Você está na página 1de 113

The Electrical and Mechanical 

Performance Evaluation of a 
Roof‐Mounted, One‐Kilowatt 
Wind Turbine 
 
 
 
 
 
 
PREPARED BY: 
Matthew Seitzler  
California Wind Energy Collaborative 
University of California, Davis 
 
 
DATE: 
March 2009 
 
 
REPORT NUMBER: 
CWEC‐2009‐003 
 
SUMMARY

Experimentation and theoretical analysis was performed on the roof-mounted, Small

Wind Energy Demonstration System at the University of California, Davis. This work

consisted of the performance prediction and experimental measurement of the electrical

and mechanical performance of the Bergey XL.1, a 1 kW, horizontal axis, tower-mounted

small wind turbine. The performance prediction entailed the use of Blade Element

Momentum Theory computational methods; previously published wind turbine efficiency

data; and the use of the commercially available Wind Turbine Performance Analysis

software. Experimentation consisted of the measurement of meteorological conditions,

turbine electrical characteristics, turbine/support tower mechanical loads, and site specific

turbulence measurements. Specific measurements were made of electrical power, rotor

thrust, and turbine rotor speed as well as the non-dimensionalized values of coefficient of

power and coefficient of thrust. The methodologies employed produced the desired

results for electrical performance; however, for the mechanical load assessment several

discrepancies were found which were related to both the thrust model assumptions and

tower load calibration factors. Findings included strong correlation between the modeled

values and measured values for electrical performance and previously published data on

the specific wind turbine used, as well as preliminary data on rotor thrust.

ii
TABLE OF CONTENTS

SUMMARY ...................................................................................................................... II
TABLE OF CONTENTS ...............................................................................................III
LIST OF FIGURES ......................................................................................................... V
NOMENCLATURE...................................................................................................... VII
CHAPTER 1- INTRODUCTION ................................................................................ 1
1.0 Small Wind Energy Overview ........................................................................ 1
1.1 Experimentation Project Background ............................................................. 4
1.2 Research Scope ............................................................................................... 6
1.3 Contribution of Research ................................................................................ 7
CHAPTER 2- EXPERIMENT EQUIPMENT & MATERIALS .............................. 8
2.0 System Overview ............................................................................................ 8
2.1 Site Selection ................................................................................................ 12
2.2 Wind Turbine ................................................................................................ 13
2.3 Support Tower .............................................................................................. 16
2.4 Electrical System .......................................................................................... 18
2.5 System Instrumentation ................................................................................ 21
2.6 Data Acquisition System............................................................................... 23
2.7 Turbine Customization.................................................................................. 25
CHAPTER 3- EXPERIMENTAL METHODS ........................................................ 28
3.0 Experimentation Overview ........................................................................... 28
3.1 Electrical System Measurement Methods..................................................... 29
3.1.1 Turbine RPM Measurement...................................................................... 29
3.1.2 Turbine Power........................................................................................... 32
3.1.3 Load Bank................................................................................................. 32
3.1.4 Bergey PowerCenter® Overview ............................................................. 33
3.2 Mechanical Load Experimentation Methods ................................................ 35
3.2.1 Thrust Experiment Overview.................................................................... 35
3.2.2 Thrust Experiment Mathematics............................................................... 36
3.2.3 Thrust Experiment Corrections................................................................. 39
3.2.4 Tower Modification .................................................................................. 41
3.2.5 Load Cell Calibration................................................................................ 43
3.3 Data Collection ............................................................................................. 45
3.4 Performance Calculation Methods................................................................ 46
3.4.1 Data Normalization................................................................................... 46
3.4.2 Data Sorting and Final Computations....................................................... 47
3.4.3 Final Computations................................................................................... 47
3.4.4 Uncertainty Evaluation ............................................................................. 48
3.5 Turbulence Effects ........................................................................................ 51
CHAPTER 4 – PERFORMANCE PREDICTION METHODS ............................. 55
4.0 Airfoil Characteristics................................................................................... 55
4.1 XFOIL........................................................................................................... 56

iii
4.2 Blade Elemental Momentum Analysis ......................................................... 57
4.3 Wind Turbine Performance Analysis, WT_Perf........................................... 57
4.4 Experimental Corrections ............................................................................. 58
4.5 Computational Results .................................................................................. 59
CHAPTER 5- EXPERIMENTAL RESULTS .......................................................... 66
5.0 Turbine RPM Measurement.......................................................................... 66
5.1 Turbine Power and Coefficient of Power ..................................................... 69
5.2 Rotor Thrust and Coefficient of Thrust ........................................................ 73
CHAPTER 6- EXPERIMENT AND MODEL DISCUSSION ................................ 77
6.0 Turbine Power and Coefficient of Power ..................................................... 77
6.1 Rotor Thrust and Coefficient of Thrust ........................................................ 82
CHAPTER 7- CONCLUSION ................................................................................... 86
7.0 Concluding Remarks..................................................................................... 86
7.1 Experimental Improvements......................................................................... 87
7.2 Areas of Further Study.................................................................................. 88
REFERENCES................................................................................................................ 90
APPENDIX...................................................................................................................... 92
APPENDIX A- TURBINE MANUFACTURER SPECIFICATIONS........................................ 93
APPENDIX B- TURBINE SPINDLE LOAD FEA INPUTS AND RESULTS .......................... 94
APPENDIX C- THRUST CORRECTION FACTOR DETAILS ............................................. 95
APPENDIX D- TOWER LOAD CALCULATIONS ............................................................ 101
APPENDIX D- TOWER LOAD CALCULATIONS CONTINUED ....................................... 102
APPENDIX E- WT_PERF SIMULATION INPUT PARAMETERS .................................... 103
APPENDIX E- WT_PERF SIMULATION INPUT PARAMETERS CONTINUED ............... 104

iv
LIST OF FIGURES
Figure 1. Global wind energy generation capacity trends. Source: International Energy
Agency, Key World Statistics 2007 _________________________________________ 1
Figure 2. US projected wind energy growth under 20% by 2030 program. Source: U.S.
Dept. of Energy. ________________________________________________________ 2
Figure 3. A residential small wind-solar hybrid energy system. Source: Southwest
Windpower.____________________________________________________________ 3
Figure 4. Small Wind Energy Demonstration System on the rooftop of Bainer Hall on
the University of California, Davis main campus. ______________________________ 5
Figure 5. Elevation view of the UC Davis, Small Wind Energy Demonstration System. 8
Figure 6. Stand-alone small wind power system with power inverter for alternating
current (AC) loads. ______________________________________________________ 9
Figure 7. Diagram of the UC Davis Small Wind Energy Demonstration System
configured with resistive load bank. ________________________________________ 10
Figure 8. System monitoring and control station located in the room adjacent to the
turbine. ______________________________________________________________ 11
Figure 9. Satellite view of the demonstration system location on the roof of Bainer Hall,
at the University of California, Davis. Source Google Earth. ____________________ 12
Figure 10. Bergey Windpower, 1 kW, XL.1, horizontal, auto-furling wind turbine shown
mounted on tubular tower. _______________________________________________ 13
Figure 11. Power curve for the Bergey XL.1 turbine. Included are the wind speeds for
cut-in, rated power, and auto-furling. Source: Supplied by Bergey Windpower ______ 14
Figure 12. Bergey XL.1 shown furling at high wind conditions. _________________ 15
Figure 13. Turbine/tower assembly in mid-lowering position.___________________ 16
Figure 14. Solar powered winch for easy and reliable raising and lowering of
turbine/tower assembly. _________________________________________________ 17
Figure 15. Electrical wiring diagram of the Small Wind Energy Demonstration System.
_____________________________________________________________________ 19
Figure 16. Resistive load bank located at the base of the turbine used to provide constant
loads for performance evaluation.__________________________________________ 20
Figure 17. Turbine instrument locations. ___________________________________ 21
Figure 18. Wind turbine data acquisition and control, real-time system interface ____ 23
Figure 19. Data acquisition system block diagram. ____________________________ 24
Figure 20. Yaw position sensor installation shown on turbine frame with nacelle cover,
tail boom, blades, and nose cone removed. __________________________________ 25
Figure 21. FEA results of modified turbine spindle solid model, using manufacturer’s
thrust load values. ______________________________________________________ 26
Figure 22. View of turbine with modified nacelle cover. _______________________ 27
Figure 23. Turbine alternator output voltage signal shown with AC ripple content within
rectified DC signal. The signal above was measured across the resistive load bank at a
sampling rate of 360 Hz._________________________________________________ 30
Figure 24. AC coupled turbine output signal with zoom of ripple content. Oscilloscope
readings of AC signal frequency are also shown.______________________________ 31
Figure 25. PowerCenter load resistance and experimental load resistances vs. electrical
power. Source: Bergey Windpower ________________________________________ 34

v
Figure 26. Simplified turbine tower support model. ___________________________ 36
Figure 27. Free-body diagram of simplified turbine tower assembly.______________ 38
Figure 28. Top view of turbine tower with horizontal forces. ____________________ 39
Figure 29. Support tower base detail showing tilting and non-tilting axis locations.__ 40
Figure 30. Solid model of secondary guy wire attachment bracket installed on tower. 41
Figure 31. UBC wind load analysis results for tower bending moments. ___________ 42
Figure 32. Load cell location shown inline with guy wire with turbine in lowered
position.______________________________________________________________ 43
Figure 33. Load cell calibration set-up shown with hanging test load. _____________ 44
Figure 34. Raw wind speed values over a 500 second data sampling interval. _______ 52
Figure 35. Measured cross-section of the Bergey XL.1 blade.___________________ 56
Figure 36. Bergey XL.1 alternator efficiency data used in performance simulations.
Source: Martinez et al7 __________________________________________________ 58
Figure 37. Lift and drag curves for the XL.1 airfoil section at a Reynolds number of
500,000. Source: Kamisky9. ______________________________________________ 59
Figure 38. Turbine power vs. freestream wind speed using WT_Perf, BEM, and
Manufacturer’s data. ____________________________________________________ 60
Figure 39. Coefficient of power vs. wind speed for both WT_Perf and BEM models._ 61
Figure 40. Coefficient of power vs. tip speed ratio of the XL.1 rotor from WT_Perf, UC
Davis BEM, and the alternator efficiency corrected UC Davis BEM analysis. _______ 62
Figure 41. Turbine WT_Perf and BEM thrust model data for maximum rpm. _______ 63
Figure 42. Coefficient of thrust vs. wind speed using BEM model. _______________ 64
Figure 43. WT_Perf and BEM, Ct vs. TSR model data. ________________________ 65
Figure 44. Turbine power output versus RPM for various loading conditions._______ 67
Figure 45. Average turbine rotor RPM vs. wind speed for 2.1 and 8.1 ohm loads. ___ 68
Figure 46. Raw electrical power vs. wind speed data.__________________________ 69
Figure 47. Measured power curves for the XL.1 under high, medium, and low loading
conditions, including manufacturer’s supplied power curve. _____________________ 70
Figure 48. Average coefficient of power (Cp) versus binned wind speed. __________ 71
Figure 49. Experimental coefficient of power vs. tip speed ratio results. ___________ 72
Figure 50. Raw thrust vs. wind speed data. __________________________________ 73
Figure 51. Average thrust, in lbs, vs. wind speed for 1.56 ohm load. ______________ 74
Figure 52. Average coefficient of thrust vs. binned wind speed. _________________ 75
Figure 53. Average coefficient of thrust vs. binned TSR. _______________________ 76
Figure 54. Comparison of Bergey XL.1 electric power curves. __________________ 78
Figure 55. Comparison of measured and predicted electric power coefficient for the
Bergey XL.1.__________________________________________________________ 79
Figure 56. Average coefficient of power vs. tip speed ratio (TSR) compared with
Clarkson and BEM model data. ___________________________________________ 80
Figure 57. Average turbine thrust vs. binned wind speed compared with BEM model
data._________________________________________________________________ 82
Figure 58. Average coefficient of thrust (Ct) and BEM model data vs. binned wind
speed. _______________________________________________________________ 84
Figure 59. Average coefficient of thrust and BEM data vs. binned TSR. ___________ 85

vi
NOMENCLATURE
MW Megawatt = 106 Watts
kW Kilowatt = 103 Watts
kWh Kilowatt Hour
CWEC California Wind Energy Collaborative
PIER Public Interest Energy Research
NREL National Renewable Energy Laboratory
NWTC National Wind Technology Center
UC University of California
PWM Pulse Width Modulation
BEM Blade Elemental Momentum
IEC International Electrotechnical Committee
DC Direct current
AC Alternating current
TTL Transistor-Transistor Logic
DAQ Data acquisition and control
RPM Revolutions Per Minute
A Area swept by a turbine rotor [meter2]
fm Turbine Output Frequency [hz]
PT Number of Alternator Poles
PR Number of Pulses or Diodes in the rectification circuit
emf Electromagnetic Force
FTi Thrust Force Parallel to Guy Wires [lbs]
TForce Total Thrust Force [lbs]
RTi Load Cell Force [lbs]
Өi Guy Wire Cable Angle [degress]
L Height of Attachment Bracket [in]
S Bending Arm Length of Turbine to Support [in]
PRotor Aerodynamic power captured by the turbine rotor [Watts]
VTurbine Turbine Output Voltage [V]
RLoad Load Resistance [ohms]
U Freestream Wind Speed [m/s]
PWind Total power in the wind [Watts]
Cp Coefficient of Performance = PTurbine / PWind

vii
Ct Coefficient of Thrust
ρ Air density [kg/m3]
U Free stream wind velocity [m/s]
Ω Rotational Speed [rad/s]
R Rotor Radius [m]
ui Initial Design Stage Uncertainty
UD,i Total Design Stage Uncertainty
S,i Data Reduction Uncertainty
Uni Total Measurement Uncertainty
U Instantaneous Wind Speed [m/s]
Ū Average Wind Speed [m/s]
*
U Turbulent Fluctuation in Wind Speed [m/s]
σ Standard Deviation of Wind Speed
I Turbulence Intensity

viii
1

CHAPTER 1- INTRODUCTION

1.0 Small Wind Energy Overview

As humanity becomes increasingly reliant upon sources of renewable energy, wind

energy – in particular – is poised to play a significant role in the global energy puzzle. As

depicted by Figure 1, wind energy is increasingly being used worldwide, and in 2007

alone, the global wind energy conversion capacity grew by 27%.

Figure 1. Global wind energy generation capacity trends. Source: International Energy

Agency, Key World Statistics 2007

With over 15,616 megawatts (MW) in capacity in the US and 94,000 MW in capacity

installed worldwide, as of 20072,1, the wind energy industry has demonstrated its

potential to be a long-standing and potent part of the energy sector. Within the United

States, the Department of Energy outlined an ambitious project to have over 20% of

country’s electricity provided by wind energy by the year 2030 further emphasizes the

presence that wind energy will have in our future3. Projections for growth in the wind
2

energy generation capacity of the United States (as determined by this project) are shown

in Figure 2, below.

Figure 2. US projected wind energy growth under 20% by 2030 program. Source: U.S. Dept. of Energy.

Here, Figure 2 shows that the United States wind energy sector must undergo large and

sustained growth throughout the next two decades in order to meet the project goal of

20% wind energy electricity generation by 2030.

Today, the American wind energy sector includes applications that range from

small scale systems (100 kilo watt (kW) and below) to the more common large scale

systems (100 kW and above).

Although much of the projected growth in wind energy capacity is estimated to

come from the deployment of large wind systems, an important amount will be provided

by the small wind sector. In 2007 the small wind energy sector had a small, but not

insignificant market share of 0.22% of the cumulative US wind energy sector production

capacity.4 Between 2006 and 2007 the small wind market grew 14% with a cumulative
3

installed capacity of 55-60 MW.4 Despite this modest market share, small wind energy

systems offer costs of energy that are competitive4. Small wind systems have costs of
4
energy of $0.10-$0.15 $/kilowatt-hour (kWh) compared to $0.15 $/kWh for

photovoltaics and approximately $0.10 $/kWh for natural gas fired power plant,

depending on the current price of natural gas.5 Combining operational costs with the cost

of installation (being $3-5/watt for small wind versus $6-9/watt for solar4) small wind

systems make economic sense in areas where the wind resource is available.

Figure 3. A residential small wind-solar hybrid energy system. Source: Southwest

Windpower.

Unlike utility scale turbines used in large wind farms, small turbines are usually

installed individually at sites where their power will be consumed. An emerging niche
4

for small wind energy systems is the deployment of small systems (with a rated power of

several kW or less) in urban areas (e.g. Figure 3) usually atop buildings or in open areas

of the urban environment. These rooftop/urban installations in the US account for only

1% of the total small wind market share in the US4; however, with the increasing interest

in renewable energy sources and the co-location of urban/rooftop installations to points-

of-use, their market share is expected to increase.

Small wind systems are typically used in residential, small business, agricultural, and

industrial applications. Aside from capacity (i.e., rated power output), the most obvious

attributes that distinguish small wind systems from their larger counterparts are (1)

smaller rotor diameters and tower heights, and (2) lower system capital costs. Small wind

systems are typically found in capacities ranging from 1 kW to 50 kW with two to seven

meter rotor diameters and can have tower heights in excess of 30 meters.

The hybrid small wind and solar power system shown in Figure 3 may be what the

future of home power will look like for those who live in areas where renewable

resources are available.

1.1 Experimentation Project Background

The California Wind Energy Collaborative (CWEC) partnered with the California Energy

Commission’s Public Interest Energy Research (PIER) program to implement a Small

Wind Energy Demonstration System at the University of California at Davis, campus.

The system is used for outreach and education, providing the general public up-close

opportunities to learn about the operation and ownership of a small turbine. Since the

installation of the system in June of 2006, it has been used as:


5

• A hands-on environment for a Small Wind Energy Systems short course.

• A demonstration system for the general public, including students from

elementary through college levels.

• A data collection station for meteorological data for public use.

Turbine
Instrument and
Control Room

Figure 4. Small Wind Energy Demonstration System on the rooftop of Bainer Hall

on the University of California, Davis main campus.

In addition to the primary goals of education and outreach, the system serves as a

research platform for independent wind turbine performance analysis and testing. The

rooftop testing facility created for this project (Figure 4) includes an instrumentation and

control system, which allows for accurate measurement of atmospheric conditions,


6

turbine performance, and turbine loads. These measurements are used for a variety of

education and research projects.

1.2 Research Scope

The study of small wind energy systems has been underway for many years. Specific

research on small wind electrical and aeronautical performance has been performed by

several institutions, some of which are used as references for our experiments. These

include Clarkson University6, The Center for Energy Studies in Monterrey, Mexico,7 and

the National Renewable Energy Laboratory8 (NREL). In addition, previous research has

also been done at the University of California, Davis.9 Research into the mechanical

loads of small wind turbines has also been performed by several institutions and is used

as reference material. These include the National Wind Technology Center10 (NWTC)

and NREL8. In addition to the various independent institutions, data from the wind

turbine manufacturer,11 whose product is used in our testing, is also used for reference in

our experiments.

All of the aforementioned small wind researchers performed tests in areas that

were considered rural and not rooftop mounted. The focus of this paper is to summarize

the methods and results of the electrical and load performance experiments performed on

the rooftop mounted, UC Davis Small Wind Demonstration System. The summary of this

experimentation consists of: (1) the methodology for the electrical system performance

measurements, (2) the methodology for mechanical load measurements, and (3) the use

of computational models to predict the system performance.


7

Wherever possible methods detailed in both the International Electrotechnical

Committee (IEC) standards IEC 61400-12-112, for power performance, and IEC 61400-

1313, for measurement of mechanical loads, were followed.

The principal results of this study show good correlations to the published

electrical system performance data. However, differences where found between the

preliminary experimental values of thrust and with both of the data obtained from

computational predictions and by NREL for mechanical load testing. These differences

will be discussed in detail in this report.

1.3 Contribution of Research

The novelties and contributions to the field of small wind energy resulting from this

research are in several areas. These include:

1. The investigation of the parameters of turbine electrical power and coefficient of

power as well as turbine rotor thrust and coefficient of thrust. Typically, the

detailed measurement of turbine mechanical loads is reserved for larger utility

scale turbines and not small scale systems.

2. The investigation of a small wind energy system located on the roof of a three

story building. Although not the most ideal setting for turbine testing, due to the

high level of turbulence, the co-location to electrical loads is ideal.

3. The testing of turbine rotor thrust via the measurement of guy wire support

tension, rather than the more common tower strain gauge measurement.
8

CHAPTER 2- EXPERIMENT EQUIPMENT & MATERIALS

The following chapter details the equipment used in the experimentation with the Small

Wind Energy Demonstration System. In addition, any alterations that were made to the

original equipment manufacturer’s (OEM) stock or standard configurations are discussed.

2.0 System Overview

The Small Wind Energy Demonstration System is deployed on the roof of the

engineering building, Bainer Hall, on the University of California (UC), Davis main

campus, as shown in Figure 5.

Wind Turbine

Instrumentation

Data Acquisition Guy Wire


and Control Room Cable

Tower Safety
Disconnect

Resistive
Data Lines Load Bank

Figure 5. Elevation view of the UC Davis, Small Wind Energy Demonstration System.
9

Small wind power systems can be configured as stand-alone systems (using a

battery system for energy storage), grid interconnected systems (using the grid to offset

energy surplus and energy needs), and battery back-up systems (grid connected with a

battery bank for grid power failures). Figure 6 shows a typical stand-alone small system

with a charge controller, battery bank, diversion load, and inverter.

Figure 6. Stand-alone small wind power system with power inverter for alternating current
(AC) loads.

The UC Davis Small Wind Energy Demonstration System is a stand-alone system

that has been modified for demonstration and experimental purposes. This system,

detailed in Figure 7, consists of a 1-kW horizontal axis wind turbine, the turbine tower,

instrumentation, a system disconnect, an adjustable resistive load bank, and a data

acquisition/control system.
10

Figure 7. Diagram of the UC Davis Small Wind Energy


Demonstration System configured with resistive load bank.

The primary difference between the standard stand-alone system, shown in Figure

6, and the demonstration system in Figure 7 is the replacement of the battery storage

system and the manufacturer’s battery charge controller with a resistive load bank. The

use of a load bank allows for precise control of the electrical loading of the turbine,

simplifying any performance analyses and research.

An additional feature of the demonstration system is the ability to easily raise and

lower the wind turbine. The turbine can be quickly and safely brought down for closer

viewing and to perform tasks necessary for maintenance or research.

For the monitoring and control of the system, a data acquisition unit is connected

to a personal computer, enabling the remote monitoring, autonomous data measurement,

and control of the resistive load bank. The computer system and data logger/controller

are shown in Figure 8. They are located in the room adjacent to the turbine, shown in

Figure 5.
11

Data Logger/ Personal


Controller 
Computer

Figure 8. System monitoring and control station located in the


room adjacent to the turbine.
12

2.1 Site Selection

The demonstration system is located on the roof of the engineering building, Bainer Hall,

on the University of California, Davis campus, as shown in Figure 9, below.

Wind Power
Demonstration Area

Figure 9. Satellite view of the demonstration system location on the roof of Bainer Hall, at the
University of California, Davis. Source Google Earth.

The location atop Bainer Hall has proven to be conveniently accessible for the Small

Wind Energy Systems short course, other education and outreach events, and the research

and experimentation detailed in this report. It has also been effective as a secure and

visible location for the turbine. However, the California Wind Resource Maps estimate

the site to be only a Class 2 wind resource (4.4 m/s - 5.1 m/s average wind speed at a

height of 30 meters); this is not ideal for large power production. In addition, the
13

turbulence induced from the building roofline, adjacent buildings, and structures in the

rooftop is not desirable. However, the wind resource is adequate for demonstration and

testing purposes.

2.2 Wind Turbine

The wind turbine selected for this project was the Bergey Windpower XL.1. This three

bladed, upwind, auto-furling, and horizontal axis wind turbine is shown in Figure 10. The

diameter of the XL.1 rotor is 2.5 m (8.2 ft) and utilizes composite vinyl ester and

fiberglass blades designed with a proprietary airfoil shape. The rotor is attached to a

permanent magnet alternator with conversion circuitry to produce direct current (DC)

power from the turbine.

Figure 10. Bergey Windpower, 1 kW, XL.1, horizontal, auto-furling wind turbine shown mounted
on tubular tower.
14

Bergey XL.1 Power Curve

1400

1200

Power Output [Watts] 1000


Furling
800 Speed

Rated
600
Power
Cut-in Speed
400 Speed

200

0
0 2 4 6 8 10 12 14 16 18 20
Wind Speed [m/s]

Figure 11. Power curve for the Bergey XL.1 turbine. Included are the wind speeds for cut-in, rated
power, and auto-furling. Source: Supplied by Bergey Windpower

The manufacturer’s power curve of the Bergey XL.1, using the stock configuration with

the PowerCenter controller, is shown in Figure 11. In this power curve the rated power

output is 1 kW at a wind speed of 11 m/s (24.6 mph). The rated cut-in speed, or the

lowest wind speed at which power generation occurs, is 2.5 m/s (5.6 mph).

The Bergey XL.1 is equipped with an over-speed mechanism designed to protect

the turbine’s mechanical and electrical components from overloading during high winds.

This mechanism is referred to as an auto-furling mechanism and functions by positioning,

or yawing the turbine rotor away from the incoming wind as shown in Figure 12. The

Bergey XL.1 furls at 13 m/s (29 mph) but has a maximum design wind speed of 54 m/s

(120 mph). Additional specifications for the Bergey XL.1 are listed in the manufacturer’s

specification sheet in Appendix A.


15

An additional reason for the selection of the Bergey XL.1 is that it has been the

subject of several studies, so an existing body of performance and operation data is

available for comparison and validation.

Incoming
wind
direction

Figure 12. Bergey XL.1 shown furling at high wind conditions.


16

2.3 Support Tower

The tower selected for the demonstration system is a tubular, tilt-up, guy wire supported

tower produced by Bergey Windpower. The tilt-up tower uses a counterbalancing gin

pole to easily raise and lower the entire tower, as shown in Figure 13. They are

particularly useful during maintenance and for lowering a turbine in the event of extreme

wind conditions. For the demonstration system, it provides convenient access to the

turbine for inspection, demonstrations, or experimentation.

Gin Pole
Tower

Figure 13. Turbine/tower assembly in mid-lowering position.

The turbine tower is 9 meters (30 feet) tall and can be raised and lowered by two people.

This height is a compromise between tower raising ability, safety, and power production.

The tower is constructed of 114 mm (4.5 inch) diameter, 2.1 mm (0.083 inch) thick

galvanized steel pipe. It is supported by four 4.76 mm (0.187 inch), 7 x 7 galvanized


17

aircraft cable guy wires, and an assortment of additional brackets and connection

hardware.

The demonstration system uses a custom designed power winch assembly to raise

and lower the turbine for frequent maintenance and experimentation. The stock winch

was replaced by a 12 volt DC winch with a full load brake. The winch system, shown in

Figure 14, is charged via a solar module to ensure a constant system charge. The winch is

mounted to one of the custom fabricated guy wire anchors and is removable.

Solar Module

Winch Battery
and Charger

Power Winch

Guy Wire

Guy Wire Anchor

Figure 14. Solar powered winch for easy and reliable raising and lowering of turbine/tower
assembly.
18

2.4 Electrical System

As mentioned previously, the demonstration system is configured as a stand-alone system

which can be upgraded to a grid connected battery back-up system in the future. The

stock system purchased includes Bergey’s PowerCenter® controller which takes the

output from the turbine generator (and optional solar array) and charges a 24 volt battery

bank. The controller is designed to maximize the amount of energy stored in the batteries

through the use of Pulse Width Modulation (PWM) to prevent battery overcharging. In

normal operation, the batteries cannot always absorb all of the energy that the turbine can

produce. The controller has the ability to reduce the power produced by the turbine in

order to protect the batteries from damage. Because of this effect, the performance of the

stock turbine system is dependent not only on the atmospheric conditions, but also the

state of charge of the battery bank.

For experimentation with the demonstration system, it was desirable to isolate the

turbine performance from the rest of the system for performance testing and other

experimentation. To accomplish this goal, the stock PowerCenter® controller was

replaced by a load bank and a data acquisition and control system. The resulting system is

shown in the wiring schematic below in Figure 15.


19

Figure 15. Electrical wiring diagram of the Small Wind Energy Demonstration System.

The custom load bank (Figure 16) was built in-house by Kamisky9 to dissipate energy

generated by the turbine. High, medium, and low load conditions are applied by 2.1 ohm,

7.9 ohm, and 25.1 ohm, high power resistors. These loads are wired in parallel and are

individually switched on and off by Kilovac EV250 series relay contactors that are

controlled by the data acquisition unit.


20

Figure 16. Resistive load bank located at the base of the turbine used to provide constant
loads for performance evaluation.
 

Additional modifications to the stock system include a disconnect switch that separates

the turbine from the load bank and a safety stop mode with a 60 amp breaker. The

breaker short circuits the positive and negative output from the turbine, allowing the rotor

to be stopped for raising and lowering the tower. Additionally, a lightning arrestor is

installed in the disconnect box, grounding the turbine in the event of a lightning strike.

The base plate and tower are also connected directly to a high capacity, dedicated earth

ground.
21

2.5 System Instrumentation

The instruments used in the experiment are located on the turbine tower, as shown below

in Figure 17, and within the resistive load bank at the base of the tower. Additional

sensors and circuitry for rotational frequency measurement are located in the data

acquisition/controller enclosure within the equipment utility shed. In Table 2.1, the

system sensors are listed with their manufacturer, operating range, and accuracy. Details

of several of these sensors will be discussed in the following sections.

Yaw Encoder

Wind
Vane

Load Cells Air


Anemometer Temperature
& Relative
Humidity
Sensor

Pressure
Sensor

Upper Guy
Wires
Lower Guy Wires

Figure 17. Turbine instrument locations.


22

Table 2.1 List of sensors and their specifications, currently deployed on the turbine.
Sensor Manufacturer Range Precision
Cup Anemometer R.M. Young Wind Sentury 0-50 m/s +/- 0.5 m/s
Wind Direction Vane R.M. Young Wind Sentury 360˚ +/- 5˚
435E Relative Humidity Vaisala 0-100% RH +/- 1% RH
435E Air Temperature Vaisala -50 - +60 ˚C +/- 0.6 ˚C
BP20 Barometric
NRG 15-115 kPa +/- 1.5kPa
Pressure Sensor
Yaw Sensor Encoder Michigan Scientific 360 deg. +/- 1 deg.
1000 lbs
Omega 1000 lb, S-Beam,
Omega Engineering Inc. tension or +/- 0.1 lbs
Load Cells
compression
Frequency Transducer Phoenix Contactor 1-10kHz +/- 0.5 Hz

Current Shunt Omega Engineering Inc. 0-50A +/- 0.25A


23

2.6 Data Acquisition System

Experimental measurements were made via a control and data acquisition system built

upon the National Instruments Compact FieldPoint system, running LabView. As shown

in Figure 18 below, the user interface allows for the real-time control of the load bank

and the turbine as well as the monitoring of the various sensors located at the site.

Figure 18. Wind turbine data acquisition and control, real-time system interface

System measurements were the temperature [°C], wind speed [m/s], relative humidity

[%], absolute pressure [kPa], wind direction [degrees], equivalent load bank resistance
24

[ohms], turbine voltage [V], turbine power [W], turbine rotational speed [rpm], rotor yaw

position [degrees], and tower loads [lbs].

A block diagram of the DAQ system is shown in Figure 19 below. In this

configuration the Compact FieldPoint® controller can be utilized for data acquisition and

control via a computer or directly via the controller itself using embedded LabView Real-

Time programming. 

Figure 19. Data acquisition system block diagram.

The Compact FieldPoint controller acquires real-time data from the various input

modules. This data is then either averaged over a specified period of time or analyzed by

the controller. It is then shared with a computer running a separate program to record and

post the data from the controller. This system allows for customized autonomous

recording of data as well as remote access to both the FieldPoint controller and computer

via the Internet.


25

2.7 Turbine Customization

To install the yaw position sensor required for the experiment, several modifications were

made to the turbine spindle and nacelle cover geometry. The Michigan Scientific slip ring

encoder used for the yaw position measurement mounts to the top of the turbine spindle

as shown in Figure 20 below. The spindle, attached to the tower, is stationary while the

rotating portion of the encoder is attached to the free yawing turbine frame.

Turbine
Alternator
Yaw Encoder Windings

Turbine Frame

Turbine Hub
Turbine Spindle Turbine Rectifier

Diodes

Yaw Encoder
Output Cable

Figure 20. Yaw position sensor installation shown on turbine frame with nacelle cover, tail boom, blades,

and nose cone removed.

In order to mount the yaw position encoder on the top of the turbine spindle it was

necessary to drill a 3/8” routing hole through the entire length of spindle for the encoder

output cable. To ensure that this modification would not have any detrimental effects on
26

the turbine assembly, a structural analysis was performed on the turbine spindle. This

analysis consisted of the solid modeling of the turbine spindle and the application of a

finite element load analysis (FEA) using SolidWorks CosmosExpress. Using the

manufacturers specified turbine mechanical load of 240 lbs, similar material properties,

and the measured spindle geometry, a load analysis was performed.

Figure 21. FEA results of modified turbine spindle solid model, using
manufacturer’s thrust load values.

The graphical results, shown in Figure 21, are from an analysis which used a high

resolution grid of 15,824 elements with 25,956 nodes. Under the manufacturer’s rated
27

load of 240 lbs, the factor of safety with the modified turbine spindle geometry is 7.4. For

more specific information the FEA inputs and results are in Appendix B.

To accommodate for the height of the yaw position encoder the nacelle cover

geometry was changed and a new nacelle cover was designed and fabricated, by the

author. This cover was designed to minimize the airflow disturbance while protecting the

encoder and turbine electronics from environmental elements. The turbine with modified

nacelle cover is shown below in Figure 22.

Figure 22. View of lowered turbine with nose cone removed and with custom nacelle cover.
28

CHAPTER 3- EXPERIMENTAL METHODS

In this chapter, the electrical and mechanical system experimentation methods, are

presented. In addition to the experimentation methods, the methods for data analysis,

procedures for determining measurement uncertainty, and methods for the assessment of

site specific turbulence are also presented.

3.0 Experimentation Overview

The Small Wind Energy System experimentation consists of three phases. In the first

phase, the turbine data logger/controller and sensors for wind speed and turbine output

voltage were installed. Preliminary turbine performance was conducted with this level of

instrumentation and is presented in Kamisky9. The second phase, and focus of this paper,

consists of the installation and application of additional instrumentation to assess the

electrical performance and mechanical loads associated with the turbine. In the third

phase (to be completed in the future) additional research will be conducted and data from

the system including wind speed and turbine power will be made available via the

Internet.

The electrical performance and mechanical load experimentation includes the

measurement of: (1) meteorological condition of wind direction, air humidity, and air

temperature, (2) turbine specific parameters including rotor rotational speed (RPM),

output voltage, and yaw position, and (3) the loads induced on the tower from the

operation of the turbine. The integration of these three major measurement regimes

allows for the assessment of the overall system, placed in the non-ideal rooftop

environment.
29

3.1 Electrical System Measurement Methods

For the electrical performance evaluation of the Small Wind Energy System, several key

measurements were taken. These included the measurement of the turbine rotational

speed in revolutions per minute (RPM), the turbine output voltage, and the load bank

resistance.

3.1.1 Turbine RPM Measurement

To obtain the rotational velocity of the turbine rotor, or turbine RPM, a signal analysis of

the turbine alternator output signal was used. If the turbine were mounted in a fixed

position, the direct measurement of RPM could have made through a tachometer.

However, because of the need of the turbine to yaw a full 360 degrees, the addition of a

slip-ring assembly was found to be not practical during this round of testing.

Alternatively, the rotor RPM was obtained via the frequency of the alternator output

signal.

The Bergey XL.1 uses a six pole pair (twelve poles total), permanent magnet,

asynchronous alternator which is directly coupled to the turbine rotor. During operation

the turbine alternator outputs a three-phase alternating current (AC) electrical signal with

a frequency and voltage proportional to the rotational speed.15 This three-phase electricity

is rectified at the turbine in a six-pulse rectifier, which adds an additional 6 pulses per

pole pair in the output signal.16 The resulting final output of the turbine is a direct current

(DC) signal (Figure 23) that contains an AC ripple that is also proportional to the turbine

RPM.
30

Figure 23. Turbine alternator output voltage signal shown with AC ripple content within rectified DC
signal. The signal above was measured across the resistive load bank at a sampling rate of 360 Hz.

Subsequently, for every single rotation of the turbine rotor, 36 pulses appear in the AC

ripple content of the rectified DC output signal. As a result, the scaling factor used to

compute the turbine rotational frequency (in RPM) from the AC ripple in the rectified DC

signal is defined as:

60 * f m
RPM =
PT Equation 1
* PR
2

where fm is the measured frequency of the turbine output signal,

PT is the number of alternator poles or magnets (twelve

for the XL.1), and

PR is the number of pulses (or diodes) in the rectification

circuit (six for the XL.1).17


31

In the typical turbine configuration using a battery and PowerCenter charge controller,

the battery bank absorbs the AC ripple and acts like a low-pass frequency filter.15

Without a battery, as in this experimental set-up, the measurement of the rotational

frequency was able to be made because of the presence of the AC ripple in the turbine

signal.

Due to the high frequency range for the expected rotational speeds, a frequency

transducer was necessary to measure the AC ripple frequency. In order to apply the

frequency transducer, the turbine alternator signal had to be digitized via a custom circuit.

This custom circuit modified the turbine signal by the capacitive or AC coupling of the

output DC signal to remove the DC offset leaving the only the AC ripple in the signal

(Figure 24). Then this AC signal was converted into a digital (TTL) signal for the

frequency transducer input. The analog output of the frequency transducer was then

measured via the data acquisition system.

Figure 24. AC coupled turbine output signal with zoom of ripple content.
Oscilloscope readings of AC signal frequency are also shown.
32

After the turbine AC ripple was digitized and input into the frequency transducer,

Equation 1 was used to compute the turbine RPM from the frequency transducer output.

3.1.2 Turbine Power

To obtain the raw power output of the turbine, PElectrical, Ohm’s Law was used in

conjunction with the turbine voltage and load bank resistance. Thus the turbine electrical

power is defined as:

PElectrical =
(VTurbine )
2

Equation 2
RLoad

where VTurbine is the average turbine output voltage within each

1-second sampling period and RLoad is the load bank

resistance.21

Note that the turbine voltage was measured directly across the restive load bank. In

addition, the turbine negative was wired to ground to prevent galvanic isolation (or

dissimilar grounding) between the turbine alternator and the data acquisition electrical

ground.

3.1.3 Load Bank

As mentioned previously, the energy produced from the turbine is dissipated through an

adjustable resistive load bank. In the load bank, three high power resistors are wired in

parallel via three contactor relay switches.


33

For the computation of power, the known nominal resistance and the voltage drop

across the load bank were used. The load bank resistance was assessed for temperature

effects which were found to be minimal for the temperature range encountered at the test

site.

3.1.4 Bergey PowerCenter® Overview

Being that the Small Wind Energy System experiment did not utilize the stock

PowerCenter controller provided by Bergey Windpower, a brief comparison between our

load bank and the PowerCenter is worth noting because of the effects on the electrical

and mechanical performance results.

The primary difference between the stock set-up and our configuration derives

from the differences in electrical loads. The alteration of the turbine electrical resistances,

or loads, changes the net torque in the alternator.15 This occurs because when the

electrical resistance changes, the current in the overall turbine electrical circuit, including

the alternator windings, changes as well. The current in the alternator windings creates an

opposing electromagnetic field (emf) within the turbine alternator windings which must

be overcome by the electromagnetic field induced from the movement of the rotor.18 This

emf manifests as a resistive torque in the alternator, which changes the cut-in speed of the

turbine as well as the power output. The extreme use of this phenomenon (i.e. when the

turbine current is allowed to be greatest) is used in the braking mechanism of the Small

Wind Energy System. When the “brake” is on, the turbine alternator input and output

leads are shorted, causing: (1) a minimum amount of load resistance, (2) a maximum

amount of system current, and (3) the maximum resistive torque which causes the rotor to
34

slow down and stop. The modification of the load resistance is a critical part of the

overall turbine performance optimization.

In Figure 25, the electrical power versus load resistance for the PowerCenter®,

along with a secondary plot of the UC Davis system resistances is shown.11 The

secondary plot values were obtained by recording the intersection of the stock power

curve and the experimental power curve for each given resistance and recording the

power value.

34
32
30
28
26
24
Resistance [ohms]

22
20
18
16
14
52.08, 11.81
12
10
8
Bergey XL.1 w/ PowerCenter
6
204.00, 3.19 UCD Bergey XL.1 w/ Load Bank
4
447.81, 1.43
2 787.50, 0.79
1459.14, 0.42 2124.30, 0.40
0
0.00 250.00 500.00 750.00 1000.00 1250.00 1500.00 1750.00 2000.00 2250.00
Power [watts]

Figure 25. PowerCenter load resistance and experimental load resistances vs. electrical power.

Source: Bergey Windpower

As shown in Figure 25, the Bergey PowerCenter alters the “load” resistance during

operation. According to the manufacturer, the modulation of this load resistance results in

conditions that are optimal for turbine power production.15 For example, in the
35

PowerCenter, the load resistance is higher initially allowing for lower cut-in speeds, but

then it decreases as the system voltage increases aiding in the prevention of turbine over-

speeding. The decrease in system load resistance also provides larger power outputs. As

the resistance values used in the load bank are constant, the maximum power outputs

from the turbine will differ from those obtained using the manufacturer’s PowerCenter

controller.

3.2 Mechanical Load Experimentation Methods

The following section details the methods and rationale used for the mechanical

measurement facet of our experimentation. The load measurements of interest for the

experiment consist of the forces induced by the turbine; these being the thrust forces. In

the following sections the methods for the computation of the thrust forces from the

measured guy wire forces are presented. These consist of the explanation of mathematical

equations and assumptions used, the corrections made to account for the experimental

set-up, the physical alterations to the tower, as well as the calibration procedures used for

the measurement equipment.

3.2.1 Thrust Experiment Overview

To asses the thrust loads induced by the XL.1 turbine on the support tower, an upper set

of supporting guy wires (Figure 17) was installed and used as the load measurement point

for the tower. Typically strain gauges are used to measure thrust loads. In tests performed

at the NWTC, by Hushy et al,10 a 1 kW small wind system, using a guy wire supported

tubular tower, was measured for thrust loading. This experiment used strain gauges in the

top tower section to determine tower bending which were then used to compute turbine
36

thrust. A different approach, with respect to the measurement of the tower reaction

forces, was used in our experiment with the Bergey XL.1. This approach followed the

guidelines set by the IEC Standard for the Measurement of Mechanical Loads.13

3.2.2 Thrust Experiment Mathematics

The theory behind the utilization of the guy wires for tower thrust measurement was built

upon both the IEC testing standards and the guy wire tower testing work by Madugula et

al at the University of Ontario, Canada.19 In Madugula’s and our experiment, the primary

assumption is that the single guy-wire-set supported tower, on which our turbine is held,

has a ball and socket joint at the base as shown below in Figure 26.

Figure 26. Simplified turbine tower support model.

This assumption allowed for the four guy wires to be the main supports of the tower,

while the lower ball and socket joint was assumed to resist translational movement,
37

allowing for tilting or rotational movement at the base. It is worth noting that in order for

the system to be mathematically determinant and congruent to the ball and socket

assumption, only the upper set of guy wires in our dual guy wire arrangement (as shown

in Figure 17) were used during the tests.20 During testing the lower guy wire tensions

were decreased so as to not influence the thrust measurement in the upper guy wires, but

still were present to provide a secondary, emergency tower support if needed.

Using the known tower, turbine, and anchor location geometries, the horizontal

reaction force, RA, at the upper guy wire attachment point, A, is then found via the free

body diagram in Figure 27. In the loading scenario shown in Figure 27, the turbine thrust

is resisted by the upwind cables while the downwind cables are in compression and slack,

subsequently adding no support to the tower. Thus under the thrust load, at most two

cables are supporting the tower providing reaction forces to the turbine thrust.
38

Figure 27. Free-body diagram of simplified turbine tower assembly.

From the cable reaction forces, the thrust force (FTi) components in each plane are

determined and then the vector sum of these values yields the total thrust (TForce) of the

turbine as shown in Equation 3 and Equation 4:

⎛ L ⎞ ⎛ L ⎞
FTi = R A ⋅ ⎜ ⎟ = RTi ⋅ cosθ i ⋅ ⎜ ⎟
⎝S +L⎠ ⎝S +L⎠ Equation 3

TForce = (FTz )2 + (FTy )2


Equation 4

where RTi is the measured reaction load in the cable, i, and

i= z and y, are in the upwind direction of the applied

thrust load.
39

In Figure 28, a top view of the wind turbine and tower assembly is shown with the

horizontal thrust force and subsequent reaction forces in the direction of the wind, Vwind.

Figure 28. Top view of turbine tower with horizontal forces.

3.2.3 Thrust Experiment Corrections

In reality the turbine base is a single axis hinge- as shown in Figure 29, instead of the

assumed ball and socket joint. This hinge allows for the turbine tower to tilt freely in one

direction (about the hinge/tower tilting axis), while providing resistance to tilting in the

direction of the hinge axis or perpendicular to the tilting axis in the non-tilting plane (the

non-tilting, resistive axis).


40

Hinge Bolt
E
Base plate
upright

Hinge/Tower
N
Non-Tilting
(Resistive) Axis Tilting Axis

Figure 29. Support tower base detail showing tilting and non-tilting axis locations.

Because of this resistance at the base of the tower, a resistive moment is created on the

tower which changes the magnitude of the reaction forces in the guy wires in the north

and south directions of the non-tilting (east/west) axis. However, in the plane of tilting

there is essentially no resistive force from the base and the thrust measurements are

assumed to be greatest in that direction, as presented in the previous sub-section. In the

north and south directions, the cable tension components were theoretically determined to

be affected by a factor of 1.17 due to the tilting resistances at the base. In the final data a

correction factor was used for loads measured in the north and south directions. More

information on this thrust correction factor is presented in Appendix C.


41

3.2.4 Tower Modification

As a part of the load analysis experimentation, the turbine support tower was modified.

These alterations were made in order to both accommodate for the mechanical load

measurement equipment and to reduce the tower bending during operation. In order to

measure the mechanical loads induced from the turbine, a secondary guy wire attachment

bracket was created. This bracket shown in Figure 30, was designed to resist twice the

amount of the manufacturer’s specified maximum loads. The primary alteration to the

tower for the installation of the thrust bracket is a single ½” hole through the tower to

accommodate for the single grade 8 anchor bolt.

Single Attachment
Bolt

Figure 30. Solid model of secondary guy wire attachment bracket installed on tower.

The second alteration to the turbine support tower was the location of the

secondary guy wire support bracket along the height of the tower. The placement of the

bracket coincides with the manufacturer’s suggested locations; however, to ensure that

the bracket, additional connection hardware (guy wire, turnbuckles, and rod ends), and
42

the turbine tower were able to perform under estimated conditions at the site, an

additional analysis was performed. This analysis was based upon the static wind load

analysis process for structures found in the Uniform Building Code (UBC)14 and specific

details of this analysis are shown in the Appendix D. The resulting placement of the

attachment bracket was in a region of the tower to best decrease the bending of the upper-

most section of the tower when the secondary guy wire support cables were the primary

means of support. The resulting calculations for the tower bending moments along the

length of the tower, are shown in Figure 31.

4000.0
2000.0
Bending Moment [in*lbs]

0.0
-2000.0 0 50 100 150 200 250 300 350 400

-4000.0
-6000.0
-8000.0
-10000.0
-12000.0
-14000.0
Tow er Height [in]

Figure 31. UBC wind load analysis results for tower bending moments.

As a result of the higher mounting location, the secondary set of guy wires

yielded a reduction in the tower bending moments of 52% over the previous arrangement

of using only the lower guy wire supports.


43

3.2.5 Load Cell Calibration

For the measurement of the reaction forces, the guy wire tensions were measured using

an S-Beam load cell inline with each guy wire cable. The orientation of the load cells

were aligned with the tilting and resistive axes to ensure the recording of the free-tilting

and resisted tilting forces. Each of these load cells (Figure 32) were tensioned to 100 lbs

and the tension and cable weight values were recorded and used to set the thrust force

measurement to zero prior to thrust measurement. For the additional upper guy wire

cables and mounting hardware, specifications (e.g. cable diameter, support plate

thickness, etc.) were increased to ensure safe operation as well as to minimize flexion

during testing.

Figure 32. Load cell location shown inline with guy wire with turbine in lowered position.
44

For each of the four S-Beam type load cells used in the experiment, a calibration

process was used. This consisted of the static loading of each of the load cells in tension

as shown in Figure 33 with a known amount of weight. The data from the calibration was

compared to the data given from the manufacturer and any differences in output voltage

due to additional electrical resistances from the final wiring configuration, were

accounted for.

Figure 33. Load cell calibration set-up shown with hanging test load.
45

3.3 Data Collection

For the acquisition of data, a raw data sampling rate of 1 Hertz was used for each of the

eleven different measurements. These data were then saved for the duration of the test –

which usually was one day – and then the data were post processed and sorted.

The post processing of the data consisted of the averaging of the data over a period of

10 seconds to yield a 10-second data point. The IEC standard allows for a maximum data

sample averaging of 10 minutes, but because of the fairly fast response of the turbine

rotor and the load cell measurements as well as the desire to measure the effects from the

turbulent rooftop environment, the shorter averaging period of 10 seconds was chosen.

In an example of this process, the calculation of the turbine average power, Pavg, is

based upon the time averaging of the raw 1-second data of the turbine electrical power,

PElectrical, and over a period of 10 seconds. Thus the turbine 10-second turbine electrical

power is defined as:

N
1
Pavg , j =
N
∑P
i =1
Electrical ,i Equation 5

where N is the number samples in the averaging period, which

is equal to 10 in our case, j, is the reference number of

the data point in the 10-second data set, and i, is the

reference of the 1-second data point in the raw data.12


46

3.4 Performance Calculation Methods

After collecting the raw measurement data, the data was corrected for differences

between the test site air density and standard air density, ρ0. This process is referred to in

the IEC as data normalization and was applied to the wind speed measurements because

of the variable speed and resistive load control of the turbine. After the data was

normalized, the data was sorted by wind speed using the method of bins, and then the

final calculations for electrical and mechanical performance were made.

3.4.1 Data Normalization

The process of data normalization begins first with the calculation of the air density for

the averaged sample data; in our case this was 10 seconds. For the calculation of the

average air density, ρavg, in kg/m3, the following relationship was used:

1 pavg
ρ avg = ⋅ Equation 6
R Tavg

where R is the specific gas constant equal to 287.05 J/(K kg),

pavg is the air pressure, in Pa, and

Tavg is the air temperature, in Kelvin.21

Once the average air density was determined it is then used to compute the normalized

wind speed. The normalized wind speed is defined as:


47

1
⎛ ρ avg ⎞3
U n = U avg ⋅ ⎜⎜ ⎟⎟ Equation 7
⎝ ρ0 ⎠

where Uavg is the time averaged wind speed, in m/s.12

3.4.2 Data Sorting and Final Computations

With the normalized wind speed and the averaged values (including averaged electrical

power) the data was then sorted using the method of bins. The method of bins used in the

sorting of data follows the IEC standard and uses wind speed, at a resolution of 0.5 m/s,

per bin, to organize and average data as a function of the wind speed.12

The definition for the binned turbine electrical power is shown in Equation 8. For

the quantities of turbine rotational speed, in RPM, and thrust load, TForce, the same

binning process was used. The average binned turbine electrical power, Pi, in watts, is

defined as:

Ni
1
Pi =
Ni
∑P
j =1
avg ,i , j Equation 8

where Ni is the number samples in the wind speed bin, i, and j is

the reference number of the data set.12

3.4.3 Final Computations

Once all the appropriate values were binned, the coefficient of power, CP,i, and

coefficient of thrust, CT,i were then computed. The coefficient of power is defined as:
48

Pi
C P ,i = Equation 9
0.5 ⋅ ρ 0 ⋅ A ⋅ (U i ) 3

where Pi is the normalized turbine electrical power for bin i, A

is the rotor swept area, in m2, and Ui is the average wind

speed within bin, i, in m/s.12

Similarly, the coefficient of thrust is defined as:

TForce,i
CT , i =
0.5 ⋅ ρ 0 ⋅ A ⋅ (U i )
2 Equation 10

Lastly, for the computation of the dimensionless value of the tip-speed-ratio (TSR), the

following definition was used:

TSRi =
(RPM i ) ⋅ 2 ⋅ π ⋅ R / 60
Equation 11
Ui

Where R is the radius of the rotor, in meters.21

3.4.4 Uncertainty Evaluation

To account for the bias (or systematic) and precision (or random) error of the

measurements, several steps were taken to quantify each type were they occur. These two

types of error occur in our measurements before any measurement has been taken, in the

design stage, and as a result of processing the data, in data-reduction stage.23 Once each
49

of these two types of error was determined for each of the stages of the experiment, they

were then combined to obtain the total measurement uncertainty.

The process to determine the total design stage uncertainty uses the resolution

uncertainties of each instrument, the analog to digital conversion uncertainty of the data

acquisition system, and the sensitivity of each parameter in the final equation (e.g. in the

case of power within the coefficient of power, Cp). The definition of the initial design

stage uncertainty, ui, is defined as:

ui = ((u inst )2 + (u adc )2 ) Equation 12

where uinst is the instrument resolution or precision error as

listed in Table 2.1, uadc is the analog to digital converter

error, and i is the measurement type.23

The initial design stage uncertainty is then combined with the individual parameter

sensitivities (defined as the partial derivative of each parameter within its final equation)

to obtain the total design stage uncertainty, UD,i. For each of the performance calculations

(Pi, Ti, RPMi, Cp,i, and Ct,i), the design stage uncertainty were determined and are shown

below in Equation 13-17.


50

⎛ ⎛ ∂P ⎞
2

U D ,PRotor ⎜
= ± ⎜⎜ Rotor
*u ⎟ ⎟
⎜ ⎝ ∂VTurbine VTurbine ⎟⎠ ⎟
Equation 13
⎝ ⎠

U D ,TForce = ± ((u ) ) = ±u
TForce
2
TForce
Equation 14

U D ,RPM = ± ((u RPM )2 ) = ±u RPM Equation 15

⎛⎛ ∂C ⎞
2
⎛ ∂C ⎞
2

⎜⎜ P
*u ⎟ +⎜ P
*u ⎟ ⎟
⎜⎜⎝ ∂VTurbine VTurbine ⎟⎠ ⎜⎝ ∂PAir PAir ⎟⎠ ⎟
UD,Cp =± ⎜ ⎟ Equation 16
⎜ ⎛ ∂C 2
⎞ ⎛ ∂C ⎞
2

⎜⎜ + ⎜⎜ P *uTAir ⎟⎟ + ⎜ P *uU ⎟ ⎟⎟
⎝ ⎝ ∂TAir ⎠ ⎝ ∂U ⎠ ⎠

⎛ ⎛ ∂C ⎞
2
⎛ ∂C ⎞
2

⎜⎜ T
*u ⎟ + ⎜ T
*u ⎟ ⎟
⎜ ⎜⎝ ∂TForce TForce ⎟⎠ ⎜⎝ ∂TAir TAir ⎟⎠ ⎟
U D,Ct =± ⎜ ⎟ Equation 17
⎜ ⎛ ∂C 2
⎞ ⎛ ∂C ⎞ ⎟
2

⎜⎜ + ⎜⎜ T * u PAir ⎟⎟ + ⎜ T * uU ⎟ ⎟⎟
⎝ ⎝ ∂PAir ⎠ ⎝ ∂U ⎠ ⎠

Once the total design stage uncertainty for each measurement (UD,i) was determined,

the data reduction uncertainty was then calculated. The error accrued through the data

reduction process is the result of the random scatter of the data. These precision errors

were computed for electrical power, thrust, and rotor velocity following the procedure for

the calculation of standard category A uncertainty as outlined in the IEC 61400-12-1.12

For example, the data reduction error for turbine electrical power error, is defined as:
51

σ P ,i
S P ,i = Equation 18
Ni

where σP,i is the standard deviation of the electrical power data per

bin, i, and Ni, is the number of samples within the bin.12

For the computation of the total uncertainty, Un, including the design stage and data

reduction errors, Equation 19 was used (as in the example of turbine electrical power):

UnP ,i = ((U ) + (S ) )
D , P ,i
2
P ,i
2
Equation 19

where i, in this case is the corresponding wind speed bin, in m/s.

3.5 Turbulence Effects

Due to the rooftop location of the installation, a brief discussion of turbulence is

necessary. The study of turbulence in itself is highly complex due to the chaotic and

stochastic nature and has been the focus of much more involved research than will be

presented. However, there are several details involving turbulence that will aid in the

discussion of the results observed for this rooftop installation.

Turbulence is defined as: fluctuations in wind speed on a relatively fast time scale

typically less than about 10 minutes.24 Shown in Figure 34 is a plot of wind speed vs.

time as measured by the anemometer on the roof of Bainer Hall for one 500 second

section of the data from a 24 hour data set.


52

18

16

14

12
Wind Speed (m/s)

10

Wind Speed
0
9000 9050 9100 9150 9200 9250 9300 9350 9400 9450 9500
Time (Seconds)

Figure 34. Raw wind speed values over a 500 second data sampling interval.

As can be seen in Figure 34, fluctuations or turbulence are present in many wind speed

measurements, particularly in regions of high roughness like that found in an urban area.

With the inclusion of turbulence, U*, the instantaneous wind speed measurement is

defined as:

*
Ui = Ui + Ui Equation 20

where U i is the mean or non-fluctuating value and Ui* is the

fluctuation value of wind speed.23

Turbulence intensity (I) is often used to assess the degree of turbulence within a given

data set and is defined as:


53

σ
I= Equation 21
U Σi

where σ is the standard deviation and U Σi is the mean of the

measured wind speed data set.24

The effects of turbulence can be significant because of the energy stored in the

fluctuating flows.23 These effects can become amplified when applied to the calculations

of power in the wind, as defined as:

PWind ,i = 0.5 ⋅ ρ 0 ⋅ A ⋅ (U i ) 3 Equation 22

where Pi is the normalized turbine electrical power, A is the

rotor swept area, in m2, and Ui is the average,

wind speed within the bin, i, in m/s.21

When the wind speed is cubed, using the expanded definition of wind speed, this is

represented by:

(
U 3 = U +U * )
3

3 2
U 3 = U + 3U U * + 3U U * ( ) + (U )
2 * 3

Due to the symmetrical nature of U*, it equals zero when averaged over time.23 Thus, the

simplified definition of the time averaged wind speed cubed, including turbulence, is:

3
U3 =U +3U* U ( ) 2
Equation 23
54

As shown in Equation 23, the turbulence contribution U* can have significant effects on

the calculation of the power in the wind especially at lower wind speeds where:

U ≤ U*

when the fluctuations in the wind speed are greater than the average wind speed

calculated via the method of bins. In other words, at low average wind speeds, the

contribution of power due to turbulence, can be significant.


55

CHAPTER 4 – PERFORMANCE PREDICTION METHODS

In addition to direct experimentation, a theoretical performance analysis of the turbine

system was performed. This analysis included the determination of the aerodynamic

properties of the turbine rotor as well as the incorporation of the previously published

turbine alternator performance data.

For the aerodynamic performance prediction of the turbine rotor, there were three

main procedures employed. First, the simulation parameters for the rotor blades were

obtained from previous work by Kamisky.9 This consisted of the airfoil characteristics

including the airfoil geometry and the published operating conditions. Next, lift and drag

properties (or polars) for the airfoil were obtained over a range of conditions using the

software XFOIL. Finally, a numerical analysis using blade element momentum (BEM)

theory was conducted to predict the performance of the rotor. Within the BEM

calculations, the mechanical power input to the turbine alternator was calculated from the

torque and speed of the rotor shaft. Using both the mechanical power input to the

alternator and published alternator efficiency data, the electrical power output of the

turbine system was calculated. Lastly, to validate the BEM analysis, the commercially

available Wind Turbine Performance program, WT_Perf, was used.

4.0 Airfoil Characteristics

The geometry of the XL.1 rotor blade airfoil was determined by Kamisky9 using a

combination of methods. First the blade was precisely measured using a Mititoyo
56

coordinate measuring machine and then the remaining geometry was estimated using a

curve fitting algorithm. The blades of the XL.1 turbine have constant chord and no twist,

so only a single two-dimensional cross section of the blade was required for the

aerodynamic analysis. Figure 35 shows the two dimensional cross-section measured from

one of the Bergey XL.1 turbine blades. This geometric data was found to correlate well

with the manufacturers stock geometry obtained after the use of the CMM.

 
Figure 35. Measured cross-section of the Bergey XL.1 blade.

Once the turbine blade airfoil geometry was measured it was then normalized to the

length of the major cord and then the data was used as inputs into XFOIL to calculate the

aerodynamic properties of lift and drag.

4.1 XFOIL

XFOIL version 6.94 was used by Kamisky9 to generate lift and drag curves for the airfoil

measured from the XL.1 turbine blades. XFOIL is a freely distributed program for the

design and analysis of airfoils in uniform subsonic flow. Its ability to rapidly generate

lift/drag properties for the subsonic flow conditions under consideration and the fact that

it is freely distributed were the main selection factors. For a complete explanation of the

technical details, see the XFOIL 6.94 User Guide.24

These resulting lift and drag properties for the XL.1 blade geometry were then used

as inputs in the two BEM based turbine performance models.


57

4.2 Blade Elemental Momentum Analysis

Blade elemental momentum theory was applied to the sectional airfoil data to determine

the torque and thrust of the Bergey XL.1 rotor. The roots of the BEM theory can be

traced to analysis of ship propellers in the late 1800s and was extended to airplane

propellers and turbine rotors in the 1920s. BEM theory combines blade element and

momentum theories to predict loads on a rotor or propeller. These loads are then used to

compute the forces of lift and drag on the blade which can then be used to determine

torque and thrust. The primary outputs of the BEM analysis used are the torque, power,

and thrust on the rotor at various rotor RPMs and include corrections of additional

attributes of the turbine’s rotor/alternator assembly (e.g. alternator efficiencies).

However, not included are the resistive torques due to loads within the alternator and

turbulent air flow conditions, as mentioned previously in Chapter 3.

For additional detail, a complete explanation of BEM theory applied to wind turbine

rotors can be found in Manwell21 or Hansen.25 The BEM code used was developed in-

house by Raymond Chow and configured for the Bergey XL.1 operational parameters.

4.3 Wind Turbine Performance Analysis, WT_Perf

For a second perspective on the computational methods employed for the performance

prediction of the Bergey XL.1, the National Wind Technology Center (NWTC) supplied

code, Wind Turbine Performance Analysis (WT_Perf) was used. This BEM based code

originally created by Aeroenvironment Inc. and since been improved, is used in industry

as a code for analyzing the performance of existing turbine systems.26 The utilization of

this code provided a validation for the in-house developed BEM code. For more
58

information regarding the specific simulation input parameters for this analysis please see

Appendix E.

4.4 Experimental Corrections

Once the theoretical performance of the turbine was obtained and validated, the data was

corrected using the known experimental XL.1 alternator DC efficiencies determined by

Martinez et al7. As shown in Figure 36, the efficiency of the XL.1 alternator is a function

of rotational velocity. This efficiency data was used in the determination of the final

turbine power performance predictions.

100

90

80

70
Efficiency [%]

60

50

40

30

20

10

0
0 100 200 300 400 500 600

RPM

Figure 36. Bergey XL.1 alternator efficiency data used in performance simulations. Source: Martinez et al7
59

4.5 Computational Results

In the next section the computational results for the blade aerodynamic properties, turbine

power, coefficient of power, rotor thrust, and coefficient of thrust, are presented. These

results were determined from the use of XFOIL, the Blade-Element Momentum analysis

(BEM), and the program, Wind Turbine Performance (WT_Perf).

4.5.1 Blade Lift and Drag Curves

Figure 37 shows the coefficients of lift and drag as a function of angle of attack generated

by XFOIL from the measured airfoil section.


2.5 0.12

0.1
2

0.08

Coefficient of Drag
Coefficient of Lift

1.5

0.06

0.04

0.5
0.02

Coefficient of Lift
Coefficient of Drag
0 0
-10 -5 0 5 10 15 20

Angle of Attack [degrees]


 
Figure 37. Lift and drag curves for the XL.1 airfoil section at a Reynolds number of
500,000. Source: Kamisky9.
60

4.5.2 BEM Analysis and WT_Perf, Power vs. Wind Speed

In Figure 38 the results for the turbine mechanical power production versus freestream

wind speed are shown for a range of turbine rotational speeds. Included in the figure is

the power curve provided by the manufacturer.

1500

1400

1300

1200

1100

1000

900
Power [watts]

800 BEM, 100 rpm


BEM, 200 rpm
700 BEM, 300 rpm
600 Bergey XL.1 Data
WT_Perf, 100 rpm
500 WT_Perf, 200 rpm
400 WT_Perf, 300 rpm

300

200

100

0
0 2 4 6 8 10 12 14
Wind Speed [m/s]

Figure 38. Turbine power vs. freestream wind speed using WT_Perf, BEM, and Manufacturer’s

data.

The theoretical power output of the turbine is presented as a compilation of curves

differing by the rotational velocity of the turbine rotor for both performance models used.

In Figure 38, the BEM model results correlate well with those provided by WT_Perf.

These results are useful in the general comparison of the turbine power performance
61

versus wind speed; however, due to the asynchronous nature of the XL.1 turbine (varying

rotational speed) these curves are not as useful as the non-dimensionalized power

coefficient, Cp, and tip-speed ratio, TSR, discussed below.

0.6
BEM, 100 RPM*
BEM, 200 RPM*
BEM, 300 RPM*
0.5
WT_Perf, 100 RPM
WT_Perf, 200 RPM
WT_Perf, 300 RPM
Coefficent of Power (Cp)

0.4
*Corrected with Efficiency Data

0.3

0.2

0.1

0
0 2 4 6 8 10 12 14
Wind Speed [m/s]

Figure 39. Coefficient of power vs. wind speed for both WT_Perf and BEM models.

In Figure 39, the coefficient of power, Cp, is shown versus the freestream wind

speed for both the WT_Perf and BEM models. Using the experimental alternator

efficiencies of Martinez et al, the BEM code predictions were corrected to more

accurately reflect the electromechanical performance of the XL.1. No coefficient of

power versus wind speed data was available from the manufacturer.
62

0.6

0.5

0.4
Coefficient of Power (Cp)

0.3

0.2

0.1 WT_Perf
UCD BEM
UCD BEM w/ Generator Eff.

0
0 1 2 3 4 5 6 7 8 9 10
TSR

Figure 40. Coefficient of power vs. tip speed ratio of the XL.1 rotor from WT_Perf,
UC Davis BEM, and the alternator efficiency corrected UC Davis BEM analysis.

In Figure 40, the coefficient of power, Cp, versus tip speed ratio, TSR, is shown.

The prediction including alternator efficiency is compared to the experimental results.

This prediction agrees with the data provided by the manufacturer, in terms of peak

efficiency condition, which states a maximum Cp at a TSR of 5.5.11 The differences

between the WT_Perf and BEM Cp values are attributed to the various correction models

employed within each model.


63

4.5.3 BEM Analysis and WT_Perf, Thrust vs. Wind Speed

In Figure 41 the results for turbine thrust versus wind speed are shown for several

rotational speeds. Here the turbine rotor thrust is shown as a compilation of constant rpm

curves with the BEM model and WT_Perf curves correlating well. However, the BEM

model predicts values of rotor thrust at wind speed of 1 m/s whereas the manufacturer’s

data, shown in Figure 11, shows a cut-in speed of 2.5 m/s. This is attributed to the lack of

alternator resistive torque within the model because the model is more like a free

spinning rotor unattached to an alternator.

100

90

80
BEM, 100 RPM
70 BEM, 200 RPM

BEM, 300 RPM


60
Thrust [lbs]

WT_Perf, 100 RPM

50 WT_Perf, 200 RPM

WT_Perf, 300 RPM


40

30

20

10

0
0 2 4 6 8 10 12 14
Wind Speed [m/s]

Figure 41. Turbine WT_Perf and BEM thrust model data for maximum rpm.
64

In addition to the average thrust versus wind speed, the coefficient of thrust (Ct) versus

wind speed was used to asses the thrust measurements. The plot of Ct versus wind speed

is show in Figure 42. In this figure the BEM model and WT_Perf data curves correlate

well at higher wind speeds, but differ at low wind speeds. This is most likely attributed to

the several different correction models employed in each model.

1.5

1.4

1.3
BEM, 100 RPM
1.2
BEM, 200 RPM
1.1 BEM, 300 RPM
Coefficient of Thrust (Ct)

1 WT_Perf,100 RPM
WT_Perf, 200 RPM
0.9
WT_Perf, 300 RPM
0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 2 4 6 8 10 12 14
Wind Speed [m/s]

Figure 42. Coefficient of thrust vs. wind speed using BEM model.

Similar to the rationale for power versus wind speed, a more detailed analysis considering

the variable rotational speed of the turbine, using dimensionless parameters, was used.

The resulting Ct vs. TSR plot was made from the performance model data and is shown

in Figure 43.
65

1.2

1.1

0.9
Coefficient of Thrust (Ct)

0.8

0.7

0.6

0.5
UCD BEM Ct
0.4
WT_Perf Ct
0.3

0.2

0.1

0
0 1 2 3 4 5 6 7 8 9 10 11
TSR

Figure 43. WT_Perf and BEM, Ct vs. TSR model data.

As can be seen in both plots the value of the coefficient of thrust increases with increase

in TSR. In theory, Ct can never be greater than one as shown in the BEM code. The

differences between the WT_Perf and BEM Ct values are again attributed to the various

correction models employed within each model.


66

CHAPTER 5- EXPERIMENTAL RESULTS

As mentioned previously, one of the factors in selecting the Bergey XL.1 was that several

ongoing research projects have published data using it as the wind energy conversion

system. Visser et al6 at Clarkson University in Potsdam, New York have installed two

Bergey XL.1 wind turbines to study rotor aerodynamics, including field tests, numerical

studies, and wind tunnel tests of the baseline and modified rotors. Visser et al6 used a

turbine loaded by a variable resistor bank with resistances that are the congruent with our

2 ohm, 8 ohm, and 25 ohm loading levels. In addition to the Clarkson University work,

Martínez et al7 at the University of Mexico performed an analysis of the

electromechanical system including the performance and efficiency of the XL.1

alternator. Lastly, for validation of our small wind system thrust experimentation, data is

used from Jeroen Van Dam, from the National Renewable Energy Laboratory,8 with his

research in the area of small wind thrust experimentation.

In the following chapter the experimental results for the turbine RPM, electrical

performance, and rotor thrust are presented.

5.0 Turbine RPM Measurement

In Figure 44, the electrical power versus rotor rotational velocity (RPM) is shown for

three constant load conditions and the variable PowerCenter load control.
67

2500

2000

1500
Power [watts]

Bergey PowerCenter Load


Control
Bergey Data, 1 Ohm Load Only

1000 UCD 1.56 Ohm

UCD 7.9 Ohm

500

0
0 100 200 300 400 500 600 700 800 900

Rotor RPM

Figure 44. Turbine power output versus RPM for various loading conditions.

As shown in Figure 44, the power output curve increases as the load resistance decreases

for the constant resistance values. This plot supports the methods used for the RPM

measurement by showing the correlation between the smallest load available in our

testing (1.56 ohm) and the Bergey test data of 1 ohm. In addition, the PowerCenter®

controlled turbine data shows the effects on the turbine power output with the modulation

of the turbine’s effective load.


68

600

500

400
Rotor RPM

300

200

UCD 1.56 Ohm


100
UCD 7.9 Ohm

0
0 2 4 6 8 10 12 14
Wind Speed [m/s]

Figure 45. Average turbine rotor RPM vs. wind speed for 1.56 and 7.9 ohm loads.

In Figure 45, the average turbine rotor RPM is shown for two loading conditions. Here it

can be seen that as the turbine resistive load is increased, the maximum rotor speeds, at a

given wind speed, also increase. This is due to a decrease in resistive torque in the

alternator because of the lower currents present in higher resistive load settings. In other

words the higher the resistive load, the lower the alternator winding current, the lower the

resistive torque, and the higher rotor RPM, for a given wind condition.
69

5.1 Turbine Power and Coefficient of Power

Shown in Figure 46 is the raw plot of the 10 sec averaged values- as required in the IEC

61400-12-1 testing standard- of power vs. wind speed, for the minimum resistive load of

1.56 ohms. The scatter plot illustrates the scatter in the power data and that there were

several data points measured, as expected, above and beyond the nominal rating of 1000

watts.

Figure 46. Raw electrical power vs. wind speed data.


70

In Figure 47, the average power for each 0.5 m/s wind speed bin is shown. As

expected the lower resistive load yielded a higher power output from the turbine based on

Equation 5. In addition, the turbine power is shown to decrease at the furling wind speed

of 13 m/s for the 2.1 ohm load, but at 12 m/s for the 7.9 ohm load. This discrepancy is

attributed to the difference in the net torque in the rotor at the higher resistance.

1400

Bergey XL.1 Data


1200
UCD 2.1 Ohm

UCD 7.9 Ohm


1000 UCD 25.1 Ohm
Power [watts]

800

600

400

200

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Wind Speed [m/s]
 
Figure 47. Measured power curves for the XL.1 under high, medium, and low loading
conditions, including manufacturer’s supplied power curve.

Of particular interest in Figure 47 is the increase in power at low wind speeds by the

lower 7.9 ohm setting compared to the 2.1 ohm setting. This correlates with the

measurements of RPM versus wind speed shown in Figure 45, where the turbine is
71

shown to have higher rotor speeds at the low wind speeds. In other words, the turbine cut-

in wind speed decreases with an increase in resistive load.

0.5

0.45

0.4

0.35
Coefficient of Power (Cp)

0.3

0.25

0.2

0.15

0.1

UCD 2.1 Ohm


0.05

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Wind Speed [m/s]

Figure 48. Average coefficient of power (Cp) versus binned wind speed.

In Figure 48, the average values of Cp are shown versus the binned wind speed,

for the power settings of 2.1 ohms. Here the maximum average value of 0.280 falls well

below the maximum value of 0.59 based on the Betz limit.18


72

0.5

0.45

0.4

0.35
Coefficient of Power (Cp)

0.3

0.25

0.2

0.15

0.1
UCD 2.1 Ohm

0.05

0
1 2 3 4 5 6 7 8 9 10

TSR

Figure 49. Experimental coefficient of power vs. tip speed ratio results.

Lastly, in Figure 49, the average values of Cp versus TSR at bins of 0.5 TSR. Due

to the averaging during the binning process the peak values of Cp were found to be

slightly lower than when Cp is presented versus wind speed. Here, the maximum value of

Cp is measured to occur at a TSR of 5.25, which correlates with the information provided

by the manufacturer as the optimum tip-speed-ratio for the turbine.


73

5.2 Rotor Thrust and Coefficient of Thrust

In Figure 50 the raw 10-second averaged thrust values (in pounds) are shown for the

maximum resistive load of 1.56 ohms. For the assessment of rotor thrust, the minimum

resistive load of 1.56 ohms was used because it allows for the maximum resistive torque,

which correlates to maximum thrust production.

Figure 50. Raw thrust vs. wind speed data.

In Figure 51, the average thrust loads are shown for the minimum resistive load of 1.56

ohms. Note that because the thrust measurement is a category A measurement, as defined

in the IEC standard, the precision error also includes a contribution from the standard
74

deviation as defined in Equation 18. As can be seen in Figure 51 the thrust load increases

with wind speed until it reaches the furling condition.

100

90

80

70

60
Thrust [lbs]

50

40

UCD 1.56 Ohm


30

20

10

0
0 2 4 6 8 10 12 14 16
Wind Speed [m/s]

Figure 51. Average thrust, in lbs, vs. wind speed for 1.56 ohm load.
75

0.5

0.4
Coefficient of Thrust (Ct)

0.3

0.2

UCD 1.56 Ohm


0.1

0
0 2 4 6 8 10 12 14
Wind Speed [m/s]

Figure 52. Average coefficient of thrust vs. binned wind speed.

In Figure 52, the average coefficient of thrust, Ct, is shown versus wind speed at 0.5 m/s

bins. At low wind speeds, Ct is shown to increase sharply where then the values remain

slightly above 0.3, as wind speeds increase. Near the rated furling speed Ct values start to

decline as expected.

In Figure 53, the average coefficient of thrust versus TSR is shown. Here again,

the maximum value of Ct is shown to initially increase with the increase in TSR, but then

remain constant at higher values of TSR.


76

0.5

0.4
Coefficient of Thrust (Ct)

0.3

0.2

UCD 1.56 Ohm

0.1

0
2 4 6 8 10
TSR

Figure 53. Average coefficient of thrust vs. binned TSR.


77

CHAPTER 6- EXPERIMENT AND MODEL DISCUSSION

In the following sections the experimental results presented previously in Chapter 5 are

compared with previously published results, and the predicted results presented in

Chapter 4. As with all experiments, similarities and differences are bound to exist

between computational and experimental results, so the following sections attempt to

detail the similarities and explain the differences.

6.0 Turbine Power and Coefficient of Power

In Figure 54 the experimental power curve, for a 2.1 ohm load, is compared with

previously published data from Clarkson University, for a 2.0 ohm load, and the

manufacturer’s provided data for the Bergey XL.1 turbine (Bergey PowerCenter® data).

At low wind speeds the turbine power was expected to follow the similar 2.0 ohm load

values presented by Clarkson University. Instead, our electrical power values were

observed to be higher than Clarkson and similar to that of the PowerCenter® at low wind

speeds. The difference in electrical power output can be most likely attributed to high

turbulence levels at the present site (turbulence intensity of 35% for the current data set).

As explained in Chapter 3, this could create a higher power capture due to energy in the

turbulent flow. Additional influences could be from the anemometer position or

differences in data acquisition and analysis methodologies. Lastly, the peak power values

are shown to be less compared to the manufacturer’s values, which were expected due to

the limited discrete resistance settings in the present experimental setup, as explained in

Chapter 3.
78

1400

1200

1000
Power [watts]

800

600

Bergey XL.1 Data

400 Clarkson 2.0 Ohm

UCD 2.1 Ohm

200

0
0 2 4 6 8 10 12 14 16
Wind Speed [m/s]

Figure 54. Comparison of Bergey XL.1 electric power curves.

In Figure 55, the experimental power coefficient (Cp) curve for a load of 2.1 ohm is

compared with the theoretically predicted curves (BEM model). As before, the BEM

results are obtained at constant rotor RPM (100, 200, and 300) whereas the experimental

results were obtained at varying RPM. The result is that the variable RPM

(experimental) curve traverses the constant RPM (BEM model) curves at the

corresponding RPM and Cp values. Also shown in Figure 55, are the measured Cp values

which are slightly less than those predicted. This can be attributed to both the differences

between the model conditions (mainly the idealized flow assumption of the BEM model)

and the conditions at the test site and any additional electrical inefficiencies not

accounted for in our particular load setting.


79

0.6

0.5

UCD 2.1 Ohm


Coefficient of Power (Cp)

0.4 BEM, 200 RPM

BEM, 300 RPM

BEM, 100 RPM


0.3

0.2

0.1

0
0 2 4 6 8 10 12 14 16
Wind Speed [m/s]

Figure 55. Comparison of measured and predicted electric power coefficient for the Bergey XL.1.
80

0.6

UCD BEM
0.5
Clarkson 2.0 ohm

Clarkson Numerical Results


Coefficient of Power (Cp)

0.4 UCD 2.1 Ohm

0.3

0.2

0.1

0
1 2 3 4 5 6 7 8 9 10

TSR

Figure 56. Average coefficient of power vs. tip speed ratio (TSR) compared with Clarkson and

BEM model data.

In Figure 56 the power coefficient for the turbine is plotted as function of the tip-

speed ratio TSR instead of the wind speed. Here the experimental power coefficient

curve for a load of 2.1 ohm is compared with the theoretically predicted curves (BEM

model) and the Clarkson 2.0 ohm results, as well as their theoretical prediction data. As

can be seen, the measured Cp and Clarkson results differ in magnitudes of Cp, but agree

fairly well in terms of the TSR value for maximum Cp. This value of maximum Cp

occurring at TSR equal to approximately 5.25 also correlates well with data received

from the manufacturer for the Bergey XL.1 at a TSR of 5.5.11 The Cp values shown in

Figure 56 are slightly lower than those in Figure 55 as a result of the binning process,
81

applied to tip-speed ratio, TSR. The Cp values for each bin are averaged and this results

in lower values for Cp versus TSR as compared to Cp versus wind speed. In addition, the

differences between the experimental Cp values and the BEM predictions can also be

attributed to the following factors: (1) the differences in air flow between the model and

the experiment, (2) the resistive torque of the alternator due to the applied loads, and (3)

the assumptions of the model “breaking down.” at high values of TSR.23


82

6.1 Rotor Thrust and Coefficient of Thrust

In Figure 57, the preliminary experimental rotor thrust results, at a load of 1.56 ohms, are

compared with the predicted BEM results. As before, the experimental results are for a

varying rotor RPM; whereas, the BEM results are for constant rotor RPMs. As expected,

the experimental thrust curve traverses the constant RPM thrust curves and it increases in

magnitude as the wind speed increases. Notable differences are that the BEM model

predicts values of thrust that are both greater in magnitude a lower wind speeds and

which occur at lower wind speeds than the measured values of thrust.

100

90

80
BEM, 100 RPM
70
BEM, 200 RPM

BEM, 300 RPM


60
Thrust [lbs]

UCD 1.56 Ohm


50

40

30

20

10

0
0 2 4 6 8 10 12 14 16
Wind Speed [m/s]

Figure 57. Average turbine thrust vs. binned wind speed compared with BEM model data.
83

These discrepancies are attributed to several factors not accounted for in the BEM

model. These are: (1) the resistive torque in the rotor which causes a higher cut-in speed

for the turbine (as explained in Chapter 3, Section 3.1.4), (2) the turbulent airflow of the

rooftop environment, and (3) the magnitude of the thrust correction factor. The extent of

the cause of these discrepancies between the predicted and measured thrust loads is to be

the subject of further study.

In Figure 58, the experimental rotor thrust coefficients are compared with the

predicted BEM results. Here the variable RPM data are shown to traverse the BEM

predicted constant RPM curves, as expected. However, at low wind speeds the values of

Ct are much less. This difference is also evident at higher wind speeds where the plateau

of maximum Ct values is less than those predicted. Subsequently, the differences in Ct are

a direct result of the differences in thrust mentioned previously.


84

1.2

BEM, 100 RPM


1.1
BEM, 200 RPM
1
BEM, 300 RPM
0.9 UCD 1.56 Ohm
Coefficient of Thrust (Ct)

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 2 4 6 8 10 12 14
Wind Speed [m/s]

Figure 58. Average coefficient of thrust (Ct) and BEM model data vs. binned wind speed.

For low wind speed values this difference is attributed to the same cause of the

difference in thrust, namely the model predicting higher values of thrust (also at lower

wind speeds) which causes a bias throughout the data.

In Figure 59, the experimental rotor thrust coefficients at a load of 1.56 ohm are

compared with the predicted BEM results. Here, the experimental and predicted curves

are similar in shape for given TSR values, but differ in magnitude of Ct. This is attributed

to the differences between the measured and predicted values of Ct, as mentioned

previously, which can be attributed to the differences in the measured and predicted

thrust.
85

1.2

1.1

0.9
Coefficient of Thrust (Ct)

0.8

0.7

0.6

0.5

0.4

0.3

0.2
UCD BEM
0.1
UCD 1.56 Ohm
0
0 2 4 6 8 10
TSR

Figure 59. Average coefficient of thrust and BEM data vs. binned TSR.
86

CHAPTER 7- CONCLUSION

7.0 Concluding Remarks

This research into the electrical and mechanical performance analysis of a rooftop

mounted Small Wind Energy System has yielded insight into the areas of power

performance and mechanical load testing. Using specified resistive loads, turbine electric

power, rotor thrust, and rotor rotational speed were measured for a range of wind speeds.

The measured results for electrical power were found to agree well with previously

published results by Clarkson University6 and differed with the manufacturer’s published

results because of the different way the resistive load was applied. The measured results

also differ from predicted results using Blade-Element Momentum (BEM) theory because

of the high turbulence levels at the present rooftop test site. The maximum coefficient of

power measured was 0.28 which nominally agrees with BEM prediction and the Clarkson

observed values. In addition, the value of TSR where Cp was found to be greatest was

approximately 5.25, which correlates to data supplied by the manufacturer.

Maximum rotor thrust values of approximately 65 lbs were found to be lower then to

data observed by NREL on similar machines. Also, coefficient of thrust values were

found to correlate less well with BEM predictions due to differences between the thrust

model and experimental conditions. As mentioned previously, the causes of these

discrepancies will be focus for experimental improvements as well as the subject of

further study.
87

7.1 Experimental Improvements

With experience from this round of testing and computational modeling, several

improvements could be made to further the research on the Small Wind Energy

Demonstration System. Experimental improvements could be made in the areas of: wind

resource assessment, instrumentation, and data acquisition timing.

In the area of wind resource assessment, the measurement of wind speed on the roof

could be improved in several ways. First adding a second, sonic anemometer, opposite of

the cup anemometer on the tower instrument boom would increase wind data resolution

and accuracy while providing redundancy in the measurement of wind speed and wind

direction. The second improvement would be the relocation of the anemometer(s) away

from the turbine and tower (but still on the tower) on a longer instrument boom to

decrease effects from both the tower and the turbine rotor. The third improvement would

consist of the use of a high speed integrated circuit, frequency-to-voltage counter, to

measure the wind speed signal. Currently, by having the data acquisition system compute

the frequency, the recording of the may have lag the wind event due to lag in CPU

prioritization. Determining the rotor frequency found via hardware instead of software

would be faster and more robust.

In the area of instrumentation, the first task would be the development of a total tower

system load calibration. Even though each individual load cell calibration curve was

verified, the total installed system calibration, including the experimental assessment of

the thrust correction factor, was not performed. This was not yet completed because of

the need for more specialized calibration equipment to allow for the accurate load

application to the 30 foot tall tower. Next the measurement of turbine current would be
88

helpful for the more accurate measurement of turbine power. This could be done by

placing a current shunt on the low end (or negative side) of the load bank to measure

system current, which would add redundancy in the measurement of power. Lastly, the

cup anemometer should be recalibrated at regular intervals per the IEC standards to

ensure its accuracy when performing the next round of tests.

In the area of data acquisition timing, the FieldPoint controller could be run

autonomously to allow for consistent data acquisition without the possible interference

from Windows.

7.2 Areas of Further Study

With a strong experimental foundation provided by Kamisky9 and the present effort,

several areas of further study are recommended. These include: (1) the experimental

verification of the thrust correction factor, (2) the ability of the load control using the

Bergey PowerCenter and a battery bank, (3) the interconnection of the turbine to the

electrical subpanel on the roof, and (4) the experimentation with different blade

geometries and materials.

As mentioned previously in the experimental improvements section, further study of

the tower load calibration and tower thrust correction factor is needed. This would allow

for the correction of either the BEM thrust model or the experimental data based upon the

total tower calibration.

The use of the Bergey PowerCenter charge controller would allow for the accurate

prediction of the manufacturer’s power curve and also begin the process of third party

testing of turbines at the Small Wind Testing site. Having an additional relay contactor to

switch between a stock configuration, as in Figure 6, and the modified load bank detailed
89

in this report, could allow for both types of testing at the site. With many small wind

systems connected to the electrical grid, the grid connection of the rooftop system has

several benefits. These include the demonstration of a rooftop grid-tied system, the

measurement of the production of the energy and its use for off-setting building electric

loads, even if small, and the calculation of the reduction of greenhouse gas emissions for

the rooftop mounted turbine system.

Lastly, blade experimentation could also include research into different shapes and

materials of blades that could increase performance, longevity, and safety of these

rooftop systems often deployed in very turbulent environments.


90

REFERENCES

1. Global Wind Energy Council, “Global Wind 2007 Report”


Source: International Energy Agency, Key World Energy Statistics 2007

2. Energy Information Administration Website, US Statistics:


http://www.eia.doe.gov/cneaf/alternate/page/renew_energy_consump/table4.html

3. http://www.20percentwind.org/Final_DOE_Executive_Summary.pdf

4. AWEA Small Wind Turbine Global Market Study 2008.

5. Energy Information Administration Website, California Energy Statistics:


http://www.eia.doe.gov/

6. Humiston, C., Visser, K., Full Scale Aerodynamic Effects of Solidity and Blade
Number on Small Horizontal Axis Wind Turbines, Department of Mechanical and
Aeronautical Engineering, Clarkson University, Potsdam, New York, 2004.

7. Martinez, J., Morales, A., Probst, O., Llamas, A., Rodriguez, C., Analysis and
Simulation of a Wind-Electric Battery Charging System, International Journal of
Energy Research, Vol.30, 2006,pp.633-649.

8. Dr. Jeroen Van Dam, National Renewable Energy Laboratory, Personal


Communication 12/12/07

9. Robert Kamisky, A Comparison of the Measured and Predicted Performance of a


1 Kilowatt Wind Turbine, University of California, Davis, CA, 2006.

10. Hushy, A. and Prascher D., Tower Design Load Verification on a 1kW Wind
Turbine, National Wind Technology Center, Golden, CO, 2005

11. Bergey XL.1 Testing Data, Provided by Tod Hanley, Bergey Windpower

12. International Standard, Power Performance Measurements of Electricity


Producing Wind Turbines, International Electrotechnical Committee, IEC 61400-
12-1, 2005

13. International Standard, Wind Generator Systems, Part 13: Measurement of


Mechanical Loads, International Electrotechnical Committee, IEC 61400-13,
November 1999
91

14. International Council on Building, Uniform Building Code, Volume 2, ICC, 1997
15. Bergey Windpower Electrical Engineer, Igor Arkshtol, Personal Communication
12/7/07

16. Six-pulse rectifier functionality, http://www.smcelectrical.com/rectifiers.htm

17. Master, G., Renewable and Efficient Electric Power Systems, Wiley and Son’s,
2004

18. Dr. Bruce Hartsough, UC Davis, Biological Systems Engineering, Personal


Communication, 10/29/08

19. Madugula, M., Wahba, Y., and Monforton, G., Dynamic Response of Guyed
Masts, College of Engineering, University of Windsor, Ontario, Canada 1998

20. Dr. Bahram Ranvani, UC Davis, Personal Communication, 4/9/08

21. Manwell, J.F., McGowan, J.G., Rogers, A.L., Wind Energy Explained, John
Wiley & Sons Ltd, UK, 2002.

22. Figliola, R., Beasely, D., Theory and Design for Mechanical Measurements, 4th
Ed., Wiley and Son’s, 2006

23. Burton, T., Sharpe, D., Jenkins, D., Bossanyi, E., Wind Energy Handbook, John
Wiley and Son’s, 2001

24. Drela, M., Youngren, H., XFOIL 6.94 User Guide, MIT, 2001.

25. Hansen, M., Aerodynamics of Wind Turbines, James & James Ltd., UK, 2000.

26. Buhl, M., WT_Perf User’s Guide, National Renewable Energy Laboratory,
NREL/EL-500-29382 2004.

27. Hibbeler, R., Mechanics of Materials, Macmillan College Publishing Company,


New York, 1994
92

APPENDIX

Appendix A- Turbine Manufacturer Specifications

Appendix B- Turbine Spindle Load FEA Results

Appendix C- Thrust Correction Factor Details

Appendix D- Tower Load Calculations

Appendix E- WT_Perf Simulation Input Parameters


93

Appendix A- Turbine Manufacturer Specifications

Source: Bergey WindPower


94

Appendix B- Turbine Spindle Load FEA Inputs and Results

Mesh Information
Mesh Type: Solid mesh
Mesher Used: Standard
Automatic Transition: Off
Smooth Surface: On
Jacobian Check: 4 Points
Element Size: 0.27895 in
Tolerance: 0.013948 in
Quality: High
Number of elements: 15824
Number of nodes: 25956

Solver Information
Quality: High
Solver Type: FFE

Name Type Min Location Max Location


(1.80962
in,
28457.8 0.135693 6.36727e+007 (-0.114597 in,
VON: von N/m^2 N/m^2
Plot1 in, 4.135 in,
Mises stress
- 0.867463 in)
0.00315338
in)
95

Appendix C- Thrust Correction Factor Details


The theoretical determination of the tower thrust correction factor for the indeterminate

loading case was calculated by using two separate methods. For the first method the

Moment-Area Method was used, while the second method used Mastan2, version 3.2; a

commercially available structural analysis program. From these two methods the more

conservative thrust correction factor was selected.

C.1 Moment-Area Method

The first determination of the thrust correction factor is based on the procedure for

solving statically indeterminate beams and shafts using the moment-area method. The

method is described as “a method which provides a semi-graphical technique for finding

the slope and displacement at specific points on the elastic curve of a beam or shaft.”27

The application of this method requires computing areas associated with the beam’s

moment diagram. Then using two moment-area theorems, the net deflection is

computed. More information on the moment-area method is available in Mechanics of

Materials, by Hibbeler.27

Several assumptions were made in this application of the moment-area method.

The first assumption is that the product of the modulus of elasticity and moment of inertia

of the beam (or in our case the tower) are constant. In other words, the tower was

assumed to only experience elastic deformation. The second assumption is that the guy

wire support cable is completely rigid.


96

Figure C1. Tower support model with base resistance.

In Figure C1, the tower reactions for the indeterminate loading case are shown. These

reactions and the elastic deformation curve (dashed line) are shown in more detail in

Figure C2 (A). Here the net deflection at point A, tA/O, is the value that is used to perform

the moment-area method about. For this analysis this value is assumed to be zero because

of the rigidity of the guy wire support cable. In other words, the cable is assumed to not

be flexible. Using the method of superposition, the combined M/EI diagram for the

reactions of RA and the load, FThrust, were determined, as shown in Figure C2 (B).
97

Figure C2. (A) Tower free body diagram with assumed tower deflection or elastic curve.

(B) Combined M/EI diagram for tower.

Again, the main assumption is that there is no displacement at point A, i.e. the distance of

the elastic curve (shown as a dotted line in Figure C2 (A)) is equal to zero at point A. The

using the moment-area theorem, and substituting FThrust for P, the relative displacement at

point A is:

⎛ R L ⎞⎛ L ⎞⎛ 2 L ⎞ ⎛ − FThrust S ⎞ ⎛ L ⎞ ⎛ − FThrust (L ) ⎞⎛ L ⎞⎛ 2 L ⎞
t A / O = 0 = ⎜ A ⎟⎜ ⎟⎜ ⎟ + ⎜ ⎟(L )⎜ ⎟ + ⎜ ⎟⎜ ⎟⎜ ⎟
⎝ EI ⎠⎝ 2 ⎠⎝ 3 ⎠ ⎝ EI ⎠ ⎝2⎠ ⎝ 2 EI ⎠⎝ 2 ⎠⎝ 3 ⎠
98

Simplifying terms and solving for FThrust then yields:

⎛ 3S ⎞ ⎛ 3S ⎞
FThrust = R A ⎜ + 1⎟ = RTi ⋅ cos θ i ⋅ ⎜ + 1⎟ Equation 24
⎝ 2L ⎠ ⎝ 2L ⎠

After applying the relationship in Equation 24 to the tower geometry configurations for

the North and South load cells, a maximum thrust correction factor was determined to be

1.24.

C.2 Mastan2, v3.2 Analysis

The second calculation of the thrust correction factor utilized the numerical analysis

software, MASTAN2, a commercially available structural analysis program and which

operates via MATLAB. The program's linear and nonlinear analysis routines are based on

the theoretical and numerical formulations presented in the text: Matrix Structural

Analysis, 2nd Edition, by McGuire, Gallagher, and Ziemian (John Wiley & Sons, Inc.

2000).

In MASTAN2 the geometry, section properties, and material properties of a given

structure are defined and then the program provides options for 1st and 2nd order elastic

and inelastic analyses. Simulation results include structure deflections, section internal

forces, and node reactions. For purposes of this analysis, the 1st order elastic analysis

option was used to obtain results for cable axial loads as a function of a load representing

rotor thrust. An additional advantage to using MASTAN2 is that the guy wire support

cable deflections can be accounted for in the analysis, thus, eliminating one of the

previous assumptions.

Shown in Figure C3 are the geometry and degree of freedom restriction inputs

(restraints) for the tower model as well as a 100 lb load applied (assumed rotor thrust) in
99

the Z-X plane to the top of the tower. As shown in the figure, the base allows for tilting in

the Y-Z plane or east/west directions, but includes a resistive moment in the Y-X plane,

or north/south directions.

B
NOTES:
A = Structure Restraints
B = Applied Thrust Load
C = Guy Wires
D = Support Tower

Figure C3. MASTAN2 tower geometry and degree of freedom restriction inputs.

In Figure C4 the net deflection results are shown for the 1st order elastic analysis using a

100 lb applied load. From this analysis the resulting cable reaction loads were determined

and a relationship between the cable load and the applied force was found. Differences

between the simplified analysis, detailed in Chapter 3, are accounted for in the thrust
100

correction factor, which for these conditions is 1.17. As a result, the more conservative

value of 1.17, from the more comprehensive MASTAN2 analysis, is used for the north

and south thrust correction factor.

Figure C4. MASTAN2 simulation results, including tower deflections.


101

Appendix D- Tower Load Calculations

Source: Matt Seitzler


102

Appendix D- Tower Load Calculations Continued


103

Appendix E- WT_Perf Simulation Input Parameters


----- WT_Perf Input File --------------------------------------------
---------
XL1 input file. Bainer Hall Installation (Dimen, Metric, Space, PROP-
PC).
Compatible with WT_Perf v3.00f
----- Input Configuration -------------------------------------------
---------
False Echo: Echo input parameters
to "<rootname>.ech"?
False DimenInp: Turbine parameters are
dimensional?
True Metric: Turbine parameters are
Metric (MKS vs FPS)?
----- Model Configuration -------------------------------------------
---------
16 NumSect: Number of
circumferential sectors.
5000 MaxIter: Max number of
iterations for induction factor.
1.0e-6 ATol: Error tolerance for
induction iteration.
1.0e-6 SWTol: Error tolerance for
skewed-wake iteration.
----- Algorithm Configuration ---------------------------------------
---------
True TipLoss: Use the Prandtl tip-
loss model?
True HubLoss: Use the Prandtl hub-
loss model?
True Swirl: Include Swirl effects?
True SkewWake: Apply skewed-wake
correction?
True AdvBrake: Use the advanced brake-
state model?
True IndProp: Use PROP-PC instead of
PROPX induction algorithm?
True AIDrag: Use the drag term in
the axial induction calculation?
True TIDrag: Use the drag term in
the tangential induction calculation?
----- Turbine Data --------------------------------------------------
---------
3 NumBlade: Number of blades.
1.25 RotorRad: Rotor radius [length].
0.112 HubRad: Hub radius [length or
div by radius].
0 PreCone: Precone angle, positive
downwind [deg].
0 Tilt: Shaft tilt [deg].
0 Yaw: Yaw error [deg].
40 HubHt: Hub height [length or
div by radius].
104

Appendix E- WT_Perf Simulation Input Parameters Continued

4 NumSeg: Number of blade segments


(entire rotor radius).
RElm Twist Chord AFfile PrntElem
0.256 0 0.0816 1 False
0.5 0 0.0816 1 False
0.7 0 0.0816 1 False
0.9 0 0.0816 1 False
----- Aerodynamic Data ----------------------------------------------
---------
1.225 Rho:
Air density [mass/volume].
1.464e-5 KinVisc:
Kinematic air viscosity
0.000 ShearExp:
Wind shear exponent (1/7 law = 0.143).
False UseCm:
Are Cm data included in the airfoil tables?
1 NumAF:
Number of airfoil files.
"C:\Documents and Settings\Administrator\Desktop\Research\Wind\Turbine
Experimentation\Engineering\Calculations\Aeronautical\WT_perf\WTPerf\xl
1_airfoil_tables.dat"
----- I/O Settings --------------------------------------------------
---------
True TabDel: Make output tab-
delimited (fixed-width otherwise).
False KFact: Output dimensional
parameters in K (e.g., kN instead on N)
True WriteBED: Write out blade element
data to "<rootname>.bed"?
False InputTSR: Input speeds as TSRs?
"mps" SpdUnits: Wind-speed units (mps,
fps, mph).
----- Combined-Case Analysis ----------------------------------------
---------
0 NumCases: Number of cases to run.
Enter zero for parametric analysis.
WS or TSR RotSpd Pitch Remove following block
of lines if NumCases is zero.
----- Parametric Analysis (Ignored if NumCases > 0 ) ----------------
---------
3 ParRow: Row parameter (1-
rpm, 2-pitch, 3-tsr/speed).
2 ParCol: Column parameter (1-
rpm, 2-pitch, 3-tsr/speed).
1 ParTab: Table parameter (1-
rpm, 2-pitch, 3-tsr/speed).
True OutPwr: Request output of rotor
power?
True OutCp: Request output of Cp?
False OutTrq: Request output of shaft
torque?
False OutFlp: Request output of flap
bending moment?
105

True OutThr: Request output of rotor


thrust?
9.5, 9.5, 0 PitSt, PitEnd, PitDel: First, last, delta
blade pitch (deg).
100, 1000, 50 OmgSt, OmgEnd, OmgDel: First, last, delta
rotor speed (rpm).
2.00, 13.00, 0.5 SpdSt, SpdEnd, SpdDel: First, last, delta
speeds.

Você também pode gostar