagram illustrated in Figure 4.5 and
calculate the flow angles and velocities. The mean flow angle βm can be calculated
using Equation (3.13), where the inlet and outlet flow angles are determined as
follows:
U
 
β1 = tan −1
, (4.18)
Va
U − Vu2
 
β2 = tan −1
. (4.19)
Va
The tangential component of absolute velocity at the outlet Vu2 highly depends on
hydraulic efficiency and the tip pressure head Ht of the fan, and can be calculated
using
gHt
Vu2 = . (4.20)
ηhyd U
As shown in Figure 4.5, the tangential component of relative velocity at the outlet
Wu2 can be calculated by taking the difference between the tangential velocity and
the tangential component of absolute velocity at the fan outlet. Moreover, the
remaining velocity components such as outlet absolute velocity V2 , inlet relative

35
4. DESIGN PROCEDURE

velocity W1 , and outlet relative velocity W2 can easily be calculated by using the
Pythagorean theorem as follows:
 0.5
V2 = Vu2
2
+ Va2 , (4.21)
 0.5
W1 = U 2 + Va2 , (4.22)
 0.5
W2 = Wu2
2
+ Va2 . (4.23)
Once the blade solidity and the flow angles are determined, it is possible to com-
pute the corresponding lift coefficient using Equation (3.14). The next step is to
determine the Reynolds number at each radial position as
Wm c
Re = , (4.24)
νair
where c is the local chord length and νair is the kinematic viscosity of the air. In
addition, the Reynolds number is a function of mean relative velocity Wm which can
be found using the velocity triangles in Figure 3.6. Hence, Wm becomes
Va
Wm = . (4.25)
cos βm
The next step is to determine the stagger angle based on either angle of attack or
incidence angle, which meets the needs of required lift coefficient. The required
data needed to calculate the stagger angle can be gathered from suitable airfoil or
cascade test data. If the blade solidity is greater than unity, it is necessary to use
the cascade approach based on appropriate test data, such as those shown in Figures
A.1-A.4. On the other hand, the modified isolated approach or isolated technique
can be employed with respect to the solidity condition. In case of low or intermediate
solidities, it is required to use aerodynamic data from one of the airfoils listed in
Table A.1. This table consists of circular-arc and symmetrical airfoils which their
shape and aerodynamic characteristics for various Reynolds numbers are gathered
using XFLR5 software. In order to provide the aerodynamic characteristics, a set of
polars corresponding to each airfoil is generated using Type 1 analysis in XFLR5.
Furthermore high angles of attack are investigated with the purpose of recognizing
pre-stall and stall conditions.
Since an accurate blade design can enhance the fan performance and mitigate the
noise generated by the fan, it is important to select an airfoil which not only fulfills
the requirements for lift but also has a high lift-to-drag ratio CL /Cd [33]. Stall is
an important phenomenon which should be treated when designing a fan impeller.
In order to prevent stall in an axial-flow fan, it is recommended that lift coefficient
shall vary from 0.6 at the tip to 1.3 at the hub when designing a blade using isolated
airfoil approach [33]. In addition, it is necessary to select an airfoil which provides
the required lift without being affected by the stall [33].
After selecting the required aerodynamic data, the corresponding angle of attack
or incidence angle can be employed to calculate the stagger angle using Equation
(3.16) or Equation (3.18). Finally, the hydraulic efficiency can be determined by
1 − Rr − φδr
" #
Rr − φδr
ηhyd =φ + , (4.26)
φ + δr Rr φ + δr (1 − Rr )

36
4. DESIGN PROCEDURE

where Rr is the local reaction for the rotor blade and δr is the drag-to-lift ratio
corresponding to the chosen airfoil which can be computed with

Rr = φ tan βm , (4.27)

Cd
δr = . (4.28)
CL
Moreover, contribution of additional pressure losses to the total drag shall be consid-
ered by using Equation (3.23) for the purpose of predicting the hydraulic efficiency.
The initial value of the hydraulic efficiency should be compared with the value esti-
mated from Equation (4.26) in order to check the accuracy of the model. The design
process needs to be repeated until the desired convergence tolerance is reached. This
procedure runs for all chosen radial positions on the blade in order to select appropri-
ate airfoil profiles which fulfill the requirements for lift, stagger angle and hydraulic
efficiency. The design process steps mentioned above are summarized in the flow
diagram shown in Appendix B.2.

37
5
RESULTS

5.1 Design Method 1


The following results, describe the design procedure used for model 1. The dimen-
sions for the other models are summarized in Table 5.2.

5.1.1 Impeller dimensions


The first step is to define the fan performance requirements. For this method, the
models’ requirements are based on data provided by Scania.

Table 5.1: Fan performance requirements.

N (rpm) Qd (m3 /s) ∆pt−s (Pa)


2000 0.833 180

As mentioned before, the pressure difference required for the Cordier diagram is the
total pressure difference, while the pressure rise specified in Table 5.1 is in form of
total to static. The approximation of the total pressure difference is done with a
script in Matlab and results in a value of 238 Pa. Therefore, the specific head Y
and fan rotational speed n are 194.28 Pa/(kg/m3 ) and 33.33 rps, respectively. Using
Equation (4.1):

√ Qd
σ = 2n π = 1.232.
(2Y )0.75
Then, from Figure 4.1, the diameter number is δ = 1.586. Next, by solving Equation
(4.2), the diameter of the fan Dt is obtained as
!−0.25
2δ 2Y
Dt = √ = 368 mm.
π Q2d

After obtaining the diameter, the specific speed Ns is calculated with Equation (4.3)
to find the hub-to-tip ratio ν.

N Qd
Ns = = 86895
∆p0.75

38
5. RESULTS

From Figure 4.2 and taking the lowest dashed line, the hub-to-tip ratio is ν = 0.44.
Therefore, the hub diameter is Dh = 0.16 m. Finally, according to Equation (4.4),
the fan must have 5 blades.

5.1.2 Blade characteristics


To calculate the chord length, the parameters required are the flow coefficient φ, the
tangential velocity u, the pitch distance s, and relative velocity wm . From Figure
4.3 and specific speed Ns , the value obtained is φ(s/c) = 0.61. Applying Equations
(3.3), (4.5), (4.6), and (4.7):
Qd
wm = = 9.75 m/s,
π(rt − rh2 )
2

s wm 2πr wm 2πr
c=φ = = = 96 mm.
0.61 u (0.61)z 2πrn (0.61)z
After finding the fan dimensions, the shape of the blade section can be found. The
mathematical routine is depicted in Appendix B.1. The blade is separated into 20
equidistant sections from the hub radius rh to the tip radius rt . The blade cross
section shape and blade angle distribution for two different radial positions are shown
in Figures 5.1 and 5.2, respectively.

40

30

20

10

0
0 10 20 30 40 50 60 70 80 90

Figure 5.1: Blade cross section shape at two radial positions.

40

35

30

25

20

15

0 10 20 30 40

Figure 5.2: Blade angle distribution at two radial positions.

39
5. RESULTS

The distribution of γ2 was slightly corrected to decrease the diffusion factor shown
in Figure 5.3. As stated before, this is done to reduce the total pressure loss due to
diffusion in the boundary layer.

0.8

0.6

0.4

0.2

0
80 100 120 140 160 180

Figure 5.3: Diffusion Factor along the blade.

5.1.3 Fan dimensions


In order to test how particular factors affect the performance, several models are
compared. In Table 5.2 the main characteristics of 6 models are listed. Model 1 and
2 have straight blades, but with different number of blades. Model 3 and 4 have the
same number of blades, but the chord length distribution varies. As for model 5,
it is compared with model 6, which is designed with method 2. Both models 5 and
6 have the same dimensions, the only difference between them is the cross-section
shape of the blades. As stated in Section 2.2.4, one method to reduce the fan’s
sound pressure level is by an uneven blade spacing [15]. According to Zhai et al.
[34], the optimum circumferential angles of a 7-blade fan are 53.5°, 103.5°, 149.5°,
210.5°, 256.5°, 306.5°, and 360°. They found that this configuration can reduce the
sound pressure level by 2.96 dBs with a slight compromise in performance. Model
11 has the same dimensions and blade characteristics as model 2 with the exception
that the blade spacing is uneven with the previous circumferential angles.

Table 5.2: Models’ characteristics based on method 1.

Model 1 Model 2 Model 3 Model 4 Model 5 Model 11


Dt (mm) 368 368 368 368 374 368
ν 0.44 0.44 0.44 0.44 0.51 0.44
c (mm) 96 68 * ** 82 68
z 5 7 5 5 7 7

* The chord distribution varies linearly from -10 to 10 extra units to the base chord
length (model 1) from root to tip.

40
5. RESULTS

** The chord distribution varies linearly from -20 to 20 extra units to the base chord
length (model 1) from root to tip.

5.2 Design Method 2


Fan model 9, which is designed based on the second development procedure described
in chapter 4, will be analyzed in detail, and the characteristics corresponding to other
models will be presented in this section.

5.2.1 Impeller dimensions


As presented in Table 5.3, the design flow rate, static pressure difference, and ro-
tational speed are considered as the required input parameters needed to start the
process. Design parameters shown in Table 5.3 are chosen to be similar to the
operating point data corresponding to a commercially available fan.

Table 5.3: Performance input parameters.

N (rpm) Qd (m3 /s) ∆pt−s (Pa)


3000 1.11 400

Since the pressure rise used in all calculations is in the form of total pressure, an
iterative process is employed to estimate the total pressure difference. Hence, the
corresponding value of total pressure rise is approximated to 523 Pa. Thereafter,
the rotational velocity is converted to 314 rad/s and used in Equation (4.11) to
determine the non-dimensional specific speed as

ω Qd
ωs = = 3.52,
(∆p/ρair )0.75
where the air density is ρair = 1.225 kg/m3 at 15 ◦C. Hence, the non-dimensional
specific diameter ∆s is obtained from Figure 4.4, which is equal to 1.55. It is now
possible to calculate the fan tip diameter using Equation (4.13).

∆s Qd
Dt = = 360 mm
(∆p/ρair )0.25
The specific speed Ns is computed with Equation (4.3), and has a value of 83365.
Next, the hub-to-tip ratio ν is obtained from Figure 4.2. Since the upper and lower
limits are defined in Figure 4.2, it is possible to choose different hub-to-tip values
which fall within the corresponding domain. The hub-to-tip value which is selected
for this model is ν = 0.48. Hence, the hub diameter is Dh = νDt = 172 mm, and
the corresponding number of blades becomes five when applying Equation (4.4).
The next step after determining the diameters is to calculate the blade span length
b by taking the difference between tip and hub radii which results in a value of 94
mm.

41
5. RESULTS

5.2.2 Blade characteristics


The first step in designing the blade is to determine the axial velocity Va as follows:
4Qd
Va = = 14.16 m/s.
π (Dt2 − Dh2 )

Next, the blade is divided into five identical elements, meaning that the blade should
be analyzed at six independent radial positions from hub to tip. The following steps
present how the blade is modeled at the tip, i.e. r = rt . Then, the overall results
will be provided in Table 5.4 at the end of this section.
In order to select an appropriate design method, the local solidity should be calcu-
lated based on other parameters such as local chord length and pitch distance as
shown below.
• Tangential velocity:
U = ωr = 56.52 m/s.
• Flow coefficient:
Va
φ= = 0.25.
U
• Pitch distance:
2πr
s= = 205.9 mm.
z
• Chord length:
s
 
φ = 0.59 ⇒ c = 87.8 mm.
c
• Solidity:
c
= 0.43.
σ=
s
Given the value of blade solidity, the isolated airfoil technique becomes the most
suitable alternative for designing the blade at the tip. After selecting the best
approach, an iterative process with a convergence tolerance  = 0.001 is employed.
As an initial measure, a value of 0.95 is assigned to ηhyd . Then, the required lift
coefficient can be calculated as follows.
• Tangential component of outlet absolute velocity:
gHt
Vu2 = = 8.28 m/s.
ηhyd U

• Inlet flow angle:


U
 
β1 = tan−1 = 75.9°.
Va
• Outlet flow angle:
U − Vu2
 
β2 = tan −1
= 73.6°.
Va
• Mean flow angle:
1
tan βm = (tan β1 + tan β2 ) ⇒ βm = 74.9°.
2
42
5. RESULTS

• Lift coefficient:
s
 
CL = 2 (tan β1 − tan β2 ) cos βm = 0.72.
c
The next step is to compute the tip Reynolds number with respect to the mean
relative velocity.
• Mean relative velocity:

Va
Wm = = 54.26 m/s.
cos βm

• Reynolds number:
Wm c
Re = = 3.24 · 105 ,
νair
where the kinematic viscosity of air is νair = 1.47 · 10−5 m2 /s at 15 ◦C. Since the
isolated airfoil data are only gathered for Reynolds numbers from 5·104 to 5·105
with a step size of 5·104 , an interpolation technique with high accuracy is used to
select the most reasonable aerodynamic data corresponding to each airfoil profile at
different Reynolds numbers. Furthermore, the transition criterion number Ncrit is
set to nine, meaning that the airfoils are analyzed in a wind tunnel under average
disturbance condition [35].
According to the proposed method presented in the previous chapter, NACA 4412
is chosen as the best airfoil profile which not only fulfills the requirements for the
lift but also provides the highest lift-to-drag ratio among all the airfoils in the first
iteration.

1.5 100

80
1

60
0.5
40
0
20

-0.5
0

-1 -20
-10 -5 0 5 10 15 -10 -5 0 5 10 15

(a) CL vs. α plot (b) L/D vs. α plot

Figure 5.4: Aerodynamic characteristics of NACA 4412 at Re = 3.24 · 105 .

The performance of NACA 4412 airfoil shown in Figure 5.4 provides the aerodynamic
information required in order to calculate stagger angle and hydraulic efficiency in
the first iteration. According to the plots illustrated in Figure 5.4, the angle of
attack of the blade tip section becomes α = 2.2° when the lift coefficient is 0.72. In
addition, the profile drag coefficient Cdp has a value of 0.009 since the lift-to-drag

43
5. RESULTS

ratio is 78.7 at α = 2.2°. After collecting the aerodynamic data corresponding to


the best performing airfoil, it is necessary to calculate the stagger angle by using

γ = βm − α = 72.7°.

Moreover, the effect of pressure loss components should be considered with the pur-
pose of estimating the fan performance. Hence, the total drag coefficient becomes:

Cd = Cdp + Cds + Cdt + Cda = 0.066,

where tip clearance CT is set to 2 mm. It is notable that the annulus loss associated
with the annular surfaces has the highest influence on the total drag. Finally, the
hydraulic efficiency can be determined as follows.
• Drag-to-lift ratio:
Cd
δr = = 0.092.
CL
• Reaction:
Rr = φ tan βm = 0.93.
• Hydraulic efficiency:

1 − Rr − φδr
" #
Rr − φδr
ηhyd =φ + = 0.72.
φ + δr Rr φ + δr (1 − Rr )

Since the preliminary and calculated values of the hydraulic efficiency do not match
the convergence criteria in the first iteration, it is necessary to use the calculated
value of the hydraulic efficiency as a new input for the next iteration. The iterative
process must continue until all the requirements are satisfied. After 12 iterations,
the hydraulic efficiency and the stagger angle converge, and therefore the process
is accomplished. Since the lift coefficient determined in the last iteration is 0.92,
NACA 4412 is no longer the best performing airfoil. It is notable that NACA 4412
is replaced with NACA 6409 airfoil in order fulfill the requirements for the hydraulic
efficiency and the stagger angle at the blade tip position, i.e. r = rt .

1.5 120

100
1
80

60
0.5
40

20
0
0

-0.5 -20
-10 -5 0 5 10 15 -10 -5 0 5 10 15

(a) CL vs. α plot (b) L/D vs. α plot

Figure 5.5: Aerodynamic characteristics of NACA 6409 at Re = 3.18 · 105 .

44
5. RESULTS

As shown in Figure 5.5, the lift coefficient of 0.92 is achieved at an angle of attack
of α = 2.1, meaning that the corresponding lift-to-drag ratio has a value of 98.7. By
evaluating the aerodynamic data corresponding to NACA 4412 and NACA 6409, it
is possible to conclude that the latter must be employed. The design procedure is
implemented for all selected radial positions from the hub rh to the tip rt with a
step size of 18.8 mm, and the outcomes are presented in Table 5.4.

Table 5.4: Main design parameters.

c s σ βm α γ ηhyd
Radius (mm) (mm) (-) (deg) (deg) (deg) (%) Airfoil
rh 87.8 98.4 0.89 50.7 11.1 39.6 85 S 1223
r2 87.8 119.9 0.73 60.6 7.1 53.5 86 NACA 6409
r3 87.8 141.4 0.62 66.2 5.3 60.9 84 NACA 6409
r4 87.8 162.9 0.54 69.9 3.8 66.1 82 NACA 6409
r5 87.8 184.4 0.48 72.6 2.8 69.8 79 NACA 6409
rt 87.8 205.9 0.43 74.6 2.1 72.5 76 NACA 6409

The efficiency values listed in Table 5.4 must be integrated from hub to tip by taking
the area average in order to provide the total-to-static efficiency of the impeller
ηt−s , which will be presented in Section 5.3.1. As explained in Section 4.2, the same
procedure is applied on each radial position along the blade span, which leads to the
selection of the airfoil profiles shown in Table 5.4. As illustrated in Figure 5.6, S 1223
airfoil profile is used to design the blade at the hub and NACA 6409 is employed
for forming the blade at other radial positions. Furthermore, it can be noticed that
the blade stagger angle increases from 39.6° at the hub to 72.5° at the tip.

Figure 5.6: Two-dimensional blade sections.

45
5. RESULTS

5.2.3 Fan dimensions


The performance of a fan is influenced by a variety of elements, such as impeller
dimensions and blade shape. Eleven different models are analyzed to identify the
factors that improve the performance of a fan. Five out of eleven models presented
in this work are designed using airfoil-shaped blades. According to the data provided
in Table 5.5, models 6, 7 and 8 are identical except for the shape of their blades.
Model 6 has straight blades while models 7 and 8 are equipped with two different
types of forward-swept blades. Finally, model 9 has the same tip diameter as model
10, however, other characteristics, including the blade configuration, are different.
It should be noted that Models 6, 7, and 8 are designed based on requirements
illustrated in Table 5.1, while data from Table 5.3 are used to develop fan models 9
and 10.

Table 5.5: Models’ characteristics based on method 2.

Model 6 Model 7 Model 8 Model 9 Model 10


Dt (mm) 374 374 374 360 360
ν 0.51 0.51 0.51 0.48 0.54
c (mm) 82 82 82 88 77
z 7 7 7 5 7

5.3 Fan Performance


5.3.1 Theoretical
In order to determine the theoretical pressure rise at the design point, Equation
(3.25) is used. In addition, the performance data including shut-off and free delivery
points, and ideal pressure difference curve can be estimated for all fan models with
Equations (3.27), (3.29), and (3.30), respectively. The total-to-static efficiency can
be calculated using Equation (3.33). In Figure 5.7, the performance graphs for 3
models are shown according to Equation (3.30). Furthermore, the performance data
corresponding to all fan models are presented in Table 5.6.

Figure 5.7: Theoretical performance graph.

46
5. RESULTS

Table 5.6: Fan models’ theoretical performance.

Model 1 Model 2 Model 3 Model 4 Model 5


∆pt−s,axial (Pa) 360.3 360.3 360.3 360.3 416
∆pt−s,max,axial (Pa) 543.9 543.9 543.9 543.9 592.1
Qt−s,max,axial (m3 /s) 1.44 1.44 1.44 1.44 1.54
ηt−s (%) 60 60 60 60 59
Model 6 Model 7 Model 8 Model 9 Model 10
∆pt−s,axial (Pa) 241.4 241.4 241.4 502.1 515.2
∆pt−s,max,axial (Pa) 592.1 592.1 592.1 1200 1265
Qt−s,max,axial (m3 /s) 1.10 1.10 1.10 1.46 1.44
ηt−s (%) 46 46 46 47 47

5.3.2 CFD
In Table 5.7, partial results are shown. The table includes the results for pressure
difference ∆pt−s , volume flow rate Q, torque required at the shaft τshaf t , and total-
to-static efficiency ηt−s . The efficiency is calculated by rewriting Equation (3.32) as
ηt−s = (Q∆pt−s )/(ωτshaf t ).

Table 5.7: CFD results at/near the peak efficiency point.

Model 1 Model 2 Model 3 Model 4 Model 5


N (rpm) 2000 2000 2000 2000 2000
∆pt−s (Pa) 130 130 127 122 162
Q (m3 /s) 0.75 0.76 0.74 0.74 0.86
τshaf t (Nm) 0.81 0.86 0.76 0.73 1.24
ηt−s (%) 57 55 59 59 54
Model 6 Model 7 Model 8 Model 9 Model 10
N (rpm) 2000 2000 2000 3000 3000
∆pt−s (Pa) 152 143 146 281 286
Q (m3 /s) 0.83 0.79 0.8 1.11 1.09
τshaf t (Nm) 1.07 1.01 1.03 1.66 1.81
ηt−s (%) 56 54 54 60 55
Model 11
N (rpm) 2000
∆pt−s (Pa) 114
Q (m3 /s) 0.71
τshaf t (Nm) 0.76
ηt−s (%) 51

47
5. RESULTS

The complete table of the performance results along with the pressure distribution
figures are presented in Appendix D.
In order to gain a deeper understanding of the aerodynamic performance, models 3
and 7 are analyzed in detail. The reason to choose these models is to compare two
fans with various blade profiles. Moreover, model 3 has one of the highest efficiencies
among the fan models, while model 7 has a lower efficiency compared to the other
models. Furthermore, these fan models have two of the main noise reduction factors,
such as varying chord along the span and forward-swept blades.

300 60
Performance Fan 3
Performance Fan 7
250 Efficiency Fan 3 50
Efficiency Fan 7
System High
200 40
System Mid
System Low
150 30

100 20

50 10

0 0
0 0.2 0.4 0.6 0.8 1

Figure 5.8: Fan performance and efficiency curves at 2000 rpm.

The performance curves illustrated in Figure 5.8 are derived based on the data
presented in Tables D.2 and D.3. Moreover, the efficiency curves are estimated by
assuming that the torque values corresponding to models 3 and 7 in Table 5.7 are
constant in the operating range.

5.4 Noise
In this section, the sound pressure levels Lp of the 11 models are estimated with
the two approaches previously introduced. Then, the sound pressure levels Lp cor-
responding to the models are compared with the values estimated for two Scania’s
fans.

5.4.1 Approach 1
According to approach 1, the noise power level depends on the volume flow rate and
pressure difference at/near the peak efficiency. With the values of model 1 from
Table 5.7 and Equation (3.34):

LW = 57.3 + log10 (Qd ) + 20 log10 (∆p) − 10 log10 (η) = 81.87 dB.

48
5. RESULTS

Assuming a distance of 1 m and a directivity factor of Qθ = 2, i.e. the fan standing


on a reflective floor, the sound pressure level can be obtained with Equation (2.3):


 
Lp = LW + 10 log10 = 73.89 dBA.
4πr2

The sound pressure level results for the fan models are provided in Table 5.8.

Table 5.8: Sound pressure level at 2000 rpm using approach 1.

Model 1 Model 2 Model 3 Model 4 Model 5


Lp (dBA) 73.89 74.08 73.52 73.22 76.15
Model 6 Model 7 Model 8 Model 9 Model 10
Lp (dBA) 75.38 75.051 75.16 73.42 73.98
Model 11
Lp (dBA) 73.26

5.4.2 Approach 2
From Table 3.1, the specific sound power levels for this fan correspond to the tubeax-
ial fans values. For model 1, the blade pass frequency is calculated to BPF=166,
therefore 7 extra units are added to the frequency band of 125 Hz. Then, by adding
10 log10 Q + 20 log10 ∆p to each specific sound power H, the spectrum is obtained.
Next, the values must be transformed to SPL and weighted for A scale.

Frequencies (Hz) 63 125 250 500 1000 2000 4000 8000


H 93 92 94 98 97 96 88 85
BPF addition 93 99 94 98 97 96 88 85
LW (dB) 74.03 80.03 75.03 79.03 78.03 77.03 69.03 66.03
Lp (dB) 66.05 72.05 67.05 71.05 70.05 69.05 61.05 58.05
A scale (dBA) 39.84 55.94 58.44 67.84 70.04 70.24 62.04 56.94

To obtain the overall sound pressure level, Equation (3.36) is used with the values
in the last row.
OASP L = 10 log10 10Lpi /10 = 74.77 dBA
X

The results for the other models are listed in Table 5.9.

49
5. RESULTS

Table 5.9: Overall sound pressure level at 2000 rpm using approach 2.

Model 1 Model 2 Model 3 Model 4 Model 5


OASP L (dBA) 74.77 75.17 74.5 74.15 77.62
Model 6 Model 7 Model 8 Model 9 Model 10
OASP L (dBA) 76.91 76.17 76.40 74.31 74.67
Model 11
OASP L (dBA) 73.73

5.4.3 Scania’s fans comparison


In this sections, the 11 models are compared to two commercially available fans by
using approach 1. The main characteristics of the Scania’s fans are not specified
due to confidentiality reasons.
In order to have a fair comparison, the performance values of the 11 models and
the Scania’s fans are scaled to a pressure difference of ∆p = 183 Pa with the fan
laws listed in Table 2.1. This means that, for the 11 models, the volume flow rate
Q and the fan speed N will increase, since the pressure difference at N = 2000 rpm
is lower than 183. In Table 5.10, the new values for the fans are shown.
Then, assuming that the efficiency ηt−s is the same, the sound pressure level Lp is
calculated with Equations (3.34) and (2.3). The results for the Scania’s fans and
the 11 models are shown in Table 5.11.

Table 5.10: Performance of the fans at ∆p = 183 Pa.

Model 1 Model 2 Model 3 Model 4 Model 5


N (rpm) 2373 2373 2400 2449 2125
∆p (Pa) 183 183 183 183 183
Q (m3 /s) 0.89 0.90 0.89 0.91 0.91
Model 6 Model 7 Model 8 Model 9 Model 10
N (rpm) 2194 2262 2239 2420 2400
∆p (Pa) 183 183 183 183 183
Q (m3 /s) 0.91 0.89 0.90 0.88 0.85
Model 11 Scania’s Fan 1 Scania’s Fan 2
N (rpm) 2534 2000 2745
∆p (Pa) 183 183 183
Q (m3 /s) 0.90 0.72 0.39

50
5. RESULTS

Table 5.11: Sound pressure level at ∆p = 183 Pa using approach 1.

Model 1 Model 2 Model 3 Model 4 Model 5


Lp (dB) 76.94 77.13 76.78 76.83 77.24
Model 6 Model 7 Model 8 Model 9 Model 10
Lp (dB) 77.04 77.23 77.17 76.74 77.12
Model 11 Scania’s Fan 1 Scania’s Fan 2
Lp (dB) 77.48 78.3 77.63

51
6
DISCUSSION

By evaluating the values of the fan models in Table 5.6, it is possible to notice
the limitations of the theoretical equations. Since the equations to estimate the
performance are only dependent on the outlet blade angle and fan dimensions, some
of them have the same results, even though the number of blades and/or chord
distribution are different. Another aspect to note is the efficiency values. In models
1 to 5 and model 11, the losses are not considered, meaning that the efficiency is
the ideal efficiency, overestimating performance. But even if these equations do
not provide exact results, they provide a good estimation on the magnitude of the
performance. For these reasons, a more complex method is required, such as CFD
analysis.
From Table D.1, it is possible to notice how different factors affect the performance
of the fans. By comparing models 1 and 2, the effect of number of blades can
be observed. Model 1 has 5 blades, while model 2 has 7 blades. At low rotational
speeds, i.e. 1000 and 2000 rpm, it seems that the performance is not greatly affected,
but the torque required is higher for model 2, decreasing the efficiency slightly. But
at a rotational speed of 3000 rpm, even though the performance is almost the same,
the torque required for model 2 is lower, increasing the efficiency. In general, and
according to previous studies, a higher number of blades increased the performance
of the fan, but reduced the efficiency. On the other hand, model 11 has the same
geometry as model 2, but with uneven blade spacing. As stated before, this can
reduce the sound pressure level by 2.96 dBs, but given that the approaches used
to estimate the sound pressure level don’t take into account these types of factors,
it is not possible to observe their influence. At the same time, the performance is
compromised, reducing both the volume flow rate and efficiency.
Another efficient technique to reduce the noise is to design a blade with a larger
tip chord. Both models 3 and 4 have a larger tip chord and a shorter root chord
compared to model 1. This can be observed in the increase in efficiency. While the
performance characteristics are not greatly affected, the torque required for models
3 and 4 is considerably lower than the torque required for model 1, especially at
high rotational speeds.
As mentioned previously, models 5 and 6 are equipped with straight blades and have
larger tip diameters compared to some of the other models. These two models share
an identical number of blades and overall geometry such as same tip diameter and
hub-to-tip ratio, but they differ in the shape of the blades. According to the data
shown in Table D.1, model 5 is able to generate high volume flow rates at higher
pressure rises relative to other fan models, due to its large diameter. In addition,
model 6, which is equipped with airfoil-shaped blades, requires a lower torque, which

52
6. DISCUSSION

in turn enhances the efficiency in comparison with model 5, while they both provide
nearly the same volume flow rate. A possible reason to explain the lower required
torque in model 6 relative to model 5 can be related to less pressure losses generated
by airfoil-shaped blades at higher hub-to-tip ratios.
Fan models 7 and 8 are the modified versions of model 6. These two models are
designed with forward-swept blades with the purpose of reducing the noise. It is
notable that the models have almost the same performance level. Since model 7 is
equipped with forward-curved blades at higher sweep angle relative to model 8, its
performance features are being negatively affected.
In order to analyze the effect of the tip diameter on the performance of a fan,
model 9 and 10 with smaller diameters relative to the other models are examined.
Model 9 has straight and airfoil-shaped blades which can be compared with model
3 since they both have similar performance characteristics. Although fan model 3
is developed based on the first blade design technique and has significantly different
size and dimension, its performance features are identical to model 9. They are not
only the most efficient fans among all the models but also generate almost the same
volume airflow at comparable pressure rise. This is due to various reasons, such as
chord length distribution along the blade span, blade shape, and hub-to-tip ratio.
On the other hand, performance of model 10 is compromised due to the usage of a
high hub-to-tip ratio and forward-curved blades compared to model 9.
Based on all the results in Table 5.7, it is noticeable that there is no clear distinction
between cambered plates and airfoils at low pressure differences, since all of the
models have relative high efficiencies. An experimental test could provide a deeper
insight into the differences.
As illustrated in Figure 5.8, three fictive system curves are used to estimate the
operating point at different system resistances, i.e. high, medium, and low, for fan
models 3 and 7. The results indicate that model 3 provides lower volume flow rate
at relatively lower pressure difference relative to model 7. On the other hand, it
operates at higher efficiencies than model 7, which matches the predicted results
shown in Table D.1.
The noise level estimations are solely done with empirical equations, for this reason
it is not possible to conclude how some factors affect the noise level. On the one
hand, it can be argued that the noise level is reduced by increasing the efficiency of
the fan, this can be observed by comparing model 1 with model 3. On the other
hand, the uneven blade spacing affects directly the efficiency, this is a well-known
design feature to reduce the sound pressure level. Therefore, increasing the efficiency
does not necessarily corresponds to lower noise levels. This was also the conclusion
in the investigation done by Robert C. Mellin [36]. In his work, he also concluded
that increasing the diameter and decreasing the speed will not necessarily reduce
the noise level for a given aerodynamic performance, as shown in the fan sound law
in Equation 2.5.
As stated earlier, the performance values are adjusted in order to have a reasonable
comparison in sound pressure levels between Scania’s and the developed fan models.
As illustrated in Table 5.11, the sound pressure levels of the 11 models are approx-
imately the same as the levels corresponding to the Scania’s fans. Nevertheless, all
the designed models have less number of blades compared to the Scania’s fan 1,

53
6. DISCUSSION

which in turn increases the sound quality. Additionally, the 11 fan models have a
smaller tip diameter relative to the Scania’s fan 1. This is beneficial when there is
a space limitation in the cooling system installation.

54
7
CONCLUSION

Throughout this study, it has been found that several factors influence the perfor-
mance and noise pollution of the electric axial fans. Based on the results discussed
and the studies cited in Section 2.2.4, it can be concluded:
• A larger chord length at the tip relative to the root of the impeller blades
enhances the efficiency and reduces the noise emitted by the fan.
• Irregular spacing between the rotor blades has a negative impact on the fan
performance and efficiency. However, it can significantly reduce the sound
pressure level.
• A forward-swept and curved configuration decreases the efficiency, but also
has a positive effect on mitigating the noise.
• A larger tip diameter and a higher number of blades increment the torque
required which in turn increases the energy consumption of the electric motor
used to run the fan.
• An airfoil-shaped impeller presents similar performance as a cambered-plate
rotor in low pressure differences.
• A small tip clearance leads to a less leakage at the blade tip which in turn
enhances the efficiency and reduces the sound pressure level.

7.1 Future Work


In this project, 11 fan models are developed by considering the noise reduction meth-
ods introduced. There are a number of recommendations that can be implemented
to improve the precision of this investigation.
Due to time constraints, the only noise control technique which is not applied on
the fan models is the trailing edge serrations. This method is considered to be a
robust noise reduction approach since it reduces the vortex-shedding at the trailing
edge of the blades. It is recommended to investigate the influence of this factor on
the noise and the performance in the future.
In addition, more accurate noise estimation methods or software can be used in
order to provide a better understanding of noise behavior. It is required to take into
account the shape of the blades in the noise calculation.
Finally, by combining the relevant factors and creating a final model, it is possible
to build a physical prototype, which can be experimentally tested and evaluated.

55
Bibliography

[1] A. Barber, Pneumatic Handbook. Oxford, UK: Elsevier Science, 1997.


[2] W.T.W. Cory, “Short history of mechanical fans and the measurement of their
noise,” in Fan Noise: An International INCE Symposium (A. Guedel, ed.),
(Senlis, France), pp. 3–44, SFA, CETIM, 9 1992.
[3] W.T.W. Cory, Fans and ventilation: a practical guide. Amsterdam, The
Netherlands: Elsevier Science, 2005.
[4] Scania CV AB, “Electrification.” https://www.scania.com/group/en/home/
about-scania/innovation/technology/electrification.html, 2020. On-
line, accessed 2020-05-21.
[5] American Society of Heating, Refrigerating and Air-Conditioning Engineers,
2008 Ashrae Handbook: HVAC Systems and Equipment. Atlanta, GA, USA:
ASHRAE, 2008.
[6] S. Pelonis, “Axial vs. centrifugal fans.” https://www.pelonistechnologies.
com/blog/axial-vs.-centrifugal-fans, 11 2015. Online, accessed 2020-03-
25.
[7] J. Yu, T. Zhang, and J. Qian, Electrical Motor Products. Cambridge, UK:
Woodhead Publishing, 2011.
[8] AMCA, ASHRAE, “Laboratory methods of testing fans for aerodynamic per-
formance rating.” https://law.resource.org/pub/us/cfr/ibr/001/amca.
210.1999.pdf, 10 1999. Online, accessed 2020-03-26.
[9] W.W. Peng, Fundamentals of Turbomachinery. Hoboken, NJ, USA: John Wiley
& Sons,Inc, 2008.
[10] T. Wright, Fluid Machinery: Performance, Analysis and Design. Boca Raton,
FL, USA: CRC Press, 1999.
[11] Twin City Fan Companies, Ltd., “Understanding fan
curves.” https://www.tcf.com/wp-content/uploads/2018/06/
Understanding-Fan-Curves-FE-2000.pdf, 2018. Online, accessed 2020-
03-26.
[12] M. J. Crocker, Handbook of Noise and Vibration Control. Hoboken, NJ, USA:
John Wiley & Sons, Inc., 2008.
[13] J. Lalanne, H. Bodén, and M. Abom, “Aeroacoustic response of a six-bladed ax-
ial fan to installation effects,” The Journal of the Acoustical Society of America,
vol. 105, no. 2, pp. 945–945, Feb. 1999.
[14] P.-O. Berglund, Investigation of acoustic source characterisation and installa-
tion effects for small axial fans. PhD thesis, Department of Vehicle Engineering,
KTH Royal Institute of Technology, Stockholm, Sweden, 2003.

56
Bibliography

[15] W. Neise, “Review of fan noise generation mechanisms and control methods,”
in Fan Noise: An International INCE Symposium (A. Guedel, ed.), (Senlis,
France), pp. 45–56, SFA, CETIM, 9 1992.
[16] S. Castegnaro and A. Lazzaretto and M. Masi, “Effectiveness of blade forward
sweep in a small industrial tube-axial fan,” in FAN 2018 International Confer-
ence on Fan Noise, Aerodynamics, Applications and and Systems, (Darmstadt,
Germany), CETIAT, AMCA, 4 2018.
[17] Y. Kumon and M. Ohtsuka, “Development of electric fan propeller featuring
chestnut tiger butterfly wing characteristics,” in Fifth International Symposium
on Aero Aqua Bio-Mechanisms, vol. 3, (Tokyo, Japan), pp. 103–108, J-STAGE,
6 2013.
[18] D. J. Moreau and L. A. Brooks and C. J. Doolan, “On the noise reduction
mechanism of a flat plate serrated trailing edge at low-to-moderate reynolds
number,” in 18th AIAA/CEAS Aeroacoustics Conference (33rd AIAA Aeroa-
coustics Conference), (Colorado Springs, CO, USA), pp. 1–20, AIAA, 6 2012.
[19] G. Ingram, “Wind turbine blade analysis using the blade element momentum
method,” 2011. Durham University.
[20] P.M. Sforza, Theory of Aerospace Propulsion. Woburn, MA, USA: Elsevier
Science, 2016.
[21] M. Pascu, Modern Layout and Design Strategy for Axial Fans. PhD thesis,
Institute of Fluid Mechanics LSTM Erlangen-Nuremberg University, Erlangen,
Germany, 2008.
[22] M. Pascu, M. Miclea, P. Epple, A. Delgado, and F. Durst, “Analytical and nu-
merical investigation of the optimum pressure distribution along a low-pressure
axial fan blade,” Proceedings of the Institution of Mechanical Engineers, Part
C: Journal of Mechanical Engineering Science, vol. 223, no. 3, pp. 643–657,
2009.
[23] S. Lieblein, F. C. Schwenk, and R. L. Broderick, “Diffusion factor for estimating
losses and limiting blade loadings in axial-flow compressor blades,” research
memo, Lewis Research Center NASA, 1953.
[24] D.R. Kirk, “Air-breathing engines: Overview of axial compressors.” https:
//www.slideserve.com/bernad/mae-4261-air-breathing-engines. PPT
Presentation, Online, accessed 2020-05-01.
[25] S. Castegnaro, “Fan blade design methods: Cascade versus isolated airfoil ap-
proach—experimental and numerical comparison,” in ASME Turbo Expo 2016:
Turbomachinery Technical Conference and Exposition (American Society of Me-
chanical Engineers, ed.), vol. 1, ASME, 2016.
[26] S.L. Dixon and C.A. Hall, Fluid Mechanics and Thermodynamics of Turboma-
chinery. Oxford, UK: Butterworth-Heinemann Inc, 2014.
[27] E. Andreadis, “Design of a low speed vane-axial fan,” Master’s thesis, Depart-
ment of Power Engineering and Propulsion, Cranfield University, Cranfield,
UK, 2011.
[28] R.A. Wallis, Axial Flow Fans and Ducts. Malabar, FL, USA: Krieger Publishing
Company, 1993.
[29] A.R. Howell, “Fluid dynamics of axial compressors,” Proceedings of the Insti-
tution of Mechanical Engineers, vol. 153, no. 1, pp. 441–452, 1945.

57
Bibliography

[30] J.D. Denton, “Loss Mechanisms in Turbomachines ,” Journal of Turbomachin-


ery, vol. 115, no. 4, pp. 621–656, 1993.
[31] T.Fukano, Y.Kodama, and Y.Takamatsu, “Noise generated by low pressure
axial flow fans, II: Effects of number of blades, chord length and camber of
blade,” Journal of Sound and Vibration, vol. 50, no. 1, pp. 75–88, Jan 1977.
[32] P. Epple, A. Delgado, and F. Durst, “A theoretical derivation of the Cordier
diagram for turbomachines,” Proceedings of the Institution of Mechanical En-
gineers, Part C: Journal of Mechanical Engineering Science, vol. 225, no. 2,
pp. 354–368, 2009.
[33] H.Ö. Keklikoğlu, “Design, construction and performance evaluation of axial
flow fans,” Master’s thesis, Department of Mechanical Engineering, Middle East
Technical University, Ankara, Turkey, 2019.
[34] S. Zhai and Q. Yao, “Design and research of unequal distance fan,” IOP Con-
ference Series: Earth and Environmental Science, vol. 310, p. 032010, 09 2019.
[35] XFLR5, “XFLR5 guidelines: Analysis of foils and wings operating at low
Reynolds numbers.” http://www.xflr5.com/xflr5.htm, 2013. Online, ac-
cessed 2020-05-11.
[36] R. C. Mellin, “Noise and performance of automotive cooling fans,” SAE Trans-
actions, vol. 89, pp. 141–171, 1980.

58
A
Design Method 2 Data

Cascade test data

Figure A.1: Recommended incidence angle i based on camber angle θ [28].

Figure A.2: Stagger angle γ vs. deviation angle coefficient m [28].

I
A. Design Method 2 Data

Figure A.3: Recommended blade solidity σ based on flow coefficient φ [28].

Figure A.4: Lift data corresponding to cambered plate [28].

II
A. Design Method 2 Data

Airfoils

Table A.1: List of airfoils used in the blade design [33].

# Name # Name # Name # Name


1 E 423 2 E 591 3 E 1211 4 MH 112
5 N 0006 6 N 0008 7 N 0009 8 N 0010
9 N 0012 10 N 0015 11 N 0018 12 N 0021
13 N 0024 14 N 1408 15 N 1410 16 N 1412
17 N 2408 18 N 2410 19 N 2411 20 N 2412
21 N 2415 22 N 2418 23 N 2421 24 N 2424
25 N 4412 26 N 4415 27 N 4418 28 N 4421
29 N 4424 30 N 6409 31 N 6412 32 S 1210
MH: Martin Hepperle
N : NACA
33 S 1223 RTL 34 S 1223
E: Eppler
S: Selig

III
B
Flow Diagrams

Figure B.1: Flow diagram for design method 1 [21].

Start

Cascade Parameters
rhub , rtip ,
n, Qdesign , z, l1

Velocity triangles
u = 2πrn
Qdesign
wm =
π(rt2 − rh2 )
wm
wu1 = u(axial entry) → tan γ1 =
wu1

1 2πn 2
"  2 # !
rt
= r − 2 (r − rh )
1.35
+1 rh2
tan γ2r rwm rh

wm
tan γ2 = → wu2
wu2

(γm , ym )

Continued

IV
B. Flow Diagrams

γ(y) = Ay 2 + By + C
y = y1 → γ = γ2
y = ym → γ = γm
y = y2 → γ = γ1

 
γ2
γ(y) =  γi 
 

γ1

∆x = ∆y tan γ(y)

xi = xi−1 + ∆x

q
l2 = (x2 − x1 )2 + (y2 − y1 )2

Change No
l2 = l1
γm , ym

Yes

Cascade Performance
wm
w(y) =
sin γ(y)
max w(y) − w2
DF =
max w(y)

Stop

V
B. Flow Diagrams

Figure B.2: Flow diagram for design method 2.

Start

ωs Qdesign , ∆p
Ns
N

∆s Dt Dh ν

b, Va z

Stop Va

rt , ..., rh U φ, β1
Yes

Converged No ηh,i s, c

Vu2 , β2
σ
βm

ηh,f

βm CL , Re

Choose an
appropriate
design approach
γ α, Cd
i, θ

VI
C
CAD Models

(a) Fan geometry (b) Impeller geometry

Figure C.1: Model 1.

(a) Fan geometry (b) Impeller geometry

Figure C.2: Model 2.

VII
C. CAD Models

(a) Fan geometry (b) Impeller geometry

Figure C.3: Model 3.

(a) Fan geometry (b) Impeller geometry

Figure C.4: Model 4.

VIII
C. CAD Models

(a) Fan geometry (b) Impeller geometry

Figure C.5: Model 5.

(a) Fan geometry (b) Impeller geometry

Figure C.6: Model 6.

IX
C. CAD Models

(a) Fan geometry (b) Impeller geometry

Figure C.7: Model 7.

(a) Fan geometry (b) Impeller geometry

Figure C.8: Model 8.

X
C. CAD Models

(a) Fan geometry (b) Impeller geometry

Figure C.9: Model 9.

(a) Fan geometry (b) Impeller geometry

Figure C.10: Model 10.

XI
C. CAD Models

(a) Fan geometry (b) Impeller geometry

Figure C.11: Model 11.

XII
D
CFD Results

Table D.1: CFD results at/near the peak efficiency point at different rotational
speeds.

Model 1 Model 2 Model 3


N (rpm) 1000 2000 3000 1000 2000 3000 1000 2000 3000
∆pt−s (Pa) 33 130 293 34 130 291 32 127 284
Q (m3 /s) 0.36 0.75 1.15 0.37 0.76 1.16 0.36 0.74 1.13
τshaf t (Nm) 0.21 0.81 1.98 0.22 0.86 1.93 0.20 0.76 1.69
ηt−s (%) 54 57 54 54 55 56 56 59 60
Model 4 Model 5 Model 6
N (rpm) 1000 2000 3000 1000 2000 3000 1000 2000 3000
∆pt−s (Pa) 32 122 277 41 162 368 38 152 342
Q (m3 /s) 0.36 0.74 1.11 0.42 0.86 1.32 0.4 0.83 1.27
τshaf t (Nm) 0.19 0.73 1.64 0.32 1.24 2.78 0.27 1.07 2.40
ηt−s (%) 57 59 60 51 54 56 53 56 58
Model 7 Model 8 Model 9
N (rpm) 1000 2000 3000 1000 2000 3000 2000 3000 4000
∆pt−s (Pa) 36 143 324 37 146 328 125 281 500
Q (m3 /s) 0.38 0.79 1.2 0.39 0.8 1.22 0.73 1.11 1.5
τshaf t (Nm) 0.26 1.01 2.26 0.26 1.03 2.29 0.74 1.66 2.93
ηt−s (%) 51 54 55 53 54 56 59 60 61
Model 10 Model 11
N (rpm) 2000 3000 4000 1000 2000 3000
∆pt−s (Pa) 127 286 507 29 114 258
Q (m3 /s) 0.71 1.09 1.46 0.34 0.71 1.08
τshaf t (Nm) 0.81 1.81 3.21 0.20 0.76 1.72
ηt−s (%) 53 55 55 48 51 52

XIII
D. CFD Results

Table D.2: CFD performance results for model 3.

Porosity Q (m3 /s) ∆p (Pa)


5 0.84 102
25 0.81 110
50 0.79 117
100 0.74 126
200 0.66 141
400 0.56 162
800 0.44 192
1200 0.36 200

Table D.3: CFD performance results for model 7.

Porosity Q (m3 /s) ∆p (Pa)


5 0.91 121
25 0.88 129
50 0.83 136
100 0.76 148
200 0.70 155
400 0.59 169
800 0.48 198
1200 0.42 225

XIV
D. CFD Results

Figure D.1: Pressure distribution on the suction side of model 1 at 2000 rpm.

Figure D.2: Pressure distribution on the suction side of model 2 at 2000 rpm.

XV
D. CFD Results

Figure D.3: Pressure distribution on the suction side of model 3 at 2000 rpm.

Figure D.4: Pressure distribution on the suction side of model 4 at 2000 rpm.

XVI
D. CFD Results

Figure D.5: Pressure distribution on the suction side of model 5 at 2000 rpm.

Figure D.6: Pressure distribution on the suction side of model 6 at 2000 rpm.

XVII
D. CFD Results

Figure D.7: Pressure distribution on the suction side of model 7 at 2000 rpm.

Figure D.8: Pressure distribution on the suction side of model 8 at 2000 rpm.

XVIII
D. CFD Results

Figure D.9: Pressure distribution on the suction side of model 9 at 3000 rpm.

Figure D.10: Pressure distribution on the suction side of model 10 at 3000 rpm.

XIX
D. CFD Results

Figure D.11: Pressure distribution on the suction side of model 11 at 2000 rpm.

XX

You might also like