Vous êtes sur la page 1sur 123

CHARACTERIZATION OF CUP ANEMOMETER

DYNAMICS

AND

CALCULATION OF THE ACOUSTIC NOISE

PRODUCED BY A NREL PHASE VI WIND TURBINE

BLADE

by

YNG-RU CHEN

Submitted in partial fulfillment of the requirements

For the degree of Doctor of Philosophy

Department of Mechanical and Aerospace Engineering

CASE WESTERN RESERVE UNIVERSITY

May, 2016
CASE WESTERN RESERVE UNIVERSITY
SCHOOL OF GRADUATE STUDIES

We hereby approve the dissertation of

Yng-Ru Chen

candidate for the degree of Doctor of Philosophy*.

Committee Chair
Dr. J. Iwan D. Alexander

Committee Member
Dr. Jaikrishnan R. Kadambi

Committee Member
Dr. Paul Barnhart

Committee Member
Dr. David H. Matthiesen

Committee Member
Dr. Michael Hölling

Date of Defense

February 29, 2016

*We also certify that written approval has been obtained

for any proprietary material contained therein


DEDICATION

感謝父母,兄姊,兒子和愛犬的包容、支持和陪伴

To my parents, sister, brother, son and dog

for their unyielding love, support and encouragement.


TABLE OF CONTENTS

TABLE OF CONTENTS......................................................................................................................... i
LIST OF TABLES ................................................................................................................................ iii
LIST OF FIGURES .............................................................................................................................. iv
LIST OF SYMBOLS .......................................................................................................................... viii
ABSTRACT........................................................................................................................................ ix
INTRODUCTION ................................................................................................................................ 1
Part I: ................................................................................................................................................ 6
I. OVERVIEW AND BACKGROUND ........................................................................................... 7
1. THE CUP ANEMOMETER .................................................................................................. 9
2. WIND RESOURCE MEASUREMENT................................................................................. 10
II. OBJECTIVE .......................................................................................................................... 14
III. THEORY .......................................................................................................................... 15
IV. CALIBRATION.................................................................................................................. 20
V. OVERSPEEDING ANALYSIS OF DIFFERENT CUP ANEMOMETERS ....................................... 26
VI. OVERSPEEDING ANALYSIS OF UNBALANCED CUP ANEMOMETERS .............................. 37
VII. PRELIMINARY FIELD DATA EXAMINATION ..................................................................... 42
VIII. CONCLUSIONS AND FUTURE RESEARCH ........................................................................ 44
Part II .............................................................................................................................................. 46
I. OVERVIEW AND OBJECTIVES ............................................................................................. 47
II. NOISE AND SOUND FUNDAMENTALS ................................................................................ 50
1. SOUND AND NOISE ........................................................................................................ 50
2. NOISE STANDARDS AND REGULATIONS ........................................................................ 52
III. WIND TURBINE AERODYNAMICS ................................................................................... 54
1. LIFT, DRAG AND MOMENT COEFFICIENTS ..................................................................... 54
2. TIP SPEED RATIO ............................................................................................................ 54
3. AIRFOIL BEHAVIOR ......................................................................................................... 55
4. TWISTED BLADE ............................................................................................................. 55
IV. NOISE MECHANISMS OF WIND TURBINES ..................................................................... 57
1. WIND TURBINE AERODYNAMIC SOUND GENERATION ................................................. 57

i
2. CLASSIFICATION OF NOISE MECHANISMS ..................................................................... 58
3. ACOUSTIC SOURCES ....................................................................................................... 61
V. NUMERICAL ANALYSIS ....................................................................................................... 65
1. PHYSICAL MODEL ........................................................................................................... 65
2. MESH GENERATION ....................................................................................................... 65
3. TURBULENCE MODEL ..................................................................................................... 69
4. FAR FIELD NOISE PREDCITION USINF THE FFOWCS WILLIAMS AND HAWKINGS
ACOUSTIC MODEL .................................................................................................................. 70
5. BROADBAND NOISE SOURCE MODEL ............................................................................ 72
VI. RESULTS ......................................................................................................................... 74
1. FLOW FIELD AND STREAMLINES .................................................................................... 74
2. SURFACE ACOUSTIC POWER .......................................................................................... 75
3. FAR-FIELD AERODYNAMIC NOISE AT DIFFERENT WIND SPEEDS ................................... 82
4. SURFACE ACOUSTIC POWER AND FAR-FIELD AERODYNAMIC NOISE PREDICTION FOR A
YAWED TURBINE .................................................................................................................... 83
VII. DISCUSSION.................................................................................................................... 87
VIII. CONCLUSIONS ................................................................................................................ 89
1. SUMMARY ...................................................................................................................... 89
2. FUTURE WORK ............................................................................................................... 90
APPENDIX A .................................................................................................................................... 91
APPENDIX B .................................................................................................................................... 93
APPENDIX C: ................................................................................................................................... 98
APPENDIX D:................................................................................................................................. 101
APPENDIX E: ................................................................................................................................. 102
BIBLIOGRAPHY ............................................................................................................................. 104

ii
LIST OF TABLES

Table IV-1: Various cup anemometer configurations .................................................................... 20

Table V-1: Measured k values of each cup anemometer .............................................................. 35

Table VI-1: Various experimental configurations .......................................................................... 38

Table VI-2: Calibration Functions ................................................................................................... 39

Table VI-3: Various k values under different configurations ......................................................... 40

Table II-1: Health and safety executives noise regulation guidelines on exceeding safe level ..... 52

Table II-2: Recommended noise level standards and guidelines by WHO and selected countries

(Yuen, 2014) [46] ........................................................................................................................... 53

Table IV-1: Wind Turbine Aerodynamic Noise Mechanisms (From Wagner, Bareib, & Guidati,

1996) .............................................................................................................................................. 61

Table V-1: Receiver Positions ......................................................................................................... 71

Table VI-1: Sound Pressure Level (dB) of different rpm at receiver locations............................... 82

Table VI-2: Surface Acoustic Power (dB) of different yawed angles at 72 rpm ............................. 84

Table VI-3: The far-field aerodynamic noise of a yawed turbine at 72 rpm at receiver locations 86

iii
LIST OF FIGURES

Figure I-1: Global Energy Consumption 2014; (Data taken from in the BP Statistical Review of

World Energy 2105) [1] .................................................................................................................... 1

Figure I-2: Annual and Cumulative Growth in U.S. wind power Capacity [11] ................................ 3

Figure I-1: Schematic of a turbulent flow with a mean horizontal unidirectional component and

showing a dye trace marking typical particle trajectories in such a flow [24] .............................. 11

Figure I-2: Two examples of velocity (U) distributions varying in time (t) [24] ............................. 12

Figure I-3: Ultrasonic wind speed measurements from the FINO platform in North Sea (Dr. M.

Hoelling, U. of Oldenburg, personal communication) ................................................................... 13

Figure III-1: Wind passing a cup anemometer ............................................................................... 15

Figure III-2: A schematic diagram of one cup perpendicular to the wind direction ...................... 16

Figure III-3: Schematic model of a cup anemometer in wind field ................................................ 17

Figure IV-1: Cup anemometers used for the current study (a): ThiesTM First Class Cup

Anemometer, (b): Second Wind C3TM Cup Anemometer, (c): ThiesTM Sensor Compact Cup

Anemometer, (d): ThiesTM Classic Cup Anemometer..................................................................... 21

Figure IV-2: The calibration function of ThiesTM First Class Cup anemometer .............................. 23

Figure IV-3: The calibration function of ThiesTM First Class Cup anemometer according to the

manufacturer and experimental results ........................................................................................ 24

Figure V-1: Schematic top view of the experiment setup ............................................................. 26

Figure V-2: Wind tunnel experiment setup (looking at outlet of the wind tunnel)....................... 27

Figure V-3: Schematic orientation of the cup anemometer operation ......................................... 28

Figure V-4: Wind speed measured by ThiesTM First Class Cup anemometer while decreasing wind

velocity situation in wind tunnel (15 m/sec wind speed) .............................................................. 30

iv
Figure V-5: Two cases of overspeeding situation of ThiesTM First Class Cup anemometer at 10

m/sec ............................................................................................................................................. 31

Figure V-6: Comparison of measured and computed response to stepwise changes in wind speed

....................................................................................................................................................... 32

Figure V-7: Overspeeding curves for the ThiesTM First Class Cup Anemometer at various wind

speeds ............................................................................................................................................ 33

Figure V-8: Overspeeding curves for the Second WindTM C3 Cup Anemometer at various wind

speeds ............................................................................................................................................ 33

Figure V-9: Overspeeding curves for the ThiesTM Compact Cup Anemometer at various wind

speeds ............................................................................................................................................ 34

Figure V-10: Overspeeding curves for the ThiesTM Classic Cup Anemometer at various wind

speeds ............................................................................................................................................ 34

Figure V-11: Overspeed curves for the four anemometers at V = 10 m/sec (* denotes the mean

uncertainty).................................................................................................................................... 36

Figure VI-1: Weighted cup anemometers ...................................................................................... 38

Figure VI-2: Calibration function of the Thies™ Classic Cup Anemometers with different weighted

cups ................................................................................................................................................ 39

Figure VI-3: Overspeed curves comparison of the Thies™ First Class Cup Anemometers with

different weighted cups at wind speed of 10 m/sec (* denotes “mean total uncertainty”) ........ 40

Figure VII-1: Measured wind speed vs. time (upper graph) and the inverse of the wind speed vs.

time (bottom graph) ...................................................................................................................... 42

Figure I-1: Atmospheric boundary layer with wind turbine .......................................................... 47

Figure III-1: Forces and moment on an airfoil section, α, angle of attack; c, chord [22] ............... 54

Figure III-2: Blade Twist.................................................................................................................. 56

v
Figure IV-1: Horizontal axis wind turbine noise sources (Romero-Sanz & Matesanz, 2008)......... 59

Figure IV-2: Aerodynamic noise sources around a turbine blade due to wind flow, U (Brooks,

Pope, & Marcolini, 1989) ............................................................................................................... 60

Figure IV-3: Pressure field produced by a monopole source ......................................................... 62

Figure IV-4: Pressure field produced by a dipole source ............................................................... 63

Figure IV-5: Pressure field produced by a quadrupole source ...................................................... 64

Figure V-1: Structured and unstructured grids .............................................................................. 65

Figure V-2: An example of unstructured grids ............................................................................... 66

Figure V-3: Structured grids consisting of triangular, polygonal meshes and orthogonal meshes66

Figure V-4: Structured grids near the blade .................................................................................. 67

Figure V-5: Grids in the computational domain ............................................................................ 67

Figure V-6: Boundary conditions ................................................................................................... 68

Figure V-7: Wind forces on the blade ............................................................................................ 69

Figure V-8: Receiver Positions [15] ................................................................................................ 71

Figure VI-1: Streamlines on the blade............................................................................................ 75

Figure VI-2: Streamlines at z=5 m .................................................................................................. 75

Figure VI-3: Surface Acoustic Power Level (dB) on blades at 36 RPM ........................................... 77

Figure VI-4: Surface Acoustic Power Level (dB) on blades at 54 RPM ........................................... 78

Figure VI-5: Surface Acoustic Power Level (dB) on blades at 60 RPM ........................................... 79

Figure VI-6: Surface Acoustic Power Level (dB) on blades at 72 RPM ........................................... 80

Figure VI-7: Maximum Sound power levels for wind speeds 7 m/s through 25 m/s at different

rotation speeds .............................................................................................................................. 81

Figure VI-8: A wind turbine yawed to the wind direction [22] ...................................................... 83

vi
Figure VI-9: Surface Acoustic Power Level (dB) of 72 rpm at wind speed of 7 m/sec on blades at

different yawed angles .................................................................................................................. 85

Figure VII-1: Sound power levels as a function of standardized wind speed for the ARE 442 wind

turbine (rotor diameter is 7.2 m) (Huskey & Dam, 2010) [63] ...................................................... 87

Figure VII-2: Measured sound pressure levels of a Southwest Whisper 900 wind turbine (rotor

diameter is 2.1 m) (Rogers, Manwell, & Wright, 2006) [64].......................................................... 88

Figure VII-3: Sound power levels as a function of standardized wind speed for the Gai Wind 11-

kW wind turbine (Huskey A., 2011) [65] ........................................................................................ 88

vii
LIST OF SYMBOLS

Symbol Definition Units

u, V Wind speed m/s

ω Cup anemometer angular speed rad/s

𝐶𝐷 Drag Coefficient dimensionless

f Rotor Pulse Frequency pulses/s

fr Rotor Revolution Frequency revolutions/s

𝐼𝑡 Turbulence Intensity dimensionless

 Air density kg/m3

M Aerodynamic torque N·m

I Moment of inertia kg·m2

𝐿𝑝 Sound Pressure Level dB

Lw Sound Power Level dB

𝜆 Wind Turbine tip speed ratio dimensionless

Ω Wind turbine rotational speed revolutions/m

𝑅𝑒 Reynolds Number dimensionless

𝑘- ω Turbulence model

k-ε Turbulence model

viii
Characterization of Cup Anemometer Dynamics

and Calculation of the Acoustic Noise Produced by a NREL

Phase VI Wind Turbine Blade

ABSTRACT

by

YNG-RU CHEN

The results of two separate investigations into two issues related to the generation of

electrical power from wind are reported in this dissertation. The first, an experimental

investigation of cup anemometer dynamics, is motivated by the need to accurately

measure local wind conditions at wind turbine array locations. This need is two-fold, one

being the need to assess wind-turbine performance (in terms of the conversion of wind

power to electrical power, the second is to ensure that predictions of the wind power

capacity of a given area of interest are based on reliable wind speed measurements. The

purpose of the investigation is to shed light on possible approaches to mitigate potential

shortcomings of wind data collected using cup anemometers by a systematic

experimental investigation of their performance under simulated “gusts” in a wind

tunnel. It is found that the response of the anemometer to changing wind conditions is

determined by two parameters both of which represent the ratio of drag forces to

ix
inertial forces. In the recent of a sudden decrease in wind speed to zero, one of the

parameters alone controls the rate of decay of the anemometers speed to zero. This

parameter is measured experimentally for 4 different anemometers at three different

initial wind speeds

The second part of this dissertation deals with the subject of aeroacoustic noise. An

important factor to take into consideration in the development of wind farms near

populated areas is the potential for noise disturbance. In this dissertation, the topic of

aeroacoustic noise is addressed using a computational fluid dynamic model to compute

the wind velocity and pressure fields through the turbine. The Ffowcs-Williams and

Hawkings acoustic analogy is used to calculate the acoustic pressure at a far field

receiver location. The results obtained are compared with experimental data. The

purpose of the investigation is, by extending earlier work of Ranft et al., to predict the

aeroacoustic noise at different wind speeds and at different rotational speeds of the

turbine. Both rotational noise and broadband noise were analyzed. It was found that the

leading edge and the blade tip are the primary source of the broadband noise. In all

simulations, the loudest noise (80.8 dB) generated by the wind turbine blade tip and the

leading edge are higher than the World Health Organization recommended noise level

standards but decay with increasing distance from the turbine.

x
INTRODUCTION

One of today’s greatest global energy challenges is the need for more affordable,

reliable, clean, secure and renewable sources of electricity. There is still heavy reliance

on coal, oil, natural gas and nuclear fuels to generate electricity. The 2015 BP Statistical

Review of World Energy shows fossil fuels (Oil, gas, coal and nuclear) accounted for 86%

of global primary energy consumption [1].

Figure I-1: Global Energy Consumption 2014; (Data taken from in the BP Statistical

Review of World Energy 2105) [1]

Note: the unit ‘millions of tonnes of oil’ was converted to TeraWatt-hours using the

conversion factor 1 TW h ≈ 4.4 x 106 million (metric) tonnes of oil.

1
Fossil fuels are finite and non-renewable on a human socio-economic time scale [2].

They are limited in spatial extent and are expected in the long term, to become

increasingly more expensive to extract [3]. Furthermore, they are not renewable [3].

Thus, the use of fossil fuels for energy generation is not sustainable. This, along with

increasing concern about anthropogenic effects on climate change with associated CO₂

and other greenhouse gas emissions is driving governments and power companies

across the globe to diversify fuel sources for electrical power generation and, in

particular, to seek alternatives to fossil fuels. [4] [5] [6].

Wind power is a commercially proven, rapidly growing form of electricity generation. It

provides clean, renewable, and cost-effective electricity around the world. According to

the International Renewable Energy Associate, global installed wind energy capacity

reached nearly 370 GW in 2014 [7] [8]. Some countries (Germany, Demark, Spain…)

have supported wind industry heavily since the end of the twentieth century. Recently,

China and India have started to invest more and more in wind energy development and

showed the highest growing rates of wind energy installed power in the past years.

According to 2015 data reported by the Global Wind Energy Council [9] and the US

Department of Energy’s Energy Information Administration (EIA), China now has the

largest installed wind power capacity (115 GW) followed by the USA (66 GW) [10] [11].

The U.S. wind power market represented over 9% of global installed capacity in 2014.

The annual growth of global installed wind capacity has been increasing steadily for

nearly 20 years. In contrast, US installed capacity has tended to be more sporadic over

the same time period (See Figure I-2). This reflects, perhaps, the lack of a clear national

2
policy to develop wind power and a reliance on incentives on a state-by-state basis.

Nevertheless, US wind energy capacity has more than doubled over the past five years.

Recent data [6], [8] indicates that US installed capacity had reached 66 GW in 2014.

Figure I-2: Annual and Cumulative Growth in U.S. wind power Capacity [11]

The continued increase in the use of wind as a source for electrical power generation

has resulted in a variety of research activities in the public and private sectors that

address a range of issues from performance, safety, efficiency and cost to public

acceptance.

The results of two separate investigations into two issues related to the generation of

electrical power from wind are reported in this dissertation. The first, an experimental

investigation of cup anemometer dynamics, is motivated by the need to accurately

measure local wind conditions at wind turbine array locations. This need is two-fold, one

being the need to assess wind-turbine performance (in terms of the conversion of wind

3
power to electrical power, the second is to ensure that predications of the wind power

capacity of a given area of interest are based on reliable wind speed measurements.

The latter is easily understood by considering that the power produced by the wind

turbine is proportional to the cube of the wind speed. Thus, a 1% error in the measured

wind speed translates to a 3% error in the estimated power, a 5% error translates to a

15% error in the power and so on. As will be shown in the following chapters, there are

still outstanding questions regarding the reliability of the 3-cup-anemometer, the

standard instrument used industry wide to measure wind speed for wind energy

applications. The purpose of the investigation is to shed light on approaches to mitigate

potential shortcomings of wind data collected using cup anemometers by a systematic

experimental investigation of their performance under simulated “gusts” in a wind

tunnel.

The second part of this dissertation deals with the subject of aeroacoustic noise. An

important factor to take into consideration in the development of wind farms near

populated areas is the potential for noise disturbance. In a number of countries, wind

power capacity has been significantly limited by the so-called “Not In My Back Yard”

attitude to specific wind power projects, even when polls show that in fact the public

supports the idea wind power as an energy source [12] [13]. Noise disturbance is among

the issues raised by wind project opponents. There are two prevalent noise sources that

arise in connection with wind turbines, mechanical [12] and aerodynamic [14]. In this

dissertation, the topic of aeroacoustic noise is addressed using a computational fluid

dynamic model to compute the wind velocity and pressure fields through the turbine.

4
The Ffowcs-Williams and Hawkings (FW–H) acoustic analogy is used to calculate the

acoustic pressure at a far field receiver location. The results obtained are compared with

experimental data. The purpose of the investigation is, by extending the work of Ranft

et al. [15], to predict the aeroacoustic noise at different wind speeds and at different

rotational speeds of the turbine.

5
Part I:

CHARACTERIZATION OF CUP ANEMOMETER DYNAMICS

6
I. OVERVIEW AND BACKGROUND

There are many options that can be exercised to quantify wind resources at a given

wind farm site but the traditional meteorological mast equipped with anemometers,

thermometers and other sensors are still the most accurate way to obtain such data.

While limited to measurements at discrete locations (and heights) this approach is used

at a minimum to quantify wind resources. The objective is to obtain a reliable estimate

of the amount of electrical power that could be produced, say annually, at a specific

location. To make such an assessment there is a specific procedure that must be

followed according to the IEC 61400-12 standard. The standard recommends measuring

wind speed using a calibrated cup anemometer installed within 2.5% of hub height of

the wind turbine. The anemometer may be installed on a boom attached to the tower

pointed into the prevailing wind direction. Flow distortion, say from buildings or other

structures should be minimized.

 Data must be collected continuously for a minimum of 12 months

 Only anemometers satisfying the IEC 61400-12 standard should be used.

 An approved institution in accordance with national and international standards

and regulations must calibrate anemometers.

The purpose of the IEC standard is to

 Provide a basis for comparing different Wind Turbines

 Minimize uncertainty in wind turbine power performance tests

 Improve repeatability between sites

7
The importance of the accuracy of the anemometer can be illustrated by considering the

following. If the actual wind speed is 10 m/s but the measured speed is 10.5 m/s the

error is 0.5 m/s or 5%. However, the power produced by the wind turbine varies as the

cube of the wind speed. It follows that the error in the estimate of the power produced

would, in this case, be over 15%. Note that, in addition to providing a quantitative

assessment of a potential wind field, anemometers are used to develop the so-called

power curve of a wind turbine and reliable data is essential to obtain the turbine’s

power curve and to monitor the performance of installed wind turbines. The most

commonly used anemometers for wind power applications are three-cup anemometers.

As will be shown in subsequent sections, cup anemometers are prone to a phenomenon

known as “overspeeding” [16] [17] [18] [19]. Because of overspeeding, the mean wind

speed measured by the anemometer will be overestimated.

The objective of the experimental investigation described in the subsequent sections is

to characterize the anemometer dynamics related to overspeeding with the intent of

developing an understanding of the process that will stimulate the development of

techniques to mitigate overspeeding errors and thereby obtain a more accurate

description of wind speeds used to calculate power production.

8
1. THE CUP ANEMOMETER

Dr. T.R. Robinson invented the first cup anemometer in 1846. It had four hemispherical

cups. Each cup was fixed on one end of four arms that were mounted at equal angles to

each other on the vertical shaft [20]. While the air flows past the cups, the cups rotation

corresponds to the wind speed. In 1926, the Canadian J. Patterson noticed that each cup

produced the maximum torque when the cup was at 45 degree to the incoming wind.

He developed the three-cup anemometer that responded to the gusty winds more

quickly than the four-cup anemometer [16]. The simple design of a cup anemometer

consists of three cups mounted on a vertical shaft. The differences of the wind pressure,

aerodynamic drag force and acting torque between the concave side and the convex

side of the cup causes it to turn in the direction from the convex side to the concave

side of next cup. The revolution speed is proportional to the wind speed irrespective of

wind direction. Wind speed signals are generated from either a generator or a pulse

generator coupled to the rotating axis of the anemometer.

The cups were conventionally made of brass for its qualities of rigidity and rust

resistance. In recent years, however, cups made of light alloy or carbon fiber thermo-

plastic have become the mainstream, allowing significant reductions in weight.

Cup anemometers are most commonly used for measuring wind speeds in the

atmospheric boundary layer. Typically, cup anemometers can be divided into three

types [21], Generator-type Cup Anemometers, Pulse Generator-type Cup anemometers

and Mechanical-type Cup Anemometers.

9
In a permanent regime, a cup anemometer turns at a frequency proportional to the

wind speed irrespective to the wind direction. One downside of cup anemometers is

their asymmetric dynamic response to accelerating and decelerating wind speeds under

turbulent wind conditions. Due to inertia the anemometer will respond faster in

response to an increase in wind speed than to a decrease in speed of the same

magnitude. Therefore, in a turbulent flow, the cup anemometer will overestimate the

mean wind speed, especially if there are many sudden large decelerations. This

systematic overestimation under turbulent wind conditions is called over-speeding [16]

[17] [18] [19].

2. WIND RESOURCE MEASUREMENT

Wind is variable in space and time. It varies non-uniformly with altitude, local

topography and because air density is a function of temperature and humidity [22] the

wind is also affected by these quantities (stratification, stability, etc.).

Commercial wind turbines stand tall with typical hub heights of between 50-120 meters.

Situated in the lower reaches of the atmospheric boundary layer (ABL) wind turbines

experience wind that are influenced directly by the earth’s surface. One would expect

the horizontal wind speed to be zero at the earth’s surface and to increase with height

in the atmospheric boundary layer. The variation of wind speed with elevation, so called

vertical wind profile or vertical wind shear, is very important in wind energy engineering

since it determines the productivity of a wind turbine and also impacts the lifespan of a

turbine blade [23].

10
Turbulence in the wind is caused by the kinetic energy dissipation of wind into thermal

energy via the creation and destruction of eddies (or gusts). Turbulent eddies create

fluctuations in velocity and the velocity record includes both a mean and a turbulent

component.

Figure I-1: Schematic of a turbulent flow with a mean horizontal unidirectional

component and showing a dye trace marking typical particle trajectories in such a flow

[24]

As discussed in the previous section, while cup anemometers are the most commonly

used instrument for wind speed measurement their response to turbulent fluctuations

in wind speed often results the overestimation of the actual wind speed through over

speeding. Consider the turbulent flow shown schematically in Figure I-1 with a mean

̅ = (𝒖, 0), where 𝒖 = (𝑢, 𝑣), and 𝑢, 𝑣 are the horizontal and vertical
wind speed 𝒖

velocity components, respectively. Here the important quantities that are used to

̅ , the horizontal and vertical turbulent


characterize the flow are the mean velocity, 𝒖
11
fluctuations, 𝑢′ , 𝒗′ , the turbulence strength 𝑢𝑟𝑚𝑠 , and the turbulence intensity 𝐼𝑡 . These

quantities are calculated as follows

𝑡+𝑇 1
𝑢̅ = ∫𝑡 𝑢(𝑡) 𝑑𝑡 = ∑𝑁
1 𝑢𝑖 (1)
𝑁

𝑢′ (𝑡) = 𝑢(𝑡) − 𝑢̅, 𝑣 ′ = 𝑣 (2)

𝑢𝑖′ = 𝑢𝑖 − 𝑢̅, 𝑣′ = 𝑣′𝑖 (3)

̅̅̅̅̅̅̅̅2 + ̅̅̅̅̅̅̅̅ 1 2 2
𝑢𝑟𝑚𝑠 = √𝑢′(𝑡) 𝑣′(𝑡)2 = √𝑁 ∑𝑁
1 (𝑢′𝑖 + 𝑣′𝑖 ) (4)

𝐼𝑡 = 𝑢𝑟𝑚𝑠 /𝑢̅ (5)

A larger value of turbulence intensity indicates a higher level of turbulence. The figure

below shows an example where two time series have the same mean velocity but the

one on the left has a higher level of turbulence [24].

Figure I-2: Two examples of velocity (U) distributions varying in time (t) [24]

In general, the highest turbulence intensities happen at the lowest wind speed but the

lower limiting values at a given location will depend on the specific terrain features and

12
surface conditions at the site. Turbulent wind may have a relatively constant mean over

time periods of an hour or more, but over shorter times in may be quite variable.

When atmospheric wind data is being analyzed, the mean velocity, turbulence strength

and turbulence intensity are usually calculated for 10-minute bins. The turbulence

intensity computed over this interval of time captures a measure of the turbulent

fluctuations but does not capture the particular nature of the turbulence. The details

may be important in that for two flows with the same turbulence intensity the specific

nature of the fluctuations may lead to more or less overspeeding. For example, two

wind measurements shown in Figure I-3 were recorded with an ultrasonic anemometer.

Both have the same turbulence intensity (about 4%). The blue curve has several sharp

decreases in wind speed (three are indicated by red arrows). It will be shown later that

such conditions can result in higher estimations of the overall wind speed.

Figure I-3: Ultrasonic wind speed measurements from the FINO platform in North Sea

(Dr. M. Hoelling, U. of Oldenburg, personal communication)

13
II. OBJECTIVE

The objective of the research is to characterize the dynamical response of cup

anemometers for fluctuating wind speeds and for decreasing speeds in particular. The

effect of different torques of an unbalanced system on the anemometer’s dynamics is

investigated by adding rare earth magnets as weight changes on the cups. These

characteristics are basis for the post processing of measured data under turbulent wind

conditions. Regions of the measured time series that are dominated by the specific

dynamics of the cup anemometer are typically affected by over-speeding and this

should be accounted for in any analysis of actual measured wind conditions.

14
III. THEORY

The response and accuracy of a cup anemometer are determined by its weight, physical

dimension (such as the cups’ shape, size and the rotation radius) and internal friction.

Sanz-Andrés, Pindado & Sorribes-Palmer [25] summarizes some research on cup

anemometers using different techniques and mathematical models since the early work

of Robinson in 1846.

To illustrate the behavior of a cup anemometer consider a three-cup anemometer that

is placed into a uniform flow, consider Figure III-1. The anemometer is inserted into the

flow in a direction perpendicular to the air flow. As it is moved into the flow, the

surfaces of the cups that are concave toward wind have higher wind resistance than

convex surface. The force of the wind causes the cups to rotate and the rate of change

of the angular speed of the rotor is equal to sum of the moments acting on each cup.

Figure III-1: Wind passing a cup anemometer

15
Figure III-1 illustrates the idea that the flow of wind around the cups of an anemometer

produces a higher force on the outer side of the cup than it does on the inside of the

cup, causing a torque that rotates the system [26].

Over the years, scientists and engineers have investigated various shapes of cups with

different lengths of arms seeking to develop a cup anemometer that suffers neither

from overspeeding nor deviations to inclined flows. Each proposed design has its

advantages and disadvantages. For example, the larger the cups with longer arms, the

slower the changes in measured wind speed. It also means the mean measured wind

speed is higher than the real wind speed [27].

Overspeeding can be better understood upon consideration of the following argument.

Consider only one single cup that is oriented perpendicular to a free wind speed, V, so

that the convex part of the cup surface is toward the oncoming wind (Figure III-2).

Figure III-2: A schematic diagram of one cup perpendicular to the wind direction

The drag force on a cup can be expressed as:

16
1
𝐹 = 2 𝐶𝐷 𝜌𝑉 2 𝜋𝑟 2 (6)

Here CD is the drag coefficient of the cup, ρ is the air density, V is the uniform wind

speed, and r is the radius of the cup. For these conditions, the torque on one cup is

given by

1
𝑀= 𝐶 𝜌𝑉 2
𝜋𝑟 2 𝑅 = 𝐹𝑅 (7)
2 𝐷

Here R is the distance from the center of the cup to center of the rotor (i.e. the axis of

rotation) and V is the wind speed.

Now consider the whole cup anemometer (Figure III-3). The rotational speed, 𝜔, of the

anemometer is proportional to the wind speed V . The incident wind speed, V, is

uniform. The angle Φi is the instantaneous angle made by the rotor arm of the ith cup

and a line that is perpendicular to the wind direction.

Figure III-3: Schematic model of a cup anemometer in wind field

17
The drag coefficient for the cups depends on the shape of the cup, the cup surface

roughness and the cup orientation relative to the wind direction. As a first

approximation, the drag coefficient of the open (concave) cup facing the incoming wind

is set to be 𝐶𝐷1 while the drag coefficient of the closed (convex) cup facing the incoming

wind is 𝐶𝐷2 .

The frictional torque on the anemometer depends on air temperature and the rotation

speed. It is normally small in comparison to the aerodynamic torque [28] [29], and can

be neglected except at very low wind speeds (close to the ‘cut-in speed of the

anemometer). In this case the aerodynamic torque T is equal to the inertial force of

the rotor which can be expressed as

3
dw
I e = T=
dt z
å r (F i )´ F (F i ),
i= 1
(8)
dF i dF
= = w, i = 1, 2, 3
dt dt

Here r (F i )= R er (F i ) is the position vector along the rotor axis to the midpoint of the

ith cup, er (F i ) is a unit vector, I is the moment of inertia of the rotor-cup assembly

and F (F i ) is instantaneous aerodynamic force on the ith cup and e z is the unit

position vector parallel to the axis of rotation. Rearranging (8) and taking the dot

product with e z we obtain

3
dw 1
dt
=
I
å k ×r (F i )´ F (F i ) (9)
i= 1

18
where

ìï æV ö2 æV ö
ïï çç cos (F )- w ÷
ïï K ÷ , if çç cos (F i )- w ÷ ÷> 0 , cos (F i )> 0

è R i ÷
ø çR
è ÷
ø
ïï
ïï
ïï
ï æ ö2 æV ö
ez ×r ´ F (F i )= ïí - K 2 çç w - V cos (F i ) ÷ ÷ , if çç cos (F )- w ÷ ÷< 0, cos (F i )> 0
(10)
ïï çè R ÷
ø èç R i
ø÷
ïï
ïï
ïï
ïï æ V ö2
÷
ç
ïï - K 2 çç w + cos (F i ) ÷ , if cos (F i )< 0
è R ÷
ø
ïî

and

r A C D 1R 3
K1 =
2I
(11)
3
r A C D 2R C
K2 = = K 1 D2
2I C D1

Here A = p r 2 and K 1, K 2 represent the ratio of aerodynamic to inertial forces for the

single cup case illustrated in Figure III-3 and the case where the rotor is spinning in zero

wind, respectively. Note that K 1 > K 2 because C D 1 > C D 2 which highlights the inherent

asymmetry of the anemometer response to wind forcing.

19
IV. CALIBRATION

Calibrations of the anemometers used in the experiments were performed at the

University of Oldenburg, Germany. Four cup anemometers from different

manufacturers were used. The dimensions of the wind tunnel outlet (1m x 0.8m) was

large enough, compared with the dimensions of the cup anemometers, to allow the

neglect of any boundary layer effects on the cup anemometer and measured velocity

data, respectively [30].

Four types of cup anemometers were calibrated. (Figure IV-1, Table IV-1). Only the

Second Wind C3TM had not been used previously. The specifications of the four

anemometers are listed below.

Shape of the cup Cup Material R (mm) r (mm)


Carbon-Fiber- 80 40
ThiesTM First Class Cone
Reinforced Plastic
69.5 25.5
Second WindTM C3 Cone Polycarbonate

Carbon-Fiber- 45 22
ThiesTM Compact Hemisphere
Reinforced Plastic
119 39
ThiesTM Classic Hemisphere Anodized Aluminum

Table IV-1: Various cup anemometer configurations

20
Figure IV-1: Cup anemometers used for the current study (a): ThiesTM First Class Cup

Anemometer, (b): Second Wind C3TM Cup Anemometer, (c): ThiesTM Sensor Compact

Cup Anemometer, (d): ThiesTM Classic Cup Anemometer

Because each anemometer has different cup shapes and different material roughness,

their drag coefficients (CD) are different. For a given wind condition, each anemometer

will experience a different aerodynamic torque because of the differences in drag

coefficient and the different sizes of the anemometers. As a result, each cup

anemometer responds differently to the same wind conditions.

21
Each cup anemometer has a manufacturer’s linear calibration function. Most calibration

procedures specify that the rotational frequency output of the anemometer or another

output variable such as voltage must be measured at certain specific wind speeds. The

calibration curve is acquired by performing a linear fit of the measured frequency f to

the actual wind speed.

V=Af+B (12)

Here V is the wind speed (m/s), A is the slope of the calibration curve (m), f is the

anemometer’s frequency output in pulses per second and B is the offset (m/s). If the

bearing is friction-free, the friction moment should be zero and B would be zero. In fact,

the bearing is not friction-free, so the friction moment is non-zero and the cup

anemometer will remain motionless until the wind reaches a finite non-zero speed. The

typical starting velocity of cup anemometer is between 0.3 - 0.5 m/sec [31]. Some

authors claim that a non-linear expression should be used instead of a linear one,

especially at low wind speeds [32] [33]. However, the linear expression is accurate in

normal working conditions and recommended in standard calibration processes [34] [35]

[36].

For some cup anemometers the raw data output is standard AC sine wave

corresponding to the turning frequency of the cup anemometer’s rotor. The cup

anemometer’s pulse frequency output, f, is the result of multiplying the rotational

frequency of the anemometer, fr, by the number of pulses per revolution given by the

cup anemometer, Np. The calibration function in the datasheet has the same format as

Equation (12) but it can be rewritten as below. [37] [31] [38]


22
V = A f + B = A Npfr + B (13)

Np is the number of pulses per revolution and fr is the rotational frequency

(revolutions/s)

Figure IV-2 shows two calibration functions from ThiesTM First Class and Second Wind

C3TM cup anemometers. The output of Second WindTM C3 is standard AC output and

ThiesTM First Class has frequency output. The calibration function of Second Wind C3 has

been converted into frequency output in the equation. Both cup anemometers have the

same conical-shaped cup but cup sizes and arm lengths are different. The calibration

function for the ThiesTM First Class shows a large offset which suggests that it’s the static

friction of its bearings is greater than the Second Wind C3. (Thus, the wind speed

needed to start a static ThiesTM First Class cup anemometer will also be greater. [28].)

Figure IV-2: The calibration function of ThiesTM First Class Cup anemometer

23
Even though the cup anemometers were calibrated by the manufacturers before

shipping, they are re-calibrated individually before the experiment. The calibration

function was obtained for each cup anemometer. Figure IV-3 shows the linear

calibration of ThiesTM First Class cup anemometer in the wind tunnel vs typical

calibration function from the manufacturer. The slopes of the two curves are almost

identical but the offsets are different. The offset in the calibration by the manufacturer

is smaller than the offset in the current calibration equation obtained from the wind

tunnel experiment. The anemometer requires higher wind speeds to initiate than the

manufacturers specification. The reason could be that the bearing is probably worn.

25

20

V = 0.0468f + 0.4246
V = 0.0462f + 0.21
Velocity (m/sec)

15

Results from
10 Experiment

Results from
Datasheet
5

0
0 100 200 300 400 500 600
f (pulses/sec)

Figure IV-3: The calibration function of ThiesTM First Class Cup anemometer according

to the manufacturer and experimental results

24
It is much easier to analyze the raw data for the frequency output than AC output. The

AC output was also observed to be noisier (results will be presented in the next section)

and so required more effort to analyze.

25
V. OVERSPEEDING ANALYSIS OF DIFFERENT CUP ANEMOMETERS

The analysis of overspeeding was accomplished by conducting a set of wind tunnel

experiments in which each anemometer was subjected to simulated rapidly changing

wind conditions by moving the anemometer out of and into a steady wind stream. After

each cup anemometer was calibrated and the calibration formulas are acquired, the

next step involved using the wind tunnel to simulate a rapid wind change. Figure V-1

shows a schematic of the experimental setup. The cup anemometer was smoothly but

rapidly moved in and out of the test section along a direction perpendicular to the wind

direction (Figure V-1, Figure V-2).

Figure V-1: Schematic top view of the experiment setup

26
Figure V-2: Wind tunnel experiment setup (looking at outlet of the wind tunnel)

Upon removal of the anemometer from the wind tunnel, the velocity V in equation (10)

is zero and the equation (9) for the rate of change of angular speed of the rotor

becomes

dw
= - 3K 2 w2 = - k w2, t > t 0 , w (t 0 )= w0 (14)
dt

For a given cup anemometer, k will be a constant but will generally be different for

different anemometers

The solution to (14) is

w0
w (t ) = (15)
1 + w0k (t - t 0 )

27
From Equation (15), it is clear that a cup anemometer having the same initial speed (t0)

but with a smaller k value will require less time to reach a given lower rotational speed.

In Figure V-1, the cup anemometer turns counter-clockwise while the generated wind

flows from left to right. If the cup anemometer is moved to the right (relative to the

wind direction) and removed from the wind stream, the rotation rate will slow gradually

and the cup will behave according to equation (15). This is the case where the

anemometer has an initial non-zero rotation tare and the wind suddenly stops blowing.

Note that, in the experiments, if the cup anemometer is moved out of the test section in

the opposite direction (i.e., to the left relative to the wind direction), rotation will stop

immediately. The reason for this is explained with the aid of Figure V-3.

(A) (B) (C)

Figure V-3: Schematic orientation of the cup anemometer operation

Consider Equation (9) in the form

dw 3 Mi

dt
= å 1 I
(16)

28
If the cups with moments M1 and M3 are pushed out the wind tunnel test section first

(moving downward from the top view in Figure V-3 (A)), the torque on these cups is

2
æ V ö
÷
M 1 = M 3 = - K 2w 2 and the torque on the remaining cup is - K 2 çç w + cos (F i ) ÷
÷
çè R ø

so anemometer will slow rapidly due the high drag force on the remaining cup. This

does not, however, represent the situation of how a cup anemometer responds to a

suddenly decreasing wind in the field. On the other hand, if the M2 leaves the wind

tunnel test section first (moving upward from the top view in Figure V-3 (B)), only one

cup has a negative torque, M 2 < 0 , while the remaining two cups are still driven by the

wind so d w / dt > 0 . In the latter case, the cup anemometer will initially speed up upon

removal from the tunnel before the velocity starts to decay after the entire

anemometer is removed from the test section. The circles in Figure V-4 and Figure V-5

show the transient increase in velocity as the anemometer is removed in this way.

29
Figure V-4: Wind speed measured by ThiesTM First Class Cup anemometer while

decreasing wind velocity situation in wind tunnel (15 m/sec wind speed)

30
Figure V-5: Two cases of overspeeding situation of ThiesTM First Class Cup anemometer

at 10 m/sec

Figure V-6 shows measurements made during a sequence that involves the removal of

the anemometer from the wind tunnel for a finite time, reinsertion into the test section,

removal and reinsertion. The cup anemometer responds more rapidly to an increase in

wind speed, faster than the response to a decrease in wind speed, overspeeding. The

cup anemometer underestimates the wind velocity as it is inserted into the test section

wind speed and overestimates the speed during the removal. This can be seen clearly

from the asymmetry in the measured and computed response to sudden changes in the

wind speed. It appears that some consideration of this asymmetry in response must be

accounted for to obtain a reliable measurement. Since in these experiments, both under

and over estimates of the wind speed are observed depending on whether the wind

31
speed increases or decreases, the corresponding curves will be referred to as under- or

overspeeding curves.

5
Wind Speed (m/sec)

3
Measured
wind speed
2
Actual wind
speed
1

0
0 20 40 60 80 100 120 140

Overestimate Underestimate Time (sec)

Figure V-6: Comparison of measured and computed response to stepwise changes in

wind speed

Figure V-7~10 shows the results of an overspeeding experiment performed at different

wind speeds (5m/s, 10m/s and 15m/s) for the four different cup anemometers listed in

Table IV-1. The goal here is to compare the degree of overspeeding for the four different

types. The experiment under the same setup is repeated at least three times and only

one set is plotted in the figures. The overspeeding curve is taken out from the set of

data and is aligned with other overspeeding curves for the comparison.

32
Figure V-7: Overspeeding curves for the ThiesTM First Class Cup Anemometer at various

wind speeds

Figure V-8: Overspeeding curves for the Second WindTM C3 Cup Anemometer at

various wind speeds

33
Figure V-9: Overspeeding curves for the ThiesTM Compact Cup Anemometer at various

wind speeds

Figure V-10: Overspeeding curves for the ThiesTM Classic Cup Anemometer at various

wind speeds

34
The value of k in equation (15) for each cup anemometer was obtained from

experimental data as follows. Each wind velocity data set obtained for anemometers

suddenly removed from the wind tunnel (i.e. the wind speed undergoes a step change

from a finite speed to zero) was processed as follows. The calibration offset velocity for

each anemometer was subtracted from the actual velocity. To convert this to a

rotational speed the resulting difference was then divided by the anemometer rotor

radius R. The k in equation (15) was then fit to these data for each anemometer. The

results are shown below in Table V-1.

Anemometer k value

ThiesTM First Class 0.027

Second WindTM C3 0.028

ThiesTM Classic 0.021

ThiesTM Compact 0.027

Table V-1: Measured k values of each cup anemometer

Figure V-11 shows the overspeeding curves of all four-cup anemometers at wind speed

of 10 m/sec. The graphs show that ThiesTM Classic Cup Anemometer is less sensitive

than the other three-cup anemometers at responding to the decreasing wind and takes

longer time to decay to zero speed. The mean measurement uncertainty for each cup

anemometer is shown in the graph in m/sec and is calculated based on the combined

uncertainty in the wind speed, uncertainty in the test anemometer and uncertainty in

human error [39].

35
Figure V-11: Overspeed curves for the four anemometers at V = 10 m/sec (* denotes

the mean uncertainty)

The same trend is also observed for the wind speeds of 5 m/sec and 15 m/sec in

Appendix A.

36
VI. OVERSPEEDING ANALYSIS OF UNBALANCED CUP

ANEMOMETERS

According to the IEC 61400-12-1 standard, mechanical anemometers should be

calibrated prior to and after the measurement campaign or after one-year. The cup

anemometers are calibrated and typically installed at heights of 40 or even 60 meters.

As a result it is difficult to frequently check the physical integrity of the instrument after

the installation. A rotational imbalance could occur when dirt has accumulated on the

cup or there is uneven wear of bearings. This generally would not be identified quickly,

if at all, during the measurement campaign. If characteristic ‘signatures’ of unbalanced

anemometers could be identified in the wind speed times series, it might be possible to

detect such abnormalities in the output data in real time. This is the motivation of this

part of the investigation. The objective of the experiments is to assess the effect of

imbalance on overspeeding.

To simulate the unbalanced cups in the laboratory setup, four modifications of the

ThiesTM First Class cup anemometers (Figure VI-1) were tested in the wind tunnel at

different velocities (Table VI-1). To obtain an imbalance, rare earth magnets were placed

on one or two cups to change the weight of the cups, and, thus, obtain an imbalance

that would affect the rotation. Two different magnets were used, 0.578 g and 1.12 g.

ThiesTM manufactured a cup anemometer with a metal ball (2.23 g/each) glued in the

center of each cup to change the weight of the cup without altering its aerodynamics

like other configurations. The moment of inertia of the cup is altered for all cases.

37
Additional weight (g)

One-cup weighted 3.468

One-cup weighted (heavy) 6.72

Two-cup weighted 2.89

Three-cup weighted 2.23

Table VI-1: Various experimental configurations

Figure VI-1: Weighted cup anemometers

The “new” calibration functions for each cup anemometer with various weighted cups

are shown in Figure VI-2. Their differences are hardly noticeable as expected. The

calibration function of one-cup weighted is almost identical with the regular cup

anemometer and the calibration functions of other three conditions are off more (Table

VI-2).

38
25
Regular

20
One-Cup
Weighted
Velocity (m/sec)

15
One-Cup
Weighted
10 (Heavy)
Two-Cup
Weighted
5
Three-Cup
Weighted
0
0 100 200 300 400 500
f (pulse/sec)

Figure VI-2: Calibration function of the Thies™ Classic Cup Anemometers with different

weighted cups

Regular V = 0.0468 * f + 0.4246

One-Cup Weighted V = 0.0468 * f + 0.4258

One-Cup Weighted (Heavy) V = 0.0474 * f + 0.3785

Two-Cup Weighted V = 0.046 * f + 0.4102

Three-Cup Weighted V = 0.0469 * f + 0.426

Table VI-2: Calibration Functions

The overspeeding curves of different weighted cup anemometer are shown in Figure

VI-3. Comparison of a balanced (unweighted) cup anemometer with an unbalanced-cup

anemometer shows that for the lighter weight the overspeeding curves are essentially

the same. When heavier weights (6.72 grams) are used, the one-cup weighted curve

shows noticeable fluctuations. Compared the regular curve with the two-cup weighted

curve and three-cup weighted curve, it is observed that it takes longer time for the cup

39
anemometer to reach the same ending speed. The one-to-one comparison can be found

in Appendix B.

Figure VI-3: Overspeed curves comparison of the Thies™ First Class Cup Anemometers

with different weighted cups at wind speed of 10 m/sec (* denotes “mean total

uncertainty”)

Anemometer k value

Regular 0.027

One-cup Weighted 0.024

Two-Cup Weighted 0.021

Three-Cup Weighted 0.016

One-Cup Weighted (Heavy) 0.023


Table VI-3: Various k values under different configurations

40
Even though curves show only small deviations from the unweighted cup curve, it can

be seen from Table VI-3 above that the k values are different for each weighted case cup

anemometer. The largest deviation is being for the Three-cup weighted case. It takes a

longer time for the three-cup weighted anemometer to stop compared with other

unbalanced cup anemometers. By frequently obtaining k values from field data and

comparing it with the nominal value, the health of the anemometer can be determined.

41
VII. PRELIMINARY FIELD DATA EXAMINATION

Figure VII-1: Measured wind speed vs. time (upper graph) and the inverse of the wind

speed vs. time (bottom graph)

Figure VII-1 is a sample of actual measured wind data (u). From the experiments to

measure the constant k discussed in the previous section it is seen that the wind velocity

measured by an anemometer under a large and rapid decrease in wind speed can be

expressed in the form

(17)

Using the idea that, under conditions when there is a large magnitude and sudden drop

in wind speed the slope of the inverse wind speed time series 1/u should be

approximately k. The 1/u graph shown below the time series for u exhibits

42
three regions where the slope corresponds to k. It is speculated that this approach could

be used to identify regions in wind time series that may have recorded large over

speeding values of the actual wind.

43
VIII. CONCLUSIONS AND FUTURE RESEARCH

The present study has shown that the dynamical properties of the cup anemometer can

result in over or underestimate of the wind speed depending on the behavior of the

wind. In addition, the time constant for the instrument (proportional to 1/k) is shown to

be important. In particular, overspeeding is seen to be a concern in that, following a

gust that would correspond to an increase in wind speed, the recorded speed spends a

‘long’ time at a higher speed as the disturbance decays back to the initial speed. For

turbine power curve measurements, the IEC 61400-12 standard recommends averaging

the wind data over 10-minute periods. For that reason, the 10-minute averaged wind

speed from the wind data that cup anemometer collected is higher than the actual

mean wind speed over this time period. This is results in an overestimation of power

output.

Because the overspeeding curves from the cup anemometer could be used to pinpoint

when the overspeeding happened during the post wind data analysis. It could be used

to identify where the data is corrupted due to the overspeeding during the data post-

processing. If over a certain amount of data in 10 minutes window is corrupted, the data

should be removed.

However, the current experiment assumes there is no vertical component of wind speed

which is too ideal for what really happens in the natural environment. For future work,

the same experiment should be repeated when vertical component of wind speed is

considered. Re-evaluate and compare the overspeeding curves and k values.

44
Even though cup anemometers react to an increase in wind speed more rapidly than for

a decrease in wind speed, the experiment for decreasing wind speed scenario should

also be performed and the curves should be obtained and analyzed.

The further step is to repeat experiments under the turbulent flow conditions. It would

be useful to know the impact of turbulence intensity on anemometer’s dynamic

behavior.

45
Part II

CALCULATION OF THE ACOUSTIC NOISE PRODUCED BY A

NREL PHASE VI WIND TURBINE BLADE

46
I. OVERVIEW AND OBJECTIVES

A wind turbine transforms the kinetic energy in the wind to mechanical energy in a shaft

and converts it into electricity. The power generation directly depends on characteristics

of the wind turbine blades and the interactions between the wind turbine blades and

the wind.

Image: Michael Hoelling

Figure I-1: Atmospheric boundary layer with wind turbine

Wind turbines generate sound in different ways, both mechanical and aerodynamic.

Even though wind turbine technology has advanced of the last few decades and wind

turbines have become much quieter, sound from wind turbines is still one of the most

studied from the point of view of environmental impact and plays an important role in

siting criterion. The challenge for wind turbines to be installed near cities is public

acceptance and acoustic noise from wind turbines remains an issue frequently cited by

opponents of wind energy.

47
Wind turbine noise consists of mechanical noise that includes the noise from the fans,

generator, gear box, etc., and aerodynamic noise which is originated from the

interaction between the rotor and the wind. Manufacturers have been able to

significantly reduce the mechanical noise. Aerodynamic noise remains an issue. This part

of the dissertation focuses on the computational simulation of aerodynamic noise

propagation. The Ffowcs-Williams and Hawkings (FW–H) acoustic analogy is used to

calculate the acoustic pressure at a far field receiver location. The results obtained are

compared with experimental data. The purpose of the investigation is, to predict

aeroacoustic noise at different wind speeds and at different rotational speeds of the

turbine.

Unsteady operating conditions such as variable angles of attack produces acoustic noise

and can also cause erosion or even splitting of the blade along the leading and trailing

edges [40]. Recently, some researchers have used the sound radiated from operating

wind turbines to monitor their structural health [41]. Wind turbine designers continue

to seek noise-reduction concepts that might offset concerns of the public.

Sound measurements of actual wind turbines are impractical to perform at scale due

the size of the wind turbine. Field measurements are also difficult because it is hard to

separate wind turbine noise from other sounds generated by nearby objects. Modeling

appears to be a good option but the numerical simulation of sound generated from

turbulent wind is also difficult. Nevertheless, for certain geometries and idealized flow

situations, modern computational techniques permit the simulation of aerodynamic

noise produced by wind turbines.

48
Wind turbines are operated at different rotational speeds depending on the incoming

wind speeds and angles. Therefore, the magnitude and extent of acoustic noise

emission varies. The current study extends the work of Ranft, Ameri, Kaltschmit, &

Alexander, [15] and focuses on the prediction of aeroacoustic noise generated from

NREL Phase VI Wind Turbine Blade at different incoming wind speeds with various

rotational speeds. The goal of this study is gain a better understanding of the

aeroacoustic noise as a function of wind speed and rotation rate.

The simulations are accomplished using the commercial software ANSYS FLUENT

following the approach of Ranft et al. (2009).

49
II. NOISE AND SOUND FUNDAMENTALS

1. SOUND AND NOISE

Sound is a vibration of the air. It can be generated by a multitude of mechanisms, like a

vibrating surface. Even flows generated by a constant pressure difference can result in

pressure fluctuations that we hear as sound [42]. A vibrating surface is a source that

emits pressure fluctuations and the fluctuations propagate as sound waves. The speed

of sound in air is approximately 340 m/s at standard temperature and pressure. Sound

waves can propagate through any phase of matter. The SI unit for pressure is Pascals

and for sound pressure denotes the force per unit area perpendicular to the direction of

the sound wave propagation. The lowest audible sound pressure for the human ear is 20

mPa (10 billionths of an atmosphere) and occurs in the frequency range 3000-4000 Hz.

Pain is experienced at sound pressures of 60 Pascals. The decibel scale (not an SI unit be

accepted by most national standards organizations) is a logarithmic scale that is a

convenient way to describe sound relative to the lowest audible sound for a normal

human ear. The human ear is less sensitive to low audio frequencies, therefore, decibel

A-Weighting, dB(A), is often used to express of the relative loudness of sounds in air as

perceived by the human ear.

The sound power or acoustic power is the sound energy constantly transferred from the

sound source. A sound source has a given constant sound power that does not change if

it is placed in a different environment. Sound power is a theoretical value that is not

measurable. It is calculated and expressed as sound power level in decibels.

50
A sound source produces sound power and this generates a sound pressure fluctuation

in the air. Sound power is the distance independent cause of this, whereas sound

pressure is the distance-dependent effect. Sound Power is the energy of sound per unit

of time from a sound source and the Sound Power Level (Lw) can be expressed as a

relation to the threshold of hearing (10-12 W) in a logarithmic scale.

𝑁
𝐿𝑤 = 10 𝑙𝑜𝑔10 (𝑁 ) (1)
𝑟𝑒𝑓

where, Lw is the Sound Power level (dB), N is the sound power (W), and Nref = 10-12 W.

The Sound Pressure Level (SPL, Lp) is a property of sound at a given observer location

and is a ratio of the absolute Sound Pressure and a reference level (the lowest human

hearable sound), it can be expressed as

𝑝2
𝐿𝑝 = 10 𝑙𝑜𝑔10 (𝑝2 ) (2)
𝑟𝑒𝑓

where, Lp is the sound pressure level (dB), p is the sound level (Pa), and pref = 2×10-5 Pa.

Any unwanted sound is termed noise. Acoustic noise can be anything from quiet to

annoying to physically harmful. Inadequately controlled noise presents a growing

danger to the health and welfare of the public [43]. Human perception of noise varies.

Whether sound is identified as noise or not, it does not only depends on the sensitivity

of the listener, but also on the sound duration and the surroundings. For examples,

people can sleep in a moving train or a flying jet but get annoyed while hearing the

sound from a dripping faucet.

51
2. NOISE STANDARDS AND REGULATIONS

There are direct links between noise and health [44] [45]. Problems caused by noise

include stress related illnesses, high blood pressure, hearing loss, sleep disruption and

low productivity [43]. Although human perception of sound varies from person to

person, not only could high-decibel damage hearing, but long term-exposure to a low-

decibel environment can also cause damage to hearing (Table II-1).

http://www.hearingcarecentre.co.uk/Info_page_two_pic_2_det.asp?art_id=6052&sec_id=3052

Table II-1: Health and safety executives noise regulation guidelines on exceeding safe

level

There are currently no common international noise standards or regulations. In most

countries, noise regulations defined upper bounds for the noise to which people may be

exposed and it also depends on the areas and time (Table II-2).

52
Table II-2: Recommended noise level standards and guidelines by WHO and selected

countries (Yuen, 2014) [46]

53
III. WIND TURBINE AERODYNAMICS

The majority of wind turbines is horizontal-axis wind turbines (HAWT) [23] and is the

focus of this investigation. .

1. LIFT, DRAG AND MOMENT COEFFICIENTS

There are two forces, lift and drag, and one momentum, pitch, that act on a HAWT wind

turbine blade (Figure III-1). The lift force is perpendicular to direction of the oncoming

air flow and is the result of unequal pressure on the upper and lower airfoil surfaces.

The drag force is parallel to the direction of the oncoming air flow and is due to viscous

friction forces at the surfaces and to unequal pressure on the airfoil surfaces facing

toward and away from the oncoming flow.

Figure III-1: Forces and moment on an airfoil section, α, angle of attack; c, chord [22]

2. TIP SPEED RATIO

The tip speed ratio is the ratio of the blade tip speed over wind speed.

Blade tip speed 𝛺𝑅


𝜆= Wind speed
= 𝑉
(3)

where Ω is the angular velocity (in radians/sec) of the tip of the blade, V is the wind

speed and R is the radius of the wind turbine rotor.

54
Generally a low speed wind turbine has the tip speed ratio value between 1 and 4 and a

high speed wind turbine has the tip speed ratio value between 5 and 9. Burton’s [22]

research suggests that the wind turbine can run at near maximum power coefficient

with the best range of tip speed ratio of 7.

At the same tip speed ratio, a blade with a large span has a low rotational speed.

3. AIRFOIL BEHAVIOR

There are three basic flow regimes associated with airfoils: the attached flow regime,

the high lift/stall development regime and the flat plate/fully stalled regime [22]. In the

attached flow regime, lift increases with the angle of attack. In the high lift/ stall regime,

the lift coefficient increases and reaches the peak when it stalls. When the angle of

attack exceeds a certain value, stall occurs and flow separates on the upper surface. It is

essential to understand blade performance in terms of the airfoil behavior since an

applicable airfoil for wind turbine blade will improve the efficiency.

4. TWISTED BLADE

The blade tips of some modern wind turbines are designed using a thick airfoil for high

lift to drag ratio. If there were no structural requirements, this is how a wind turbine

blade would be proportioned. However, the blade needs to support the lift, drag and

gravitational forces acting on it. The root region of a wind turbine blade is designed

using a thick version of the same airfoil. The crucial factors for choosing airfoil are to

have the maximum lift to drag ratio and low pitch moment. The closer to the tip of the

blade, the faster the blade moves through the air and the greater is the apparent wind

55
angle. Therefore, to optimize the angle of attack along the entire blade, the blade must

twist from root to the tip. In order to achieve the maximum lift and efficiency for some

long blades, the chord length, thickness, twisted angle and the shape of airfoil varied

alone the blade (Figure III-2). These requirements are a challenge for manufacturers.

Figure III-2: Blade Twist

56
IV. NOISE MECHANISMS OF WIND TURBINES

1. WIND TURBINE AERODYNAMIC SOUND GENERATION

The sound generated by wind turbine operation can be divided into four categories:

tonal, broadband, low frequency, and impulsive.

a) TONAL

Tonal sound is defined as sound at discrete frequencies. It is caused by non-

aerodynamic sources such as gears, generators, converters, etc.,) as well as wind

interaction with a rotor blade surface, vortex shedding from a blunt trailing edge,

or unstable flows over holes or slits or a blunt trailing edge. The Department of

Environmental Protection of the State of Maine has, for the purposes of wind

turbine noise assessment defined an occurrence of tonal noise when: “…at a

protected location, the 10 minute one-third octave band sound pressure level in

the band containing the tonal sound exceeds the arithmetic average of the

sound pressure levels of the tow contiguous one-third octave bands by 5dB for

center frequencies between 500 Hz and 10,000 Hz and by 8dB for center

frequencies between 160 Hz and 400 Hz, and 15 dB for center frequencies

between 25 Hz and 125 Hz. (An octave band is a frequency band where the

highest frequency is twice the lowest frequency. A third octave band is a third of

the frequency width of the octave band.)

57
b) BROADBAND

Broadband sound is characterized by a continuous distribution of sound pressure

with frequencies greater than 100Hz. It is often caused by the interaction of

wind turbine blades with atmospheric turbulence, and also described as a

characteristic “swishing or “whooshing” sound.

c) LOW FREQUENCY

Low frequency noise (20 Hz -120 Hz) was the focus of a recent study in Denmark

[47]. It was shown that large wind turbines (> 2 MW) produce low frequency

noise at higher levels than small turbines (< 2 MW).

d) IMPULSIVE

Impulsive noise from wind turbines is characterized by short acoustic impulses or

thumping sounds that vary in amplitude with time. There are variety of proposed

causes ranging from blade-tower interaction, blade noise directivity, variation of

wind speed over the rotor area, and interactions between sound fields from

neighboring turbines [48]. The mechanism for impulsive noise production is not

well understood [49].

2. CLASSIFICATION OF NOISE MECHANISMS

The noise generated from the operating wind turbines can be classified into two types,

mechanical noise and aerodynamic noise. Figure IV-1 summarizes the mechanical and

aerodynamic sound sources of a wind turbine and indicates their respective sound

power levels; a/b refers to airborne noise and s/b refers to structural noise.

58
Figure IV-1: Horizontal axis wind turbine noise sources (Romero-Sanz & Matesanz,

2008)

a) Mechanical noise

Mechanical noise is generated mainly from the metal components moving or

knocking against each other, such as rotating components in the gear box and

the generator, the cooling fans, yaw system, pumps and compressors.

Mechanical noise tends to be more tonal and narrowband that bother humans

more than broadband sound [50]. Therefore, many countries have regulations

which stipulate distance increases between wind turbines and the nearest

buildings. Mechanical noise can be transmitted either through the air or through

the wind turbine structure. In the latter, sound travels along the structure of the

wind turbine then into the surroundings through different surfaces, such as the

nacelle cover or the rotor blades. [51]

59
b) Aerodynamic noise

Aerodynamic noise is the largest source of wind turbine noise. Aerodynamic

sound generally increases with rotor speed. Figure IV-2 shows the six main

regions along the blade that are considered to independently generate acoustic

noise. The noises from these six regions are produced different at various

locations along the blade.

Figure IV-2: Aerodynamic noise sources around a turbine blade due to wind flow, U

(Brooks, Pope, & Marcolini, 1989)

The six sources of aerodynamic noise identified in Figure IV-2 can be divided into

three groups: Low frequency sound, Inflow turbulence and airfoil self-noise

(Table IV-1).

60
Table IV-1: Wind Turbine Aerodynamic Noise Mechanisms (From Wagner, Bareib, &

Guidati, 1996)

3. ACOUSTIC SOURCES

Lighthill [52] [53] proposed that aerodynamic sound sources can be modeled as series of

monopoles, dipoles, and quadrupoles generated by the turbulence in an ideal fluid

region surrounded by a large fluid region at rest. Physically, monopoles result from a

fluctuating mass flow [54]. Dipoles are formed when there are fluctuating forces. When

61
fluctuating Reynolds stresses appear, quadrupoles are formed. It is referred to as

Lighthill’s acoustic analogy (Appendix C).

a) MONOPOLE

A monopole radiates sound equally in all directions and is the simplest acoustic

source (Figure IV-3). In aeroacoustics, monopoles normally result from pulsating

flow. The order of acoustic power of monopole is similar to U3M (U is a typical

flow velocity in the flow field and M is Mach number). Examples include tire, and

compressor noise. One example of a monopole source is a pulsating sphere. The

siren on a moving ambulance is also a monopole sound source.

http://en.wikibooks.org/wiki/File:Mono.GIF#/media/File:Mono.GIF

Figure IV-3: Pressure field produced by a monopole source

b) DIPOLE

A dipole is the superposition of two monopole sources that are out of phase.

This results a sound radiation field such as that shown in Figure IV-4. The

simplest model of a dipole is a small oscillating sphere. In aeroacoustics, dipoles

62
are normally the result of vortex shedding as would occur in turbulent flow over

a rod or cavity. The order of the acoustic power of dipole is similar to U3M3.

http://en.wikibooks.org/wiki/File:Dipole.GIF#/media/File:Dipole.GIF

Figure IV-4: Pressure field produced by a dipole source

c) QUADRUPOLES

A dipole source consists of two opposite monopole sources, a simple quadrupole

source can be obtained by the superposition of two dipole sources of the same

strength that are out-of-phase (Figure IV-5). Quadrupoles arise from turbulence

and the order of the acoustic power of quadrupole is similar to U 3M5. One

example is the jet stream. Depending on the distribution of the dipoles,

quadrupoles can be further classified as longitudinal and lateral. Quadrupole

sources are induced by fluctuating moments or viscous forces.

63
http://en.wikibooks.org/wiki/File:Quadpole.GIF#/media/File:Quadpole.GIF

Figure IV-5: Pressure field produced by a quadrupole source

64
V. NUMERICAL ANALYSIS

1. PHYSICAL MODEL

Ranft et al. (2009) carried out computational simulations of the acoustic noise emitted

by S809 blade, and validated the computations by comparing the results with

experiments carried out at NASA AMES Research Center [55]. The S809 blade airfoil

blade has a baseline span of 5.029 m and the chord lengths vary from 73.7 cm at 25% to

35.6 cm at the blade tip (Appendix E). The numerical model was validated. The same

model is used in this investigation.

2. MESH GENERATION

Two types of meshes were used corresponding to structured and unstructured grids.

Figure V-1: Structured and unstructured grids

A structured grid consists of regular mesh lines while an unstructured grid is identified

by its irregular connectivity. The disadvantage of structured grids is that it takes more

effort to apply it to complex geometries.

65
Figure V-2: An example of unstructured grids

An unstructured grid is more flexible and adapts well to complex geometries. The mesh

can be any shape (Figure V-3). Due to the irregularity of the structures there is, however,

some computational cost.

Figure V-3: Structured grids consisting of triangular, polygonal meshes and orthogonal

meshes

It is usual practice to have a refined mesh close to solid surfaces in the near-field region

so that the flow structures (usually involving large velocity gradients) in these areas can

be sufficiently well resolved. Depending on the type of flow, the mesh can be gradually

coarsened away from boundaries. To reduce computational costs for this study, an

unstructured grid is used for the calculation of the mean flow away from the blade and a

66
structured grid is used in the region close to the blade to permit resolution of the

turbulent flow occurring near the wall.

Figure V-4: Structured grids near the blade

The grid was created using the mesh generator "Pointwise". The computational domain

is extended to 10 blade lengths in all direction and contains approximately 1.5 x 10 6

tetrahedral and prismatic elements as displayed in Figure V-4. The mesh is refined

towards the blade in the x and y directions as shown.

Figure V-5: Grids in the computational domain

67
To be able to compare the simulation results with the experimental results, one blade of

the two-bladed turbine is modeled in a 180° sector with the boundary conditions as

show in Figure V-5. The velocity vectors are set to be perpendicular to x-z place as the

upper stream boundary condition. The ambient pressure of 101,325 Pa is defined by the

pressure outlet boundary condition in z-direction and downstream boundary condition.

A non-slip condition is applied on the wall of the blade (Figure V-6). In this model, the

blade is not moving so a moving frame of reference is set to simulate the rotating wind

turbine blade.

Figure V-6: Boundary conditions

68
3. TURBULENCE MODEL

Figure V-7 shows a schematic of the wind velocity impinging on the blade.

Figure V-7: Wind forces on the blade

Here Vrot is the rotational velocity of the blade, Vw is the incoming wind speed

V  Vrot
2
 Vw2 is the resultant velocity, ω is the angular velocity (RPM).

The Reynolds number is taken to be 𝑅𝑒 = 𝑉∞ 𝐿/𝜐, where L is the blade’s chord length

and υ is the kinematic viscosity. The lowest wind speed used in the simulation is 7 m/sec.

which, for a blade chord of 0.5 m, gives Re = 1.3 x 106. Hence, the flow is assumed to be

turbulent.

There are several turbulence models in FLUENT package [56]. The k-ε model proposed

by W. P. Jones and B. K. Launder is widely used in industry due to its stability and

convergence [57]. This model includes two equations, the k (turbulence kinetic energy)

69
equation and ε (dissipation rate) equation. The model is good for high Reynolds flow

when the flow is fully developed.

The k-ω turbulence model, first introduced by Kolmogorov in 1942 [58], is also a two-

parameter turbulence model with the turbulence kinetic energy term (k) and the

dissipation per unit turbulence kinetic energy (ω). The k-ω turbulence model performs

better at near wall layers than the k-ε turbulence model and has been shown to fit best

for the type of application considered here [59]. To avoid the effects of pressure

induced separation a version of the k-ω model known as Shear Stress Transport k-ω or

SST k-ω turbulence model was employed in this study. The SST k-ω model is a two

equation eddy-viscosity model that combines both k-ε and k-ω turbulence models. The

k-ε model part is in effect in the free stream while the k-ω model part accounts better

for the viscous effects near a wall.

Simulations were performed in a steady state simulation and the Moving Reference

Frame (MRF) option was used to model wind turbine rotation.

4. FAR FIELD NOISE PREDCITION USINF THE FFOWCS WILLIAMS AND

HAWKINGS ACOUSTIC MODEL

There are two steps for computing sound using the FW-H (Ffowcs Williams and

Hawkings, Appendix D) acoustics model in ANSYS FLUENT. First, a time-accurate flow

solution must be obtained so the pressure, velocity and density on the selected source

surface are obtained. Next, sound pressure signals are computed using the source data

collected during the first step at the user-specified receiver locations. For the current

70
study, the model has a single rotating reference frame; therefore a steady state FW-H

model can be used.

The wind turbine used in the actual experiment has a hub height of 16 m. The ground

effects of sound propagation were not taken into account in the computation. Table V-1

& Figure V-8 shows the four receiver locations related to the wind turbine.

Name x- coord. (m) y- coord. (m) z- coord. (m)


Receiver - 1 0 10 -10
Receiver - 2 5 5 -16
Receiver - 3 5 10 -16
Receiver - 4 0 20 -16
Table V-1: Receiver Positions

Figure V-8: Receiver Positions [15]

The wind turbine was set as the acoustic source and four acoustic receivers were

located to measure the noise from near-to far-field using the results of the FW-H

71
acoustic model. Upon completing of the simulation a Fast Fourier Transform (FFT)

analysis was applied to interpret the spectral distribution of the sound pressure level.

5. BROADBAND NOISE SOURCE MODEL

Broadband noise occurs of the frequency range of 5-6 and 10 kHz and includes the

inflow turbulence noise and the airfoil self- noise [60]. In many practical applications

involving turbulent flow, the noise has no distinct tones and the sound energy is

continuously distributed over a broad range of frequencies.

Using the broadband noise source model available in the ANSYS FLUENT software

package enables the quantification of the local contribution to the total acoustic power

level generated by the flow. The source model can be used to extract useful diagnostics

on the noise source to determine which portion of the flow is primarily responsible for

the noise generation. However, in contrast to the FW-H model, the broadband source

model does not predict the sound at receivers.

The acoustic power generated by isotropic turbulence without mean flow is calculated

using Proudman’s formula which has been derived from Lighthill’s acoustic analogy

(quadrupole sources). Proudman’s formula gives the acoustic power (in W/m 3) per unit

volume of isotropic turbulent fluid as

PA    0M t5 (5)

2k
where M t  ,
a0

72
βε is a model constant that is set to 0.1 in ANSYS FLUENT based on the calibration of [61]

using direct numerical simulation of isotropic turbulence. k and ε are constants from the

turbulence model described by Lilley [62].

ANSYS FLUENT can also show acoustic power (Lp), in dB, which is computed from

 P 
L p  10 log A  (6)
P 
 ref 

where Pref = 10-12W/m3 is the reference acoustic power. Proudman’s formula gives an

approximate measure of the local contribution to total acoustic power per unit volume

in a given turbulence field.

73
VI. RESULTS

In this section results are presented first for the sound pressure distribution on the wind

turbine blades using the broadband model and then the far field results using the FW-H

model. The simulations were performed for 5 different wind speeds, 7m/s, 10m/s, 15

m/s, 20 m/s and 25 m/s, at 36, 54, 60 and 72 rpm respectively.

1. FLOW FIELD AND STREAMLINES

The flow field and streamlines at different wind speeds and rotational speeds were

computed using and are presented below. At lower wind speeds (lower angle of attack)

of 72 rpm, the flow is fully attached on the active part of the rotor blade. However,

when the wind speed increases, the central portion of the blade begins to stall. Blade

surface streamlines in Figure VI-1 show that the boundary layer separates at the leading

edge from root to tip. When the wind speed is 25 m/s, the streamlines show it is a fully

detached flow. It is also observed that the boundary layer flow is not only spanwise

from root to tip, but also streamwise from trailing edge to leading edge.

Figure VI-2 exhibits the flow at section z = 5 m is attached at 7 m/sec but starts to

separate at 10 m/sec.

74
10 m/s

25 m/s

Figure VI-1: Streamlines on the blade

Figure VI-2: Streamlines at z=5 m

2. SURFACE ACOUSTIC POWER

The surface acoustic power level in Figure VII-3~6 describes the acoustic power per unit

area generated by boundary layer turbulence on the surface of the blade using the

75
Broadband Sound Source model described in Section V-5. The strength of the sound

sources on the suction sides of the blade increases with higher wind speeds. This

explains why flow separation occurs on the suction side of the blade at higher wind

speeds. It can also be observed that the sound sources on the pressure side move

towards the trailing edge at higher wind speeds. A maximum in the Surface Acoustic

Power Level always occurs at the leading edge of the blade tip which can be explained

by higher velocity and pressure values occurred at the blade top due to the rotation.

76
Figure VI-3: Surface Acoustic Power Level (dB) on blades at 36 RPM

77
Figure VI-4: Surface Acoustic Power Level (dB) on blades at 54 RPM

78
Figure VI-5: Surface Acoustic Power Level (dB) on blades at 60 RPM

79
Figure VI-6: Surface Acoustic Power Level (dB) on blades at 72 RPM

80
The maximum surface acoustic power levels under different conditions are shown in

Figure VI-7. It is observed that the maximum surface acoustic power level increases with

the increasing wind speed in the beginning as expected. At low rotation rates, a local

minimum occurs as the wind speed increases and then the surface power level rises

again.

Figure VI-7: Maximum Sound power levels for wind speeds 7 m/s through 25 m/s at

different rotation speeds

There is currently no way to measure the acoustic noise caused by boundary layer

turbulence on blade surface in the lab or in the field and so validation of the code in this

sense is not possible.

81
3. FAR-FIELD AERODYNAMIC NOISE AT DIFFERENT WIND SPEEDS

The noise radiated to the observer in a far-field is investigated by using the integral

formulation of the Ffowcs Williams and Hawking equation (Section V-4).

The observer locations are shown and listed in Figure V-8 and Table V-1. The simulation

results of aerodynamic noise under different wind speeds at the various rotational

speeds are shown in Table VI-1.

Receiver -1 Receiver-2 Receiver-3 Receiver-4


7 m/s 57.1 50.4 48.4 36.3
10 m/s 59.8 52.5 50.8 38.8
36 rpm 15 m/s 63.4 55.6 54.1 42.2
20 m/s 66.7 58.0 57.2 45.4
25 m/s 69.0 61.3 60.1 48.2
7 m/s 60.7 54.6 52.3 40.4
10 m/s 62.6 56.2 54.0 42.1
54 rpm 15 m/s 66.0 58.8 57.0 45.2
20 m/s 68.3 60.7 59.1 47.4
25 m/s 70.5 62.6 61.2 49.5
7 m/s 61.7 55.8 53.5 41.6
10 m/s 63.9 57.7 55.5 43.7
60 rpm 15 m/s 67.0 60.0 58.2 46.4
20 m/s 69.1 61.7 60.1 48.4
25 m/s 71.3 63.6 62.1 50.4
7 m/s 64.0 58.4 56.0 44.5
10 m/s 66.3 60.3 58.1 46.5
72 rpm 15 m/s 68.6 62.1 60.1 48.5
20 m/s 71.1 64.0 62.3 50.8
25 m/s 72.8 65.3 63.8 52.3
Table VI-1: Sound Pressure Level (dB) of different rpm at receiver locations

It is expected that the emitted Sound Pressure Level (SPL) is louder for higher wind

speeds and rotational speeds. The receiver 4 has the longest liner distance from the

82
noise source (blade), the acoustic noise in the minimum compared with noise at other

receiver locations.

4. SURFACE ACOUSTIC POWER AND FAR-FIELD AERODYNAMIC NOISE

PREDICTION FOR A YAWED TURBINE

Wind direction changes continuously and so it is often the case that the rotor axis of a

wind turbine is not always aligned with the wind direction that is, the wind turbine is

yawed. When a wind turbine is in the yawed condition, even in a steady wind, the angle

of attack on each blade is continuously changing and the loads on the blades are varied.

Figure VI-8: A wind turbine yawed to the wind direction [22]

There are six (6) yawed angles are simulated for a wind turbine at 72 rpm. The Surface

Acoustic Power results are shown in Table VI-2. It is noted that the Surface Acoustic

83
Power is the lowest at lower wind speeds when the wind turbine is not yawed. This is

not the case when the wind speed is at 25 m/sec.

7 m/sec 10 m/sec 15 m/sec 20 m/sec 25 m/sec


-15 Degree 96.6 97.4 95.4 94.0 96.1
-10 Degree 96.9 97.5 95.6 94.7 96.2
-5 Degree 97.0 97.7 95.9 93.7 96.4
0 Degree 94.2 96.0 95.9 95.7 97.1
5 Degree 97.4 98.1 96.6 95.7 95.6
10 Degree 97.4 98.2 95.8 95.1 95.9
15 Degree 97.4 98.8 96.3 94.7 96.7

Table VI-2: Surface Acoustic Power (dB) of different yawed angles at 72 rpm

Figure VI-9 shows the Surface Acoustic Power Level (dB) on blades at different yawed

angles. There is no obvious difference. The maximum Acoustic Power Level always

occurs at the leading edge of the blade tip as expected.

Table VI-3 shows the far-field aerodynamic noise of a yawed turbine at 72 rpm. The

receiver locations are shown in Table V-1 and Figure V-8. Even though the yawed angles

vary; there is no significant difference in Sound Pressure Level at the receiver locations.

84
Figure VI-9: Surface Acoustic Power Level (dB) of 72 rpm at wind speed of 7 m/sec on

blades at different yawed angles

85
Receiver -1 Receiver-2 Receiver-3 Receiver-4
7 m/s 64.2 58.5 56.1 44.6
10 m/s 66.2 60.3 58.0 46.4
15 Degree 15 m/s 68.5 62.1 60.0 48.4
20 m/s 71.1 64.1 62.3 50.8
25 m/s 72.8 65.5 63.9 52.4
7 m/s 64.0 58.4 56.0 44.5
10 m/s 66.3 60.4 58.1 46.5
10 Degree 15 m/s 68.7 62.2 60.1 48.6
20 m/s 71.2 64.1 62.4 50.8
25 m/s 72.9 65.5 63.9 52.4
7 m/s 64.0 58.4 56.0 44.5
10 m/s 66.4 60.4 58.1 46.5
5 Degree 15 m/s 68.7 62.2 60.2 48.6
20 m/s 71.2 64.1 62.4 50.9
25 m/s 72.9 65.5 63.9 52.4
7 m/s 64.0 58.4 56.0 44.5
10 m/s 66.3 60.3 58.1 46.5
0 Degree 15 m/s 68.6 62.1 60.1 48.5
20 m/s 71.1 64.0 62.3 50.8
25 m/s 72.8 65.3 63.8 52.3
7 m/s 63.9 58.3 55.9 44.3
10 m/s 65.8 59.9 57.6 46.0
-5 Degree 15 m/s 68.6 62.1 60.0 48.5
20 m/s 70.6 63.5 61.7 50.2
25 m/s 72.8 65.3 63.7 52.2
7 m/s 63.8 58.2 55.8 44.3
10 m/s 65.6 59.7 57.4 45.8
-10 Degree 15 m/s 68.3 61.8 59.8 48.3
20 m/s 70.5 63.4 61.7 50.1
25 m/s 72.3 64.8 63.2 51.8
7 m/s 63.6 58.1 55.7 44.1
10 m/s 65.3 59.5 57.2 45.6
-15 Degree 15 m/s 68.0 61.6 59.5 48.0
20 m/s 70.1 63.1 61.3 49.8
25 m/s 71.8 64.4 62.8 51.4
Table VI-3: The far-field aerodynamic noise of a yawed turbine at 72 rpm at receiver

locations

86
VII. DISCUSSION

There are not many acoustic measurement experiments involving wind turbine that can

be found in the open literature. The wind turbines tested in the literature have different

sizes and the blades are not representative of an S809 airfoil. Also, the acoustic noise

was measured at a distance from the turbine and not on the blade surface. For that

reason, the experiment results could not be used to quantitatively validate the acoustic

simulation results. Nevertheless a qualitative comparison is useful.

In Figure VII-1 and Figure VII-2, Huskey & Dam (2010) [63] and Rogers, Manwell &

Wright (2006) [64] reports that the sound pressure levels at the receiver locations do

not increase with wind speeds. The sound pressure level displays a minimum at

intermediate wind speeds which is in qualitative agreement with the 36 RPM rotor

speed case shown in Figure VI-7.

100
Sound Power Level (dB)

95

90

85

80

75
4 5 6 7 8 9 10 11 12
Standardized Wind Speed (m/s)

Figure VII-1: Sound power levels as a function of standardized wind speed for the ARE

442 wind turbine (rotor diameter is 7.2 m) (Huskey & Dam, 2010) [63]

87
Figure VII-2: Measured sound pressure levels of a Southwest Whisper 900 wind

turbine (rotor diameter is 2.1 m) (Rogers, Manwell, & Wright, 2006) [64]

In contrast, another field study by Huskey (2011) [65] for which the turbine rotor

diameter is 13 m, sound pressure levels increase with the wind speed (Figure VII-3) and

no decrease in progress is observed. This result is qualitatively similar to the

computational results the different receiver locations shown in Table VI-1.

Figure VII-3: Sound power levels as a function of standardized wind speed for the Gai

Wind 11-kW wind turbine (Huskey A., 2011) [65]

88
VIII. CONCLUSIONS

1. SUMMARY

The noise generated by the S809 airfoil has been investigated using a combination of

computational fluid dynamic simulation and two acoustic models. While a qualitative

comparison with experimental measurements suggests that the computational results

are reasonable, there is no acoustic experiment result of S809 airfoil available to

validate aeroacoustic simulation results.

Both rotational noise and broadband noise were analyzed. As in the earlier study (Ranft

et al., 2009) it was found that the leading edge and the blade tip are the primary source

of the broadband noise (Section VI-2). In all simulations, the wind turbine with the

slowest rotational speed (36 rpm) and lowest incoming wind speed (7 m/sec), the

loudest noises (80.8 dB) generated by the wind turbine blade tip and the leading edge

are still higher than the World Health Organization recommended noise level standards

(Table II-2). An increase of 3dB in Acoustic Power Level will double the sound source

energy. Therefore, the noise will increase with higher wind speed and/or faster rotation.

Modifications of the blade tip and the leading edge are recommended to reduce the

noise generated by a working wind turbine. Modifications of the leading edge could be

made to avoid or delay the flow separation that adds additional sound sources on the

blade surfaces and the separated flow field.

Table VI-2 shows that the sound pressure level at Receiver 4 is below the WHO

recommended noise level (55 dB in daytime) even with the fastest rotational speed (72

89
rpm) and incoming wind speed (25 m/sec). This result could be used as a baseline to

define a zone around the wind turbine that where excessive noise will occur.

2. FUTURE WORK

In this work the wind profile is assumed to be uniform and the interaction of the wind

with the ground is not considered. The effect of the ground will be a velocity profile that

increases with height and, thus, the loading conditions on each blade will be different

depending on their position. The ground can also absorb and reflect sound waves

although this is expected to decrease rather than amplify the propagating sound waves.

The noise produced by the wind turbine tower also needs to be assessed since the flow

is also separated by the tower. This produces wakes behind the tower and low-

frequency noise (Table IV-1) is generated. Three-blade wind turbines are the most

frequently used wind turbines. Future modeling of a three-blade wind turbine would be

useful for comparison.

90
APPENDIX A

Overspeed curves comparison of the Thies™ Classic, Second Wind™ C3, Thies™ First

Class and Thies™ Compact Cup Anemometers at wind speed of 5 m/sec (* indicates

“mean total uncertainty”)

91
Overspeed curves comparison of the Thies™ Classic, Second Wind™ C3, Thies™ First

Class and Thies™ Compact Cup Anemometers at wind speed of 15 m/sec (* indicates

“mean total uncertainty”)

92
APPENDIX B

The overspeed curves comparison of manufactured ThiesTM First Class and one-cup

weighted unbalanced anemometer

The overspeed curves comparison of manufactured ThiesTM First Class and one-cup

(heavy) weighted unbalanced anemometer

93
The overspeed curves comparison of manufactured ThiesTM First Class and two-cup

weighted unbalanced anemometer

The overspeed curves comparison of manufactured ThiesTM First Class and three-cup

weighted unbalanced anemometer

94
The overspeed curves comparison of one-cup weighted and one-cup weighted (heavy)

unbalanced ThiesTM First Class anemometer

The overspeed curves comparison of one-cup weighted and two-cup weighted

unbalanced ThiesTM First Class anemometer

95
The overspeed curves comparison of one-cup weighted and three-cup weighted

unbalanced ThiesTM First Class anemometer

The overspeed curves comparison of one-cup weighted and three-cup weighted

unbalanced ThiesTM First Class anemometer

96
The overspeed curves comparison of two-cup weighted and one-cup weighted (heavy)

unbalanced ThiesTM First Class anemometer

The overspeed curves comparison of three-cup weighted and one-cup weighted (heavy)

unbalanced ThiesTM First Class anemometer

97
APPENDIX C:

LIGHTHILL’S ACOUSTIC ANALOGY

In 1952, Dr. Michael James Lighthill published a paper, “On Sound Generated

Aerodynamically, I. General Theory”. In the paper, he estimated the radiated sound

from a given fluctuating fluid flow with two major assumptions. The first assumption is

that the acoustic propagation of fluctuations in the flow is not considered. The second

assumption is the preclusion of the back-reaction of the sound produced on the flow

field itself. Thus the effects of solid boundaries are neglected.

The Continuity and momentum equations are a fluid can be expressed as:

 ( ui )
 0 (C.1)
t xi

( ui ) ( ui u j ) p  ij
    g i (C.2)
t x j xi x j

p is the statistic pressure of the flow field,  ij stands for the stress tensor referred to as

the momentum j transported by fluid particles in the direction i per unit time and unit

area. It is defined for a Newtonian fluid as

 ui u j  2
 ij         ij u k (C.3)
 x j xi  3 xk

where  ij is the Kronecker delta function and μ the dynamic viscosity.

(C.1) multiplied by ui

 ( ui )
uj  uj 0 (C.4)
t xi

98
Adding (C.4) to the momentum equation (C.2) leads to

 u j ( ui ) u j p  ij
uj   uj  ui   
t t xi xi x j xi
(C.5)
 ( u j )  ( ui u j )

t xi

Separate the variables and merge terms, it becomes

( u j ) Pij
 (C.6)
t xi

    0 
Add c 2 to (C.6) then subtract
xi

with Pij  ui u j  p ij   ij

For an isentropic condition,

 dp 
c 2    (C.7)
 d  is

c: the speed of sound

Substitute p with either c2(ρ-ρ0) or (p-p0) and the molecular momentum transport

assumes to be zero if it is an ideal fluid.

 ij  0 (C.8)

~
P ij  ( p  p0 ) ij (C.9)

~
P ij  c 2 (    0 ) ij (C.10)

The Lighthill tensor, Lij, is the external force acting on the fluid

 
Lij  ui u j   p  p0   c 2   0   ij   ij (C.11)

99
    0 
The acoustic wave equation is derived from (C.6). So add c 2 to both sides of
xi

(C.6) and replace P ij  ( p  p0 ) ij . Equation (C.6) becomes

 u j      0  Lij
 c2  (C.12)
t xi xi

In order to find an inhomogeneous form of the wave equation, differentiate equation

(C.1) with respect to the time t

 2   2 ui
 0 (C.13)
t 2 xi t

Differentiated (C.12) with respect to xj

 2 u j  2    0   2 Lij
 c2  (C.14)
x j t xi x j xi x j

Substract (C.13) from (C.14)

2     0 
2 2
 2 Lij
c  (C.15)
t 2 xi x j xi x j

This is the famous Lighthill equation.

It can also be rewritten in terms of pressure perturbations as

1  2 p  2 ( p  po )  2 Lij
  (C.16)
c02 t 2 xi2 xi x j

The right hand side represents the source term, where Lighthill’s stress tensor, Lij, is. The

left hand side described the propagation of the acoustic wave in both the spatial and

temporal domains. The equation was used to predict aircraft jet noise originally.

100
APPENDIX D:

FFOWCS WILLIAMS AND HAWKINGS [66]

Ffowcs Williams and Hawkings extended Lighthill’s equation and applied the motion of a

moving surface into the theory. The Ffowcs Williams-Hawkings (FW-H) equation does a

very good job of calculating the airfoil self-noise of a moving body. There are three

source terms in FW-H equation that represent, quadrupole, monopole and dipole

sources.

 2 ( H ) 2  ( H )
 2 ( HLij )     
 Pij  ( f ) f      0 u i  ( f ) f 
2
 c   
t 2
xi
2
xi x j xi  x j  t  xi 
   
         
Quadrupole Dipole Monopole

Where

0, 𝑓 < 0
𝐻(𝑓) = { Heaviside function
1, 𝑓 ≥ 0

1, 𝑓 = 0
𝛿(𝑓) = { Dirac delta function
0, 𝑒𝑙𝑠𝑒𝑤ℎ𝑒𝑟𝑒

Lij: Lighthill’s tensor

For a Newtonian fluid,

 u u j 2uk 
Pij  p ij    i    ij 
 x j xi 3xk 

101
APPENDIX E:

S809 BLADE CHORD AND TWIST DISTRIBUTIONS [55]

1. The blade radius is modified by changing the tip piece.

102
2. Twist convention is positive towards feather. Values listed are relative to zero twist at the 3.772-

m station [75% span on a 5.029-m blade]. Twist is 2.5 degree toward stall at the tip [on a 5.532-m

blade].

3. Each blade attaches to the hub at a point 0.508 m from the center of rotation.

4. There is a cylindrical section at the root that extends from 0.508 to 0.883 m. The airfoil transition

begins at approximately the 0.883-m radial station.

5. There is a transition from the cylindrical section to the S809 airfoil along the 0.883- to 1.257-m

region. The transition ends with a 0.737-m chord S809 airfoil at the 1.257-m span station.

103
BIBLIOGRAPHY

[1] "BP Statistical Review of World Energy June 2015," 2015.

[2] I. Capellan-Perez, M. Mediavilla, O. Carpintero and L. J. Miguel, "Fossil Fuel Depletion and

Socio-Economic Scenarios: An Integrated Approach," Energy, pp. 641-666, 2014.

[3] S. Shafiee and E. Topal, "When Will Fossi Fuel Reserves Be Diminished?," Energy Policy, pp.

181-189, 2009.

[4] A. Kharrazi, M. Sato, M. Yarime, H. Nakayama, Y. Yu and S. Kraines, "Examining the

Resilence of National Energy Systems: Measurements of Diversity in Production-Based and

Consumption-Based Electricity in the Globalization of Trade Networks," Energy Policy, pp.

455-464, 2015.

[5] S. Pfenninger and J. Keirstead, "Renewables, Nuclear, or Fossil Fuels? Scenarios for Great

Britain's Power System Considering Costs, Emissions and Energy Security," Applied Energy,

pp. 83-93, 2015.

[6] E. Haas, "Renewable Energy Data Book," DOE/GO-102015-4724, 2015.

[7] "Renewable Energy Capacity Statistics 2015," International Renewable Energy Agency

Report, 2015.

[8] L. Fried, L. Qiao, S. Sawyer and S. Shukla, "Global Wind Report, Annual Market Update

2014," Global Wind Energy Council, 2015.

[9] Global Wind Energy Council, "Global Wind Report 2013 - Annual market update".

[10] "Annual Energy Outlook 2015 with Projections to 2040," United States Department of

Energy, Energy Information Adminstration, DOE/EIA-0383, 2015.

104
[11] R. Wiser and M. Bolinger, "Wind Technologies Market Report," Department of Energy,

DOE/GO-102015-470271522014, 2015.

[12] M. Wolsink, "Wind Power and the NIMBY-myth: Institutional Capacity and the Limited

Significance of Public Support," Renewable Energy, pp. 49-64, 2000.

[13] "http://www.usatoday.com/money/industries/energy/2008-11-03-windturbines_N.htm".

[14] N. Pinder, "Mechanical Noise from Wind Turbines," Wind Energy, pp. 158-167, 1992.

[15] J. I. D. Alexander, A. Ameri, M. Kaltschmit and K. Ranft, Acoustic Analysis of the NREL Phase

VI Wind Turbine, 2009.

[16] J. Patterson, "The Cup Anemometer," Trans.Roy. Soc. Canada, pp. 1-54, 1926.

[17] P. Coppin, "Cup Anemometer Overspeeding," Meteorol Rdsch, pp. 1-11, 1982.

[18] L. Kristensen, "Cup Anemometer Behavior in Turbulent Environments," J. Atmos. Ocean.

Technol., pp. 5-17, 1998.

[19] N. Busch and L. Kristensen, "Cup Anemometer Overspeeding," Journal of Applied

Meteorology, pp. 1328-1332, 1976.

[20] T. Robinson, "On the Theory of the Rob Anemometer, and the Determination of its

Constants," Irish Academy, 1855.

[21] G. . P. Srivastava, Surface Meteorological Instruments and Measurement Practices, Atlantic,

2008.

[22] T. Burton, D. Sharpe, N. Kenkins and E. Bossanyi, Wind Energy Handbook, Chichester: John

Wiley & Sons Ltd., 2001.

[23] J. F. Manwell, J. G. McGowan and A. I. Rogers, Wind Energy Explained, John Wiley & Sons

Ltd., 2009.

105
[24] "Basics of Turbulent Flow," 2005. [Online]. Available:

http://www.mit.edu/course/1/1.061/OldFiles/www/dream/SEVEN/SEVENTHEORY.PDF.

[25] Á. Sanz-Andrés, S. Pindado and F. Sorribes-Palmer, "Mathematical Analysis of the Effect of

Rotor Geometry on Cup Anemometer Response," The Scientific World Journal, 2014.

[26] I. Strangeways, Back to basics: The ‘met. enclosure’: Part 6 — Wind. Weather, Blackwell

Publishing Ltd, 2001.

[27] O. Schrenk, "Errors due to inertia with cup anemometers in fluctuating winds," NZ Tech.

Phys, pp. 57-77, 1929.

[28] Y. P. Solov'ev, A. I. Korovushkin and Y. N. Toloknov, "Characteristics of A Cup Anemometer

and A Procedure of Measureing the Wind Velocity," Physical Oceanography, pp. 173-186,

2004.

[29] E. Kaganov and A. Yaglom, "Errors in wind speed measurements by rotation anemometers,"

Boundary-Layer Meteorology, Vol. 10, pp. 15-34, 1976.

[30] T. Lutz and A. Sayers, "The Effect of the Boundary Layer Present in Wind Tunnels on the

Aerodynamic Drag of a Model Truck," University of Cape Town, South Africa, 1997.

[31] S. Pindado, J. Cubas and F. Sorribes-Palmer, "The Cup Anemometer, a Fundamental

Meteorological Instrument for the Wind Energy Industry," in International Electronic

Conference on Sensors and Applications, 2014.

[32] L. Makkonen, P. Lehtonen and L. Helle, "Anemometry in icing conditions," Journal of

Atmospheric and Oceanic Technology, vol. 18, pp. 1467-1469, 2001.

[33] L. Makkonen and L. Helle, "Calibration of anemometers - An uncertainty in wind energy

resource assessment.," in In Proceedings of the Fifth European Wind Energy Association

Conference, Thessaloniki, Greece, 1994.

106
[34] "MEASNET. Anemometer Calibration Procedure, Version 2," MEASNET, Madrid, Spain,

2009.

[35] "Standard Test Method for Determining the Performance of a Cup Anemometer or

Propeller Anemometer (ASTM D 5096-02)," ASTM International, West Conshohocken, PA,

2002.

[36] IEC, "Wind Turbine Generator Systems, Part 12: Wind Turbine Power Perfrmance Testing.

International Standard 61400-12-1," International Electrotechnical Commission, 1998.

[37] S. Pindado, J. Cubas and Á. Sanz-Andrés, "Aerodynamic Analysis of Cup Anemometers

Performance: The Stationary Harmonic Response," The Scientific World Journal, 2013.

[38] S. Pindado, E. Vega, A. Martínez, E. Meseguer, S. Franchini and I. P. Sarasola, "Analysis of

calibration results from cup and propeller anemometers. Influence on wind turbine Annual

Energy Production (AEP) calculations," Wind Energy, pp. 119-132, 2011.

[39] R. Coquilla and J. Obermeier, "Calibration Uncertainty Comparisons Between Various

Anemometers," in American Wind Energy Association, Windpower 2008, 2008.

[40] K. Gharali and D. A. Johnson, "Numerical modeling of an S809 airfoil under dynamic stall,

erosion and high reduced frequencies," Applied Energy, 2011.

[41] C. Niezrecki, P. Poozesh, K. Aizawa and G. Heilmann, "Wind Turbine Blade Health

Monitoring using Acoustic Beamforming Techniques," in Acoustical Society of America

Meeting, Providence, RI, 2014.

[42] J. Ffowcs-Williams and M. J. Lighthill, Film Notes for the Aerodynamic Generation of Sound,

Chicago, Illinois: Encyclopedia Britannica Educational Corporation, 1971.

[43] United States Environmental Protection Agency, "Noise Control Act," 1972.

107
[44] S. A. Stansfeld and M. P. Matheson, "Noise Pollution: Non-Auditory Effects on Health,"

British Medical Bulletin, vol. 68, pp. 243-257, 2003.

[45] D. Halperin, "Environmental Noise and Sleep Disturbances: A Threat to Health?," Sleep

Science, no. 7, pp. 209-212, 2014.

[46] F. K. Yuen, "A Vision of the Environmental and Occupational Noise Pollution in Malaysia,"

Noise & Health, vol. 1673, pp. 427-436, 2014.

[47] H. Moller and C. Pedersen, "Low-Frequency Noise from Large Wind Turbines," Journal of

the Acoustical Society of America, pp. 3727-3744, 2011.

[48] R. Tonin, "Sources of Wind Turbine NOise and Sound Propagation," Acoustics Australia, vol.

40, pp. 20-27, 2012.

[49] S. Oerlemans and J. Schepers, "Prediction of Wind Turbine Noise and Validation Against

Experiment," International Journal of Aeroacoustics, vol. 6, pp. 555-584, 2009.

[50] H. Klug, "Noise from Wind Turbines: Standards and Noise Reduction Procedure," Sevilla,

Spain, 2002.

[51] I. Romero-Sanz and A. Matesanz, "Noise management n odern wind turbines," Wind

Engineering, no. 32, pp. 27-44, 2008.

[52] M. J. Lighthill, "On Sound Generated Aerodynamically. I. General Theory," Proceeding of the

Royal Society A: Mathematical, Physical and Engineering Sciences, vol. 211, pp. 564-587,

1952.

[53] M. J. Lighthill, "On Sound Generated Aerodynamically. II. Turbuence as a Source of Sound,"

Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, vol.

222, pp. 1-32, 1954.

108
[54] S. Wagner, R. Bareib and G. Guidati, Wind Turbine Noise, Berlin: Springer, 1996.

[55] M. M. Hand, D. A. Simms, L. J. Fingersh, D. W. Jager, J. R. Cotrell, S. J. Schreck and S. M.

Larwood, "Unsteady Aerodynamics Experiment Phase VI: Wind Tunnel Test Configurations

and Available Data Campaigns," 2001.

[56] "FLUENT 14.0 User's Guide," ANSYS Inv., 2011.

[57] B. E. Launder and D. B. Spalding, Lectures in Mathematical Models of Turbulence, Academic

Press, 1972.

[58] A. N. Kolmogorov, "Equations of Turbulent Motion of an Incompressible Fluid," Izvetia

Academy of Sciences; USSR; Physics, pp. 56-58, 1942.

[59] Y. L. Chang, S. L. Yang and O. Arici, "Flow Field Computation of the NREL S809 Airfoil using

Various Turbulence Model," ASME Energy Wee-96 Book VIII, pp. 172-178, 1996.

[60] R. Camussi, Noise Sources in Turbulent Shear Flows: Fundamentals and Applications,

Springer, 2013.

[61] S. Sarkar and M. Y. Hussaini, "Computation of the Sound Generated by Isotropic

Turbulence," NASA Contract Report 93-74, NASA Langley Research Center, Hampton, VA,

1993.

[62] G. M. Lilley, "The radiated noise from isotropic turbuence revisited," NASA Langley

Research Center, Hampton, VA, 1993.

[63] A. Huskey and J. v. Dam, "Wind Turbine Generator System Acoustic Noise Test Report for

the ARE 442 Wind Turbine," National Renewable Energy Laboratory, 2010.

[64] A. L. Rogers, J. F. Manwell and S. Wright, "Wind Turbine Acoustic Noise," Renewable Energy

Research Laboratory, 2006.

109
[65] A. Huskey, "Wind Turbine Generator System Acoustic Noise Test Report for the Gaia Wind

11-kW Wind Turbine," National Renewable Energy Laboratory, 2011.

[66] J. F. Williams and D. Hawkings, "Sound Generation by Turbuence and Surface in Arbitrary

Motion," Philosophical Transactions of the Royal Society, pp. 321-342, 1969.

[67] J. Wyngaard, J. Bauman and R. Lynch, "Cup Anemometer Dynamics," Proc. Flow, Its

Measurements and Control in Science and Industry, pp. 701-708, 1974.

[68] L. Kristensen, "Can A Cup Anemometer 'Underspeed'?," Boundary-Layer Meteorology, pp.

163-172, 2002.

[69] S. Yassine, "Cup Anemometer Overspeeding," 2011.

110

Vous aimerez peut-être aussi